SPECIAL REPORT OF THE
INTERGOVERNMENTAL PANEL
ON CLIMATE CHANGE
MANAGING THE RISKS OF EXTREME
EVENTS AND DISASTERS TO ADVANCE
CLIMATE CHANGE ADAPTATION
Managing the Risks of Extreme Events
and Disasters to Advance
Climate Change Adaptation
Special Report of the
Intergovernmental Panel on Climate Change
Extreme weather and climate events, interacting with exposed and vulnerable human and natural systems, can lead to disasters. This
Special Report explores the challenge of understanding and managing the risks of climate extremes to advance climate change adaptation.
Weather- and climate-related disasters have social as well as physical dimensions. As a result, changes in the frequency and severity of
the physical events affect disaster risk, but so do the spatially diverse and temporally dynamic patterns of exposure and vulnerability.
Some types of extreme weather and climate events have increased in frequency or magnitude, but populations and assets at risk have
also increased, with consequences for disaster risk. Opportunities for managing risks of weather- and climate-related disasters exist or
can be developed at any scale, local to international. Some strategies for effectively managing risks and adapting to climate change
involve adjustments to current activities. Others require transformation or fundamental change.
The Intergovernmental Panel on Climate Change (IPCC) is the leading international body for the assessment of climate change, including
the physical science of climate; impacts, adaptation, and vulnerability; and mitigation of climate change. The IPCC was established by the
United Nations Environment Programme (UNEP) and the World Meteorological Organization (WMO) to provide the world with a
comprehensive assessment of the current state of knowledge of climate change and its potential environmental and socioeconomic
impacts.
Managing the Risks of Extreme Events
and Disasters to Advance
Climate Change Adaptation
Special Report of the
Intergovernmental Panel on Climate Change
Edited by
Christopher B. Field
Co-Chair Working Group II
Carnegie Institution
for Science
Vicente Barros
Co-Chair Working Group II
CIMA / Universidad de
Buenos Aires
Thomas F. Stocker
Co-Chair Working Group I
University of Bern
Qin Dahe
Co-Chair Working Group I
China Meteorological
Administration
David Jon Dokken
Gian-Kasper Plattner
Kristie L. Ebi
Simon K. Allen
Michael D. Mastrandrea
Melinda Tignor
Katharine J. Mach
Pauline M. Midgley
CAMBRIDGE UNIVERSITY PRESS
Cambridge, New York, Melbourne, Madrid, Cape Town,
Singapore, São Paulo, Delhi, Tokyo, Mexico City
Cambridge University Press
32 Avenue of the Americas, New York, NY 10013-2473, USA
www.cambridge.org
Information on this title: www.cambridge.org/9781107607804
© Intergovernmental Panel on Climate Change 2012
This publication is in copyright. Subject to statutory exception
and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without the written
permission of Cambridge University Press.
First published 2012
Printed in the United States of America
A catalog record for this publication is available from the British Library.
ISBN 978-1-107-02506-6 Hardback
ISBN 978-1-107-60780-4 Paperback
Cambridge University Press has no responsibility for the persistence or accuracy of URLs
for external or third-party Internet Web sites referred to in this publication and does not
guarantee that any content on such Web sites is, or will remain, accurate or appropriate.
This book was printed on acid-free stock that is from SFI (Sustainable Forestry Initiative)
certified mills and distributors. It is FSC chain-of-custody certified.
Use the following reference to cite the entire volume:
IPCC, 2012: Managing the Risks of Extreme Events and Disasters to Advance Climate
Change Adaptation. A Special Report of Working Groups I and II of the
Intergovernmental Panel on Climate Change [Field, C.B., V. Barros, T.F. Stocker,
D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen,
M. Tignor, and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, UK,
and New York, NY, USA, 582 pp.
v
Section I Foreword ............................................................................................................................................................vi
Preface
...............................................................................................................................................................vii
Section II Summary for Policymakers
...............................................................................................................................3
Section III Chapter 1
Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience ....
...
25
Chapter 2 Determinants of Risk: Exposure and Vulnerability ....
....................................................................
65
Chapter 3
Changes in Climate Extremes and their Impacts on the Natural Physical Environment ....
........
109
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems ....
.........................
231
Chapter 5
Managing the Risks from Climate Extremes at the Local Level ....
.............................................
291
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters....
..................
339
Chapter 7
Managing the Risks: International Level and Integration across Scales....
.................................
393
Chapter 8 Toward a Sustainable and Resilient Future....
.............................................................................
437
Chapter 9
Case Studies....
............................................................................................................................
487
Section IV Annex I
Authors and Expert Reviewers....
................................................................................................
545
Annex II
Glossary of Terms ....
....................................................................................................................
555
Annex III
Acronyms.....................................................................................................................................565
A
nnex IV
List of Major IPCC Reports....
......................................................................................................
569
Index
................................................................................................................................................................573
Contents
I Foreword and Preface
viii
This Special Report on Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation
(SREX) has been jointly coordinated by Working Groups I (WGI) and II (WGII) of the Intergovernmental Panel on
Climate Change (IPCC). The report focuses on the relationship between climate change and extreme weather and
climate events, the impacts of such events, and the strategies to manage the associated risks.
The IPCC was jointly established in 1988 by the World Meteorological Organization (WMO) and the United Nations
Environment Programme (UNEP), in particular to assess in a comprehensive, objective, and transparent manner all the
relevant scientific, technical, and socioeconomic information to contribute in understanding the scientific basis of risk
of human-induced climate change, the potential impacts, and the adaptation and mitigation options. Beginning in
1990, the IPCC has produced a series of Assessment Reports, Special Reports, Technical Papers, methodologies, and
other key documents which have since become the standard references for policymakers and scientists.
This Special Report, in particular, contributes to frame the challenge of dealing with extreme weather and climate
events as an issue in decisionmaking under uncertainty, analyzing response in the context of risk management. The
report consists of nine chapters, covering risk management; observed and projected changes in extreme weather and
climate events; exposure and vulnerability to as well as losses resulting from such events; adaptation options from the
local to the international scale; the role of sustainable development in modulating risks; and insights from specific
case studies.
Success in developing this report depended foremost on the knowledge, integrity, enthusiasm, and collaboration of
hundreds of experts worldwide, representing a very wide range of disciplines. We would like to express our gratitude
to all the Coordinating Lead Authors, Lead Authors, Contributing Authors, Review Editors, and Expert and Government
Reviewers who devoted considerable expertise, time, and effort to produce this report. We are extremely grateful for
their commitment to the IPCC process and we would also like to thank the staff of the WGI and WGII Technical
Support Units and the IPCC Secretariat, for their unrestricted commitment to the development of such an ambitious
and highly significant IPCC Special Report.
We are also very grateful to the governments which supported their scientists’ participation in this task, as well as to
all those that contributed to the IPCC Trust Fund, thereby facilitating the essential participation of experts from the
developing world. We would also like to express our appreciation, in particular, to the governments of Australia,
Panama, Switzerland, and Vietnam for hosting the drafting sessions in their respective countries, as well as to the
government of Uganda for hosting in Kampala the First Joint Session of Working Groups I and II which approved the
report. Our thanks are also due to the governments of Switzerland and the United States of America for funding the
Technical Support Units for WGI and WGII, respectively. We also wish to acknowledge the collaboration of the
government of Norway – which also provided critical support for meetings and outreach – and the United Nations
International Strategy for Disaster Reduction (ISDR), in the preparation of the original report proposal.
We would especially wish to thank the IPCC Chairman, Dr. Rajendra Pachauri, for his direction and guidance of the
IPCC process, as well as the Co-Chairs of Working Groups II and I, Professors Vicente Barros, Christopher Field, Qin
Dahe, and Thomas Stocker, for their leadership throughout the development of this Special Report.
Foreword
Foreword
M. Jarraud
Secretary-General
World Meteorological Organization
A. Steiner
Executive Director
United Nations Environment Programme
ix
Preface
This volume, Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation, is a Special
Report of the Intergovernmental Panel on Climate Change (IPCC). The report is a collaborative effort of Working Group I
(WGI) and Working Group II (WGII). The IPCC leadership team for this report also has responsibility for the IPCC Fifth
Assessment Report (AR5), scheduled for completion in 2013 and 2014.
The Special Report brings together scientific communities with expertise in three very different aspects of managing
risks of extreme weather and climate events. For this report, specialists in disaster recovery, disaster risk management,
and disaster risk reduction, a community mostly new to the IPCC, joined forces with experts in the areas of the physical
science basis of climate change (WGI) and climate change impacts, adaptation, and vulnerability (WGII). Over the
course of the two-plus years invested in assessing information and writing the report, scientists from these three
communities forged shared goals and products.
Extreme weather and climate events have figured prominently in past IPCC assessments. Extremes can contribute to
disasters, but disaster risk is influenced by more than just the physical hazards. Disaster risk emerges from the interaction
of weather or climate events, the physical contributors to disaster risk, with exposure and vulnerability, the contributors
to risk from the human side. The combination of severe consequences, rarity, and human as well as physical determinants
makes disasters difficult to study. Only over the last few years has the science of these events, their impacts, and
options for dealing with them become mature enough to support a comprehensive assessment. This report provides a
careful assessment of scientific, technical, and socioeconomic knowledge as of May 2011, the cut-off date for literature
included.
The Special Report introduced some important innovations to the IPCC. One was the integration, in a single Special
Report, of skills and perspectives across the disciplines covered by WGI, WGII, and the disaster risk management com-
munity. A second important innovation was the report’s emphasis on adaptation and disaster risk management. A
third innovation was a plan for an ambitious outreach effort. Underlying these innovations and all aspects of the
report is a strong commitment to assessing science in a way that is relevant to policy but not policy prescriptive.
The Process
The Special Report represents the combined efforts of hundreds of leading experts. The Government of Norway and
the United Nations International Strategy for Disaster Reduction submitted a proposal for the report to the IPCC in
September 2008. This was followed by a scoping meeting to develop a candidate outline in March 2009. Following
approval of the outline in April 2009, governments and observer organizations nominated experts for the author team.
The team approved by the WGI and WGII Bureaux consisted of 87 Coordinating Lead Authors and Lead Authors, plus
19 Review Editors. In addition, 140 Contributing Authors submitted draft text and information to the author teams. The
drafts of the report were circulated twice for formal review, first to experts and second to both experts and governments,
resulting in 18,784 review comments. Author teams responded to every comment and, where scientifically appropriate,
modified drafts in response to comments, with Review Editors monitoring the process. The revised report was presented
for consideration at the First Joint Session of WGI and WGII, from 14 to 17 November 2011. At the joint session,
delegates from over 100 countries evaluated and approved, by consensus, the Summary for Policymakers on a
line-by-line basis and accepted the full report.
Structure of the Special Report
This report contains a Summary for Policymakers (SPM) plus nine chapters. References in the SPM point to the
supporting sections of the technical chapters that provide a traceable account of every major finding. The first two
chapters set the stage for the report. Chapter 1 frames the issue of extreme weather and climate events as a challenge
Preface
x
in understanding and managing risk. It characterizes risk as emerging from the overlap of a triggering physical event
with exposure of people and assets and their vulnerability. Chapter 2 explores the determinants of exposure and
vulnerability in detail, concluding that every disaster has social as well as physical dimensions. Chapter 3, the major
contribution of WGI, is an assessment of the scientific literature on observed and projected changes in extreme weather
and climate events, and their attribution to causes where possible. Chapter 4 assesses observed and projected
impacts, considering patterns by sector as well as region. Chapters 5 through 7 assess experience and theory in
adaptation to extremes and disasters, focusing on issues and opportunities at the local scale (Chapter 5), the national
scale (Chapter 6), and the international scale (Chapter 7). Chapter 8 assesses the interactions among sustainable
development, vulnerability reduction, and disaster risk, considering both opportunities and constraints, as well as the
kinds of transformations relevant to overcoming the constraints. Chapter 9 develops a series of case studies that
illustrate the role of real life complexity but also document examples of important progress in managing risk.
Acknowledgements
We wish to express our sincere appreciation to all the Coordinating Lead Authors, Lead Authors, Contributing Authors,
Review Editors, and Expert and Government Reviewers. Without their expertise, commitment, and integrity, as well as
vast investments of time, a report of this quality could never have been completed. We would also like to thank the
members of the WGI and WGII Bureaux for their assistance, wisdom, and good sense throughout the preparation of
the report.
We would particularly like to thank the remarkable staffs of the Technical Support Units of WGI and WGII for their
professionalism, creativity, and dedication. In WGI, thanks go to Gian-Kasper Plattner, Simon Allen, Pauline Midgley,
Melinda Tignor, Vincent Bex, Judith Boschung, and Alexander Nauels. In WGII, which led the logistics and overall
coordination, thanks go to Dave Dokken, Kristie Ebi, Michael Mastrandrea, Katharine Mach, Sandy MacCracken, Rob
Genova, Yuka Estrada, Eric Kissel, Patricia Mastrandrea, Monalisa Chatterjee, and Kyle Terran. Their tireless and very
capable efforts to coordinate the Special Report ensured a final product of high scientific quality, while maintaining an
atmosphere of collegiality and respect.
We would also like to thank the staff of the IPCC Secretariat: Renate Christ, Gaetano Leone, Mary Jean Burer, Sophie
Schlingemann, Judith Ewa, Jesbin Baidya, Joelle Fernandez, Annie Courtin, Laura Biagioni, and Amy Smith Aasdam.
Thanks are also due to Francis Hayes (WMO), Tim Nuthall (European Climate Foundation), and Nick Nutall (UNEP).
Our sincere thanks go to the hosts and organizers of the scoping meeting, the four lead author meetings, and the
approval session. We gratefully acknowledge the support from the host countries: Norway, Panama, Vietnam,
Switzerland, Australia, and Uganda. It is a pleasure to extend special thanks to the government of Norway, which
provided untiring support throughout the Special Report process.
Preface
Vicente Barros and Christopher B. Field
IPCC WGII Co-Chairs
Qin Dahe and Thomas F. Stocker
IPCC WGI Co-Chairs
II Summary for Policymakers
3
Drafting Authors:
Simon K. Allen (Switzerland), Vicente Barros (Argentina), Ian Burton (Canada),
Diarmid Campbell-Lendrum (UK), Omar-Dario Cardona (Colombia), Susan L. Cutter (USA),
O. Pauline Dube (Botswana), Kristie L. Ebi (USA), Christopher B. Field (USA),
John W. Handmer (Australia), Padma N. Lal (Australia), Allan Lavell (Costa Rica),
Katharine J. Mach (USA), Michael D. Mastrandrea (USA), Gordon A. McBean (Canada),
Reinhard Mechler (Germany), Tom Mitchell (UK), Neville Nicholls (Australia),
Karen L. O’Brien (Norway), Taikan Oki (Japan), Michael Oppenheimer (USA), Mark Pelling
(UK), Gian-Kasper Plattner (Switzerland), Roger S. Pulwarty (USA), Sonia I. Seneviratne
(Switzerland), Thomas F. Stocker (Switzerland), Maarten K. van Aalst (Netherlands),
Carolina S. Vera (Argentina), Thomas J. Wilbanks (USA)
This Summary for Policymakers should be cited as:
IPCC, 2012: Summary for Policymakers. In: Managing the Risks of Extreme Events and Disasters to Advance
Climate Change Adaptation [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea,
K.J. Mach, G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups
I and II of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, and
New York, NY, USA, pp. 3-21.
SPM
Summary
for Policymakers
4
A.
Summary for Policymakers
Context
This Summary for Policymakers presents key findings from the Special Report on Managing the Risks of Extreme
Events and Disasters to Advance Climate Change Adaptation (SREX). The SREX approaches the topic by assessing the
scientific literature on issues that range from the relationship between climate change and extreme weather and
climate events (‘climate extremes’) to the implications of these events for society and sustainable development. The
assessment concerns the interaction of climatic, environmental, and human factors that can lead to impacts and
disasters, options for managing the risks posed by impacts and disasters, and the important role that non-climatic
factors play in determining impacts. Box SPM.1 defines concepts central to the SREX.
The character and severity of impacts from climate extremes depend not only on the extremes themselves but also on
exposure and vulnerability. In this report, adverse impacts are considered disasters when they produce widespread
damage and cause severe alterations in the normal functioning of communities or societies. Climate extremes,
exposure, and vulnerability are influenced by a wide range of factors, including anthropogenic climate change, natural
climate variability, and socioeconomic development (Figure SPM.1). Disaster risk management and adaptation to
climate change focus on reducing exposure and vulnerability and increasing resilience to the potential adverse impacts
of climate extremes, even though risks cannot fully be eliminated (Figure SPM.2). Although mitigation of climate
change is not the focus of this report, adaptation and mitigation can complement each other and together can
significantly reduce the risks of climate change. [SYR AR4, 5.3]
Figure SPM.1 | Illustration of the core concepts of SREX. The report assesses how exposure and vulnerability to weather and climate events determine impacts and the likelihood
of disasters (disaster risk). It evaluates the influence of natural climate variability and anthropogenic climate change on climate extremes and other weather and climate events
that can contribute to disasters, as well as the exposure and vulnerability of human society and natural ecosystems. It also considers the role of development in trends in exposure
and vulnerability, implications for disaster risk, and interactions between disasters and development. The report examines how disaster risk management and adaptation to climate
change can reduce exposure and vulnerability to weather and climate events and thus reduce disaster risk, as well as increase resilience to the risks that cannot be eliminated.
Other important processes are largely outside the scope of this report, including the influence of development on greenhouse gas emissions and anthropogenic climate change,
and the potential for mitigation of anthropogenic climate change. [1.1.2, Figure 1-1]
5
Summary for Policymakers
Box SPM.1 | Definitions Central to SREX
Core concepts defined in the SREX glossary
1
and used throughout the report include:
Climate Change: A change in the state of the climate that can be identified (e.g., by using statistical tests) by changes in the mean
and/or the variability of its properties and that persists for an extended period, typically decades or longer. Climate change may be due
to natural internal processes or external forcings, or to persistent anthropogenic changes in the composition of the atmosphere or in
land use.
2
Climate Extreme (extreme weather or climate event): The occurrence of a value of a weather or climate variable above (or below)
a threshold value near the upper (or lower) ends of the range of observed values of the variable. For simplicity, both extreme weather
events and extreme climate events are referred to collectively as ‘climate extremes.The full definition is provided in Section 3.1.2.
Exposure: The presence of people; livelihoods; environmental services and resources; infrastructure; or economic, social, or cultural
assets in places that could be adversely affected.
Vulnerability: The propensity or predisposition to be adversely affected.
Disaster: Severe alterations in the normal functioning of a community or a society due to hazardous physical events interacting with
vulnerable social conditions, leading to widespread adverse human, material, economic, or environmental effects that require immediate
emergency response to satisfy critical human needs and that may require external support for recovery.
Disaster Risk: The likelihood over a specified time period of severe alterations in the normal functioning of a community or a society
due to hazardous physical events interacting with vulnerable social conditions, leading to widespread adverse human, material,
economic, or environmental effects that require immediate emergency response to satisfy critical human needs and that may require
external support for recovery.
Disaster Risk Management: Processes for designing, implementing, and evaluating strategies, policies, and measures to improve the
understanding of disaster risk, foster disaster risk reduction and transfer, and promote continuous improvement in disaster preparedness,
response, and recovery practices, with the explicit purpose of increasing human security, well-being, quality of life, resilience, and
sustainable development.
Adaptation: In human systems, the process of adjustment to actual or expected climate and its effects, in order to moderate harm or
exploit beneficial opportunities. In natural systems, the process of adjustment to actual climate and its effects; human intervention may
facilitate adjustment to expected climate.
Resilience: The ability of a system and its component parts to anticipate, absorb, accommodate, or recover from the effects of a
hazardous event in a timely and efficient manner, including through ensuring the preservation, restoration, or improvement of its
essential basic structures and functions.
Transformation: The altering of fundamental attributes of a system (including value systems; regulatory, legislative, or bureaucratic
regimes; financial institutions; and technological or biological systems).
____________
1
Reflecting the diversity of the communities involved in this assessment and progress in science, several of the definitions used in this Special Report differ in breadth or
focus from those used in the Fourth Assessment Report and other IPCC reports.
2
This definition differs from that in the United Nations Framework Convention on Climate Change (UNFCCC), where climate change is defined as: “a change of climate
which is attributed directly or indirectly to human activity that alters the composition of the global atmosphere and which is in addition to natural climate variability
observed over comparable time periods.” The UNFCCC thus makes a distinction between climate change attributable to human activities altering the atmospheric
composition, and climate variability attributable to natural causes.
6
Summary for Policymakers
This report integrates perspectives from several historically distinct research communities studying climate science,
climate impacts, adaptation to climate change, and disaster risk management. Each community brings different
viewpoints, vocabularies, approaches, and goals, and all provide important insights into the status of the knowledge
base and its gaps. Many of the key assessment findings come from the interfaces among these communities. These
interfaces are also illustrated in Table SPM.1. To accurately convey the degree of certainty in key findings, the report
relies on the consistent use of calibrated uncertainty language, introduced in Box SPM.2. The basis for substantive
paragraphs in this Summary for Policymakers can be found in the chapter sections specified in square brackets.
Exposure and vulnerability are key determinants of disaster risk and of impacts when risk is realized.
[1.1.2, 1.2.3, 1.3, 2.2.1, 2.3, 2.5] For example, a tropical cyclone can have very different impacts depending on where
and when it makes landfall. [2.5.1, 3.1, 4.4.6] Similarly, a heat wave can have very different impacts on different
populations depending on their vulnerability. [Box 4-4, 9.2.1] Extreme impacts on human, ecological, or physical
systems can result from individual extreme weather or climate events. Extreme impacts can also result from non-
extreme events where exposure and vulnerability are high [2.2.1, 2.3, 2.5] or from a compounding of events or their
impacts. [1.1.2, 1.2.3, 3.1.3] For example, drought, coupled with extreme heat and low humidity, can increase the risk
of wildfire. [Box 4-1, 9.2.2]
Extreme and non-extreme weather or climate events affect vulnerability to future extreme events by modifying
resilience, coping capacity, and adaptive capacity. [2.4.3] In particular, the cumulative effects of disasters at local
Figure SPM.2 | Adaptation and disaster risk management approaches for reducing and managing disaster risk in a changing climate. This report assesses a wide range of
complementary adaptation and disaster risk management approaches that can reduce the risks of climate extremes and disasters and increase resilience to remaining risks as they
change over time. These approaches can be overlapping and can be pursued simultaneously. [6.5, Figure 6-3, 8.6]
7
B.
or sub-national levels can substantially affect
livelihood options and resources and the capacity
of societies and communities to prepare for and
respond to future disasters. [2.2, 2.7]
A changing climate leads to changes in the
frequency, intensity, spatial extent, duration,
and timing of extreme weather and climate
events, and can result in unprecedented
extreme weather and climate events. Changes
in extremes can be linked to changes in the mean,
variance, or shape of probability distributions, or all
of these (Figure SPM.3). Some climate extremes (e.g.,
droughts) may be the result of an accumulation of
weather or climate events that are not extreme
when considered independently. Many extreme
weather and climate events continue to be the
result of natural climate variability. Natural variability
will be an important factor in shaping future
extremes in addition to the effect of anthropogenic
changes in climate. [3.1]
Observations of
Exposure, Vulnerability,
Climate Extremes,
Impacts, and Disaster
Losses
The impacts of climate extremes and the potential
for disasters result from the climate extremes
themselves and from the exposure and vulnerability
of human and natural systems. Observed changes
in climate extremes reflect the influence of
anthropogenic climate change in addition to natural
climate variability, with changes in exposure and
vulnerability influenced by both climatic and non-
climatic factors.
Exposure and Vulnerability
Exposure and vulnerability are dynamic, varying across temporal and spatial scales, and depend on
economic, social, geographic, demographic, cultural, institutional, governance, and environmental factors
(high confidence). [2.2, 2.3, 2.5] Individuals and communities are differentially exposed and vulnerable based on
inequalities expressed through levels of wealth and education, disability, and health status, as well as gender, age,
class, and other social and cultural characteristics. [2.5]
Settlement patterns, urbanization, and changes in socioeconomic conditions have all influenced observed
trends in exposure and vulnerability to climate extremes (high confidence). [4.2, 4.3.5] For example, coastal
Summary for Policymakers
Without climate change
With climate change
extreme cold
extreme hot
cold
h
o
t
Probability of OccurrenceProbability of OccurrenceProbability of Occurrence
less
extreme cold
weather
more
extreme cold
weather
less
cold
weather
near constant
extreme cold
weather
near constant
cold
weather
more
extreme hot
weather
more
extreme hot
weather
more
extreme hot
weather
more
cold
weather
more
h
ot
weather
more
h
ot
weather
more
h
ot
weather
a)
b)
c)
Shifted Mean
Increased Variability
Changed Symmetry
Mean:
without and with weather change
Figure SPM.3 | The effect of changes in temperature distribution on
extremes. Different changes in temperature distributions between present and
future climate and their effects on extreme values of the distributions:
(a) effects of a simple shift of the entire distribution toward a warmer climate;
(b) effects of an increase in temperature variability with no shift in the mean;
(c) effects of an altered shape of the distribution, in this example a change in
asymmetry toward the hotter part of the distribution. [Figure 1-2, 1.2.2]
8
Summary for Policymakers
settlements, including in small islands and megadeltas, and mountain settlements are exposed and vulnerable to
climate extremes in both developed and developing countries, but with differences among regions and countries.
[4.3.5, 4.4.3, 4.4.6, 4.4.9, 4.4.10] Rapid urbanization and the growth of megacities, especially in developing countries,
have led to the emergence of highly vulnerable urban communities, particularly through informal settlements and
inadequate land management (high agreement, robust evidence). [5.5.1] See also Case Studies 9.2.8 and 9.2.9.
Vulnerable populations also include refugees, internally displaced people, and those living in marginal areas. [4.2, 4.3.5]
Climate Extremes and Impacts
There is evidence from observations gathered since 1950 of change in some extremes. Confidence in
observed changes in extremes depends on the quality and quantity of data and the availability of studies
analyzing these data, which vary across regions and for different extremes. Assigning ‘low confidence’ in
observed changes in a specific extreme on regional or global scales neither implies nor excludes the
possibility of changes in this extreme. Extreme events are rare, which means there are few data available to make
assessments regarding changes in their frequency or intensity. The more rare the event the more difficult it is to identify
long-term changes. Global-scale trends in a specific extreme may be either more reliable (e.g., for temperature
extremes) or less reliable (e.g., for droughts) than some regional-scale trends, depending on the geographical uniformity
of the trends in the specific extreme. The following paragraphs provide further details for specific climate extremes
from observations since 1950. [3.1.5, 3.1.6, 3.2.1]
It is very likely that there has been an overall decrease in the number of cold days and nights,
3
and an overall increase
in the number of warm days and nights,
3
at the global scale, that is, for most land areas with sufficient data. It is likely
that these changes have also occurred at the continental scale in North America, Europe, and Australia. There is medium
confidence in a warming trend in daily temperature extremes in much of Asia. Confidence in observed trends in daily
temperature extremes in Africa and South America generally varies from low to medium depending on the region. In
many (but not all) regions over the globe with sufficient data, there is medium confidence that the length or number
of warm spells or heat waves
3
has increased. [3.3.1, Table 3-2]
There have been statistically significant trends in the number of heavy precipitation events in some regions. It is likely
that more of these regions have experienced increases than decreases, although there are strong regional and
subregional variations in these trends. [3.3.2]
There is low confidence in any observed long-term (i.e., 40 years or more) increases in tropical cyclone activity (i.e.,
intensity, frequency, duration), after accounting for past changes in observing capabilities. It is likely that there has been
a poleward shift in the main Northern and Southern Hemisphere extratropical storm tracks. There is low confidence in
observed trends in small spatial-scale phenomena such as tornadoes and hail because of data inhomogeneities and
inadequacies in monitoring systems. [3.3.2, 3.3.3, 3.4.4, 3.4.5]
There is medium confidence that some regions of the world have experienced more intense and longer droughts, in
particular in southern Europe and West Africa, but in some regions droughts have become less frequent, less intense,
or shorter, for example, in central North America and northwestern Australia. [3.5.1]
There is limited to medium evidence available to assess climate-driven observed changes in the magnitude and
frequency of floods at regional scales because the available instrumental records of floods at gauge stations are
limited in space and time, and because of confounding effects of changes in land use and engineering. Furthermore,
there is low agreement in this evidence, and thus overall low confidence at the global scale regarding even the sign of
these changes. [3.5.2]
____________
3
See SREX Glossary for definition of these terms: cold days / cold nights, warm days / warm nights, and warm spell – heat wave.
9
Summary for Policymakers
It is likely that there has been an increase in extreme coastal high water related to increases in mean sea level.
[3.5.3]
There is evidence that some extremes have changed as a result of anthropogenic influences, including
increases in atmospheric concentrations of greenhouse gases. It is likely that anthropogenic influences have led
to warming of extreme daily minimum and maximum temperatures at the global scale. There is medium confidence
that anthropogenic influences have contributed to intensification of extreme precipitation at the global scale. It is
likely that there has been an anthropogenic influence on increasing extreme coastal high water due to an increase in
mean sea level. The uncertainties in the historical tropical cyclone records, the incomplete understanding of the physical
mechanisms linking tropical cyclone metrics to climate change, and the degree of tropical cyclone variability provide
only low confidence for the attribution of any detectable changes in tropical cyclone activity to anthropogenic
influences. Attribution of single extreme events to anthropogenic climate change is challenging. [3.2.2, 3.3.1, 3.3.2,
3.4.4, 3.5.3, Table 3-1]
Disaster Losses
Economic losses from weather- and climate-related disasters have increased, but with large spatial and
interannual variability (high confidence, based on high agreement, medium evidence). Global weather- and
climate-related disaster losses reported over the last few decades reflect mainly monetized direct damages to assets,
and are unequally distributed. Estimates of annual losses have ranged since 1980 from a few US$ billion to above
200 billion (in 2010 dollars), with the highest value for 2005 (the year of Hurricane Katrina). Loss estimates are lower-
bound estimates because many impacts, such as loss of human lives, cultural heritage, and ecosystem services, are
difficult to value and monetize, and thus they are poorly reflected in estimates of losses. Impacts on the informal or
undocumented economy as well as indirect economic effects can be very important in some areas and sectors, but are
generally not counted in reported estimates of losses. [4.5.1, 4.5.3, 4.5.4]
Economic, including insured, disaster losses associated with weather, climate, and geophysical events
4
are
higher in developed countries. Fatality rates and economic losses expressed as a proportion of gross
domestic product (GDP) are higher in developing countries (high confidence). During the period from 1970 to
2008, over 95% of deaths from natural disasters occurred in developing countries. Middle-income countries with rapidly
expanding asset bases have borne the largest burden. During the period from 2001 to 2006, losses amounted to about
1% of GDP for middle-income countries, while this ratio has been about 0.3% of GDP for low-income countries and
less than 0.1% of GDP for high-income countries, based on limited evidence. In small exposed countries, particularly
small island developing states, losses expressed as a percentage of GDP have been particularly high, exceeding 1% in
many cases and 8% in the most extreme cases, averaged over both disaster and non-disaster years for the period from
1970 to 2010. [4.5.2, 4.5.4]
Increasing exposure of people and economic assets has been the major cause of long-term increases in
economic losses from weather- and climate-related disasters (high confidence). Long-term trends in economic
disaster losses adjusted for wealth and population increases have not been attributed to climate change,
but a role for climate change has not been excluded (high agreement, medium evidence). These conclusions
are subject to a number of limitations in studies to date. Vulnerability is a key factor in disaster losses, yet it is not well
accounted for. Other limitations are: (i) data availability, as most data are available for standard economic sectors in
developed countries; and (ii) type of hazards studied, as most studies focus on cyclones, where confidence in observed
trends and attribution of changes to human influence is low. The second conclusion is subject to additional limitations:
(iii) the processes used to adjust loss data over time, and (iv) record length. [4.5.3]
____________
4
Economic losses and fatalities described in this paragraph pertain to all disasters associated with weather, climate, and geophysical events.
10
C.
Summary for Policymakers
Disaster Risk Management and Adaptation to Climate
Change: Past Experience with Climate Extremes
Past experience with climate extremes contributes to understanding of effective disaster risk management and
adaptation approaches to manage risks.
The severity of the impacts of climate extremes depends strongly on the level of the exposure and
vulnerability to these extremes (high confidence). [2.1.1, 2.3, 2.5]
Trends in exposure and vulnerability are major drivers of changes in disaster risk (high confidence). [2.5]
Understanding the multi-faceted nature of both exposure and vulnerability is a prerequisite for determining how
weather and climate events contribute to the occurrence of disasters, and for designing and implementing effective
adaptation and disaster risk management strategies. [2.2, 2.6] Vulnerability reduction is a core common element of
adaptation and disaster risk management. [2.2, 2.3]
Development practice, policy, and outcomes are critical to shaping disaster risk, which may be increased
by shortcomings in development (high confidence). [1.1.2, 1.1.3] High exposure and vulnerability are generally
the outcome of skewed development processes such as those associated with environmental degradation, rapid and
unplanned urbanization in hazardous areas, failures of governance, and the scarcity of livelihood options for the poor.
[2.2.2, 2.5] Increasing global interconnectivity and the mutual interdependence of economic and ecological systems
can have sometimes contrasting effects, reducing or amplifying vulnerability and disaster risk. [7.2.1] Countries more
effectively manage disaster risk if they include considerations of disaster risk in national development and sector plans
and if they adopt climate change adaptation strategies, translating these plans and strategies into actions targeting
vulnerable areas and groups. [6.2, 6.5.2]
Data on disasters and disaster risk reduction are lacking at the local level, which can constrain improvements
in local vulnerability reduction (high agreement, medium evidence). [5.7] There are few examples of national
disaster risk management systems and associated risk management measures explicitly integrating knowledge of and
uncertainties in projected changes in exposure, vulnerability, and climate extremes. [6.6.2, 6.6.4]
Inequalities influence local coping and adaptive capacity, and pose disaster risk management and adaptation
challenges from the local to national levels (high agreement, robust evidence). These inequalities reflect
socioeconomic, demographic, and health-related differences and differences in governance, access to livelihoods,
entitlements, and other factors. [5.5.1, 6.2] Inequalities also exist across countries: developed countries are often better
equipped financially and institutionally to adopt explicit measures to effectively respond and adapt to projected
changes in exposure, vulnerability, and climate extremes than are developing countries. Nonetheless, all countries face
challenges in assessing, understanding, and responding to such projected changes. [6.3.2, 6.6]
Humanitarian relief is often required when disaster risk reduction measures are absent or inadequate
(high agreement, robust evidence). [5.2.1] Smaller or economically less-diversified countries face particular
challenges in providing the public goods associated with disaster risk management, in absorbing the losses caused by
climate extremes and disasters, and in providing relief and reconstruction assistance. [6.4.3]
Post-disaster recovery and reconstruction provide an opportunity for reducing weather- and climate-related
disaster risk and for improving adaptive capacity (high agreement, robust evidence). An emphasis on rapidly
rebuilding houses, reconstructing infrastructure, and rehabilitating livelihoods often leads to recovering in ways that
recreate or even increase existing vulnerabilities, and that preclude longer-term planning and policy changes for
enhancing resilience and sustainable development. [5.2.3] See also assessment in Sections 8.4.1 and 8.5.2.
Risk sharing and transfer mechanisms at local, national, regional, and global scales can increase resilience
to climate extremes (medium confidence). Mechanisms include informal and traditional risk sharing mechanisms,
11
micro-insurance, insurance, reinsurance, and national, regional, and global risk pools. [5.6.3, 6.4.3, 6.5.3, 7.4] These
mechanisms are linked to disaster risk reduction and climate change adaptation by providing means to finance relief,
recovery of livelihoods, and reconstruction; reducing vulnerability; and providing knowledge and incentives for reducing
risk. [5.5.2, 6.2.2] Under certain conditions, however, such mechanisms can provide disincentives for reducing disaster
risk. [5.6.3, 6.5.3, 7.4.4] Uptake of formal risk sharing and transfer mechanisms is unequally distributed across regions
and hazards. [6.5.3] See also Case Study 9.2.13.
Attention to the temporal and spatial dynamics of exposure and vulnerability is particularly important
given that the design and implementation of adaptation and disaster risk management strategies and
policies can reduce risk in the short term, but may increase exposure and vulnerability over the longer
term (high agreement, medium evidence). For instance, dike systems can reduce flood exposure by offering
immediate protection, but also encourage settlement patterns that may increase risk in the long term. [2.4.2, 2.5.4,
2.6.2] See also assessment in Sections 1.4.3, 5.3.2, and 8.3.1.
National systems are at the core of countries’ capacity to meet the challenges of observed and projected
trends in exposure, vulnerability, and weather and climate extremes (high agreement, robust evidence).
Effective national systems comprise multiple actors from national and sub-national governments, the private sector,
research bodies, and civil society including community-based organizations, playing differential but complementary
roles to manage risk, according to their accepted functions and capacities. [6.2]
Closer integration of disaster risk management and climate change adaptation, along with the incorporation
of both into local, sub-national, national, and international development policies and practices, could provide
benefits at all scales (high agreement, medium evidence). [5.4, 5.5, 5.6, 6.3.1, 6.3.2, 6.4.2, 6.6, 7.4] Addressing
social welfare, quality of life, infrastructure, and livelihoods, and incorporating a multi-hazards approach into planning
and action for disasters in the short term, facilitates adaptation to climate extremes in the longer term, as is increasingly
recognized internationally. [5.4, 5.5, 5.6, 7.3] Strategies and policies are more effective when they acknowledge multiple
stressors, different prioritized values, and competing policy goals. [8.2, 8.3, 8.7]
Future Climate Extremes, Impacts, and Disaster Losses
Future changes in exposure, vulnerability, and climate extremes resulting from natural climate variability, anthropogenic
climate change, and socioeconomic development can alter the impacts of climate extremes on natural and human
systems and the potential for disasters.
Climate Extremes and Impacts
Confidence in projecting changes in the direction and magnitude of climate extremes depends on many
factors, including the type of extreme, the region and season, the amount and quality of observational
data, the level of understanding of the underlying processes, and the reliability of their simulation in
models. Projected changes in climate extremes under different emissions scenarios
5
generally do not strongly diverge
in the coming two to three decades, but these signals are relatively small compared to natural climate variability over
this time frame. Even the sign of projected changes in some climate extremes over this time frame is uncertain. For
projected changes by the end of the 21st century, either model uncertainty or uncertainties associated with emissions
scenarios used becomes dominant, depending on the extreme. Low-probability, high-impact changes associated with
Summary for Policymakers
D.
____________
5
Emissions scenarios for radiatively important substances result from pathways of socioeconomic and technological development. This report uses
a subset (B1, A1B, A2) of the 40 scenarios extending to the year 2100 that are described in the IPCC Special Report on Emissions Scenarios
(SRES) and that did not include additional climate initiatives. These scenarios have been widely used in climate change projections and
encompass a substantial range of carbon dioxide equivalent concentrations, but not the entire range of the scenarios included in the SRES.
12
Summary for Policymakers
2
18
24
7
17
3
6
26
22
9
15
5
1
10
23
25
14
4
11
16
13
19
8
21
12
20
Full model range
Central 50%
intermodel range
Median
B1 A1B A2
Scenarios:
204665 208100
1
2
5
10
20
Return period (Years)
Decrease in return period implies more frequent extreme temperature events (see caption)
Legend
204665 208100
1
2
5
10
20
22
Alaska/N.W. Canada - 1
204665 208100
1
2
5
10
20
E. Canada/Greenl./Icel. - 2
204665 208100
1
2
5
10
20
W. North America - 3
204665 208100
1
2
5
10
20
C. North America - 4
204665 208100
1
2
5
10
20
E. North America - 5
204665 208100
1
2
5
10
20
Central America/Mexico - 6
204665 208100
1
2
5
10
20
Amazon - 7
204665 208100
1
2
5
10
20
N.E. Brazil - 8
204665 208100
1
2
5
10
20
W. Coast South America - 9
204665 208100
1
2
5
10
20
S.E. South America - 10
204665 208100
1
2
5
10
20
31
24
23
N. Europe - 11
204665 208100
1
2
5
10
20
C. Europe - 12
204665 208100
1
2
5
10
20
S. Europe/Mediterranean - 13
204665 208100
1
2
5
10
20
Sahara - 14
204665 208100
1
2
5
10
20
W. Africa - 15
204665 208100
1
2
5
10
20
E. Africa - 16
204665 208100
1
2
5
10
20
S. Africa - 17
204665 208100
1
2
5
10
20
N. Asia - 18
204665 208100
1
2
5
10
20
E. Asia - 22
204665 208100
1
2
5
10
20
Tibetan Plateau - 21
204665 208100
1
2
5
10
20
C. Asia - 20
204665 208100
1
2
5
10
20
W. Asia - 19
204665 208100
1
2
5
10
20
S. Asia - 23
204665 208100
1
2
5
10
20
S.E. Asia - 24
204665 208100
1
2
5
10
20
N. Australia - 25
204665 208100
1
2
5
10
20
S. Australia/New Zealand - 26
204665 208100
1
2
5
10
20
Globe (Land only)
Figure SPM.4A | Projected return periods for the maximum daily temperature that was exceeded on average once during a 20-year period in the late 20th century (1981–2000). A decrease in return period implies more
frequent extreme temperature events (i.e., less time between events on average). The box plots show results for regionally averaged projections for two time horizons, 2046 to 2065 and 2081 to 2100, as compared to the late
20th century, and for three different SRES emissions scenarios (B1, A1B, A2) (see legend). Results are based on 12 global climate models (GCMs) contributing to the third phase of the Coupled Model Intercomparison Project
(CMIP3). The level of agreement among the models is indicated by the size of the colored boxes (in which 50% of the model projections are contained), and the length of the whiskers (indicating the maximum and minimum
projections from all models). See legend for defined extent of regions. Values are computed for land points only. The ‘Globe’ inset box displays the values computed using all land grid points. [3.3.1, Figure 3-1, Figure 3-5]
13
the crossing of poorly understood climate thresholds cannot be excluded, given the transient and complex nature of
the climate system. Assigning ‘low confidence’ for projections of a specific extreme neither implies nor excludes the
possibility of changes in this extreme. The following assessments of the likelihood and/or confidence of projections are
generally for the end of the 21st century and relative to the climate at the end of the 20th century. [3.1.5, 3.1.7, 3.2.3,
Box 3-2]
Models project substantial warming in temperature extremes by the end of the 21st century. It is virtually
certain that increases in the frequency and magnitude of warm daily temperature extremes and decreases in cold
extremes will occur in the 21st century at the global scale. It is very likely that the length, frequency, and/or intensity
of warm spells or heat waves will increase over most land areas. Based on the A1B and A2 emissions scenarios, a
1-in-20 year hottest day is likely to become a 1-in-2 year event by the end of the 21st century in most regions, except
in the high latitudes of the Northern Hemisphere, where it is likely to become a 1-in-5 year event (see Figure SPM.4A).
Under the B1 scenario, a 1-in-20 year event would likely become a 1-in-5 year event (and a 1-in-10 year event in
Northern Hemisphere high latitudes). The 1-in-20 year extreme daily maximum temperature (i.e., a value that was
exceeded on average only once during the period 1981–2000) will likely increase by about 1°C to 3°C by the mid-21st
century and by about 2°C to 5°C by the late 21st century, depending on the region and emissions scenario (based on
the B1, A1B, and A2 scenarios). [3.3.1, 3.1.6, Table 3-3, Figure 3-5]
It is likely that the frequency of heavy precipitation or the proportion of total rainfall from heavy falls will
increase in the 21st century over many areas of the globe. This is particularly the case in the high latitudes and
tropical regions, and in winter in the northern mid-latitudes. Heavy rainfalls associated with tropical cyclones are likely
to increase with continued warming. There is medium confidence that, in some regions, increases in heavy precipitation
will occur despite projected decreases in total precipitation in those regions. Based on a range of emissions scenarios
(B1, A1B, A2), a 1-in-20 year annual maximum daily precipitation amount is likely to become a 1-in-5 to 1-in-15 year
event by the end of the 21st century in many regions, and in most regions the higher emissions scenarios (A1B and A2)
lead to a stronger projected decrease in return period. See Figure SPM.4B. [3.3.2, 3.4.4, Table 3-3, Figure 3-7]
Average tropical cyclone maximum wind speed is likely to increase, although increases may not occur in
all ocean basins. It is likely that the global frequency of tropical cyclones will either decrease or remain
essentially unchanged. [3.4.4]
There is medium confidence that there will be a reduction in the number of extratropical cyclones averaged
over each hemisphere. While there is low confidence in the detailed geographical projections of extratropical
cyclone activity, there is medium confidence in a projected poleward shift of extratropical storm tracks. There is low
confidence in projections of small spatial-scale phenomena such as tornadoes and hail because competing physical
processes may affect future trends and because current climate models do not simulate such phenomena. [3.3.2, 3.3.3,
3.4.5]
There is medium confidence that droughts will intensify in the 21st century in some seasons and areas, due
to reduced precipitation and/or increased evapotranspiration. This applies to regions including southern Europe
and the Mediterranean region, central Europe, central North America, Central America and Mexico, northeast Brazil,
and southern Africa. Elsewhere there is overall low confidence because of inconsistent projections of drought changes
(dependent both on model and dryness index). Definitional issues, lack of observational data, and the inability of models
to include all the factors that influence droughts preclude stronger confidence than medium in drought projections.
See Figure SPM.5. [3.5.1, Table 3-3, Box 3-3]
Projected precipitation and temperature changes imply possible changes in floods, although overall there
is low confidence in projections of changes in fluvial floods. Confidence is low due to limited evidence and
because the causes of regional changes are complex, although there are exceptions to this statement. There is medium
confidence (based on physical reasoning) that projected increases in heavy rainfall would contribute to increases in
local flooding in some catchments or regions. [3.5.2]
Summary for Policymakers
14
Summary for Policymakers
2
18
24
7
17
3
6
26
22
9
15
5
1
10
23
25
14
4
11
16
13
19
8
21
12
20
Full model range
Central 50%
intermodel range
Median
B1 A1B A2
Scenarios:
Return period (Years)
204665 208100
3
5
10
20
50
Decrease in return period implies more frequent extreme precipitation events (see caption)
Legend
204665 208100
3
5
10
20
50
Globe (Land only)
204665 208100
3
5
10
20
50
S. Australia/New Zealand - 26
204665 208100
3
5
10
20
50
N. Australia - 25
204665 208100
3
5
10
20
50
2.4
S.E. Asia - 24
204665 208100
3
5
10
20
50
S. Asia - 23
204665 208100
3
5
10
20
50
53
W. Asia - 19
204665 208100
3
5
10
20
50
C. Asia - 20
204665 208100
3
5
10
20
50
Tibetan Plateau - 21
204665 208100
3
5
10
20
50
E. Asia - 22
204665 208100
3
5
10
20
50
N. Asia - 18
204665 208100
3
5
10
20
50
S. Africa - 17
204665 208100
3
5
10
20
50
E. Africa - 16
204665 208100
3
5
10
20
50
W. Africa - 15
204665 208100
3
5
10
20
50
64
56
Sahara - 14
204665 208100
3
5
10
20
50
S. Europe/Mediterranean - 13
204665 208100
3
5
10
20
50
C. Europe - 12
204665 208100
3
5
10
20
50
N. Europe - 11
204665 208100
3
5
10
20
50
S.E. South America - 10
204665 208100
3
5
10
20
50
53
61
W. Coast South America - 9
204665 208100
3
5
10
20
50
57
N.E. Brazil - 8
204665 208100
3
5
10
20
50
Amazon - 7
204665 208100
3
5
10
20
50
Central America/Mexico - 6
204665 208100
3
5
10
20
50
E. North America - 5
204665 208100
3
5
10
20
50
C. North America - 4
204665 208100
3
5
10
20
50
W. North America - 3
204665 208100
3
5
10
20
50
E. Canada/Greenl./Icel. - 2
204665 208100
3
5
10
20
50
2.4
Alaska/N.W. Canada - 1
Figure SPM.4B | Projected return periods for a daily precipitation event that was exceeded in the late 20th century on average once during a 20-year period (1981–2000). A decrease in return period implies more frequent
extreme precipitation events (i.e., less time between events on average). The box plots show results for regionally averaged projections for two time horizons, 2046 to 2065 and 2081 to 2100, as compared to the late 20th
century, and for three different SRES emissions scenarios (B1, A1B, A2) (see legend). Results are based on 14 GCMs contributing to the CMIP3. The level of agreement among the models is indicated by the size of the colored
boxes (in which 50% of the model projections are contained), and the length of the whiskers (indicating the maximum and minimum projections from all models). See legend for defined extent of regions. Values are computed
for land points only. The ‘Globe’ inset box displays the values computed using all land grid points. [3.3.2, Figure 3-1, Figure 3-7]
15
Summary for Policymakers
It is very likely that mean sea level rise will contribute to upward trends in extreme coastal high water
levels in the future. There is high confidence that locations currently experiencing adverse impacts such as coastal
erosion and inundation will continue to do so in the future due to increasing sea levels, all other contributing factors
being equal. The very likely contribution of mean sea level rise to increased extreme coastal high water levels, coupled
with the likely increase in tropical cyclone maximum wind speed, is a specific issue for tropical small island states.
[3.5.3, 3.5.5, Box 3-4]
There is high confidence that changes in heat waves, glacial retreat, and/or permafrost degradation will
affect high mountain phenomena such as slope instabilities, movements of mass, and glacial lake outburst
floods. There is also high confidence that changes in heavy precipitation will affect landslides in some regions. [3.5.6]
There is low confidence in projections of changes in large-scale patterns of natural climate variability.
Confidence is low in projections of changes in monsoons (rainfall, circulation) because there is little consensus in climate
models regarding the sign of future change in the monsoons. Model projections of changes in El Niño–Southern
-0.6 -0.2 0.2 0.60
Standard DeviationStandard Deviation
-0.75 -0.25 0.25 0.7500.4-0.4 -0.50 0.50
2046 - 2065
Change in consecutive dry days (CDD)





 
 

  
  
  


 


  
     
   
    
 





2046 - 2065
Soil moisture anomalies (SMA)





2081 - 2100



 

 


 
 

 
 

 
 
  
   
 
  
   
    
     
  

 
  
   
    
    
   
   
   
 






2081 - 2100






Dryness
+
Dryness
+
Figure SPM.5 | Projected annual changes in dryness assessed from two indices. Left column: Change in annual maximum number of consecutive dry days (CDD: days with
precipitation <1 mm). Right column: Changes in soil moisture (soil moisture anomalies, SMA). Increased dryness is indicated with yellow to red colors; decreased dryness with
green to blue. Projected changes are expressed in units of standard deviation of the interannual variability in the three 20-year periods 1980–1999, 2046–2065, and 2081–2100.
The figures show changes for two time horizons, 2046–2065 and 2081–2100, as compared to late 20th-century values (1980–1999), based on GCM simulations under emissions
scenario SRES A2 relative to corresponding simulations for the late 20th century. Results are based on 17 (CDD) and 15 (SMA) GCMs contributing to the CMIP3. Colored shading
is applied for areas where at least 66% (12 out of 17 for CDD, 10 out of 15 for SMA) of the models agree on the sign of the change; stippling is added for regions where at least
90% (16 out of 17 for CDD, 14 out of 15 for SMA) of all models agree on the sign of the change. Grey shading indicates where there is insufficient model agreement (<66%).
[3.5.1, Figure 3-9]
E.
16
Summary for Policymakers
Oscillation variability and the frequency of El Niño episodes are not consistent, and so there is low confidence in
projections of changes in this phenomenon. [3.4.1, 3.4.2, 3.4.3]
Human Impacts and Disaster Losses
Extreme events will have greater impacts on sectors with closer links to climate, such as water, agriculture
and food security, forestry, health, and tourism. For example, while it is not currently possible to reliably project
specific changes at the catchment scale, there is high confidence that changes in climate have the potential to seriously
affect water management systems. However, climate change is in many instances only one of the drivers of future
changes, and is not necessarily the most important driver at the local scale. Climate-related extremes are also expected
to produce large impacts on infrastructure, although detailed analysis of potential and projected damages are limited
to a few countries, infrastructure types, and sectors. [4.3.2, 4.3.5]
In many regions, the main drivers of future increases in economic losses due to some climate extremes will
be socioeconomic in nature (medium confidence, based on medium agreement, limited evidence). Climate
extremes are only one of the factors that affect risks, but few studies have specifically quantified the effects of
changes in population, exposure of people and assets, and vulnerability as determinants of loss. However, the few
studies available generally underline the important role of projected changes (increases) in population and capital at
risk. [4.5.4]
Increases in exposure will result in higher direct economic losses from tropical cyclones. Losses will also
depend on future changes in tropical cyclone frequency and intensity (high confidence). Overall losses due to
extratropical cyclones will also increase, with possible decreases or no change in some areas (medium confidence).
Although future flood losses in many locations will increase in the absence of additional protection measures (high
agreement, medium evidence), the size of the estimated change is highly variable, depending on location, climate
scenarios used, and methods used to assess impacts on river flow and flood occurrence. [4.5.4]
Disasters associated with climate extremes influence population mobility and relocation, affecting host and
origin communities (medium agreement, medium evidence). If disasters occur more frequently and/or with greater
magnitude, some local areas will become increasingly marginal as places to live or in which to maintain livelihoods. In
such cases, migration and displacement could become permanent and could introduce new pressures in areas of
relocation. For locations such as atolls, in some cases it is possible that many residents will have to relocate. [5.2.2]
Managing Changing Risks
of Climate Extremes and Disasters
Adaptation to climate change and disaster risk management provide a range of complementary approaches for
managing the risks of climate extremes and disasters (Figure SPM.2). Effectively applying and combining approaches
may benefit from considering the broader challenge of sustainable development.
Measures that provide benefits under current climate and a range of future climate change scenarios,
called low-regrets measures, are available starting points for addressing projected trends in exposure,
vulnerability, and climate extremes. They have the potential to offer benefits now and lay the foundation
for addressing projected changes (high agreement, medium evidence). Many of these low-regrets strategies
produce co-benefits, help address other development goals, such as improvements in livelihoods, human well-being,
and biodiversity conservation, and help minimize the scope for maladaptation. [6.3.1, Table 6-1]
Potential low-regrets measures include early warning systems; risk communication between decisionmakers and local
citizens; sustainable land management, including land use planning; and ecosystem management and restoration.
17
Other low-regrets measures include improvements to health surveillance, water supply, sanitation, and irrigation and
drainage systems; climate-proofing of infrastructure; development and enforcement of building codes; and better
education and awareness. [5.3.1, 5.3.3, 6.3.1, 6.5.1, 6.5.2] See also Case Studies 9.2.11 and 9.2.14, and assessment in
Section 7.4.3.
Effective risk management generally involves a portfolio of actions to reduce and transfer risk and to
respond to events and disasters, as opposed to a singular focus on any one action or type of action (high
confidence). [1.1.2, 1.1.4, 1.3.3] Such integrated approaches are more effective when they are informed by and
customized to specific local circumstances (high agreement, robust evidence). [5.1] Successful strategies include a
combination of hard infrastructure-based responses and soft solutions such as individual and institutional capacity
building and ecosystem-based responses. [6.5.2]
Multi-hazard risk management approaches provide opportunities to reduce complex and compound hazards
(high agreement, robust evidence). Considering multiple types of hazards reduces the likelihood that risk reduction
efforts targeting one type of hazard will increase exposure and vulnerability to other hazards, in the present and
future. [8.2.5, 8.5.2, 8.7]
Opportunities exist to create synergies in international finance for disaster risk management and adaptation
to climate change, but these have not yet been fully realized (high confidence). International funding for
disaster risk reduction remains relatively low as compared to the scale of spending on international humanitarian
response. [7.4.2] Technology transfer and cooperation to advance disaster risk reduction and climate change adaptation
are important. Coordination on technology transfer and cooperation between these two fields has been lacking, which
has led to fragmented implementation. [7.4.3]
Stronger efforts at the international level do not necessarily lead to substantive and rapid results at the
local level (high confidence). There is room for improved integration across scales from international to local. [7.6]
Integration of local knowledge with additional scientific and technical knowledge can improve disaster
risk reduction and climate change adaptation (high agreement, robust evidence). Local populations document
their experiences with the changing climate, particularly extreme weather events, in many different ways, and this self-
generated knowledge can uncover existing capacity within the community and important current shortcomings. [5.4.4]
Local participation supports community-based adaptation to benefit management of disaster risk and climate
extremes. However, improvements in the availability of human and financial capital and of disaster risk and climate
information customized for local stakeholders can enhance community-based adaptation (medium agreement, medium
evidence). [5.6]
Appropriate and timely risk communication is critical for effective adaptation and disaster risk management
(high confidence). Explicit characterization of uncertainty and complexity strengthens risk communication. [2.6.3]
Effective risk communication builds on exchanging, sharing, and integrating knowledge about climate-related risks
among all stakeholder groups. Among individual stakeholders and groups, perceptions of risk are driven by psychological
and cultural factors, values, and beliefs. [1.1.4, 1.3.1, 1.4.2] See also assessment in Section 7.4.5.
An iterative process of monitoring, research, evaluation, learning, and innovation can reduce disaster risk
and promote adaptive management in the context of climate extremes (high agreement, robust evidence).
[8.6.3, 8.7] Adaptation efforts benefit from iterative risk management strategies because of the complexity, uncertainties,
and long time frame associated with climate change (high confidence). [1.3.2] Addressing knowledge gaps through
enhanced observation and research can reduce uncertainty and help in designing effective adaptation and risk
management strategies. [3.2, 6.2.5, Table 6-3, 7.5, 8.6.3] See also assessment in Section 6.6.
Table SPM.1 presents examples of how observed and projected trends in exposure, vulnerability, and
climate extremes can inform risk management and adaptation strategies, policies, and measures. The
Summary for Policymakers
18
Summary for Policymakers
Table SPM.1 | Illustrative examples of options for risk management and adaptation in the context of changes in exposure, vulnerability, and climate extremes. In each example, information is characterized at the
scale directly relevant to decisionmaking. Observed and projected changes in climate extremes at global and regional scales illustrate that the direction of, magnitude of, and/or degree of certainty for changes may
differ across scales.
The examples were selected based on availability of evidence in the underlying chapters, including on exposure, vulnerability, climate information, and risk management and adaptation options. They are intended
to reflect relevant risk management themes and scales, rather than to provide comprehensive information by region. The examples are not intended to reflect any regional differences in exposure and vulnerability, or in
experience in risk management.
The confidence in projected changes in climate extremes at local scales is often more limited than the confidence in projected regional and global changes. This limited confidence in changes places a focus on
low-regrets risk management options that aim to reduce exposure and vulnerability and to increase resilience and preparedness for risks that cannot be entirely eliminated. Higher-confidence projected changes in
climate extremes, at a scale relevant to adaptation and risk management decisions, can inform more targeted adjustments in strategies, policies, and measures. [3.1.6, Box 3-2, 6.3.1, 6.5.2]
Observed: Low confidence at global scale
regarding (climate-driven) observed changes in
the magnitude and frequency of floods.
Projected: Low confidence in projections of
changes in floods because of limited evidence
and because the causes of regional changes are
complex. However, medium confidence (based on
physical reasoning) that projected increases in
heavy precipitation will contribute to
rain-generated local flooding in some
catchments or regions.
[Table 3-1, 3.5.2]
Rapid expansion of poor people living
in informal settlements around
Nairobi has led to houses of weak
building materials being constructed
immediately adjacent to rivers and to
blockage of natural drainage areas,
increasing exposure and vulnerability.
[6.4.2, Box 6-2]
Observed: Low confidence regarding
trends in heavy precipitation in East
Africa, because of insufficient evidence.
Projected: Likely increase in heavy
precipitation indicators in East Africa.
[Table 3-2, Table 3-3, 3.3.2]
Limited ability to provide local flash
flood projections.
[3.5.2]
Low-regrets options that reduce exposure and
vulnerability across a range of hazard trends:
Strengthening building design and regulation
Poverty reduction schemes
City-wide drainage and sewerage improvements
The Nairobi Rivers Rehabilitation and Restoration
Programme includes installation of riparian buffers,
canals, and drainage channels and clearance of existing
channels; attention to climate variability and change in
the location and design of wastewater infrastructure; and
environmental monitoring for flood early warning.
[6.3, 6.4.2, Box 6-2, Box 6-6]
Observed: Likely increase in extreme coastal
high water worldwide related to increases in
mean sea level.
Projected: Very likely that mean sea level rise
will contribute to upward trends in extreme
coastal high water levels.
High confidence that locations currently
experiencing coastal erosion and inundation will
continue to do so due to increasing sea level, in
the absence of changes in other contributing
factors.
Likely that the global frequency of tropical
cyclones will either decrease or remain
essentially unchanged.
Likely increase in average tropical cyclone
maximum wind speed, although increases may
not occur in all ocean basins.
[Table 3-1, 3.4.4, 3.5.3, 3.5.5]
Sparse regional and temporal coverage
of terrestrial-based observation
networks and limited in situ ocean
observing network, but with improved
satellite-based observations in recent
decades.
While changes in storminess may
contribute to changes in extreme coastal
high water levels, the limited
geographical coverage of studies to date
and the uncertainties associated with
storminess changes overall mean that a
general assessment of the effects of
storminess changes on storm surge is
not possible at this time.
[Box 3-4, 3.5.3]
Low-regrets options that reduce exposure and
vulnerability across a range of hazard trends:
Maintenance of drainage systems
Well technologies to limit saltwater contamination of
groundwater
Improved early warning systems
Regional risk pooling
Mangrove conservation, restoration, and replanting
Specific adaptation options include, for instance,
rendering national economies more climate-independent
and adaptive management involving iterative learning. In
some cases there may be a need to consider relocation,
for example, for atolls where storm surges may
completely inundate them.
[4.3.5, 4.4.10, 5.2.2, 6.3.2, 6.5.2, 6.6.2, 7.4.4, 9.2.9,
9.2.11, 9.2.13]
Observed: Tides and El Niño–Southern
Oscillation have contributed to the more
frequent occurrence of extreme coastal
high water levels and associated
flooding experienced on some Pacific
Islands in recent years.
Projected: The very likely contribution
of mean sea level rise to increased
extreme coastal high water levels,
coupled with the likely increase in
tropical cyclone maximum wind speed, is
a specific issue for tropical small island
states.
See global changes column for
information on global projections for
tropical cyclones.
[Box 3-4, 3.4.4, 3.5.3]
Small island states in the Pacific,
Indian, and Atlantic Oceans, often
with low elevation, are particularly
vulnerable to rising sea levels and
impacts such as erosion, inundation,
shoreline change, and saltwater
intrusion into coastal aquifers. These
impacts can result in ecosystem
disruption, decreased agricultural
productivity, changes in disease
patterns, economic losses such as in
tourism industries, and population
displacement – all of which reinforce
vulnerability to extreme weather
events.
[3.5.5, Box 3-4, 4.3.5, 4.4.10, 9.2.9]
Flash floods in
informal
settlements in
Nairobi, Kenya
Options for risk management and
adaptation in the example
Exposure and vulnerability
at scale of risk management
in the example
Example
GLOBAL
Observed (since 1950) and projected
(to 2100) global changes
REGIONAL
Observed (since 1950) and projected
(to 2100) changes in the example
SCALE OF RISK MANAGEMENT
Available information for the
example
Information on Climate Extreme Across Spatial Scales
Inundation related
to extreme sea
levels in tropical
small island
developing states
Continued next page
19
Summary for Policymakers
Table SPM.1 (continued)
Observed: Medium confidence that the length
or number of warm spells or heat waves has
increased since the middle of the 20th century, in
many (but not all) regions over the globe.
Very likely increase in number of warm days and
nights at the global scale.
Projected: Very likely increase in length,
frequency, and/or intensity of warm spells or
heat waves over most land areas.
Virtually certain increase in frequency and
magnitude of warm days and nights at the global
scale.
[Table 3-1, 3.3.1]
Observations and projections can
provide information for specific urban
areas in the region, with increased heat
waves expected due to regional trends
and urban heat island effects.
[3.3.1, 4.4.5]
Low-regrets options that reduce exposure and
vulnerability across a range of hazard trends:
Early warning systems that reach particularly
vulnerable groups (e.g., the elderly)
Vulnerability mapping and corresponding measures
Public information on what to do during heat waves,
including behavioral advice
Use of social care networks to reach vulnerable
groups
Specific adjustments in strategies, policies, and measures
informed by trends in heat waves include awareness
raising of heat waves as a public health concern; changes
in urban infrastructure and land use planning, for
example, increasing urban green space; changes in
approaches to cooling for public facilities; and
adjustments in energy generation and transmission
infrastructure.
[Table 6-1, 9.2.1]
Observed: Medium confidence in
increase in heat waves or warm spells in
Europe.
Likely overall increase in warm days and
nights over most of the continent.
Projected: Likely more frequent, longer,
and/or more intense heat waves or
warm spells in Europe.
Very likely increase in warm days and
nights.
[Table 3-2, Table 3-3, 3.3.1]
Factors affecting exposure and
vulnerability include age, pre-existing
health status, level of outdoor
activity, socioeconomic factors
including poverty and social isolation,
access to and use of cooling,
physiological and behavioral
adaptation of the population, and
urban infrastructure.
[2.5.2, 4.3.5, 4.3.6, 4.4.5, 9.2.1]
Observed: Low confidence in any observed
long-term (i.e., 40 years or more) increases in
tropical cyclone activity, after accounting for past
changes in observing capabilities.
Projected: Likely that the global frequency of
tropical cyclones will either decrease or remain
essentially unchanged.
Likely increase in average tropical cyclone
maximum wind speed, although increases may
not occur in all ocean basins.
Heavy rainfalls associated with tropical cyclones
are likely to increase.
Projected sea level rise is expected to further
compound tropical cyclone surge impacts.
[Table 3-1, 3.4.4]
Limited model capability to project
changes relevant to specific settlements
or other locations, due to the inability of
global models to accurately simulate
factors relevant to tropical cyclone
genesis, track, and intensity evolution.
[3.4.4]
Low-regrets options that reduce exposure and
vulnerability across a range of hazard trends:
Adoption and enforcement of improved building
codes
Improved forecasting capacity and implementation of
improved early warning systems (including
evacuation plans and infrastructures)
Regional risk pooling
In the context of high underlying variability and
uncertainty regarding trends, options can include
emphasizing adaptive management involving learning
and flexibility (e.g., Cayman Islands National Hurricane
Committee).
[5.5.3, 6.5.2, 6.6.2, Box 6-7, Table 6-1, 7.4.4, 9.2.5,
9.2.11, 9.2.13]
See global changes column for global
projections.
Exposure and vulnerability are
increasing due to growth in
population and increase in property
values, particularly along the Gulf and
Atlantic coasts of the United States.
Some of this increase has been offset
by improved building codes.
[4.4.6]
Options for risk management and
adaptation in the example
Exposure and vulnerability
at scale of risk management
in the example
Example
GLOBAL
Observed (since 1950) and projected
(to 2100) global changes
REGIONAL
Observed (since 1950) and projected
(to 2100) changes in the example
SCALE OF RISK MANAGEMENT
Available information for the
example
Information on Climate Extreme Across Spatial Scales
Impacts of heat
waves in urban
areas in Europe
Increasing losses
from hurricanes in
the USA and the
Caribbean
Observed: Medium confidence that some
regions of the world have experienced more
intense and longer droughts, but in some regions
droughts have become less frequent, less intense,
or shorter.
Projected: Medium confidence in projected
intensification of drought in some seasons and
areas. Elsewhere there is overall low confidence
because of inconsistent projections.
[Table 3-1, 3.5.1]
Less advanced agricultural practices
render region vulnerable to increasing
variability in seasonal rainfall,
drought, and weather extremes.
Vulnerability is exacerbated by
population growth, degradation of
ecosystems, and overuse of natural
resources, as well as poor standards
for health, education, and
governance.
[2.2.2, 2.3, 2.5, 4.4.2, 9.2.3]
Observed: Medium confidence in an
increase in dryness. Recent years
characterized by greater interannual
variability than previous 40 years, with
the western Sahel remaining dry and the
eastern Sahel returning to wetter
conditions.
Projected: Low confidence due
to inconsistent signal in model
projections.
[Table 3-2, Table 3-3, 3.5.1]
Sub-seasonal, seasonal, and interannual
forecasts with increasing uncertainty
over longer time scales.
Improved monitoring, instrumentation,
and data associated with early warning
systems, but with limited participation
and dissemination to at-risk populations.
[5.3.1, 5.5.3, 7.3.1, 9.2.3, 9.2.11]
Low-regrets options that reduce exposure and
vulnerability across a range of hazard trends:
Traditional rain and groundwater harvesting and
storage systems
Water demand management and improved irrigation
efficiency measures
Conservation agriculture, crop rotation, and livelihood
diversification
Increasing use of drought-resistant crop varieties
Early warning systems integrating seasonal forecasts
with drought projections, with improved
communication involving extension services
Risk pooling at the regional or national level
[2.5.4, 5.3.1, 5.3.3, 6.5, Table 6-3, 9.2.3, 9.2.11]
Droughts in the
context of food
security in West
Africa
20
Summary for Policymakers
importance of these trends for decisionmaking depends on their magnitude and degree of certainty at the temporal
and spatial scale of the risk being managed and on the available capacity to implement risk management options
(see Table SPM.1).
Implications for Sustainable Development
Actions that range from incremental steps to transformational changes are essential for reducing risk from
climate extremes (high agreement, robust evidence). Incremental steps aim to improve efficiency within existing
technological, governance, and value systems, whereas transformation may involve alterations of fundamental attributes
of those systems. Transformations, where they are required, are also facilitated through increased emphasis on adaptive
management and learning. Where vulnerability is high and adaptive capacity low, changes in climate extremes can
make it difficult for systems to adapt sustainably without transformational changes. Vulnerability is often concentrated
in lower-income countries or groups, although higher-income countries or groups can also be vulnerable to climate
extremes. [8.6, 8.6.3, 8.7]
Social, economic, and environmental sustainability can be enhanced by disaster risk management and
adaptation approaches. A prerequisite for sustainability in the context of climate change is addressing the
underlying causes of vulnerability, including the structural inequalities that create and sustain poverty and
constrain access to resources (medium agreement, robust evidence). This involves integrating disaster risk
management and adaptation into all social, economic, and environmental policy domains. [8.6.2, 8.7]
The most effective adaptation and disaster risk reduction actions are those that offer development benefits
in the relatively near term, as well as reductions in vulnerability over the longer term (high agreement,
medium evidence). There are tradeoffs between current decisions and long-term goals linked to diverse values,
interests, and priorities for the future. Short- and long-term perspectives on disaster risk management and adaptation
to climate change thus can be difficult to reconcile. Such reconciliation involves overcoming the disconnect between
local risk management practices and national institutional and legal frameworks, policy, and planning. [8.2.1, 8.3.1,
8.3.2, 8.6.1]
Progress toward resilient and sustainable development in the context of changing climate extremes can
benefit from questioning assumptions and paradigms and stimulating innovation to encourage new
patterns of response (medium agreement, robust evidence). Successfully addressing disaster risk, climate
change, and other stressors often involves embracing broad participation in strategy development, the capacity to
combine multiple perspectives, and contrasting ways of organizing social relations. [8.2.5, 8.6.3, 8.7]
The interactions among climate change mitigation, adaptation, and disaster risk management may have a
major influence on resilient and sustainable pathways (high agreement, limited evidence). Interactions
between the goals of mitigation and adaptation in particular will play out locally, but have global consequences.
[8.2.5, 8.5.2]
There are many approaches and pathways to a sustainable and resilient future. [8.2.3, 8.4.1, 8.6.1, 8.7] However, limits
to resilience are faced when thresholds or tipping points associated with social and/or natural systems are exceeded,
posing severe challenges for adaptation. [8.5.1] Choices and outcomes for adaptive actions to climate events must
reflect divergent capacities and resources and multiple interacting processes. Actions are framed by tradeoffs between
competing prioritized values and objectives, and different visions of development that can change over time. Iterative
approaches allow development pathways to integrate risk management so that diverse policy solutions can be
considered, as risk and its measurement, perception, and understanding evolve over time. [8.2.3, 8.4.1, 8.6.1, 8.7]
21
Summary for Policymakers
Box SPM.2 | Treatment of Uncertainty
Based on the Guidance Note for Lead Authors of the IPCC Fifth Assessment Report on Consistent Treatment of Uncertainties,
6
this
Summary for Policymakers relies on two metrics for communicating the degree of certainty in key findings, which is based on author
teams’ evaluations of underlying scientific understanding:
Confidence in the validity of a finding, based on the type, amount, quality, and consistency of evidence (e.g., mechanistic
understanding, theory, data, models, expert judgment) and the degree of agreement. Confidence is expressed qualitatively.
Quantified measures of uncertainty in a finding expressed probabilistically (based on statistical analysis of observations or model
results, or expert judgment).
This Guidance Note refines the guidance provided to support the IPCC Third and Fourth Assessment Reports. Direct comparisons between
assessment of uncertainties in findings in this report and those in the IPCC Fourth Assessment Report are difficult if not impossible,
because of the application of the revised guidance note on uncertainties, as well as the availability of new information, improved
scientific understanding, continued analyses of data and models, and specific differences in methodologies applied in the assessed
studies. For some extremes, different aspects have been assessed and therefore a direct comparison would be inappropriate.
Each key finding is based on an author team’s evaluation of associated evidence and agreement. The confidence metric provides a
qualitative synthesis of an author team’s judgment about the validity of a finding, as determined through evaluation of evidence and
agreement. If uncertainties can be quantified probabilistically, an author team can characterize a finding using the calibrated likelihood
language or a more precise presentation of probability. Unless otherwise indicated, high or very high confidence is associated with
findings for which an author team has assigned a likelihood term.
The following summary terms are used to describe the available evidence: limited, medium, or robust; and for the degree of
agreement: low, medium, or high. A level of confidence is expressed using five qualifiers: very low, low, medium, high, and very high. The
accompanying figure depicts summary statements for evidence and agreement and their relationship to confidence. There is flexibility in
this relationship; for a given evidence and agreement statement, different confidence levels can be assigned, but increasing levels of
evidence and degrees of agreement are correlated with increasing confidence.
The following terms indicate the assessed likelihood:
Term* Likelihood of the Outcome
Virtually certain 99–100% probability
Very likely 90–100% probability
Likely 66–100% probability
About as likely as not 33–66% probability
Unlikely 0–33% probability
Very unlikely 0–10% probability
Exceptionally unlikely 0–1% probability
* Additional terms that were used in limited circumstances in the Fourth
Assessment Report (extremely likely: 95–100% probability, more likely than
not: >50–100% probability, and extremely unlikely: 0–5% probability) may
also be used when appropriate.
____________
6
Mastrandrea, M.D., C.B. Field, T.F. Stocker, O. Edenhofer, K.L. Ebi, D.J. Frame, H. Held, E. Kriegler, K.J. Mach, P.R. Matschoss, G.-K. Plattner, G.W. Yohe, and F.W. Zwiers,
2010: Guidance Note for Lead Authors of the IPCC Fifth Assessment Report on Consistent Treatment of Uncertainties. Intergovernmental Panel on Climate Change
(IPCC), Geneva, Switzerland, www.ipcc.ch.
High agreement
Limited evidence
High agreement
Medium evidence
High agreement
Robust evidence
Medium agreement
Robust evidence
Medium agreement
Medium evidence
Medium agreement
Limited evidence
Low agreement
Limited evidence
Low agreement
Medium evidence
Low agreement
Robust evidence
Evidence (type, amount, quality, consistency)
Agreement
Confidence
Scale
A depiction of evidence and agreement statements and their relationship to
confidence. Confidence increases toward the top-right corner as suggested by the
increasing strength of shading. Generally, evidence is most robust when there are
multiple, consistent independent lines of high-quality evidence.
22
Summary for Policymakers
III Chapters 1 to 9
25
Coordinating Lead Authors:
Allan Lavell (Costa Rica), Michael Oppenheimer (USA)
Lead Authors:
Cherif Diop (Senegal), Jeremy Hess (USA), Robert Lempert (USA), Jianping Li (China),
Robert Muir-Wood (UK), Soojeong Myeong (Republic of Korea)
Review Editors:
Susanne Moser (USA), Kuniyoshi Takeuchi (Japan)
Contributing Authors:
Omar-Dario Cardona (Colombia), Stephane Hallegatte (France), Maria Lemos (USA),
Christopher Little (USA), Alexander Lotsch (USA), Elke Weber (USA)
This chapter should be cited as:
Lavell, A., M. Oppenheimer, C. Diop, J. Hess, R. Lempert, J. Li, R. Muir-Wood, and S. Myeong, 2012: Climate change: new
dimensions in disaster risk, exposure, vulnerability, and resilience. In: Managing the Risks of Extreme Events and Disasters to
Advance Climate Change Adaptation [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J.
Mach, G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups I and II of the
Intergovernmental Panel on Climate Change (IPCC). Cambridge University Press, Cambridge, UK, and New York, NY, USA,
pp. 25-64.
1
Climate Change:
New Dimensions in
Disaster Risk, Exposure,
Vulnerability, and Resilience
Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
26
Executive Summary ...................................................................................................................................27
1.1. Introduction................................................................................................................................29
1.1.1. Purpose and Scope of the Special Report .........................................................................................................................29
1.1.2. Key Concepts and Definitions ............................................................................................................................................30
1.1.2.1. Definitions Related to General Concepts....................................................
.......................................................................................
.30
1.1.2.2. Concepts and Definitions Relating to Disaster Risk Management and Adaptation to Climate Change .............................................34
1.1.2.3. The Social Construction of Disaster Risk.............................................................................................................................................36
1.1.3. Framing the Relation between Adaptation to Climate Change and Disaster Risk Management...
....
..............................37
1.1.4. Framing the Processes of Disaster Risk Management and Adaptation to Climate Change .............................................38
1.1.4.1. Exceptionality, Routine, and Everyday Life ...................................................
......................................................................................38
1.1.4.2.
Territorial Scale, Disaster Risk, and Adaptation...................................................................................................................................39
1.2. Extreme Events, Extreme Impacts, and Disasters ......................................................................39
1.2.1. Distinguishing Extreme Events, Extreme Impacts, and Disasters ......................................................................................39
1.2.2. Extreme Events Defined in Physical Terms ........................................................................................................................40
1.2.2.1. Definitions of Extremes....................................................
............................................................................................
.......................40
1.2.2.2. Extremes in a Changing Climate ........................................................................................................................................................40
1.2.2.3. The Diversity and Range of Extremes .................................................................................................................................................40
1.2.3. Extreme Impacts...
....
..........................................................................................................................................................41
1.2.3.1. Three Classes of Impacts ................................................................................................................................................
....................41
1.2.3.2. Complex Nature of an Extreme ‘Event’...............................................................................................................................................42
1.2.3.3. Metrics to Quantify Social Impacts and the Management of Extremes..............................................................................................42
1.2.3.4. Traditional Adjustment to Extremes....................................................................................................................................................43
1.3. Disaster Management, Disaster Risk Reduction, and Risk Transfer ..........................................44
1.3.1. Climate Change Will Complicate Management of Some Disaster Risks ...........................................................................46
1.3.1.1. Challenge of Quantitative Estimates of Changing Risks.....................................................................................................................46
1.3.1.2.
Processes that Influence Judgments about Changing Risks ...............................................................................................................46
1.3.2. Adaptation to Climate Change Contributes to Disaster Risk Management...
....
...............................................................47
1.3.3. Disaster Risk Management and Adaptation to Climate Change Share Many Concepts, Goals, and Processes................48
1.4. Coping and Adapting .................................................................................................................50
1.4.1. Definitions, Distinctions, and Relationships.......................................................................................................................51
1.4.1.1. Definitions and Distinctions...............................................................................................................................................
.................51
1.4.1.2. Relationships between Coping, Coping Capacity, Adaptive Capacity, and the Coping Range............................................................51
1.4.2. Learning...
....
.......................................................................................................................................................................53
1.4.3. Learning to Overcome Adaptation Barriers .......................................................................................................................54
1.4.4. ‘No Regrets,’ Robust Adaptation, and Learning.................................................................................................................56
References .................................................................................................................................................56
Chapter 1
Table of Contents
27
Disaster signifies extreme impacts suffered when hazardous physical events interact with vulnerable social
conditions to severely alter the normal functioning of a community or a society (high confidence). Social
vulnerability and exposure are key determinants of disaster risk and help explain why non-extreme physical events
and chronic hazards can also lead to extreme impacts and disasters, while some extreme events do not. Extreme
impacts on human, ecological, or physical systems derive from individual extreme or non-extreme events, or a
compounding of events or their impacts (for example, drought creating the conditions for wildfire, followed by heavy
rain leading to landslides and soil erosion). [1.1.2.1, 1.1.2.3, 1.2.3.1, 1.3]
Management strategies based on the reduction of everyday or chronic risk factors and on the reduction of
risk associated with non-extreme events, as opposed to strategies based solely on the exceptional or
extreme, provide a mechanism that facilitates the reduction of disaster risk and the preparation for and
response to extremes and disasters (high confidence). Effective adaptation to climate change requires an
understanding of the diverse ways in which social processes and development pathways shape disaster risk. Disaster
risk is often causally related to ongoing, chronic, or persistent environmental, economic, or social risk factors. [1.1.2.2,
1.1.3, 1.1.4.1, 1.3.2]
Development practice, policy, and outcomes are critical to shaping disaster risk (high confidence). Disaster
risk may be increased by shortcomings in development. Reductions in the rate of depletion of ecosystem services,
improvements in urban land use and territorial organization processes, the strengthening of rural livelihoods, and
general and specific advances in urban and rural governance advance the composite agenda of poverty reduction,
disaster risk reduction, and adaptation to climate change. [1.1.2.1, 1.1.2.2, 1.1.3, 1.3.2, 1.3.3]
Climate change will pose added challenges for the appropriate allocation of efforts to manage disaster
risk (high confidence). The potential for changes in all characteristics of climate will complicate the evaluation,
communication, and management of the resulting risk. [1.1.3.1, 1.1.3.2, 1.2.2.2, 1.3.1, 1.3.2, 1.4.3]
Risk assessment is one starting point, within the broader risk governance framework, for adaptation to
climate change and disaster risk reduction and transfer (high confidence). The assessment and analysis process
may employ a variety of tools according to management context, access to data and technology, and stakeholders
involved. These tools will vary from formalized probabilistic risk analysis to local level, participatory risk and context
analysis methodologies. [1.3, 1.3.1.2, 1.3.3, Box 1-2]
Risk assessment encounters difficulties in estimating the likelihood and magnitude of extreme events and
their impacts (high confidence). Furthermore, among individual stakeholders and groups, perceptions of risk are
driven by psychological and cultural factors, values, and beliefs. Effective risk communication requires exchanging,
sharing, and integrating knowledge about climate-related risks among all stakeholder groups. [Box 1-1, 1.1.4.1,
1.2.2.1, 1.3.1.1, 1.3.1.2, Box 1-2, Box 1-3, 1.4.2]
Management of the risk associated with climate extremes, extreme impacts, and disasters benefits from
an integrated systems approach, as opposed to separately managing individual types of risk or risk in
particular locations (high confidence). Effective risk management generally involves a portfolio of actions to
reduce and transfer risk and to respond to events and disasters, as opposed to a singular focus on any one action or
type of action. [1.1.2.2, 1.1.4.1, 1.3, 1.3.3, 1.4.2]
Learning is central to adaptation to climate change. Furthermore, the concepts, goals, and processes of
adaptation share much in common with disaster risk management, particularly its disaster risk reduction
component (high confidence). Disaster risk management and adaptation to climate change offer frameworks for, and
examples of, advanced learning processes that may help reduce or avoid barriers that undermine planned adaptation
efforts or lead to implementation of maladaptive measures. Due to the deep uncertainty, dynamic complexity, and
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
Executive Summary
28
long timeframe associated with climate change, robust adaptation efforts would require iterative risk management
strategies. [1.1.3, 1.3.2, 1.4.1.2, 1.4.2, 1.4.5, Box 1-4]
Projected trends and uncertainty in hazards, exposure, and vulnerability associated with climate change
and development make return to the status quo, coping, or static resilience increasingly insufficient goals
for disaster risk management and adaptation (high confidence). Recent approaches to resilience of social-
ecological systems expand beyond these concepts to include the ability to self-organize, learn, and adapt over time.
[1.1.2.1, 1.1.2.2, 1.4.1.2, 1.4.2, 1.4.4]
Given shortcomings of past disaster risk management and the new dimension of climate change, greatly
improved and strengthened disaster risk management and adaptation will be needed, as part of
development processes, in order to reduce future risk (high confidence). Efforts will be more effective when
informed by the experience and success with disaster risk management in different regions during recent decades, and
appropriate approaches for risk identification, reduction, transfer, and disaster management. In the future, the
practices of disaster risk management and adaptation can each greatly benefit from far greater synergy and linkage in
institutional, financial, policy, strategic, and practical terms. [1.1.1, 1.1.2.2, 1.1.3, 1.3.3, 1.4.2]
Community participation in planning, the determined use of local and community knowledge and capacities,
and the decentralization of decisionmaking, supported by and in synergy with national and international
policies and actions, are critical for disaster risk reduction (high confidence). The use of local level risk and
context analysis methodologies, inspired by disaster risk management and now strongly accepted by many civil society
and government agencies in work on adaptation at the local levels, would foster greater integration between, and
greater effectiveness of, both adaptation to climate change and disaster risk management. [1.1.2.2, 1.1.4.2, 1.3.3, 1.4.2]
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
29
1.1. Introduction
1.1.1. Purpose and Scope of the Special Report
Climate change, an alteration in the state of the climate that can be
identified by changes in the mean and/or the variability of its properties,
and that persists for an extended period, typically decades or longer, is
a fundamental reference point for framing the different management
themes and challenges dealt with in this Special Report.
Climate change may be due to natural internal processes or external
forcings, or to persistent anthropogenic changes in the composition of
the atmosphere or in land use (see Chapter 3 for greater detail).
Anthropogenic climate change is projected to continue during this
century and beyond. This conclusion is robust under a wide range of
scenarios for future greenhouse gas emissions, including some that
anticipate a reduction in emissions (IPCC, 2007a).
While specific, local outcomes of climate change are uncertain, recent
assessments project alteration in the frequency, intensity, spatial extent,
or duration of weather and climate extremes, including climate and
hydrometeorological events such as heat waves, heavy precipitation
events, drought, and tropical cyclones (see Chapter 3). Such change, in
a context of increasing vulnerability, will lead to increased stress on
human and natural systems and a propensity for serious adverse effects
in many places around the world (UNISDR, 2009e, 2011). At the same
time, climate change is also expected to bring benefits to certain places
and communities at particular times.
New, improved or strengthened processes for anticipating and dealing
with the adverse effects associated with weather and climate events
will be needed in many areas. This conclusion is supported by the fact
that despite increasing knowledge and understanding of the factors
that lead to adverse effects, and despite important advances over
recent decades in the reduction of loss of life with the occurrence of
hydrometeorological events (mainly attributable to important advances
with early warning systems, e.g., Section 9.2.11), social intervention in
the face of historical climate variability has not kept pace with the rapid
increases in other adverse economic and social effects suffered during
this period (ICSU, 2008) (high confidence). Instead, a rapid growth in
real economic losses and livelihood disruption has occurred in many
parts of the world (UNISDR, 2009e, 2011). In regard to losses associated
with tropical cyclones, recent analysis has shown that, with the exception
of the East Asian and Pacific and South Asian regions, “both exposure
and the estimated risk of economic loss are growing faster than GDP
per capita. Thus the risk of losing wealth in disasters associated with
tropical cyclones is increasing faster than wealth itself is increasing”
(UNISDR, 2011, p. 33).
The Hyogo Framework for Action (UNISDR, 2005), adopted by 168
governments, provides a point of reference for disaster risk management
and its practical implementation (see Glossary and Section 1.1.2.2 for a
definition of this practice). Subsequent United Nations statements
suggest the need for closer integration of disaster risk management and
adaptation with climate change concerns and goals, all in the context of
development and development planning (UNISDR, 2008a, 2009a,b,c).
Such a concern led to the agreement between the IPCC and the United
Nations International Strategy for Disaster Reduction (UNISDR), with
the support of the Norwegian government, to undertake this Special
Report on “Managing the Risks of Extreme Events and Disasters to
Advance Climate Change Adaptation” (IPCC, 2009).
This Special Report responds to that concern by considering climate
change and its effects on extreme (weather and climate) events, disaster,
and disaster risk management; how human responses to extreme
events and disasters (based on historical experience and evolution in
practice) could contribute to adaptation objectives and processes; and
how adaptation to climate change could be more closely integrated
with disaster risk management practice.
The report draws on current scientific knowledge to address three
specific goals:
1) To assess the relevance and utility of the concepts, methods,
strategies, instruments, and experience gained from the management
of climate-associated disaster risk under conditions of historical
climate patterns, in order to advance adaptation to climate change
and the management of extreme events and disasters in the
future.
2) To assess the new perspectives and challenges that climate change
brings to the disaster risk management field.
3) To assess the mutual implications of the evolution of the disaster
risk management and adaptation to climate change fields,
particularly with respect to the desired increases in social resilience
and sustainability that adaptation implies.
The principal audience for this Special Report comprises decisionmakers
and professional and technical personnel from local through to national
governments, international development agencies, nongovernmental
organizations, and civil society organizations. The report also has relevance
for the academic community and interested laypeople.
The first section of this chapter briefly introduces the more important
concepts, definitions, contexts, and management concerns needed to
frame the content of this report. Later sections of the chapter expand on
the subjects of extreme events and extreme impacts; disaster risk
management, reduction, and transfer and their integration with
climate change and adaptation processes; and the notions of coping and
adaptation. The level of detail and discussion presented in this chapter
is commensurate with its status as a ‘scene setting’ initiative. The
following eight chapters provide more detailed and specific analysis.
Chapter 2 assesses the key determinants of risk, namely exposure and
vulnerability in the context of climate-related hazards. A particular focus is
the connection between near-term experience and long-term adaptation.
Key questions addressed include whether reducing vulnerability to
current hazards improves adaptation to longer-term climate change,
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
30
and how near-term risk management decisions and adjustments
constrain future vulnerability and enable adaptation.
Chapter 3 focuses on changes in extremes of atmospheric weather and
climate variables (e.g., temperature and precipitation), large-scale
phenomena that are related to these extremes or are themselves
extremes (e.g., tropical and extratropical cyclones, El Niño, and monsoons),
and collateral effects on the physical environment (e.g., droughts,
floods, coastal impacts, landslides). The chapter builds on and updates
the Fourth Assessment Report, which in some instances, due to new
literature, leads to revisions of that assessment.
Chapter 4 explores how changes in climate, particularly weather and
climate extremes assessed in Chapter 3, translate into extreme impacts
on human and ecological systems. A key issue is the nature of both
observed and expected trends in impacts, the latter resulting from
trends in both physical and social conditions. The chapter assesses
these questions from both a regional and a sectoral perspective, and
examines the direct and indirect economic costs of such changes and
their relation to development.
Chapters 5, 6, and 7 assess approaches to disaster risk management and
adaptation to climate change from the perspectives of local, national,
and international governance institutions, taking into consideration the
roles of government, individuals, nongovernmental organizations, the
private sector, and other civil society institutions and arrangements.
Each chapter reviews the efficacy of current disaster risk reduction,
preparedness, and response and risk transfer strategies and previous
approaches to extremes and disasters in order to extract lessons for the
future. Impacts, adaptation, and the cost of risk management are
assessed through the prism of diverse social aggregations and means
for cooperation, as well as a variety of institutional arrangements.
Chapter 5 focuses on the highly variable local contexts resulting from
differences in place, social groupings, experience, management,
institutions, conditions, and sets of knowledge, highlighting risk
management strategies involving housing, buildings, and land use.
Chapter 6 explores similar issues at the national level, where
mechanisms including national budgets, development goals, planning,
warning systems, and building codes may be employed to manage, for
example, food security and agriculture, water resources, forests,
fisheries, building practice, and public health. Chapter 7 carries this
analysis to the international level, where the emphasis is on institutions,
organizations, knowledge generation and sharing, legal frameworks and
practices, and funding arrangements that characterize international
agencies and collaborative arrangements. This chapter also discusses
integration of responsibilities across all governmental scales, emphasizing
the linkages among disaster risk management, climate change adaptation,
and development.
Chapter 8 assesses how disaster risk reduction strategies, ranging from
incremental to transformational, can advance adaptation to climate
change and promote a more sustainable and resilient future. Key
questions include whether an improved alignment between climate
change responses and sustainable development strategies may be
achieved, and whether short- and long-term perspectives may be
reconciled.
Chapter 9 closes this report by presenting case studies in order to
identify lessons and best practices from past responses to extreme
climate-related events and extreme impacts. Cases illustrate concrete
and diverse examples of disaster types as well as risk management
methodologies and responses discussed in the other chapters, providing
a key reference point for the entire report.
1.1.2. Key Concepts and Definitions
The concepts and definitions presented in this chapter and employed
throughout the Special Report take into account a number of existing
sources (IPCC, 2007c; UNISDR, 2009d; ISO, 2009) but also reflect the fact
that concepts and definitions evolve as knowledge, needs, and contexts
vary. Disaster risk management and adaptation to climate change are
dynamic fields, and have in the past exhibited and will necessarily
continue in the future to exhibit such evolution.
This chapter presents ‘skeleton’ definitions that are generic rather than
specific. In subsequent chapters, the definitions provided here are often
expanded in more detail and variants among these definitions will be
examined and explained where necessary.
A glossary of the fundamental definitions used in this assessment is
provided at the end of this study. Figure 1-1 provides a schematic
of the relationships among many of the key concepts defined here.
1.1.2.1. Definitions Related to General Concepts
In order to delimit the central concerns of this Special Report, a distinction
is made between those concepts and definitions that relate to disaster
risk and adaptation to climate change generally; and, on the other
hand, those that relate in particular to the options and forms of social
intervention relevant to these fields. In Section 1.1.2.1, consideration is
given to general concepts. In Section 1.1.2.2, key concepts relating to
social intervention through ‘Disaster Risk Management’ and ‘Climate
Change Adaptation’ are considered.
Extreme (weather and climate) events and disasters comprise the two
central risk management concerns of this Special Report.
Extreme events comprise a facet of climate variability under stable or
changing climate conditions. They are defined as the occurrence of a
value of a weather or climate variable above (or below) a threshold
value near the upper (or lower) ends (‘tails’) of the range of observed
values of the variable. This definition is further discussed and amplified
in Sections 1.2.2, 3.1.1, and 3.1.2.
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
31
Disasters are defined in this report as severe alterations in the normal
functioning of a community or a society due to hazardous physical events
interacting with vulnerable social conditions, leading to widespread
adverse human, material, economic, or environmental effects that
require immediate emergency response to satisfy critical human needs
and that may require external support for recovery.
The hazardous physical events referred to in the definition of disaster
may be of natural, socio-natural (originating in the human degradation
or transformation of the physical environment), or purely anthropogenic
origins (see Lavell, 1996, 1999; Smith, 1996; Tobin and Montz, 1997;
Wisner et al., 2004). This Special Report emphasizes hydrometeorological
and oceanographic events; a subset of a broader spectrum of physical
events that may acquire the characteristic of a hazard if conditions of
exposure and vulnerability convert them into a threat. These include
earthquakes, volcanoes, and tsunamis, among others. Any one geographic
area may be affected by one, or a combination of, such events at the
same or different times. Both in this report and in the wider literature,
some events (e.g., floods and droughts) are at times referred to as
physical impacts (see Section 3.1.1).
Extreme events are often but not always associated with disaster. This
association will depend on the particular physical, geographic, and social
conditions that prevail (see this section and Chapter 2 for discussion of
the conditioning circumstances associated with so-called ‘exposure’
and ‘vulnerability’) (Ball, 1975; O’Keefe et al., 1976; Timmerman, 1981;
Hewitt, 1983; Maskrey, 1989; Mileti, 1999; Wisner et al., 2004).
Non-extreme physical events also can and do lead to disasters where
physical or societal conditions foster such a result. In fact, a significant
number of disasters registered annually in most disaster databases
are associated with physical events that are not extreme as defined
probabilistically, yet have important social and economic impacts on
local communities and governments, both individually and in aggregate
(UNISDR, 2009e, 2011) (high confidence).
For example, many of the ‘disasters’ registered in the widely consulted
University of Louvaine EM-DAT database (CRED, 2010) are not initiated
by statistically extreme events, but rather exhibit extreme properties
expressed as severe interruptions in the functioning of local social and
economic systems. This lack of connection is even more obvious in the
DesInventar database (Corporación OSSO, 2010), developed first in
Latin America in order to specifically register the occurrence of small-
and medium-scale disasters, and which has registered tens and tens of
thousands of these during the last 30 years in the 29 countries it covers
to date. This database has been used by the UNISDR, the Inter-American
Development Bank, and others to examine disaster occurrence, scale,
and impacts in Latin America and Asia, in particular (Cardona 2005,
2008; IDEA, 2005; UNISDR, 2009e, 2011; ERN-AL, 2011). In any one
place, the range of disaster-inducing events can increase if social
conditions deteriorate (Wisner et al., 2004, 2011).
The occurrence of disaster is always preceded by the existence of
specific physical and social conditions that are generally referred to
as disaster risk (Hewitt, 1983; Lewis, 1999, 2009; Bankoff, 2001;
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
Greenhouse Gas Emissions
Vulnerability
Exposure
Weather and
Climate
Events
DISASTER
RISK
Anthropogenic
Climate Change
Disaster Risk
Management
Climate Change
Adaptation
Disaster
Natural
Variability
DEVELOPMENTCLIMATE
Figure 1-1 | The key concepts and scope of this report. The figure indicates schematically key concepts involved in disaster risk management and climate change adaptation, and
the interaction of these with sustainable development.
32
Wisner et al., 2004, 2011; ICSU, 2008; UNISDR, 2009e, 2011; ICSU-LAC,
2009).
Disaster risk is defined for the purposes of this study as the likelihood
over a specified time period of severe alterations in the normal
functioning of a community or a society due to hazardous physical
events interacting with vulnerable social conditions, leading to
widespread adverse human, material, economic, or environmental
effects that require immediate emergency response to satisfy critical
human needs and that may require external support for recovery.
Disaster risk derives from a combination of physical hazards and the
vulnerabilities of exposed elements and will signify the potential for
severe interruption of the normal functioning of the affected society once
it materializes as disaster. This qualitative statement will be expressed
formally later in this assessment (Section 1.3 and Chapter 2).
The definitions of disaster risk and disaster posited above do not include
the potential or actual impacts of climate and hydrological events on
ecosystems or the physical Earth system per se. In this assessment, such
impacts are considered relevant to disaster if, as is often the case, they
comprise one or more of the following, at times interrelated, situations:
i) they impact livelihoods negatively by seriously affecting ecosystem
services and the natural resource base of communities; ii) they have
consequences for food security; and/or iii) they have impacts on human
health.
Extreme impacts on the physical environment are addressed in Section
3.5 and extreme impacts on ecosystems are considered in detail in
Chapter 4. In excluding such impacts from the definition of ‘disaster’ as
employed here, this chapter is in no way underestimating their broader
significance (e.g., in regard to existence value) or suggesting they
should not be dealt with under the rubric of adaptation concerns and
management needs. Rather, we are establishing their relative position
within the conceptual framework of climate-related, socially-defined
‘disaster’ and ‘disaster risk’ and the management options that are
available for promoting disaster risk reduction and adaptation to climate
change (see Section 1.1.2.2 and the Glossary for definitions of these
terms). Thus this report draws a distinction between ‘social disaster,
where extreme impacts on the physical and ecological systems may or
may not play a part, and so-called ‘environmental disaster,’ where direct
physical impacts of human activity and natural physical processes on
the environment are fundamental causes (with possible direct feedback
impacts on social systems).
Disaster risk cannot exist without the threat of potentially damaging
physical events. However, such events, once they occur, are not in and of
themselves sufficient to explain disaster or its magnitude. In the search
to better understand the concept of disaster risk (thus disaster) it is
important to consider the notions of hazard, vulnerability, and exposure.
When extreme and non-extreme physical events, such as tropical
cyclones, floods, and drought, can affect elements of human systems in
an adverse manner, they assume the characteristic of a hazard. Hazard
is defined here as the potential occurrence of a natural or human-
induced physical event that may cause loss of life, injury, or other
health impacts, as well as damage and loss to property, infrastructure,
livelihoods, service provision, and environmental resources. Physical
events become hazards where social elements (or environmental
resources that support human welfare and security) are exposed to
their potentially adverse impacts and exist under conditions that could
predispose them to such effects. Thus, hazard is used in this study to
denote a threat or potential for adverse effects, not the physical event
itself (Cardona, 1986, 1996, 2011; Smith, 1996; Tobin and Montz, 1997;
Lavell, 2003; Hewitt, 2007; Wisner et al., 2004).
Exposure is employed to refer to the presence (location) of people,
livelihoods, environmental services and resources, infrastructure, or
economic, social, or cultural assets in places that could be adversely
affected by physical events and which, thereby, are subject to potential
future harm, loss, or damage. This definition subsumes physical and
biological systems under the concept of ‘environmental services and
resources,’ accepting that these are fundamental for human welfare and
security (Crichton, 1999; Gasper, 2010).
Exposure may also be dictated by mediating social structures (e.g.,
economic and regulatory) and institutions (Sen, 1983). For example,
food insecurity may result from global market changes driven by
drought or flood impacts on crop production in another location. Other
relevant and important interpretations and uses of exposure are
discussed in Chapter 2.
Under exposed conditions, the levels and types of adverse impacts will
be the result of a physical event (or events) interacting with socially
constructed conditions denoted as vulnerability.
Vulnerability is defined generically in this report as the propensity or
predisposition to be adversely affected. Such predisposition constitutes
an internal characteristic of the affected element. In the field of disaster
risk, this includes the characteristics of a person or group and their
situation that influences their capacity to anticipate, cope with, resist, and
recover from the adverse effects of physical events (Wisner et al., 2004).
Vulnerability is a result of diverse historical, social, economic, political,
cultural, institutional, natural resource, and environmental conditions
and processes.
The concept has been developed as a theme in disaster work since the
1970s (Baird et al., 1975; O’Keefe et al., 1976; Wisner et al., 1977; Lewis,
1979, 1984, 1999, 2009; Timmerman, 1981; Hewitt, 1983, 1997, 2007;
Cutter, 1996; Weichselgartner, 2001; Cannon, 2006; Gaillard, 2010) and
variously modified in different fields and applications in the interim
(Adger, 2006; Eakin and Luers, 2006; Füssel, 2007). Vulnerability has
been evaluated according to a variety of quantitative and qualitative
metrics (Coburn and Spence, 2002; Schneider et al., 2007; Cardona,
2011). A detailed discussion of this notion and the drivers or root
causes of vulnerability are provided in Chapter 2.
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
33
The importance of vulnerability to the disaster risk management
community may be appreciated in the way it has helped to highlight the
role of social factors in the constitution of risk, moving away from purely
physical explanations and attributions of loss and damage (see Hewitt,
1983 for an early critique of what he denominated the ‘physicalist’
interpretation of disaster). Differential levels of vulnerability will lead
to differential levels of damage and loss under similar conditions of
exposure to physical events of a given magnitude (Dow, 1992; Wisner
et al., 2011).
The fundamentally social connotation and ‘predictive’ value of
vulnerability is emphasized in the definition used here. The earlier
IPCC definition of vulnerability refers, however, to “the degree to which
a system is susceptible to and unable to cope with adverse effects of
climate change, including climate variability and extremes. Vulnerability
is a function of the character, magnitude, and rate of climate change and
variation to which a system is exposed, its sensitivity, and its adaptive
capacity” (IPCC, 2007c, p. 883). This definition makes physical causes and
their effects an explicit aspect of vulnerability while the social context
is encompassed by the notions of sensitivity and adaptive capacity
(these notions are defined later). In the definition used in this report, the
social context is emphasized explicitly, and vulnerability is considered
independent of physical events (Hewitt, 1983, 1997, 2007;
Weichselgartner, 2001; Cannon, 2006; O’Brien et al., 2007).
Vulnerability has been contrasted and complimented with the notion of
capacity.
Capacity refers to the combination of all the strengths, attributes, and
resources available to an individual, community, society, or organization
that can be used to achieve established goals. This includes the conditions
and characteristics that permit society at large (institutions, local groups,
individuals, etc.) access to and use of social, economic, psychological,
cultural, and livelihood-related natural resources, as well as access to
the information and the institutions of governance necessary to reduce
vulnerability and deal with the consequences of disaster. This definition
extends the definition of capabilities referred to in Sen’s ‘capabilities
approach to development’ (Sen, 1983).
The lack of capacity may be seen as being one dimension of overall
vulnerability, while it is also seen as a separate notion that, although
contributing to an increase in vulnerability, is not part of vulnerability
per se. The existence of vulnerability does not mean an absolute, but
rather a relative lack of capacity.
Promoted in disaster recovery work by Anderson and Woodrow (1989)
as a means, among other objectives, to shift the analytical balance from
the negative aspects of vulnerability to the positive actions by people,
the notion of capacity is fundamental to imagining and designing a
conceptual shift favoring disaster risk reduction and adaptation to climate
change. Effective capacity building, the notion of stimulating and
providing for growth in capacity, requires a clear image of the future
with clearly established goals.
Adaptive capacity comprises a specific usage of the notion of capacity
and is dealt with in detail in later sections of this chapter and Chapters
2 and 8 in particular.
The existence of vulnerability and capacity and their importance for
understanding the nature and extent of the adverse effects that may
occur with the impact of physical events can be complemented with a
consideration of the characteristics or conditions that help ameliorate or
mitigate negative impacts once disaster materializes. The notions of
resilience and coping are fundamental in this sense.
Coping (elaborated upon in detail in Section 1.4 and Chapter 2) is
defined here generically as the use of available skills, resources, and
opportunities to address, manage, and overcome adverse conditions
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
FAQ 1.1 | Is there a one-to-one relationship between extreme events and disasters?
No. Disaster entails social, economic, or environmental impacts that severely disrupt the normal functioning of affected communities.
Extreme weather and climate events will lead to disaster if: 1) communities are exposed to those events; and 2) exposure to potentially
damaging extreme events is accompanied by a high level of vulnerability (a predisposition for loss and damage). On the other hand,
disasters are also triggered by events that are not extreme in a statistical sense. High exposure and vulnerability levels will transform
even some small-scale events into disasters for some affected communities. Recurrent small- or medium-scale events affecting the same
communities may lead to serious erosion of its development base and livelihood options, thus increasing vulnerability. The timing (when
they occur during the day, month, or year) and sequence (similar events in succession or different events contemporaneously) of such
events is often critical to their human impact. The relative importance of the underlying physical and social determinants of disaster risk
varies with the scale of the event and the levels of exposure and vulnerability. Because the impact of lesser events is exacerbated by
physical, ecological, and social conditions that increase exposure and vulnerability, these events disproportionately affect resource-poor
communities with little access to alternatives for reducing hazard, exposure, and vulnerability. The potential negative consequences of
extreme events can be moderated in important ways (but rarely eliminated completely) by implementing corrective disaster risk
management strategies that are reactive, adaptive, and anticipatory, and by sustainable development.
34
with the aim of achieving basic functioning in the short to medium
terms.
Resilience is defined as the ability of a system and its component parts
to anticipate, absorb, accommodate, or recover from the effects of a
potentially hazardous event in a timely and efficient manner, including
through ensuring the preservation, restoration, or improvement of its
essential basic structures and functions. As Gaillard (2010) points out,
this term has been used in disaster studies since the 1970s (Torry, 1979)
and has its origins in engineering (Gordon, 1978), ecology (Holling,
1973) and child psychology (Werner et al., 1971).
Although now widely employed in the fields of disaster risk management
and adaptation, resilience has been subject to a wide range of
interpretations and levels of acceptance as a concept (Timmerman,
1981; Adger, 2000; Klein et al., 2003; Berkes et al., 2004; Folke, 2006;
Gallopín, 2006; Manyena, 2006; Brand and Jax, 2007; Gaillard 2007;
Bosher, 2008; Cutter et al., 2008; Kelman, 2008; Lewis and Kelman,
2009; Bahadur et al., 2010; Aven, 2011). Thus, for example, the term is
used by some in reference to situations at any point along the risk
‘cycle’ or ‘continuum’, that is, before, during, or after the impact of the
physical event. And, in a different vein, some consider the notions of
‘vulnerability’ and ‘capacity’ as being sufficient for explaining the ranges
of success or failure that are found in different recovery scenarios and
are thus averse to the use of the term at all (Wisner et al., 2004, 2011).
Under this latter formulation, vulnerability both potentiates original loss
and damage and also impedes recovery, while capacity building can
change this adverse balance and contribute to greater sustainability
and reduced disaster risk.
Older conceptions of resilience, as ‘bouncing back,’ and its conceptual
cousin, coping (see Section 1.4), have implicitly emphasized a return to
a previous status quo or some other marginally acceptable level, such
as ‘surviving,’ as opposed to generating a cyclical process that leads
to continually improving conditions, as in ‘bouncing forward’ and/or
eventually ‘thriving’ (Davies, 1993; Manyena, 2006). However, the
dynamic and often uncertain consequences of climate change (as well as
ongoing, now longstanding, development trends such as urbanization)
for hazard and vulnerability profiles underscore the fact that ‘bouncing
back’ is an increasingly insufficient goal for disaster risk management
(Pelling, 2003; Vale and Campanella, 2005; Pendalla et al., 2010)
(high confidence). Recent conceptions of resilience of social-ecological
systems focus more on process than outcomes (e.g., Norris et al., 2008),
including the ability to self-organize, learn, and adapt over time (see
Chapter 8). Some definitions of resilience, such as that used in this
report, now also include the idea of anticipation and ‘improvement’ of
essential basic structures and functions. Section 1.4 examines the
importance of learning that is emphasized within this more forward-
looking application of resilience. Chapter 8 builds on the importance of
learning by drawing also from literature that has explored the scope for
innovation, leadership, and adaptive management. Together these
strategies offer potential pathways for transforming existing development
visions, goals, and practices into more sustainable and resilient futures.
Chapters 2 and 8 address the notion of resilience and its importance in
discussions on sustainability, disaster risk reduction, and adaptation in
greater detail.
1.1.2.2. Concepts and Definitions Relating to Disaster Risk
Management and Adaptation to Climate Change
Disaster risk management is defined in this report as the processes
for designing, implementing, and evaluating strategies, policies, and
measures to improve theunderstanding of disaster risk, foster disaster
risk reduction and transfer, and promote continuous improvement in
disaster preparedness, response, and recovery practices, with the explicit
purpose of increasing human security, well-being, quality of life, and
sustainable development.
Disaster risk management is concerned with both disaster and disaster
risk of differing levels and intensities. In other words, it is not restricted
to a ‘manual’ for the management of the risk or disasters associated with
extreme events, but rather includes the conceptual framework that
describes and anticipates intervention in the overall and diverse patterns,
scales, and levels of interaction of exposure, hazard, and vulnerability
that can lead to disaster. A major recent concern of disaster risk
management has been that disasters are associated more and more with
lesser-scale physical phenomena that are not extreme in a physical sense
(see Section 1.1.1). This is principally attributed to increases in exposure
and associated vulnerability (UNISDR, 2009e, 2011).
Where the term risk management is employed in this chapter and
report, it should be interpreted as being a synonym for disaster risk
management, unless otherwise made explicit.
Disaster Risk Management can be divided to comprise two related but
discrete subareas or components: disaster risk reduction and disaster
management.
Disaster risk reduction denotes both a policy goal or objective, and
the strategic and instrumental measures employed for anticipating
future disaster risk, reducing existing exposure, hazard, or vulnerability,
and improving resilience. This includes lessening the vulnerability of
people, livelihoods, and assets and ensuring the appropriate sustainable
management of land, water, and other components of the environment.
Emphasis is on universal concepts and strategies involved in the
consideration of reducing disaster risks, including actions and activities
enacted pre-impact, and when recovery and reconstruction call for
the anticipation of new disaster risk scenarios or conditions. A strong
relationship between disaster risk and disaster risk reduction, and
development and development planning has been established and
validated, particularly, but not exclusively, in developing country
contexts (UNEP, 1972; Cuny, 1983; Sen, 1983; Hagman, 1984; Wijkman
and Timberlake, 1988; Lavell, 1999, 2003, 2009; Wisner et al., 2004,
2011; UNDP, 2004; van Niekerk, 2007; Dulal et al., 2009; UNISDR,
2009e, 2011) (high confidence).
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
35
Disaster management refers to social processes for designing,
implementing, and evaluating strategies, policies, and measures
thatpromote and improve disaster preparedness, response, and recovery
practices at different organizational and societal levels. Disaster
management processes are enacted once the immediacy of the disaster
event has become evident and resources and capacities are put in place
with which to respond prior to and following impact. These include the
activation of early warning systems, contingency planning, emergency
response (immediate post-impact support to satisfy critical human
needs under conditions of severe stress), and, eventually, recovery
(Alexander, 2000; Wisner et al., 2011). Disaster management is required
due to the existence of ‘residual’ disaster risk that ongoing disaster
risk reduction processes have not mitigated or reduced sufficiently or
eliminated or prevented completely (IDB, 2007).
Growing disaster losses have led to rapidly increasing concerns for post-
impact financing of response and recovery (UNISDR, 2009e, 2011). In this
context, the concept and practice of disaster risk transfer has received
increased interest and achieved greater salience. Risk transfer refers to
the process of formally or informally shifting the financial consequences
of particular risks from one party to another, whereby a household,
community, enterprise, or state authority will obtain resources from
the other party after a disaster occurs, in exchange for ongoing or
compensatory social or financial benefits provided to that other party.
Disaster risk transfer mechanisms comprise a component of both disaster
management and disaster risk reduction. In the former case, financial
provision is made to face up to the impacts and consequences of disaster
once this materializes. In the latter case, the adequate use of insurance
premiums, for example, can promote and encourage the use of disaster
risk reduction measures in the insured elements. Chapters 5, 6, 7, and 9
discuss risk transfer in some detail.
Over the last two decades, the more integral notion of disaster risk
management and its risk reduction and disaster management components
has tended to replace the unique conception and terminology of ‘disaster
and emergency management’ that prevailed almost unilaterally up to
the beginning of the 1990s and that emphasized disaster as opposed to
disaster risk as the central issue to be confronted. Disaster as such
ordered the thinking on required intervention processes, whereas with
disaster risk management, disaster risk now tends to assume an
increasingly dominant position in thought and action in this field (see
Hewitt, 1983; Blaikie et al., 1994; Smith, 1996; Hewitt, 1997; Tobin and
Montz, 1997; Lavell, 2003; Wisner et al., 2004, 2011; van Niekerk, 2007;
Gaillard, 2010 for background and review of some of the historical
changes in favor of disaster risk management).
The notion of disaster or disaster management cycle was introduced
and popularized in the earlier context dominated by disaster or emergency
management concerns and viewpoints. The cycle, and the later ‘disaster
continuum’ notion, depicted the sequences and components of so-called
disaster management. In addition to considering preparedness, emergency
response, rehabilitation, and reconstruction, it also included disaster
prevention and mitigation as stated components of ‘disaster management’
and utilized the temporal notions of before, during, and after disaster to
classify the different types of action (Lavell and Franco, 1996; van
Niekerk, 2007).
The cycle notion, criticized for its mechanistic depiction of the intervention
process, for insufficient consideration of the ways different components
and actions merge and can act synergistically with and influence each
other, and for its incorporation of disaster risk reduction considerations
under the rubric of ‘disaster management’ (Lavell and Franco, 1996;
Lewis, 1999; Wisner et al., 2004; Balamir, 2005; van Niekerk, 2007), has
tended to give way over time, in many parts of the world, to the more
comprehensive approach and concept of disaster risk management with
its consideration of distinct risk reduction and disaster intervention
components. The move toward a conception oriented in terms of disaster
risk and not disaster per se has led to initiatives to develop the notion
of a disaster risk continuum’ whereby risk is seen to evolve and
change constantly, requiring different modalities of intervention over
time, from pre-impact risk reduction through response to new risk
conditions following disaster impacts and the need for control of new
risk factors in reconstruction (see Lavell, 2003).
With regard to the influence of actions taken at one stage of the ‘cycle’
on other stages, much has been written, for example, on how the form
and method of response to disaster itself may affect future disaster risk
reduction efforts. The fostering of active community involvement, the
use of existing local and community capacities and resources, and
the decentralization of decisionmaking to the local level in disaster
preparedness and response, among other factors, have been considered
critical for also improving understanding of disaster risk and the
development of future disaster risk reduction efforts (Anderson and
Woodrow, 1989; Alexander, 2000; Lavell, 2003; Wisner et al., 2004)
(high confidence). And, the methods used for, and achievements with,
reconstruction clearly have important impacts on future disaster risk
and on the future needs for preparedness and response.
In the following subsection, some of the major reasons that explain
the transition from disaster management, with its emphasis on disaster,
to disaster risk management, with its emphasis on disaster risk, are
presented as a background for an introduction to the links and options
for closer integration of the adaptation and disaster risk management
fields.
The gradual evolution of policies that favor disaster risk reduction
objectives as a component of development planning procedures (as
opposed to disaster management seen as a function of civil protection,
civil defense, emergency services, and ministries of public works) has
inevitably placed the preexisting emergency or disaster-response-oriented
institutional and organizational arrangements for disaster management
under scrutiny. The prior dominance of response-based and infrastructure
organizations has been complemented with the increasing incorporation
of economic and social sector and territorial development agencies or
organizations, as well as planning and finance ministries. Systemic, as
opposed to single agency, approaches are now evolving in many places.
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
36
Synergy, collaboration, coordination, and development of multidisciplinary
and multiagency schemes are increasingly seen as positive attributes for
guaranteeing implementation of disaster risk reduction and disaster risk
management in a sustainable development framework (see Lavell and
Franco, 1996; Ramírez and Cardona, 1996; Wisner et al., 2004, 2011).
Under these circumstances the notion of national disaster risk
management systems or structures has emerged strongly. Such
notions are discussed in detail in Chapter 6.
Adaptation to climate change, the second policy, strategic, and
instrumental aspect of importance for this Special Report, is a notion
that refers to both human and natural systems. Adaptation in human
systems is defined here as the process of adjustment to actual or
expected climate and its effects, in order to moderate harm or exploit
beneficial opportunities. In natural systems, it is defined as the process
of adjustment to actual climate and its effects; human intervention may
facilitate adjustment to expected climate.
These definitions modify the IPCC (2007c) definition that generically
speaks of the “adjustment in natural and human systems in response to
actual and expected climatic stimuli, such as to moderate harm or
exploit beneficial opportunities. The objective of the redefinition used
in this report is to avoid the implication present in the prior IPCC
definition that natural systems can adjust to expected climate stimuli.
At the same time, it accepts that some forms of human intervention may
provide opportunities for supporting natural system adjustment to
future climate stimuli that have been anticipated by humans.
Adaptation is a key aspect of the present report and is dealt with in
greater detail in Sections 1.3 and 1.4 and later chapters. The more
ample introduction to disaster risk management offered above derives
from the particular perspective of the present report: that adaptation is
a goal to be advanced and extreme event and disaster risk management
are methods for supporting and advancing that goal.
The notion of adaptation is counterposed to the notion of mitigation
in the climate change literature and practice. Mitigation there refers to
the reduction of the rate of climate change via the management of
its causal factors (the emission of greenhouse gases from fossil fuel
combustion, agriculture, land use changes, cement production, etc.)
(IPCC, 2007c). However, in disaster risk reduction practice, ‘mitigation’
refers to the amelioration of disaster risk through the reduction of existing
hazards, exposure, or vulnerability, including the use of different disaster
preparedness measures.
Disaster preparedness measures, including early warning and the
development of contingency or emergency plans, may be considered a
component of, and a bridge between, disaster risk reduction and disaster
management. Preparedness accepts the existence of residual, unmitigated
risk, and attempts to aid society in eliminating certain of the adverse
effects that could be experienced once a physical event(s) occurs (for
example, by the evacuation of persons and livestock from exposed and
vulnerable circumstances). At the same time, it provides for better
response to adverse effects that do materialize (for example, by
planning for adequate shelter and potable water supplies for the affected
or destitute persons or food supplies for affected animal populations).
In order to accommodate the two differing definitions of mitigation, this
report presumes that mitigation is a substantive action that can be
applied in different contexts where attenuation of existing specified
conditions is required.
Disaster mitigation is used to refer to actions that attempt to limit
further adverse conditions once disaster has materialized. This refers to
the avoidance of what has sometimes been called the ‘second disaster’
following the initial physical impacts (Alexander, 2000; Wisner et al.,
2011). The ‘second disaster’ may be characterized, among other things,
by adverse effects on health (Noji, 1997; Wisner et al., 2011) and
livelihoods due to inadequate disaster response and rehabilitation plans,
inadequate enactment of existing plans, or unforeseen or unforeseeable
circumstances.
Disaster risk prevention and disaster prevention refer, in a strict sense,
to the elimination or avoidance of the underlying causes and conditions
that lead to disaster, thus precluding the possibility of either disaster
risk or disaster materializing. The notion serves to concentrate attention
on the fact that disaster risk is manageable and its materialization is
preventable to an extent (which varies depending on the context).
Prospective (proactive) disaster risk management and adaptation
can contribute in important ways to avoiding future, and not just reducing
existing, risk and disaster once they have become manifest, as is the case
with corrective or reactive management (Lavell, 2003; UNISDR, 2011).
1.1.2.3. The Social Construction of Disaster Risk
The notions of hazard, exposure, vulnerability, disaster risk, capacity,
resilience, and coping, and their social origins and bases, as presented
above, reflect an emerging understanding that disaster risk and disaster,
while potentiated by an objective, physical condition, are fundamentally
a ‘social construction,’ the result of social choice, social constraints, and
societal action and inaction (high confidence). The notion of social
construction of risk implies that management can take into account the
social variables involved and to the best of its ability work toward
risk reduction, disaster management, or risk transfer through socially
sustainable decisions and concerted human action (ICSU-LAC, 2009).
This of course does not mean that there are not risks that may be too
great to reduce significantly through human intervention, or others
that the very social construction process may in fact exacerbate (see
Sections 1.3.1.2 and 1.4.3). But in contrast with, for example, many
natural physical events and their contribution to disaster risk, the
component of risk that is socially constructed is subject to intervention
in favor of risk reduction.
The contribution of physical events to disaster risk is characterized by
statistical distributions in order to elucidate the options for risk reduction
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
37
and adaptation (Section 1.2 and Chapter 3). But, the explicit recognition
of the political, economic, social, cultural, physical, and psychological
elements or determinants of risk leads to a spectrum of potential outcomes
of physical events, including those captured under the notion of
extreme impacts (Section 1.2 and Chapter 4). Accordingly, risk
assessment (see Section 1.3) using both quantitative and qualitative
(social and psychological) measures is required to render a more
complete description of risk and risk causation processes (Section 1.3;
Douglas and Wildavsky, 1982; Cardona, 2004; Wisner et al., 2004; Weber,
2006). Climate change may introduce a break with past environmental
system functioning so that forecasting physical events becomes less
determined by past trends. Under these conditions, the processes that
cause, and the established indicators of, human vulnerability need to be
reconsidered in order for risk assessment to remain an effective tool.
The essential nature and structure of the characteristics that typify
vulnerability can of course change without climate changing.
1.1.3. Framing the Relation between Adaptation to
Climate Change and Disaster Risk Management
Adaptation to climate change and disaster risk management both seek
to reduce factors and modify environmental and human contexts that
contribute to climate-related risk, thus supporting and promoting
sustainability in social and economic development. The promotion of
adequate preparedness for disaster is also a function of disaster risk
management and adaptation to climate change. And, both practices are
seen to involve learning (see Section 1.4), having a corrective and
prospective component dealing with existing and projected future risk.
However, the two practices have tended to follow independent paths of
advance and development and have on many occasions employed
different interpretations of concepts, methods, strategies, and institutional
frameworks to achieve their ends. These differences should clearly be
taken into account in the search for achieving greater synergy between
them and will be examined in an introductory fashion in Section 1.3 and
in greater detail in following chapters of this report.
Public policy and professional concepts of disaster and their approaches
to disaster and disaster risk management have undergone very significant
changes over the last 30 years, so that challenges that are now an
explicit focus of the adaptation field are very much part of current disaster
risk reduction as opposed to mainstream historical disaster management
concerns (Lavell, 2010; Mercer, 2010).These changes have occurred under
the stimuli of changing concepts, multidisciplinary involvement, social
and economic demands, and impacts of disasters, as well as institutional
changes reflected in international accords and policies such as the UN
Declaration of the International Decade for Natural Disaster Reduction
in the 1990s, the 2005 Hyogo Framework for Action, as well as the work
of the International Strategy for Disaster Reduction since 2000.
Particularly in developing countries, this transition has been stimulated
by the documented relationship between disaster risk and ‘skewed’
development processes (UNEP, 1972; Cuny, 1983; Sen, 1983; Hagman,
1984; Wijkman and Timberlake, 1988; Lavell, 1999, 2003; UNDP, 2004;
Wisner et al., 2004, 2011; Dulal et al., 2009; UNISDR, 2009e, 2011).
Significant differentiation in the distribution or allocation of gains from
development and thus in the incidence of chronic or everyday risk,
which disproportionately affect poorer persons and families, is a major
contributor to the more specific existence of disaster risk (Hewitt, 1983,
1997; Wisner et al., 2004). Reductions in the rate of ecosystem services
depletion, improvements in urban land use and territorial organization
processes, the strengthening of rural livelihoods, and general and specific
advances in urban and rural governance are viewed as indispensable to
achieving the composite agenda of poverty reduction, disaster risk
reduction, and adaptation to climate change (UNISDR, 2009e, 2011)
(high confidence).
Climate change is at once a problem of development and also a symptom
of ‘skewed’ development. In this context, pathways toward resilience
include both incremental and transformational approaches to development
(Chapter 8). Transformational strategies place emphasis on addressing
risk that stems from social structures as well as social behavior and
have a broader scope extending from disaster risk management into
development goals, policy, and practice (Nelson et al., 2007). In this
way transformation builds on a legacy of progressive, socially informed
disaster risk research that has applied critical methods, including that of
Hewitt (1983), Watts (1983), Maskrey (1989, 2011), Blaikie et al. (1994),
and Wisner et al. (2004).
However, while there is a longstanding awareness of the role of
development policy and practice in shaping disaster risk, advances in
the reduction of the underlying causes – the social, political, economic,
and environmental drivers of disaster risk – remain insufficient to
reduce hazard, exposure, and vulnerability in many regions (UNISDR,
2009e, 2011) (high confidence).
The difficult transition to more comprehensive disaster risk management
raises challenges for the proper allocation of efforts among disaster risk
reduction, risk transfer, and disaster management efforts. Countries
exhibit a wide range of acceptance or resistance to the various challenges
of risk management as seen from a development perspective, due to
differential access to information and education, varying levels of
debate and discussion, as well as contextual, ideological, institutional,
and other related factors. The introduction of disaster risk reduction
concerns in established disaster response agencies may in some cases
have led to a downgrading of efforts to improve disaster response,
diverting scarce resources in favor of risk reduction aspects (Alexander,
2000; DFID, 2004, 2005; Twigg, 2004).
The increasing emphasis placed on considering disaster risk management
as a dimension of development, and thus of development planning, as
opposed to strict post-impact disaster response efforts, has been
accompanied by increasing emphasis and calls for proactive, prospective
disaster risk prevention as opposed to reactive, corrective disaster risk
mitigation (Lavell, 2003, 2010; UNISDR, 2009e, 2011).
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
38
The more recent emergence of integrated disaster risk management
reflects a shift from the notion of disaster to the notion of disaster risk
as a central concept and planning concern. Disaster risk management
places increased emphasis on comprehensive disaster risk reduction.
This shifting emphasis to risk reduction can be seen in the increasing
importance placed on developing resistance to the potential impacts
of physical events at various social or territorial scales, and in different
temporal dimensions (such as those required for corrective or prospective
risk management), and to increasing the resilience of affected communities.
Resistance refers to the ability to avoid suffering significant adverse
effects.
Within this context, disaster risk reduction and adaptation to climate
change are undoubtedly far closer practically than when emergency or
disaster management objectives dominated the discourse and practice. The
fact that many in the climate change and disaster fields have associated
disaster risk management principally with disaster preparedness and
response, and not with disaster risk reduction per se, contributed to the
view that the two practices are essentially different, if complementary
(Lavell, 2010; Mercer, 2010). Once the developmental basis of adaptation
to climate change and disaster risk management are considered, along
with the role of vulnerability in the constitution of risk, the temporal
scale of concerns, and the corrective as well as prospective nature of
disaster risk reduction, the similarities between and options for merging
of concerns and practices increases commensurately.
Section 1.3 examines the current status of adaptation to climate
change, as a prelude to examining in more detail the barriers and
options for greater integration of the two practices. The historical frame
offered in this subsection comprises an introduction to that discussion.
1.1.4. Framing the Processes of Disaster Risk
Management and Adaptation to Climate Change
In this section, we explore two of the key issues that should be considered
in attempting to establish the overlap or distinction between the
phenomena and social processes that concern disaster risk management
on the one hand, and adaptation to climate change on the other, and
that influence their successful practice: 1) the degree to which the focus
is on extreme events (instead of a more inclusive approach that considers
the full continuum of physical events with potential for damage, the social
contexts in which they occur, and the potential for such events to generate
‘extreme impacts’ or disasters); and 2) consideration of the appropriate
social-territorial scale that should be examined (i.e., aggregations, see
Schneider et al., 2007) in order to foster a deeper understanding of the
causes and effects of the different actors and processes at work.
1.1.4.1. Exceptionality, Routine, and Everyday Life
Explanations of loss and damage resulting from extreme events that
focus primarily or exclusively on the physical event have been referred
to as ‘physicalist’ (Hewitt, 1983). By contrast, notions developed around
the continuum of normal, everyday-life risk factors through to a linked
consideration of physical and social extremes have been defined as
‘comprehensive,‘integral,’ or ‘holistic’ insofar as they embrace the social
as well as physical aspects of disaster risk and take into consideration the
evolution of experience over time (Cardona, 2001; ICSU-LAC, 2009). The
latter perspective has been a major contributing factor in the development
of the so-called ‘vulnerability paradigm’ as a basis for understanding
disaster (Timmerman, 1981; Hewitt, 1983, 1997; Wisner et al., 2004;
Eakin and Luers, 2006; NRC, 2006).
Additionally, attention to the role of small- and medium-scale disasters
(UNISDR, 2009e, 2011) highlights the need to deal integrally with the
problem of cumulative disaster loss and damage, looking across the
different scales of experience both in human and physical worlds,
in order to advance the efficacy of disaster risk management and
adaptation. The design of mechanisms and strategies based on the
reduction and elimination of everyday or chronic risk factors (Sen, 1983;
World Bank, 2001), as opposed to actions based solely on the
‘exceptional’ or ‘extreme’ events, is one obvious corollary of this
approach. The ability to deal with risk, crisis, and change is closely
related to an individual’s life experience with smaller-scale, more
regular physical and social occurrences (Maskrey, 1989, 2011; Lavell,
2003; Wisner et al., 2004) (high confidence). These concepts point
toward the possibility of reducing vulnerability and increasing resilience
to climate-related disaster by broadly focusing on exposure, vulnerability,
and socially-determined propensity or predisposition to adverse effects
across a range of risks.
As illustrated in Box 1-1, many of the extreme impacts associated with
climate change, and their attendant additional risks and opportunities,
will inevitably need to be understood and responded to principally at
the scale of the individual, the individual household, and the community,
in the framework of localities and nations and their organizational and
management options, and in the context of the many other day-to-day
changes, including those of an economic, political, technological, and
cultural nature. As this real example illustrates, everyday life, history, and
a sequence of crises can affect attitudes and ways of approaching more
extreme or complex problems. In contrast, many agents and institutions
of disaster risk management and climate change adaptation activities
necessarily operate from a different perspective, given the still highly
centralized and hierarchical authority approaches found in many parts
of the world today.
Whereas disaster risk management has been modified based on the
experiences of the past 30 years or more, adaptation to anthropogenic
climate change is a more recent issue on most decisionmakers’ policy
agendas and is not informed by such a long tradition of immediate
experience. However, human adaptation to prevailing climate variability
and change, and climate and weather extremes in past centuries and
millennia, provides a wealth of experience from which the field of
adaptation to climate change, and individuals and governments, can
draw.
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
39
The ethnographic vignette in Box 1-1 suggests the way some individuals
may respond to climate change in the context of previous experience,
illustrating both the possibility of drawing successfully on past experience
in adapting to climate variability, or, on the other hand, failing to
comprehend the nature of novel risks.
1.1.4.2. Territorial Scale, Disaster Risk, and Adaptation
Climate-related disaster risk is most adequately depicted, measured, and
monitored at the local or micro level (families, communities, individual
buildings or production units, etc.) where the actual interaction of hazard
and vulnerability are worked out in situ (Hewitt, 1983, 1997; Lavell,
2003; Wisner et al., 2004; Cannon, 2006; Maskrey, 2011). At the same
time, it is accepted that disaster risk construction processes are not
limited to specifically local or micro processes but, rather, to diverse
environmental, economic, social, and ideological influences whose
sources are to be found at scales from the international through to the
national, sub-national and local, each potentially in constant flux (Lavell,
2002, 2003; Wisner et al., 2004, 2011).
Changing commodity prices in international trading markets and their
impacts on food security and the welfare of agricultural workers, decisions
on location and cessation of agricultural production by international
corporations, deforestation in the upper reaches of river basins, and land
use changes in urban hinterlands are but a few of these ‘extra-territorial’
influences on local risk. Moreover, disasters, once materialized, have ripple
effects that many times go well beyond the directly affected zones (Wisner
et al., 2004; Chapter 5) Disaster risk management and adaptation policy,
strategies, and institutions will only be successful where understanding
and intervention is based on multi-territorial and social-scale principles
and where phenomena and actions at local, sub-national, national, and
international scales are construed in interacting, concatenated ways
(Lavell, 2002; UNISDR, 2009e, 2011; Chapters 5 through 9).
1.2. Extreme Events, Extreme Impacts,
and Disasters
1.2.1. Distinguishing Extreme Events,
Extreme Impacts, and Disasters
Both the disaster risk management and climate change adaptation
literature define ‘extreme weather’ and ‘extreme climate’ events and
discuss their relationship with ‘extreme impacts’ and ‘disasters.’
Classification of extreme events, extreme impacts, and disasters is
influenced by the measured physical attributes of weather or climatic
variables (see Section 3.1.2) or the vulnerability of social systems (see
Section 2.4.1).
This section explores the quantitative definitions of different classes of
extreme weather events, what characteristics determine that an impact
is extreme, and how climate change affects the understanding of
extreme climate events and impacts.
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
Box 1-1 | One Person’s Experience with Climate Variability in the Context of Other Changes
Joseph is 80 years old. He and his father and his grandfather have witnessed many changes. Their homes have shifted back and forth
from the steep slopes of the South Pare Mountains at 1,500 m to the plains 20 km away, near the Pangani River at 600 m, in Tanzania.
What do ‘changes’ (mabadiliko) mean to someone whose father saw the Germans and British fight during the First World War and
whose grandfather defended against Maasai cattle raids when Victoria was still Queen?
Joseph outlived the British time. He saw African Socialism come and go after Independence. A road was constructed parallel to the old
German rail line. Successions of commercial crops were dominant during his long life, some grown in the lowlands on plantations (sisal,
kapok, and sugar), and some in the mountains (coffee, cardamom, ginger). He has seen staple foods change as maize became more
popular than cassava and bananas. Land cover has also changed. Forest retreated, but new trees were grown on farms. Pasture grasses
changed as the government banned seasonal burning. The Pangani River was dammed, and the electricity company decides how much
water people can take for irrigation. Hospitals and schools have been built. Insecticide-treated bed nets recently arrived for the children
and pregnant mothers.
Joseph has nine plots of land at different altitudes spanning the distance from mountain to plain, and he keeps in touch with his children
who work them by mobile phone. What is ‘climate change’ (mabadiliko ya tabia nchi) to Joseph? He has suffered and benefited from
many changes. He has lived through many droughts with periods of hunger, witnessed floods, and also seen landslides in the mountains.
He is skilled at seizing opportunities from changes – small and large: “Mabadiliko bora kuliko mapumziko” (Change is better than resting).
The provenance of this story is an original field work interview undertaken by Ben Wisner in November 2009 in Same District, Kilimanjaro
Region, Tanzania in the context of the U.S. National Science Foundation-funded research project “Linking Local Knowledge and Local
Institutions for the Study of Adaptive Capacity to Climate Change: Participatory GIS in Northern Tanzania.”
40
1.2.2. Extreme Events Defined in Physical Terms
1.2.2.1. Definitions of Extremes
Some literature reserve the term ‘extreme event’ for initial meteorological
phenomena (Easterling et al., 2000; Jentsch et al., 2007), some include
the consequential physical impacts, like flooding (Young, 2002), and some
the entire spectrum of outcomes for humans, society, and ecosystems
(Rich et al., 2008). In this report, we use ‘extreme (weather or climate)
event’ to refer solely to the initial and consequent physical phenomena
including some (e.g., flooding) that may have human components to
causation other than that related to the climate (e.g., land use or land
cover change or changes in water management; see Section 3.1.2 and
Glossary). The spectrum of outcomes for humans, society, and physical
systems, including ecosystems, are considered ‘impacts’ rather than part
of the definition of ‘events’ (see Sections 1.1.2.1 and 3.1.2 and the
Glossary).
In addition to providing a long-term mean of weather, ‘climate’
characterizes the full spectrum of means and exceptionality associated
with ‘unusual’ and unusually persistent weather. The World Meteorological
Organization (WMO, 2010) differentiates the terms in the following way
(see also FAQ 6.1): At the simplest level the weather is what is happening
to the atmosphere at any given time. Climate in a narrow sense is
usually defined as the ‘average weather,’ or more rigorously, as the
statistical description in terms of the mean and variability of relevant
quantities over a period of time.”
Weather and climate phenomena reflect the interaction of dynamic and
thermodynamic processes over a very wide range of space and temporal
scales. This complexity results in highly variable atmospheric conditions,
including temperatures, motions, and precipitation, a component of
which is referred to as ‘extreme events.’ Extreme events include the
passage of an intense tornado lasting minutes and the persistence of
drought conditions over decades – a span of at least seven orders of
magnitude of timescales. An imprecise distinction between extreme
‘weather’ and ‘climate’ events, based on their characteristic timescales,
is drawn in Section 3.1.2. Similarly, the spatial scale of extreme climate
or weather varies from local to continental.
Where there is sufficient long-term recorded data to develop a statistical
distribution of a key weather or climate variable, it is possible to find the
probability of experiencing a value above or below different thresholds
of that distribution as is required in engineering design (trends may be
sought in such data to see if there is evidence that the climate has not
been stationary over the sample period; Milly et al., 2008). The extremity
of a weather or climate event of a given magnitude depends on
geographic context (see Section 3.1.2 and Box 3-1): a month of daily
temperatures corresponding to the expected spring climatological daily
maximum in Chennai, India, would be termed a heat wave in France; a
snow storm expected every year in New York, USA, might initiate a
disaster when it occurs in southern China. Furthermore, according to the
location and social context, a 1-in-10 or 1-in-20 annual probability
event may not be sufficient to result in unusual consequences.
Nonetheless, universal thresholds can exist – for example, a reduction
in the incidence or intensity of freezing days may allow certain disease
vectors to thrive (e.g., Epstein et al., 1998). These various aspects are
considered in the definition of ‘extreme (weather and climate) events.
The availability of observational data is of central relevance for defining
climate characteristics and for disaster risk management; and, while data
for temperature and precipitation are widely available, some associated
variables, such as soil moisture, are poorly monitored, or, like extreme
wind speeds and other low frequency occurrences, not monitored with
sufficient spatial resolution or temporal continuity (Section 3.2.1).
1.2.2.2. Extremes in a Changing Climate
An extreme event in the present climate may become more common, or
more rare, under future climate conditions. When the overall distribution
of the climate variable changes, what happens to mean climate may
be different from what happens to the extremes at either end of the
distribution (see Figure 1-2).
For example, a warmer mean climate could result from fewer cold days,
leading to a reduction in the variance of temperatures, or more hot days,
leading to an expansion in the variance of the temperature distribution,
or both. The issue of the scaling of changes in extreme events with respect
to changes in mean temperatures is addressed further in Section 3.1.6.
In general, single extreme events cannot be simply and directly attributed
to anthropogenic climate change, as there is always a possibility the
event in question might have occurred without this contribution (Hegerl
et al., 2007; Section 3.2.2; FAQ 3.2). However, for certain classes of
regional, long-duration extremes (of heat and rainfall) it has proved
possible to argue from climate model outputs that the probability of
such an extreme has changed due to anthropogenic climate forcing
(Stott et al., 2004; Pall et al., 2011).
Extremes sometimes result from the interactions between two unrelated
geophysical phenomena such as a moderate storm surge coinciding
with an extreme spring tide, as in the most catastrophic UK storm surge
flood of the past 500 years in 1607 (Horsburgh and Horritt, 2006).
Climate change may alter both the frequency of extreme surges and
cause gradual sea level rise, compounding such future extreme floods
(see Sections 3.5.3 and 3.5.5).
1.2.2.3. The Diversity and Range of Extremes
The specification of weather and climate extremes relevant to the
concerns of individuals, communities, and governments depends on the
affected stakeholder, whether in agriculture, disease control, urban
design, infrastructure maintenance, etc. Accordingly, the range of such
extremes is very diverse and varies widely. For example, whether it falls
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
41
as rain, freezing rain (rain falling through a surface layer below freezing),
snow, or hail, extreme precipitation can cause significant damage
(Peters et al., 2001). The absence of precipitation (McKee et al., 1993) as
well as excess evapotranspiration from the soil (see Box 3-3) can be
climate extremes, and lead to drought. Extreme surface winds are
chiefly associated with structured storm circulations (Emanuel, 2003;
Zipser et al., 2006; Leckebusch et al., 2008). Each storm type, including
the most damaging tropical cyclones and mid-latitude extratropical
cyclones, as well as intense convective thunderstorms, presents a
spectrum of size, forward speed, and intensity. A single intense storm
can combine extreme wind and extreme rainfall.
The prolonged absence of winds is a climate extreme that can also be a
hazard, leading to the accumulation of urban pollution and disruptive
fog (McBean, 2006).
The behavior of the atmosphere is also highly interlinked with that of
the hydrosphere, cryosphere, and terrestrial environment so that extreme
(or sometimes non-extreme) atmospheric events may cause (or contribute
to) other rare physical events. Among the more widely documented
hydroclimatic extremes are:
Large cyclonic storms that generate wind and pressure anomalies
causing coastal flooding and severe wave action (Xie et al., 2004).
Floods, reflecting river flows in excess of the capacity of the normal
channel, often influenced by human intervention and water
management, resulting from intense precipitation; rapid thaw of
accumulated winter snowfall; rain falling on previous snowfall (Sui
and Koehler, 2001); or an outburst from an ice, landslide, moraine,
or artificially dammed lake (de Jong et al., 2005). According to the
scale of the catchment, river systems have characteristic response
times with steep short mountain streams, desert wadis, and urban
drainage systems responding to rainfall totals over a few hours, while
peak flows in major continental rivers reflect regional precipitation
extremes lasting weeks (Wheater, 2002).
Long-term reductions in precipitation, or dwindling of residual
summer snow and ice melt (Rees and Collins, 2006), or increased
evapotranspiration from higher temperatures, often exacerbated
by human groundwater extraction, reducing ground water levels
and causing spring-fed rivers to disappear (Konikow and Kendy,
2005), and contributing to drought.
Landslides (Dhakal and Sidle, 2004) when triggered by raised
groundwater levels after excess rainfall or active layer detachments
in thawing slopes of permafrost (Lewcowicz and Harris, 2005).
1.2.3. Extreme Impacts
1.2.3.1. Three Classes of Impacts
In this subsection we consider three classes of ‘impacts’: 1) changes in
the natural physical environment, like beach erosion from storms and
mudslides; 2) changes in ecosystems, such as the blow-down of forests
in hurricanes, and 3) adverse effects (according to a variety of metrics)
on human or societal conditions and assets. However, impacts are not
always negative: flood-inducing rains can have beneficial effects on the
following season’s crops (Khan, 2011), while an intense freeze may
reduce insect pests at the subsequent year’s harvest (Butts et al., 1997).
An extreme impact reflects highly significant and typically long-lasting
consequences to society, the natural physical environment, or ecosystems.
Extreme impacts can be the result of a single extreme event, successive
extreme or non-extreme events, including non-climatic events (e.g.,
wildfire, followed by heavy rain leading to landslides and soil erosion),
or simply the persistence of conditions, such as those that lead to
drought (see Sections 3.5.1 and 9.2.3 for discussion and examples).
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
Probability of OccurrenceProbability of OccurrenceProbability of Occurrence
less
record cold
weather
more
record cold
weather
less
cold
weather
near constant
record cold
weather
near constant
cold
weather
more
record hot
weather
more
record hot
weather
more
record hot
weather
more
cold
weather
more
hot
weather
more
hot
weather
more
hot
weather
Previous Climate
Future Climate
a)
b)
c)
Shifted Mean
Increased Variability
Changed Shape
Cold HotAverage
Figure 1-2 | The effect of changes in temperature distribution on extremes. Different
changes in temperature distributions between present and future climate and their
effects on extreme values of the distributions: a) effects of a simple shift of the entire
distribution toward a warmer climate; b) effects of an increased temperature variability
with no shift of the mean; and c) effects of an altered shape of the distribution, in this
example an increased asymmetry toward the hotter part of the distribution.
42
Whether an extreme event results in extreme impacts on humans and
social systems depends on the degree of exposure and vulnerability to
that extreme, in addition to the magnitude of the physical event (high
confidence). Extreme impacts on human systems may be associated
with non-extreme events where vulnerability and exposure are high
(Sections 1.1.2.1 and 9.2.3). A key weather parameter may cross some
critical value at that location (such as that associated with heat wave-
induced mortality, or frost damage to crops), so that the distribution of
the impact shifts in a way that is disproportionate to physical changes
(see Section 4.2). A comprehensive assessment of projected impacts of
climate changes would consider how changes in atmospheric conditions
(temperature, precipitation) translate to impacts on physical (e.g.,
droughts and floods, erosion of beaches and slopes, sea level rise),
ecological (e.g., forest fires), and human systems (e.g., casualties,
infrastructure damages). For example, an extreme event with a large
spatial scale (as in an ice storm or windstorm) can have an exaggerated,
disruptive impact due to the systemic societal dependence on electricity
transmission and distribution networks (Peters et al., 2006). Links between
climate events and physical impacts are addressed in Section 3.5, while
links to ecosystems and human systems impacts are addressed in 4.3.
Disaster signifies extreme impacts suffered by society, which may also
be associated with extreme impacts on the physical environment and
on ecosystems. Building on the definition set out in Section 1.1.2.1,
extreme impacts resulting from weather, climate, or hydrological events
can become disasters once they surpass thresholds in at least one of
three dimensions: spatial – so that damages cannot be easily restored
from neighboring capacity; temporal – so that recovery becomes
frustrated by further damages; and intensity of impact on the affected
population – thereby undermining, although not necessarily eliminating,
the capacity of the society or community to repair itself (Alexander,
1993). However, for the purposes of tabulating occurrences, some
agencies only list ‘disasters’ when they exceed certain numbers of killed
or injured or total repair costs (Below et al., 2009; CRED, 2010).
1.2.3.2. Complex Nature of an Extreme ‘Event’
In considering the range of weather and climate extremes, along with
their impacts, the term ‘event’ as used in the literature does not
adequately capture the compounding of outcomes from successive
physical phenomena, for example, a procession of serial storms tracking
across the same region (as in January and February 1990 and December
1999 across Western Europe, Ulbrich et al., 2001). In focusing on the social
context of disasters, Quarantelli (1986) proposed the use of the notion of
‘disaster occurrences or occasions’ in place of ‘events’ due to the abrupt
and circumstantial nature of the connotation commonly attributed to
the word ‘event,’ which belies the complexity and temporality of disaster,
in particular because social context may precondition and extend the
duration over which impacts are felt.
Sometimes locations affected by extremes within the ‘same’ large-scale
stable atmospheric circulation can be far apart, as for example the
Russian heat wave and Indus valley floods in Pakistan in the summer of
2010 (Lau and Kim, 2011). Extreme events can also be interrelated
through the atmospheric teleconnections that characterize the principal
drivers of oceanic equatorial sea surface temperatures and winds in the
El Niño–Southern Oscillation. The relationship between modes of climate
variability and extremes is discussed in greater detail in Section 3.1.1.
The aftermath of one extreme event may precondition the physical
impact of successor events. High groundwater levels and river flows can
persist for months, increasing the probability of a later storm causing
flooding, as on the Rhine in 1995 (Fink et al., 1996). A thickness reduction
in Arctic sea ice preconditions more extreme reductions in the summer
ice extent (Holland et al., 2006). A variety of feedbacks and other
interactions connect extreme events and physical system and ecological
responses in a way that may amplify physical impacts (Sections 3.1.4
and 4.3.5). For example, reductions in soil moisture can intensify heat
waves (Seneviratne et al., 2006), while droughts following rainy seasons
turn vegetation into fuel that can be consumed in wildfires (Westerling
and Swetman, 2003), which in turn promote soil runoff and landslides
when the rains return (Cannon et al., 2001). However, extremes can also
interact to reduce disaster risk. The wind-driven waves in a hurricane
bring colder waters to the surface from beneath the thermocline; for the
next month, any cyclone whose path follows too closely will have a
reduced potential maximum intensity (Emanuel, 2001). Intense rainfall
accompanying monsoons and hurricanes also brings great benefits to
society and ecosystems; on many occasions it helps to fill reservoirs,
sustain seasonal agriculture, and alleviate summer dry conditions in arid
zones (e.g., Cavazos et al., 2008).
1.2.3.3. Metrics to Quantify Social Impacts
and the Management of Extremes
Metrics to quantify social and economic impacts (thus used to define
extreme impacts) may include, among others (Below et al., 2009):
Human casualties and injuries
Number of permanently or temporarily displaced people
Number of directly and indirectly affected persons
Impacts on properties, measured in terms of numbers of buildings
damaged or destroyed
Impacts on infrastructure and lifelines
Impacts on ecosystem services
Impacts on crops and agricultural systems
Impacts on disease vectors
Impacts on psychological well being and sense of security
Financial or economic loss (including insurance loss)
Impacts on coping capacity and need for external assistance.
All of these may be calibrated according to the magnitude, rate, duration,
and degree of irreversibility of the effects (Schneider et al., 2007).
These metrics may be quantified and implemented in the context of
probabilistic risk analysis in order to inform policies in a variety of
contexts (see Box 1-2).
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
43
Information on direct, indirect, and collateral impacts is generally
available for many large-scale disasters and is systematized and provided
by organizations such as the Economic Commission for Latin America,
large reinsurers, and the EM-DAT database (CRED, 2010). Information
on impacts of smaller, more recurrent events is far less accessible and
more restricted in the number of robust variables it provides. The
Desinventar database (Corporación OSSO, 2010), now available for 29
countries worldwide, and the Spatial Hazard Events and Losses
Database for the United States (SHELDUS; HVRI, 2010), are attempts to
satisfy this need. However, the lack of data on many impacts impedes
complete knowledge of the global social and economic impacts of
smaller-scale disasters (UNISDR, 2009e).
1.2.3.4. Traditional Adjustment to Extremes
Disaster risk management and climate change adaptation may be seen
as attempts to duplicate, promote, or improve upon adjustments that
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
Box 1-2 | Probabilistic Risk Analysis
In its simplest form, probabilistic risk analysis defines risk as the product of the probability that some event (or sequence) will occur and
the adverse consequences of that event.
Risk = Probability x Consequence (1)
For instance, the risk a community faces from flooding from a nearby river might be calculated based on the likelihood that the river floods
the town, inflicting casualties among inhabitants and disrupting the community’s economic livelihood. This likelihood is multiplied by the
value people place on those casualties and economic disruption. Equation (1) provides a quantitative representation of the qualitative
definition of disaster risk given in Section 1.1. All three factors – hazard, exposure, and vulnerability – contribute to ‘consequences.
Hazard and vulnerability can both contribute to the ‘probability’: the former to the likelihood of the physical event (e.g., the river flooding
the town) and the latter to the likelihood of the consequence resulting from the event (e.g., casualties and economic disruption).
When implemented within a broader risk governance framework, probabilistic risk analysis can help allocate and evaluate efforts to
manage risk. Equation (1) implies what the decision sciences literature (Morgan and Henrion, 1990) calls a decision rule – that is, a
criterion for ranking alternative sets of actions by their ability to reduce overall risk. For instance, an insurance company (as part of a risk
transfer effort) might set the annual price for flood insurance based on multiplying an estimate of the probability a dwelling would be
flooded in any given year by an estimate of the monetary losses such flooding would cause. Ideally, the premiums collected from the
residents of many dwellings would provide funds to compensate the residents of those few dwellings that are in fact flooded (and
defray administrative costs). In another example, a water management agency (as part of a risk reduction effort) might invest the
resources to build a reservoir of sufficient size so that, if the largest drought observed in their region over the last 100 years (or some
other timeframe) occurred again in the future, the agency would nonetheless be able to maintain a reliable supply of water.
A wide variety of different expressions of the concepts in Equation (1) exist in the literature. The disaster risk management community
often finds it convenient to express risk as a product of hazard, exposure, and vulnerability (e.g., UNISDR, 2009e, 2011). In addition, the
decision sciences literature recognizes decision rules, useful in some circumstances, that do not depend on probability and consequence
as combined in Equation (1). For instance, if the estimates of probabilities are sufficiently imprecise, decisionmakers might use a criterion
that depends only on comparing estimates of potential consequences (e.g., mini-max regret, Savage, 1972).
In practice, probabilistic risk analysis is often not implemented in its pure form for reasons including data limitations; decision rules that
yield satisfactory results with less effort than that required by a full probabilistic risk assessment; the irreducible imprecision of some
estimates of important probabilities and consequences (see Sections 1.3.1.1 and 1.3.2); and the need to address the wide range of factors
that affect judgments about risk (see Box 1-3). In the above example, the water management agency is not performing a full probabilistic
risk analysis, but rather employing a hybrid decision rule in which it estimates that the consequences of running out of water would be
so large as to justify any reasonable investment needed to keep the likelihood of that event below the chosen probabilistic threshold.
Chapter 2 describes a variety of practical quantitative and qualitative approaches for allocating efforts to manage disaster risk.
The probabilistic risk analysis framework in its pure form is nonetheless important because its conceptual simplicity aids understanding
by making assumptions explicit, and because its solid theoretical foundations and the vast empirical evidence examining its application
in specific cases make it an important point of comparison for formal evaluations of the effectiveness of efforts to manage disaster risk.
44
society and nature have accomplished on many occasions spontaneously
in the past, if over a different range of conditions than expected in the
future.
Within the sphere of adaptation of natural systems to climate, among
trees, for example, natural selection has the potential to evolve
appropriate resilience to extremes (at some cost). Resistance to
windthrow is strongly species-dependent, having evolved according to
the climatology where that tree was indigenous (Canham et al., 2001).
In their original habitat, trees typically withstand wind extremes expected
every 10 to 50 years, but not extremes that lie beyond their average
lifespan of 100 to 500 years (Ostertag et al., 2005).
In human systems, communities traditionally accustomed to periodic
droughts employ wells, boreholes, pumps, dams, and water harvesting
and irrigation systems. Those with houses exposed to high seasonal
temperatures employ thick walls and narrow streets, have developed
passive cooling systems, adapted lifestyles, or acquired air conditioning.
In regions unaccustomed to heat waves, the absence of such systems,
in particular in the houses of the most vulnerable elderly or sick,
contributes to excess mortality, as in Paris, France, in August 2003
(Vandentorren et al., 2004) or California in July 2006 (Gershunov et al.,
2009).
The examples given above of ‘spontaneous’ human system adjustment
can be contrasted with explicit measures that are taken to reduce risk
from an expected range of extremes. On the island of Guam, within
the most active and intense zone of tropical cyclone activity on Earth,
buildings are constructed to the most stringent wind design code in the
world. Buildings are required to withstand peak gust wind speeds of
76 ms
-1
, expected every few decades (International Building Codes,
2003). More generally, annual wind extremes for coastal locations will
typically be highest at mid-latitudes while those expected once every
century will be highest in the 10° to 25° latitude tropics (Walshaw,
2000). Consequently, indigenous building practices are less likely to be
resilient close to the equator than in the windier (and storm surge
affected) mid-latitudes (Minor, 1983).
While local experience provides a reservoir of knowledge from which
disaster risk management and adaptation to climate change are drawing
(Fouillet et al., 2008), it may not be available to other regions yet to be
affected by such extremes. Thus, these experiences may not be drawn
upon to provide guidance if future extremes go outside the traditional
or recently observed range, as is expected for some extremes as the
climate changes (see Chapter 3).
1.3. Disaster Management, Disaster Risk
Reduction, and Risk Transfer
One important component of both disaster risk management and
adaptation to climate change is the appropriate allocation of efforts
among disaster management, disaster risk reduction, and risk transfer,
as defined in Section 1.1.2.2. The current section provides a brief survey
of the risk governance framework for making judgments about such an
allocation, suggests why climate change may complicate effective
management of disaster risks, and identifies potential synergies
between disaster risk management and adaptation to climate change.
Disaster risks appear in the context of human choices that aim to satisfy
human wants and needs (e.g., where to live and in what types of
dwelling, what vehicles to use for transport, what crops to grow, what
infrastructure to support economic activities, Hohenemser et al., 1984;
Renn, 2008). Ideally, the choice of any portfolio of actions to address
disaster risk would take into consideration human judgments about
what constitutes risk, how to weigh such risk alongside other values
and needs, and the social and economic contexts that determine whose
judgments influence individuals’ and societal responses to those risks.
The risk governance framework offers a systematic way to help situate
such judgments about disaster management, risk reduction, and risk
transfer within this broader context. Risk governance, under Renn’s
(2008) formulation, consists of four phases – pre-assessment, appraisal,
characterization/evaluation, and management – in an open, cyclical,
iterative, and interlinked process. Risk communication accompanies all
four phases. This process is consistent with those in the UNISDR Hyogo
Framework for Action (UNISDR, 2005), the best known and adhered to
framework for considering disaster risk management concerns (see
Chapter 7).
As one component of its broader approach, risk governance uses
concepts from probabilistic risk analysis to help judge appropriate
allocations in level of effort and over time and among risk reduction,
risk transfer, and disaster management actions. The basic probabilistic
risk analytic framework for considering such allocations regards risk
as the product of the probability of an event(s) multiplied by its
consequence (see Box 1-2; Bedford and Cooke, 2001). In this formulation,
risk reduction aims to reduce exposure and vulnerability as well as the
probability of occurrence of some events (e.g., those associated with
landslides and forest fires induced by human intervention). Risk transfer
efforts aim to compensate losses suffered by those who directly experience
an event. Disaster management aims to respond to the immediate
consequences and facilitate reduction of longer-term consequences (see
Section 1.1).
Probabilistic risk analysis can help compare the efficacy of alternative
actions to manage risk and inform judgments about the appropriate
allocation of resources to reduce risk. For instance, the framework
suggests that equivalent levels of risk reduction result from reducing an
event’s probability or by reducing its consequences by equal percentages.
Probabilistic risk analysis also suggests that a series of relatively smaller,
more frequent events could pose the same risk as a single, relatively less
frequent, larger event. Probabilistic risk analysis can help inform decisions
about alternative allocations of risk management efforts by facilitating
the comparison of the increase or decrease in risk resulting from the
alternative allocations (high confidence). Since the costs of available
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
45
risk reduction, risk transfer, and disaster management actions will in
general differ, the framework can help inform judgments about an
effective mix of such actions in any particular case (see UNISDR, 2011,
for efforts at stratifying different risk levels as a prelude to finding the
most adequate mix of disaster risk management actions).
Probabilistic risk analysis is, however, rarely implemented in its pure
form, in part because quantitative estimates of hazard and vulnerability
are not always available and are not numbers that are independent of
the individuals making those estimates. Rather, these estimates are
determined by a combination of direct physical consequences of an
event and the interaction of psychological, social, institutional, and
cultural processes (see Box 1-3). For instance, perceptions of the risks of a
nuclear power plant may be influenced by individuals’ trust in the people
operating the plant and by views about potential linkages between
nuclear power and nuclear weapons proliferation – factors that may not
be considered in a formal risk assessment for any given plant. Given this
social construction of risk (see Section 1.1.2.2), effective allocations of
efforts among risk reduction, risk transfer, and disaster management
may best emerge from an integrated risk governance process, which
includes the pre-assessment, appraisal, characterization/evaluation, and
ongoing communications elements. Disaster risk management and
adaptation to climate change each represent approaches that already
use or could be improved by the use of this risk governance process, but
as described in Section 1.3.1, climate change poses a particular set of
additional challenges.
Together, the implications of probabilistic risk analysis and the social
construction of risk reinforce the following considerations with regard to
the effective allocation and implementation of efforts to manage risks
in both disaster risk management and adaptation to climate change:
As noted in Section 1.1, vulnerability, exposure, and hazard are
each critical to determining disaster risk and the efficacy of actions
taken to manage that risk (high confidence).
Effective disaster risk management will in general require a
portfolio of many types of risk reduction, risk transfer, and disaster
management actions appropriately balanced in terms of resources
applied over time (high confidence).
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
Box 1-3 | Influence of Cognitive Processes, Culture, and Ideology on Judgments about Risk
A variety of cognitive, cultural, and social processes affect judgments about risk and about the allocation of efforts to address these
risks. In addition to the processes described in Section 1.3.1.2, subjective judgments may be influenced more by emotional reactions to
events (e.g., feelings of fear and loss of control) than by analytic assessments of their likelihood (Loewenstein et al., 2001). People
frequently ignore predictions of extreme events if those predictions fail to elicit strong emotional reactions, but will also overreact to
such forecasts when the events elicit feelings of fear or dread (Slovic et al., 1982; Slovic 1993, 2010; Weber, 2006). Even with sufficient
information, everyday concerns and satisfaction of basic wants may prove a more pressing concern than attention and effort toward
actions to address longer-term disaster risk (Maskrey, 1989, 2011; Wisner et al., 2004).
In addition to being influenced by cognitive shortcuts (Kahneman and Tversky, 1979), the perceptions of risk and extremes and reactions
to such risk and events are also shaped by motivational processes (Weber, 2010). Cultural theory combines insights from anthropology
and political science to provide a conceptual framework and body of empirical studies that seek to explain societal conflict over risk
(Douglas, 1992). People’s worldview and political ideology guide attention toward events that threaten their desired social order
(Douglas and Wildavsky, 1982). Risk in this framework is defined as the disruption of a social equilibrium. Personal beliefs also influence
which sources of expert forecasts of extreme climate events will be trusted. Different cultural groups put their trust into different
organizations, from national meteorological services to independent farm organizations to the IPCC; depending on their values, beliefs,
and corresponding mental models, people will be receptive to different types of interventions (Dunlap and McCright, 2008; Malka and
Krosnick, 2009). Judgments about the veracity of information regarding the consequences of alternative actions often depend on the
perceived consistency of those actions with an individual’s cultural values, so that individuals will be more willing to consider information
about consequences that can be addressed with actions seen as consistent with their values (Kahan and Braman, 2006; Kahan et al., 2007).
Factual information interacts with social, institutional, and cultural processes in ways that may amplify or attenuate public perceptions of
risk and extreme events (Kasperson et al., 1988). The US public’s estimates of the risk of nuclear power following the accident at Three
Mile Island provide an example of the socio-cultural filtering of engineering safety data. Social amplification increased public perceptions
of the risk of nuclear power far beyond levels that would derive only from analysis of accident statistics (Fischhoff et al., 1983). The public’s
transformation of expert-provided risk signals can serve as a corrective mechanism by which cultural subgroups of society augment a
science-based risk analysis with psychological risk dimensions not considered in technical risk assessments (Slovic, 2000). Evidence from
health, social psychology, and risk communication literature suggests that social and cultural risk amplification processes modify
perceptions of risk in either direction and in ways that may generally be socially adaptive, but can also bias reactions in socially
undesirable ways in specific instances (APA, 2009).
46
Participatory and decentralized processes that are linked to higher
levels of territorial governance (regions, nation) are a crucial part
of all the stages of risk governance that include identification,
choice, and implementation of these actions (high confidence).
1.3.1. Climate Change Will Complicate
Management of Some Disaster Risks
Climate change will pose added challenges in many cases for attaining
disaster risk management goals, and appropriately allocating efforts to
manage disaster risks, for at least two sets of reasons. First, as discussed in
Chapters 3 and 4, climate change is very likely to increase the occurrence
and vary the location of some physical events, which in turn will affect
the exposure faced by many communities, as well as their vulnerability.
Increased exposure and vulnerability would contribute to an increase in
disaster risk. For example, vulnerability may increase due to direct climate-
related impacts on the development and development potential of the
affected area, because resources otherwise available and directed
towards development goals are deflected to respond to those impacts,
or because long-standing institutions for allocating resources such as
water no longer function as intended if climate change affects the
scarcity and distribution of that resource. Second, climate change will
make it more difficult to anticipate, evaluate, and communicate both
probabilities and consequences that contribute to disaster risk, in
particular that associated with extreme events. This set of issues,
discussed in this subsection, will affect the management of these risks
as discussed in Chapters 5, 6, 7, and 8 (high confidence).
1.3.1.1. Challenge of Quantitative Estimates of Changing Risks
Extreme events pose a particular set of challenges for implementing
probabilistic approaches because their relative infrequency often makes
it difficult to obtain adequate data for estimating the probabilities and
consequences. Climate change exacerbates this challenge because it
contributes to potential changes in the frequency and character of such
events (see Section 1.2.2.2).
The likelihood of extreme events is most commonly described by the
return period, the mean interval expected between one such event and its
recurrence. For example, one might speak of a 100-year flood or a 50-year
windstorm. More formally, these intervals are inversely proportional to
the ‘annual exceedance probability,’ the likelihood that an event
exceeding some magnitude occurs in any given year. Thus the 100-year
flood has a 1% chance of occurring in any given year (which translates
into a 37% chance of a century passing without at least one such flood
((1-0.01)
100
= 37%). Though statistical methods exist to estimate
frequencies longer than available data time series (Milly et al., 2002),
the long return period of extreme events can make it difficult, if not
impossible, to reliably estimate their frequency. Paleoclimate records
make clear that in many regions of the world, the last few decades of
observed climate data do not represent the full natural variability of
many important climate variables (Jansen et al., 2003). In addition,
future climate change exacerbates the challenge of non-stationarity
(Milly et al., 2008), where the statistical properties of weather events
will not remain constant over time. This complicates an already difficult
estimation challenge by altering frequencies and consequences of
extremes in difficult-to-predict ways (Chapter 3; Meehl et al., 2007; TRB,
2008; NRC, 2009).
Estimating the likelihood of different consequences and their value is at
least as challenging as estimating the likelihood of extreme events.
Projecting future vulnerability and response capacity involves predicting
the trends and changes in underlying causes of human vulnerability and
the behavior of complex human systems under potentially stressful and
novel conditions. For instance, disaster risk is endogenous in the sense that
near-term actions to manage risk may affect future risk in unintended
ways and near-term actions may affect perceptions of future risks (see
Box 1-3). Section 1.4 describes some of the challenges such system
complexity may pose for effective risk assessment. In addition, disasters
affect socioeconomic systems in multiple ways so that assigning a
quantitative value to the consequences of a disaster proves difficult (see
Section 1.2.3.3). The literature distinguishes between direct losses,
which are the immediate consequences of the disaster-related physical
events, and indirect losses, which are the consequences that result from
the disruption of life and activity after the immediate impacts of the
event (Pelling et al., 2002; Lindell and Prater, 2003; Cochrane, 2004; Rose,
2004). Section 1.3.2 discusses some means to address these challenges.
1.3.1.2. Processes that Influence Judgments
about Changing Risks
Effective risk governance engages a wide range of stakeholder groups
– such as scientists, policymakers, private firms, nongovernmental
organizations, media, educators, and the public – in a process of
exchanging, integrating, and sharing knowledge and information. The
recently emerging field of sustainability science (Kates et al., 2001)
promotes interactive co-production of knowledge between experts and
other actors, based on transdisciplinarity (Jasanoff, 2004; Pohl et al.,
2010) and social learning (Pelling et al., 2008; Pahl-Wostl, 2009; see also
Section 1.4.2). The literature on judgment and decisionmaking suggests
that various cognitive behaviors involving perceptions and judgments
about low-probability, high-severity events can complicate the intended
functioning of such stakeholder processes (see Box 1-3). Climate change
can exacerbate these challenges (high confidence).
The concepts of disaster, risk, and disaster risk management have very
different meanings and interpretations in expert and non-expert contexts
(Sjöberg, 1999a; see also Pidgeon and Fischhoff, 2011). Experts acting
in formal private and public sector roles often employ quantitative
estimates of both probability and consequence in making judgments
about risk. In contrast, the general public, politicians, and the media
tend to focus on the concrete adverse consequences of such events,
paying less attention to their likelihood (Sjöberg, 1999b). As described
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
47
in Box 1-3, expert estimates of probability and consequence may also
not address the full range of concerns people bring to the consideration
of risk. By definition (if not always in practice), expert understanding of
risks associated with extreme events is based in large part on analytic
tools. In particular, any estimates of changes in disaster risk due to
climate change are often based on the results of complex climate
models as described in Chapter 3. Non-experts, on the other hand, rely
to a greater extent on more readily available and more easily processed
information, such as their own experiences or vicarious experiences from
the stories communicated through the news media, as well as their
subjective judgment as to the importance of such events (see Box 1-1).
These gaps between expert and non-expert understanding of extreme
events present important communication challenges (Weber and Stern,
2011), which may adversely affect judgments about the allocation of
efforts to address risk that is changing over time (high confidence).
Quantitative methods based on probabilistic risk analysis, such as those
described in Sections 5.5 and 6.3, can allow people operating in expert
contexts to use observed data, often from long time series, to make
systematic and internally consistent estimates of the probability of
future events. As described in Section 1.3.1.1, climate change may
reduce the accuracy of such past observations as predictors for future
risk. Individuals, including non-experts and experts making estimates
without the use of formal methods (Barke et al., 1997), often predict the
likelihood of encountering an event in the future by consulting their
past experiences with such events. The ‘availability’ heuristic (i.e., useful
shortcut) is commonly applied, in which the likelihood of an event is
judged by the ease with which past instances can be brought to mind
(Tversky and Kahneman, 1974). Extreme events, by definition, have a
low probability of being represented in past experience and thus will be
relatively unavailable. Experts and non-experts alike may essentially
ignore such events until they occur, as in the case of a 100-year flood
(Hertwig et al., 2004). When extreme events do occur with severe and
thus memorable consequences, people’s estimates of their future risks
will, at least temporarily, become inflated (Weber et al., 2004).
1.3.2. Adaptation to Climate Change
Contributes to Disaster Risk Management
The literature and practice of adaptation to climate change attempts to
anticipate future impacts on human society and ecosystems, such as
those described in Chapter 4, and respond to those already experienced.
In recent years, the adaptation to climate change literature has introduced
the concept of climate-related decisions (and climate proofing), which
are choices by individuals or organizations, the outcomes of which can
be expected to be affected by climate change and its interactions with
ecological, economic, and social systems (Brown et al., 2006; McGray et
al., 2007; Colls et al., 2009; Dulal et al., 2009; NRC, 2009). For instance,
choosing to build in a low-lying area whose future flooding risk increases
due to climate change represents a climate-related decision. Such a
decision is climate-related whether or not the decisionmakers recognize
it as such. The disaster risk management community may derive added
impetus from the new context of a changing climate for certain of its
pre-existing practices that already reflect the implementation of this
concept. In many circumstances, choices about the appropriate allocation
of efforts among disaster management, disaster risk reduction, and risk
transfer actions will be affected by changes in the frequency and
character of extreme events and other impacts of a climate change on
the underlying conditions that affect exposure and vulnerability.
Much of the relevant adaptation literature addresses how expectations
about future deviations from past patterns in physical, biological, and
socioeconomic conditions due to climate change should affect the
allocation of efforts to manage risks. While there exist differing views
on the extent to which the adaptation to climate change literature has
unique insights on managing changing conditions per se that it can
bring to disaster risk management (Lavell, 2010; Mercer, 2010; Wisner
et al., 2011), the former field’s interest in anticipating and responding
to the full range of consequences from changing climatic conditions can
offer important new perspectives and capabilities to the latter field.
The disaster risk management community can benefit from the debates
in the adaptation literature about how to best incorporate information
about current and future climate into climate-related decisions. Some
adaptation literature has emphasized the leading role of accurate
regional climate predictions as necessary to inform such decisions
(Collins, 2007; Barron, 2009; Doherty et al., 2009; Goddard et al., 2009;
Shukla et al., 2009; Piao et al., 2010; Shapiro et al., 2010). This argument
has been criticized on the grounds that predictions of future climate
impacts are highly uncertain (Dessai and Hulme, 2004; Cox and
Stephenson, 2007; Stainforth et al., 2007; Dessai et al., 2009; Hawkins
and Sutton, 2009; Knutti, 2010) and that predictions are insufficient to
motivate action (Fischhoff, 1994; Sarewitz et al., 2000; Cash et al., 2003,
2006; Rayner et al., 2005; Moser and Luers, 2008; Dessai et al., 2009;
NRC, 2009). Other adaptation literature has emphasized that many
communities do not sufficiently manage current risks and that improving
this situation would go a long way toward preparing them for any
future changes due to climate change (Smit and Wandel, 2006; Pielke et
al., 2007). As discussed in Section 1.4, this approach will in some cases
underestimate the challenges of adapting to future climate change.
To address these challenges, the adaptation literature has increasingly
discussed an iterative risk management framework (Carter et al., 2007;
Jones and Preston, 2011), which is consistent with risk governance as
described earlier in this section. Iterative risk management recognizes
that the process of anticipating and responding to climate change does
not constitute a single set of judgments at some point in time, but rather
an ongoing assessment, action, reassessment, and response that will
continue – in the case of many climate-related decisions – indefinitely
(ACC, 2010). In many cases, iterative risk management contends with
conditions where the probabilities underlying estimates of future risk
are imprecise and/or the structure of the models that relate events to
consequences are under-determined (NRC, 2009; Morgan et al., 2009).
Such deep or severe uncertainty (Lempert and Collins, 2007) can
characterize not only understanding of future climatic events but also
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
48
future patterns of human vulnerability and the capability to respond to
such events. With many complex, poorly understood physical and
socioeconomic systems, research and social learning may enrich
understanding over time, but the amount of uncertainty, as measured
by observers’ ability to make specific, accurate predictions, may grow
larger (Morgan et al., 2009, pp. 114–115; NRC, 2009, pp. 18–19; see
related discussion of ‘surprises’ in Section 3.1.7). In addition, theory and
models may change in ways that make them less, rather than more,
reliable as predictive tools over time (Oppenheimer et al., 2008).
Recent literature has thus explored a variety of approaches that can
help disaster risk management address such uncertainties (McGray et
al., 2007; IIED 2009; Schipper, 2009), in particular approaches that help
support decisions when it proves difficult or impossible to accurately
estimate probabilities of events and their adverse consequences.
Approaches for characterizing uncertainty include qualitative scenario
methods (Parson et al., 2007); fuzzy sets (Chongfu, 1996; El-Baroudy
and Simonovic, 2004; Karimi and Hullermeier, 2007; Simonovic, 2010);
and the use of ranges of values or sets of distributions, rather than single
values or single best-estimate distributions (Morgan et al., 2009; see
also Mastrandrea et al., 2010). Others have suggested managing such
uncertainty with robust policies that perform well over a wide range of
plausible futures (Dessai and Hulme, 2007; Groves and Lempert, 2007;
Brown, 2010; Means et al., 2010; Wilby and Dessai, 2010; Dessai and
Wilby, 2011; Reeder and Ranger, 2011; also see discussion in Chapter 8).
Decision rules based on the concept of robust adaptive policies go
beyond ‘no regrets’ by suggesting how in some cases relatively low-
cost, near-term actions and explicit plans to adjust those actions over
time can significantly improve future ability to manage risk (World
Bank, 2009; Hine and Hall, 2010; Lempert and Groves, 2010; Walker et
al., 2010; Brown, 2011; Ranger and Garbett-Shiels, 2011; see also
Section 1.4.5).
The resilience literature, as described in Chapter 8, also takes an interest
in managing difficult-to-predict futures. Both the adaptation to climate
change and vulnerability literatures often take an actor-oriented view
(Wisner et al., 2004; McLaughlin and Dietz, 2007; Nelson et al., 2007;
Moser 2009) that focuses on particular agents faced with a set of
decisions who can make choices based on their various preferences;
their institutional interests, power, and capabilities; and the information
they have available. Robustness in the adaptation to climate change
context often refers to a property of decisions specific actors may take
(Hallegatte, 2009; Lempert and Groves, 2010; Dessai and Wilby, 2011).
In contrast, the resilience literature tends to take a systems view (Olsson
et al., 2006; Walker et al., 2006; Berkes, 2007; Nelson et al., 2007) that
considers multi-interacting agents and their relationships in and with
complex social, ecological, and geophysical systems (Miller et al., 2010).
These literatures can help highlight for disaster risk management such
issues as the tension between resilience to specific, known disturbances
and novel and unexpected ones (sometimes referred to as the distinction
between ‘specified’ and ‘general’ resilience, Miller et al., 2010), the
tension between resilience at different spatial and temporal scales, and
the tension between the ability of a system to persist in its current state
and its ability to transform to a fundamentally new state (Section 1.4;
Chapter 8; ICSU, 2002; Berkes, 2007).
Disaster risk management will find similarities to its own multi-sector
approach in the adaptation literature’s recent emphasis, consistent with
the concept of climate-related decisions, on climate change as one of
many factors affecting the management of risks. For instance, some
resource management agencies now stress climate change as one of many
trends such as growing demand for resources, environmental constraints,
aging infrastructure, and technological change that, particularly in
combination, could require changes in investment plans and business
models (CCSP, 2008; Brick et al., 2010). It has become clear that many
less-developed regions will have limited success in reducing overall
vulnerability solely by managing climate risk because vulnerability,
adaptive capacity, and exposure are critically influenced by existing
structural deficits (low income and high inequality, lack of access to
health and education, lack of security and political access, etc.). For
example, in drought-ravaged northeastern Brazil, many vulnerable
households could not take advantage of risk management interventions
such as seed distribution programs because they lacked money to travel
to pick up the seeds or could not afford a day’s lost labor to participate
in the program (Lemos, 2003). In Burkina Faso, farmers had limited
ability to use seasonal forecasts (a risk management strategy) because
they lacked the resources (basic agricultural technology such as plows,
alternative crop varieties, fertilizers, etc.) needed to effectively respond
to the projections (Ingram et al., 2002). In Bangladesh, however, despite
persisting poverty, improved disaster preparedness and response and
relative higher levels of household adaptive capacity have dramatically
decreased the number of deaths as a result of flooding (del Ninno et al.,
2002, 2003; Section 9.2.5).
Scholars have argued that building adaptive capacity in such regions
requires a dialectic, two-tiered process in which climatic risk management
(specific adaptive capacity) and deeper-level socioeconomic and political
reform (generic adaptive capacity) iterate to shape overall vulnerability
(Lemos et al., 2007; Tompkins et al., 2008). When implemented as part of
a systems approach, managing climate risks can create positive synergies
with development goals through participatory and transparent
approaches (such as participatory vulnerability mapping or local disaster
relief committees) that empower local households and institutions (e.g.,
Degg and Chester, 2005; Nelson, 2005).
1.3.3. Disaster Risk Management and
Adaptation to Climate Change Share
Many Concepts, Goals, and Processes
The efficacy of the mix of actions used by communities to reduce,
transfer, and respond to current levels of disaster risk could be vastly
increased. Understanding and recognition of the many development-
based instruments that could be put into motion to achieve disaster risk
reduction is a prerequisite for this (Lavell and Lavell, 2009; UNISDR,
2009e, 2011; Maskrey 2011; Wisner et al., 2011). At the same time,
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
49
some aspects of disaster risk will increase for many communities due to
climate change and other factors (Chapters 3 and 4). Exploiting the
potential synergies between disaster risk management and adaptation
to climate change literature and practice will improve management of
both current and future risks.
Both fields share a common interest in understanding and reducing the
risk created by the interactions of human with physical and biological
systems. Both seek appropriate allocations of risk reduction, risk
transfer, and disaster management efforts, for instance balancing pre-
impact risk management or adaptation with post-impact response and
recovery. Decisions in both fields may be organized according to the
risk governance framework. For instance, many countries, are gaining
experience in implementing cooperative, inter-sector and multi- or
interdisciplinary approaches (ICSU, 2002; Brown et al., 2006; McGray et
al., 2007; Lavell and Lavell, 2009). In general, disaster risk management
can help those practicing adaptation to climate change to learn from
addressing current impacts. Adaptation to climate change can help
those practicing disaster risk management to more effectively address
future conditions that will differ from those of today.
The integration of concepts and practices is made more difficult because
the two fields often use different terminology, emerge from different
academic communities, and may be seen as the responsibility of different
government organizations. As one example, Section 1.4 will describe
how the two fields use the word ‘coping’ with different meanings and
different connotations. In general, various contexts have made it more
difficult to recognize that the two fields share many concepts, goals, and
processes, as well as to exploit the synergies that arise from their
differences. These include differences in historical and evolutionary
processes; conceptual and definitional bases; processes of social
knowledge construction and the ensuing scientific compartmentalization
of subject areas; institutional and organizational funding and
instrumental backgrounds; scientific origins and baseline literature;
conceptions of the relevant causal relations; and the relative importance
of different risk factors (see Sperling and Szekely, 2005; Schipper and
Pelling, 2006; Thomalla et al., 2006; Mitchell and van Aalst, 2008;
Venton and La Trobe, 2008, Schipper and Burton, 2009; Lavell, 2010).
These aspects will be considered in more detail in future chapters.
Potential synergies from the fields’ different emphases include the
following.
First, disaster risk management covers a wide range of hazardous
events, including most of those of interest in the adaptation to climate
change literature and practice. Thus, adaptation could benefit from
experience in managing disaster risks that are analogous to the new
challenges expected under climate change. For example, relocation and
other responses considered when confronted with sea level change can
be informed by disaster risk management responses to persistent or
large-scale flooding and landslides or volcanic activity and actions with
pre- or post-disaster relocation; responses to water shortages due to loss
of glacial meltwater would bear similarities to shortages due to other
drought stressors; and public health challenges due to modifications in
disease vectors due to climate change have similarities to those
associated with current climate variability, such as the occurrence of
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
FAQ 1.2 | What are effective strategies for managing disaster risk in a changing climate?
Disaster risk management has historically operated under the premise that future climate will resemble that of the past. Climate change
now adds greater uncertainty to the assessment of hazards and vulnerability. This will make it more difficult to anticipate, evaluate, and
communicate disaster risk. Uncertainty, however, is not a ‘new’ problem. Previous experience with disaster risk management under
uncertainty, or where long return periods for extreme events prevail, can inform effective risk reduction, response, and preparation, as
well as disaster risk management strategies in general.
Because climate variability occurs over a wide range of timescales, there is often a historical record of previous efforts to manage and
adapt to climate-related risk that is relevant to risk management under climate change. These efforts provide a basis for learning via
the assessment of responses, interventions, and recovery from previous impacts. Although efforts to incorporate learning into the
management of weather- and climate-related risks have not always succeeded, such adaptive approaches constitute a plausible model
for longer-term efforts. Learning is most effective when it leads to evaluation of disaster risk management strategies, particularly with
regard to the allocation of resources and efforts between risk reduction, risk sharing, and disaster response and recovery efforts, and
when it engages a wide range of stakeholder groups, particularly affected communities.
In the presence of deeply uncertain long-term changes in climate and vulnerability, disaster risk management and adaptation to climate
change may be advanced by dealing adequately with the present, anticipating a wide range of potential climate changes, and promoting
effective ‘no-regrets’ approaches to both current vulnerabilities and to predicted changes in disaster risk. A robust plan or strategy that
both encompasses and looks beyond the current situation with respect to hazards and vulnerability will perform well over a wide range
of plausible climate changes.
50
El Niño. Moreover, like disaster risk management, adaptation to climate
change will often take place within a multi-hazard locational framework
given that many areas affected by climate change will also be affected
by other persistent and recurrent hazards (Wisner et al., 2004, 2011;
Lavell, 2010; Mercer, 2010). Additionally, learning from disaster risk
management can help adaptation, which to date has focused more on
changes in the climate mean, increasing its focus on future changes in
climate extremes and other potentially damaging events.
Second, disaster risk management has tended to encourage an expanded,
bottom-up, grass roots approach, emphasizing local and community-
based risk management in the framework of national management
systems (see Chapters 5 and 6), while an important segment of the
adaptation literature focuses on social and economic sectors and macro
ecosystems over large regional scales. However, a large body of the
adaptation literature – in both developed and developing countries – is
very locally focused. Both fields could benefit from the body of work on
the determinants of adaptive capacity that focus on the interaction of
individual and collective action and institutions that frame their actions
(McGray et al., 2007; Schipper, 2009).
Third, the current disaster risk management literature emphasizes the
social conditioning of risk and the construction of vulnerability as a causal
factor in explaining loss and damage. Early adaptation literature and
some more recent output, particularly from the climate change field,
prioritizes physical events and exposure, seeing vulnerability as what
remains after all other factors have been considered (O’Brien et al.,
2007). However, community-based adaptation work in developing
countries (Beer and Hamilton, 2002; Brown et al., 2006; Lavell and
Lavell, 2009; UNISDR, 2009b,c) and a growing number of studies in
developed nations (Burby and Nelson, 1991; de Bruin et al., 2009;
Bedsworth and Hanak, 2010; Brody et al., 2010; Corfee-Morlot et al.,
2011; Moser and Eckstrom, 2011) have considered social causation.
Both fields could benefit from further integration of these concepts.
Overall, the disaster risk management and adaptation to climate change
literatures both now emphasize the value of a more holistic, integrated,
trans-disciplinary approach to risk management (ICSU-LAC, 2009).
Dividing the world up sectorally and thematically has often proven
organizationally convenient in government and academia, but can
undermine a thorough understanding of the complexity and interaction
of the human and physical factors involved in the constitution and
definition of a problem at different social, temporal, and territorial
scales. A more integrated approach facilitates recognition of the complex
relationships among diverse social, temporal, and spatial contexts;
highlights the importance of decision processes that employ participatory
methods and decentralization within a supporting hierarchy of higher
levels; and emphasizes that many disaster risk management and other
organizations currently face climate-related decisions whether they
recognize them or not.
The following areas, some of which have been pursued by governments,
civil society actors, and communities, have been recommended or
proposed to foster such integration between, and greater effectiveness
of, both adaptation to climate change and disaster risk management
(see also WRI, 2008; Birkmann and von Teichman, 2010; Lavell, 2010):
Development of a common lexicon and deeper understanding of
the concepts and terms used in each field (Schipper and Burton,
2009)
Implementation of government policymaking and strategy
formulation that jointly considers the two topics
Evolution of national and international organizations and institutions
and their programs that merge and synchronize around the
two themes, such as environmental ministries coordinating with
development and planning ministries (e.g., National Environmental
Planning Authority in Jamaica and Peruvian Ministries of Economy
and Finance, Housing, and Environment)
Merging and/or coordinating disaster risk management and
adaptation financing mechanisms through development agencies
and nongovernmental organizations
The use of participatory, local level risk and context analysis
methodologies inspired by disaster risk management that are now
strongly accepted by many civil society and government agencies
in work on adaptation at the local levels (IFRC, 2007; Lavell and
Lavell, 2009; UNISDR, 2009 b,c)
Implementing bottom-up approaches whereby local communities
integrate adaptation to climate change, disaster risk management,
and other environmental and development concerns in a single,
causally dimensioned intervention framework, commensurate
many times with their own integrated views of their own physical
and social environments (Moench and Dixit, 2004; Lavell and
Lavell, 2009).
1.4. Coping and Adapting
The discussion in this section has four goals: to clarify the relationship
between adaptation and coping, particularly the notion of coping range;
to highlight the role of learning in an adaptation process; to discuss
barriers to successful adaptation and the issue of maladaptation; and
to highlight examples of learning in the disaster risk management
community that have already advanced climate change adaptation.
A key conclusion of this section is that learning is central to adaptation,
and that there are abundant examples (see Section 1.4.5 and Chapter 9)
of the disaster risk management community learning from prior experience
and adjusting its practices to respond to a wide range of existing and
evolving hazards. These cases provide the adaptation to climate change
community with the opportunity not only to study the specifics of learning
as outlined in these cases, but also to reflect on how another community
that also addresses climate-related risk has incorporated learning into
its practice over time.
As disaster risk management includes both coping and adapting, and
these two concepts are central for adaptation to climate change in both
scholarship and practice, it is important to start by clarifying the meanings
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
51
of these terms. Without a clear conception of the distinctions between
the concepts and overlaps in their meanings, it is difficult to fully
understand a wide range of related issues, including those concerned
with the coping range, adaptive capacity, and the role of institutional
learning in promoting robust adaptation to climate change. Clarifying
such distinctions carries operational significance for decisionmakers
interested in promoting resilience, a process that relies on coping for
immediate survival and recovery, as well as adaptation and disaster
risk reduction, which entail integrating new information to moderate
potential future harm.
1.4.1. Definitions, Distinctions, and Relationships
In both the disaster risk management and climate change adaptation
literature, substantial differences are apparent as to the meaning and
significance of coping as well as its relationship with and distinction
from adaptation. Among the discrepancies, for example, some disaster
risk management scholars have referred to coping as a way to engage
local populations and utilize indigenous knowledge in disaster
preparedness and response (Twigg, 2004), while others have critiqued this
idea, concerned that it would divert attention away from addressing
structural problems (Davies, 1993) and lead to a focus on ‘surviving’
instead of ‘thriving.There has also been persistent debate over whether
coping primarily occurs before or after a disastrous event (UNISDR,
2008b,c, 2009e). This debate is not entirely resolved by the current
UNISDR definition of coping, the “ability of people, organizations, and
systems, using available skills and resources, to face and manage
adverse conditions, emergencies or disasters” (UNISDR, 2009d). Clearly,
emergencies and disasters are post facto circumstances, but ‘adverse
conditions’ is an indeterminate concept that could include negative pre-
impact livelihood conditions and disaster risk circumstances or merely
post-impact effects.
The first part of this section is focused on parsing these two concepts.
Once the terms are adequately distinguished, the focus shifts in the
second part to important relationships between the two terms and other
related concepts, which taken together have operational significance for
governments and stakeholders.
1.4.1.1. Definitions and Distinctions
Despite the importance of the term coping in the fields of both disaster
risk management and adaptation to climate change, there is substantial
confusion regarding the term’s meaning (Davies, 1996) and how it is
distinguished from adaptation.
In order to clarify this aspect, it is helpful first to look outside of the
disaster risk and adaptation contexts. The Oxford English Dictionary
defines coping as “the action or process of overcoming a problem or
difficulty” or “managing or enduring a stressful situation or condition”
and adapting as “rendering suitable, modifying” (OED, 1989). As noted
in Table 1-1, contrasting the two terms highlights several important
dimensions in which they differ – exigency, constraint, reactivity, and
orientation – relevant examples of which can be found in the literature
cited.
Overall, coping focuses on the moment, constraint, and survival;
adapting (in terms of human responses) focuses on the future, where
learning and reinvention are key features and short-term survival is less
in question (although it remains inclusive of changes inspired by
already-modified environmental conditions).
1.4.1.2. Relationships between Coping, Coping Capacity,
Adaptive Capacity, and the Coping Range
The definitions of coping and adapting used in this report reflect the
dictionary definitions. As an example, a community cannot adapt its way
through the aftermath of a disastrous hurricane; it must cope instead.
Its coping capacity, or capacity to respond (Gallopín, 2003), is a function
of currently available resources that can be used to cope, and determines
the community’s ability to survive the disaster intact (Bankoff, 2004;
Wisner et al., 2004). Repeated use of coping mechanisms without
adequate time and provisions for recovery can reduce coping capacity
and shift a community into what has been termed transient poverty
(Lipton and Ravallion, 1995). Rather than leaving resources for adaptation,
communities forced to cope can become increasingly vulnerable to
future hazards (O’Brien and Leichenko, 2000).
Adaptation in anticipation of future hurricanes, however, can limit the
need for coping that may be required to survive the next storm. A
community’s adaptive capacity will determine the degree to which
adaptation can be pursued (Smit and Pilofosova, 2003). While there is
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
Dimension Coping Adapting
Exigency
Survival in the face of immediate,
unusually significant stress, when
resources, which may have been
minimal to start with, are taxed
(Wisner et al., 2004).
Reorientation in response to
recent past or anticipated future
change, often without specific
reference to resource limitations.
Constraint
Survival is foremost and tactics are
constrained by available
knowledge, experience, and assets;
reinvention is a secondary concern
(Bankoff, 2004).
Adjustment is the focus and
strategy is constrained less by
current limits than by
assumptions regarding future
resource availability and trends.
Reactivity
Decisions are primarily tactical and
made with the goal of protecting
basic welfare and providing for
basic human security after an event
has occurred (Adger, 2000).
Decisions are strategic and
focused on anticipating change
and addressing this proactively
(Füssel, 2007), even if spurred by
recent events seen as harbingers
of further change.
Orientation
Focus is on past events that shape
current conditions and limitations;
by extension, the focus is also on
previously successful tactics
(Bankoff, 2004).
Focus on future conditions and
strategies; past tactics are
relevant to the extent they might
facilitate adjustment, though
some experts believe past and
future orientation can overlap
and blend (Chen, 1991).
Table 1-1 | The various dimensions of coping and adapting.
52
some variability in how coping capacity and adaptive capacity are
defined, the literature generally recognizes that adaptive capacity
focuses on longer-term and more sustained adjustments (Gallopín, 2006;
Smit and Wandel, 2006). However, in the same way that repeatedly
invoking coping mechanisms consumes resources available for subsequent
coping needs, it also consumes resources that might otherwise be
available for adaptation (Adger, 1996; Risbey et al., 1999).
There is also a link between adaptation and the coping range – that
is, a system’s capacity to reactively accommodate variations in climatic
conditions and their impacts (a system can range from a particular
ecosystem to a society) (IPCC, 2007b). In the adaptation literature,
Yohe and Tol (2002, p. 26) have used the term to refer to the range of
“circumstances within which, by virtue of the underlying resilience of
the system, significant consequences are not observed” in response to
external stressors. Outside the coping range, communities will “feel
significant effects from change and/or variability in their environments”
(Yohe and Tol, 2002, p. 25). Within its coping range, a community
can survive and even thrive with significant natural hazards. This is
particularly the case when the historical distribution of hazard intensity
is well known and relatively stable (see Section 1.2.3.4). A community’s
coping range is determined, in part, by prior adaptation (Hewitt and
Burton, 1971; de Vries, 1985; de Freitas, 1989), and a community is most
likely to survive and thrive when adaptation efforts have matched its
coping range with the range of hazards it typically encounters (Smit and
Pilifosova, 2003). As climate change alters future variability and the
occurrence of extreme events, and as societal trends change human
systems’ vulnerability, adaptation is required to adjust the coping range
so as to maintain societal functioning within an expected or acceptable
range of risk (Moser and Luers, 2008).
Box 1-4 provides an example of this process in the region that is now
The Netherlands. As this box illustrates, the process of shifting a society’s
coping range both depends on and facilitates further economic
development (i.e., requires adaptive capacity and enhances coping
capacity). The box also illustrates that the process requires continuous
reassessment of risk and adjustment in response to shifting hazard
distributions in order to avoid increasing, and maladaptive, hazard
exposure. Successful adjustments, facilitated in part by institutional
learning, can widen and shift a community’s coping range, promoting
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
Box 1-4 | Adaptation to Rising Levels of Risk
Before AD 1000, in the low-lying coastal floodplain of the southern North Sea and around the Rhine delta, the area that is now The
Netherlands, the inhabitants lived on dwelling mounds, piled up to lie above the height of the majority of extreme storm surges. By the
10th century, with a population estimated at 300,000 people, inhabitants had begun to construct the first dikes, and within 400 years
had ringed all significant areas of land above spring tide, allowing animals to graze and people to live in the protected wetlands. The
expansion of habitable land encouraged a significant increase in the population exposed to catastrophic floods (Borger and Ligtendag,
1998). The weak sea dikes broke in a series of major storm surge floods through the stormy 13th and 14th centuries (in particular in
1212, 1219, 1287, and 1362), flooding enormous areas (often permanently) and causing more than 200,000 fatalities, reflecting an
estimated lifetime mortality rate from floods for those living in the region in excess of 5% (assuming a 30-year average lifespan;
Gottschalk, 1971, 1975, 1977).
To adapt to increasingly adverse environmental conditions (reflecting long-term delta subsidence), major improvements in the technology
of dike construction and drainage engineering began in the 15th century. As the country became richer and population increased (to an
estimated 950,000 by 1500 and 1.9 million by 1700), it became an imperative not only to provide better levels of protection but also to
reclaim land from the sea and from the encroaching lakes, both to reduce flood hazard and expand the land available for food production
(Hoeksma, 2006). Examples of the technological innovations included the development of windmills for pumping, and methods to lift
water at least 4 m whether by running windmills in series or through the use of the wind-powered Archimedes screw. As important was
the availability of capital to be invested in joint stock companies with the sole purpose of land reclamation. In 1607, a company was
formed to reclaim the 72 km
2
Beemster Lake north of Amsterdam (12 times larger than any previous reclamation). A 50-km canal and
dike ring were excavated, a total of 50 windmills installed that after five years pumped dry the Beemster polder, 3 to 4 m below the
surrounding countryside, which, within 30 years, had been settled by 200 farmhouses and 2,000 people.
After the major investment in raising and strengthening flood defenses in the 17th century, there were two or three large floods, one in
1717 (when 14,000 people drowned) and two notable floods in 1825 and 1953; since that time the average flood mortality rate has
been around 1,000 per century, equivalent to a lifetime mortality rate (assuming a 50-year average lifetime) of around 0.01%, 500 times
lower than that which had prevailed through the Middle Ages (Van Baars and Van Kempen, 2009). This change reflects increased protection
rather than any reduction in storminess. The flood hazard and attendant risk is now considered to be rising again (Bouwer and Vellinga,
2007) and plans are being developed to manage further rises, shifting the coping range in anticipation of the new hazard distribution.
53
resilience to a wider range of future disaster risk (Yohe and Tol, 2002),
as illustrated in Box 1-4 and discussed further in Section 1.4.2 (high
confidence).
1.4.2. Learning
Risk management decisions are made within social-ecological systems
(a term referring to social systems intimately tied to and dependent on
environmental resources and conditions). Some social-ecological systems
are more resilient than others. The most resilient are characterized by
their capacity to learn and adjust, their ability to reorganize after
disruption, and their retention of fundamental structure and function in
the face of system stress (Folke, 2006). The ability to cope with extreme
stress and resume normal function is thus an important component of
resilience, but learning, reorganizing, and changing over time are also
key. As Chapter 8 highlights, transformational changes are required to
achieve a future in which society’s most important social-ecological
systems are sustainable and resilient. Learning, along with adaptive
management, innovation, and leadership, is essential to this process.
Learning related to social-ecological systems requires recognizing
their complex dynamics, including delays, stock-and-flow dynamics,
and feedback loops (Sterman, 2000), features that can complicate
management strategies by making it difficult to perceive how a system
operates. Heuristic devices and mental models can sometimes inhibit
learning by obscuring a problem’s full complexity (Kahneman et al., 1982;
Section 1.3.1.2) and complicating policy action among both experts and
lay people (Cronin et al., 2009). For instance, common heuristics (see
Section 1.3.1.2) lead to misunderstanding of the relationship between
greenhouse gas emission rates and their accumulation in atmospheric
stocks, lending credence to a ‘wait and see’ approach to mitigation
(Sterman, 2008). Through a variety of mechanisms, such factors can lead
to paralysis and failure to engage in appropriate risk management
strategies despite the availability of compelling evidence pointing to
particular risk management pathways (Sterman, 2006). The resulting
learning barriers thus deserve particular attention when exploring how
to promote learning that will lead to effective adaptation.
Given the complex dynamics of social-ecological systems and their
interaction with a changing climate, the literature on adaptation to climate
change (usually referred to here, as above, simply as ‘adaptation’)
emphasizes iterative learning and management plans that are explicitly
designed to evolve as new information becomes available (Morgan et
al., 2009: NRC, 2009). Unlike adaptation, the field of disaster risk
management has not historically focused as explicitly on the implications
of climate change and the need for iterative learning. However, the
field provides several important examples of learning, including some
presented in Chapter 9, that could be instructive to adaptation
practitioners. Before introducing these case studies in Section 1.4.5, we
will outline relevant theory of institutional learning and ‘learning loops.
Extensive literature explores both the role of learning in adaptation
(Armitage et al., 2008; Moser, 2010; Pettengell, 2010) and strategies for
facilitating institutional and social learning in ‘complex adaptive systems’
(Pahl-Wostl, 2009). Some important strategies include the use of
knowledge co-production, wherein scientists, policymakers, and other
actors work together to exchange, generate, and apply knowledge
(van Kerkhoff and Lebel, 2006), and action research, an iterative process
in which teams of researchers develop hypotheses about real-world
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
CONTEXT FRAMES
Single-Loop Learning
Double-Loop Learning
Triple-Loop Learning
Reacting
Reframing
Transforming
ACTIONS OUTCOMES
What strategies might facilitate more
effective future transboundary flood
management?
• How should vulnerability to other
climate change impacts be included in
flood management planning?
• Should dike height be increased by 10
or 20 cm?
• Should resources be allocated toward
protecting existing populations and
infrastructure at increasing risk in a
changing climate, or should these assets
be relocated or abandoned once certain
risk thresholds are crossed?
Figure 1-3 | Learning loops: pathways, outcomes, and dynamics of single-, double-, and triple-loop learning and applications to flood management. Adapted from Argyris and
Schön, 1978; Hargrove, 2002; Sterman et al., 2006; Folke et al., 2009; and Pahl-Wostl, 2009.
54
problems and revise management strategies based on the results
(List, 2006). Prior work on learning theories, for example, experiential
learning (Kolb, 1984) and transformative learning (Mezirow, 1995),
emphasize the importance of action-oriented problem-solving, learning-
by-doing, concrete learning cycles, and how these processes result in
reflection, reconsideration of meaning, and re-interpretation of value
structures. The learning loop framework (Kolb and Fry, 1975; Argyris and
Schön, 1978; Keen et al., 2005) integrates these theories and divides
learning processes into three different loops depending on the degree
to which the learning promotes transformational change in management
strategies. Figure 1-3 outlines this framework and its application to the
issue of flood management.
In single-loop learning processes, changes are made based on the
difference between what is expected and what is observed. Single-loop
learning is primarily focused on improving the efficiency of action
(Pelling et al., 2008) and answering the question of “whether things are
being done right” (Flood and Romm, 1996), that is, whether management
tactics are appropriate or adequate to achieve identified objectives. In
flood management, for example, when floodwaters threaten to breach
existing flood defenses, flood managers may ask whether dike and
levee heights are sufficient and make adjustments accordingly. As
Figure 1-3 indicates, single-loop learning focuses primarily on actions;
data are integrated and acted on but the underlying mental model used
to process the data is not changed.
In double-loop learning, the evaluation is extended to assess whether
actors are “doing the right things” (Flood and Romm, 1996), that is,
whether management goals and strategies are appropriate. Corrective
actions are made after the problem is reframed and different management
goals are identified (Pelling et al., 2008); data are used to promote
critical thinking and challenge underlying mental models of what works
and why. Continuing with the flood management example, double-loop
learning results when the goals of the current flood management
regime are critically examined to determine if the regime is sustainable
and resilient to anticipated shifts in hydrological extremes over a
particular time period. For instance, in a floodplain protected by levees
built to withstand a 500-year flood, a shift in the annual exceedance
probability from 0.002 to 0.005 (equivalent to stating that the likelihood
that a 500-year flood will occur in a given year has shifted to that seen
historically for a 200-year event) will prompt questions about whether
the increased likelihood of losses justifies different risk management
decisions, ranging from increased investments in flood defenses to
changed insurance policies for the vulnerable populations.
Many authors also distinguish triple-loop learning (Argyris and Schön,
1978; Hargrove, 2002; Peschl, 2007), or learning that questions deeply
held underlying principles (Pelling et al., 2008). In triple-loop learning,
actors question how institutional and other power relationships determine
perceptions of the range of possible interventions, allowable costs, and
appropriate strategies (Flood and Romm, 1996). In response to evidence
that management strategies are not serving a larger agreed-upon goal,
that is, they are maladaptive, triple-loop learning questions how the
social structures, cultural norms, dominant value structures, and other
constructs that mediate risk and risk management (see Box 1-3) might
be changed or transformed. Extending the flood control example, triple-
loop learning might entail entirely new approaches to governance and
participatory risk management involving additional parties, crossing
cultural, institutional, national, and other boundaries that contribute
significantly to flood risk, and planning aimed at robust actions instead of
strategies considered optimal for particular constituents (Pahl-Wostl, 2009).
Different types of learning are more or less appropriate in given
circumstances (Pahl-Wostl, 2009, p. 359). For example, overreliance
on single-loop learning may be problematic in rapidly changing
circumstances. Single-loop learning draws on an inventory of existing
skills and memories specific to particular circumstances. As a result,
rapid, abrupt, or surprising changes may confound single-loop learning
processes (Batterbury, 2008). Coping mechanisms, even those that
have developed over long periods of time and been tested against
observation and experience, may not confer their usual survival
advantage in new contexts. Double- and triple-loop learning are better
suited to matching coping ranges with new hazard regimes (Yohe and
Tol, 2002). Integrating double- and triple-loop learning into adaptation
projects, particularly for populations exposed to multiple risks and
stressors, is more effective than more narrowly planned approaches
dependent on specific future climate information (McGray et al., 2007;
Pettengell, 2010).
Easier said than done, triple-loop learning is analogous to what some
have termed ‘transformation’ (Kysar, 2004; see Section 1.1.3; Chapter
8), in that it can lead to recasting social structures, institutions, and
constructions that contain and mediate risk to accommodate more
fundamental changes in world view (Pelling, 2010). Translating double-
and triple-loop learning into policy requires not only articulation of
a larger risk-benefit universe, but also mechanisms to identify,
account for, and compare the costs associated with a wide range of
interventions and their benefits and harms over various time horizons.
Stakeholders would need also to collaborate to an unusual degree in
order to collectively and cooperatively consider the wide range of risk
management possibilities and their impacts.
1.4.3. Learning to Overcome Adaptation Barriers
Learning focused on barriers to adaptation can be particularly useful.
Resource limitations are universally noted as a significant impediment in
pursuing adaptation strategies, to a greater or lesser degree depending
on the context. In addition, some recent efforts to identify and categorize
adaptation barriers have focused on specific cultural factors (Nielsen
and Reenberg, 2010) or issues specific to particular sectors (Huang et
al., 2011), while others have discussed the topic more comprehensively
(Moser and Ekstrom, 2010). Some studies identify barriers in the specific
stages of the adaptation process. Moser and Ekstrom (2010), for instance,
outline three phases of adaptation: understanding, planning, and
management. Each phase contains several key steps, and barriers can
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
55
impede progress at each. Barriers to understanding, for instance, can
include difficulty recognizing a changing signal due to difficulty with its
detection, perception, and appreciation; preoccupation with other
pressing concerns that divert attention from the growing signal; and
lack of administrative and social support for making adaptive decisions.
While this study offers a diagnostic framework and avoids prescriptions
about overcoming adaptation barriers, other studies, such as those
mentioned above, offer more focused prescriptions relevant to particular
sectors and contexts.
Research on barriers has generally focused on adaptation as a process,
recognizing the difficulty in furnishing a universally acceptable a priori
definition of successful adaptation outcomes (Adger et al., 2005). This
skirts potentially important normative questions, however, and some
researchers have considered whether particular activities should be
considered maladaptive, defined as an “action taken ostensibly to avoid
or reduce vulnerability to climate change that impacts adversely on,
or increases the vulnerability of other systems, sectors, or social
groups” (Barnett and O’Neill, 2009, p. 211). They identify activities
that increase greenhouse gas releases, burden vulnerable populations
disproportionately, and require excessive commitment to one path of
action (Barnett and O’Neill, 2009). Other candidates include actions that
offset one set of risks but increase others, resulting in net risk increase,
for example, a dam that reduces flooding but increases the threat of
zoonotic diseases, and actions that amplify risk to those who remain
exposed (or are newly exposed as a result of a maladaptive action), of
which there are abundant examples in the public health literature
(Sterman, 2006) and other fields.
These issues have a long history in disaster risk management. For instance,
in 1942, deriving from study and work in the 1930s, Gilbert White asserted
that levees can provide a false sense of security and are eventually
fallible, ultimately leading to increased risk, and advocated, among
other ‘adjustment’ measures, land use planning and environmental
management schemes in river basins in order to face up to flooding
hazards (see Burton et al., 1978). Such findings are among the early
advances in the field of ‘human adjustment to hazards,’ which derived
from an ecological approach to human-environmental relationships. In
the case of levees for example, the distinction between adaptive and
maladaptive actions depends on the time period over which risks are
being assessed. From a probabilistic perspective, the overall likelihood
of a catastrophic flood overwhelming a levee’s protective capacity is a
function of time. The wrinkle that climate change introduces is that
many climate-related hazards may become more frequent, shrinking the
timescale over which certain decisions can be considered ‘adaptive’ and
communities can consider themselves ‘adapted’ (Nelson et al., 2007).
While frameworks that help diagnose barriers to adaptation are helpful
in identifying the origin of maladaptive decisions, crafting truly adaptive
policies is still difficult even when the barriers are fully exposed. For
instance, risk displacement is a common concern in large insurance
systems when risk is not continuously reassessed, risk management
strategies and mechanisms for distributing risk across populations (such
as risk pricing in insurance schemes) are inadequately maintained, or if
new risk management strategies are not recruited as necessary. This
was the case with the levees in New Orleans prior to Hurricane Katrina,
wherein the levees were built to make a hazardous area safer but
paradoxically facilitated the exposure of a much larger population to a
large hazard. As a result of multiple factors (Burby, 2006), inadequate
levee infrastructure increased the likelihood of flooding but no other
adequate risk reduction and management measures were implemented,
resulting in catastrophic loss of life and property when the city was hit
with the surge from a strong Category 3 storm (Comfort, 2006). Some have
suggested that, as a result of the U.S. federal government’s historical
approach to disasters, those whose property was at risk in New Orleans
anticipated that they would receive federal recovery funds in the event of
a flooding disaster. This, in turn, may have distorted the risk management
landscape, resulting in improper pricing of flooding risks, decreased
incentives to take proper risk management actions, and exposure of a
larger population to flood risk than might otherwise have been the case
(Kunreuther, 2006).
This example illustrates how an adaptation barrier may have resulted in
an ultimately maladaptive risk management regime, and demonstrates
the importance of considering how risk, in practice, is assumed and
shared. One goal of risk sharing is to properly price risk so that, in the
event risk is realized, there is an adequate pool of capital available to
fund recovery. When risk is improperly priced and risk sharing is not
adequately regulated, as can occur when risk-sharing devices are not
monitored appropriately, an adequate pool of reserves may not
accumulate. When risk is realized, the responsibility for funding the
recovery falls to the insurer of last resort, often the public.
The example also illustrates how an insurance system designed to
motivate adaptation (by individual homeowners or flood protection
agencies) can function properly only if technical rates – rates that properly
reflect empirically determined levels of risk – can be established and
matched with various levels of risk at a relatively high level of spatial
and temporal resolution. Even in countries with free-market flood
insurance systems, insurers may be reluctant to charge the full technical
rate because consumers have come to assume that insurance costs
should be relatively consistent in a given location. Without charging
technical rates, however, it is difficult to use pricing to motivate adaptation
strategies such as flood proofing or elevating the ground floor of a new
development (Lamond et al., 2009), restricting where properties can be
built, or justifying the construction of communal flood defenses. In such
a case, barriers to adaptation (in both planning and management, in
this case) can result in a strategy with maladaptive consequences in the
present. In places where risk levels are rising due to climate change
under prevailing negative conditions of exposure and vulnerability,
reconsideration of these barriers – a process that includes double- and
triple-loop learning – could promote more adaptive risk management.
Otherwise, maladaptive risk management decisions may commit collective
resources (public or private) to coping and recovery rather than successful
adaptation and may force some segments of society to cope with
disproportionate levels of risk.
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
56
1.4.4. ‘No Regrets,’ Robust Adaptation, and Learning
The mismatch between adaptation strategies and projected needs has
been characterized as the potential for regret, that is, opportunity costs
associated with decisions (and related path dependence, wherein earlier
choices constrain future circumstances and decisions) that are optimal
for one or a small number of possible climate futures but not necessarily
robust over a wider range of scenarios (Lempert and Schlesinger, 2001).
‘No regrets’ adaptation refers to decisions that have net benefits over
the entire range of anticipated future climate and associated impacts
(Callaway and Hellmuth, 2007; Heltberg et al., 2009).
To address the challenge of risk management in the dynamically
complex context of climate change and development, as well as under
conditions where probabilistic estimates of future climatic conditions
remain imprecise, several authors have advanced the concept of
robustness (Wilby and Dessai, 2010), of which ‘no regrets’ adaptation is
a special case (Lempert and Groves, 2010). Robustness is a property of
a plan or strategy that performs well over a wide range of plausible
future scenarios even if it does not perform optimally in any particular
scenario. Robust adaptation plans may perform relatively well even if
probabilistic assessments of risk prove wrong because they aim to
address both expected and surprising changes, and may allow diverse
stakeholders to agree on actions even if they disagree about values and
expectations (Brown and Lall, 2006; Dessai and Hulme; 2007; Lempert
and Groves, 2010; Means et al., 2010; see also Section 1.3.2).
As Section 1.4.3 highlights, currently, in many instances risks associated
with extreme weather and other climate-sensitive hazards are often not
well managed. To be effective, adaptation would prioritize measures
that increase current as well as future resilience to threats. Robustness
over time would increase if learning were a central pillar of adaptation
efforts, including learning focused on addressing current vulnerabilities
and enhancing current risk management (high confidence). Single-,
double-, and triple-loop learning will all improve the efficacy of
management strategies.
The case studies in Chapter 9 highlight some important examples of
learning in disaster risk management relevant to a wide range of climate-
sensitive threats and a variety of sectors. Section 9.2 provides examples
of how single- and double-loop learning processes – enhancing public
health response capacity, augmenting early warning systems, and
applying known strategies for protecting health from the threat of
extreme heat in new settings – had demonstrable impacts on heat-
related mortality, quickly shifting a region’s coping range with regard to
extreme heat (Section 9.2.1). Other case studies, examining risk transfer
(Section 9.2.13) and early warning systems (Section 9.2.11), provide
instances of how existing methods and tools can be modified and
deployed in new settings in response to changing risk profiles – examples
of both double- and triple-loop learning. Similarly, the case studies on
governance (Section 9.2.12) and on the limits to adaptation in small
island developing states (Section 9.2.9) provide examples of third-loop
learning and transformative approaches to disaster risk management.
References
A digital library of non-journal-based literature cited in this chapter that
may not be readily available to the public has been compiled as part of
the IPCC review and drafting process, and can be accessed via either the
IPCC Secretariat or IPCC Working Group II web sites.
ACC, 2010: Informing an Effective Response to Climate Change. America’s Climate
Choices, National Academies Press, Washington, DC.
Adger, W.N., 1996: Approaches to Vulnerability to Climate Change. CSERGE Working
Papers. University of East Anglia, Norwich, UK.
Adger, W.N., 2000: Social and ecological resilience: Are they related? Progress in
Human Geography, 24(3), 347-364.
Adger, W.N., 2006: Vulnerability. Global Environmental Change, 16, 268-281.
Adger, W.N., N. Arnell, and E.M. Thompkins, 2005: Successful adaptation to climate
change across scales. Global Environmental Change, 15, 77-86.
Alexander, D., 1993: Natural Disasters. UCL Press, London, 632 pp.
Alexander, D., 2000: Confronting Catastrophe. Oxford University Press, New York.
Anderson, M. and P. Woodrow, 1989: Rising from the Ashes: Development
Strategies in Times of Disasters. Westview Press, Boulder, CO.
APA, 2009: Psychology and Global Climate Change: Addressing a Multi-faceted
Phenomenon and Set of Challenges. American Psychological Association Task
Force on the Interface between Psychology and Global Climate Change,
American Psychological Association, Washington, DC.
Argyris, C. and D. Schön, 1978: Organizational Learning: A Theory of Action
Perspective. Addison-Wesley, Reading, MA.
Armitage, D., M. Marschke, and R. Plummer, 2008: Adaptive co-management and
the paradox of learning. Global Environmental Change, 18, 86-98.
Aven, T., 2011. On some recent definitions and analysis frameworks for risk,
vulnerability, and resilience. Risk Analysis, 31(4), 515-522.
Bahadur, A.V., M. Ibrahim, and T. Tanner, 2010: The resilience renaissance?
Unpacking of Resilience for Tackling Climate Change and Disasters. Institute of
Development Studies (for the Strengthening Climate Resilience (SCR) consortium),
Brighton, UK.
Baird, A., P. O’Keefe, K. Westgate, and B. Wisner, 1975: Towards an Explanation of and
Reduction of Disaster Proneness. Occasional Paper number 11, Disaster
Research Unit, University of Bradford, Bradford, UK.
Balamir, M., 2005: Ways of understanding urban earthquake risks. In: Book of
Abstracts from ‘Rethinking Inequalities,7th Conference of the European
Sociological Association, Institute of Sociology, Nicolaus Copernicus University
of Torun, Poland, p. 132.
Ball, N., 1975: The myth of the natural disaster. The Ecologist, 5(10), 368-369.
Bankoff, G., 2001: Rendering the world safe: vulnerability as western discourse.
Disasters, 25
(10), 19-35.
Bankoff, G., 2004: The historical geography of disaster: “vulnerability” and “local
knowledge” in western discourse. In: Mapping Vulnerability: Disasters,
Development, and People [G. Bankoff, G. Frerks, and D. Hillhorst (eds.)].
Earthscan, London, pp. 25-36.
Barke, R, H. Jenkins-Smith, and P. Slovic, 1997: Risk perceptions of men and women
scientists. Social Science Quarterly, 78, 167-176.
Barnett, J. and S. O’Neill, 2009: Maladaptation. Global Environmental Change, 20,
211-213.
Barron, E.J., 2009: Beyond climate science. Science, 326, 643.
Batterbury, S., 2008: Anthropology and global warming: the need for environmental
engagement. Australian Journal of Anthropology, 1, 62-67.
Bedford, T.J. and R.M. Cooke, 2001: Probabilistic Risk Analysis: Foundations and
Methods. Cambridge University Press, New York, NY and Cambridge, UK.
Bedsworth, L.W. and E. Hanak, 2010: Adaptation to climate change: A review of
challenges and tradeoffs in six areas. Journal of the American Planning
Association, 76(4), 477-495.
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
57
Beer, T. and R. Hamilton, 2002: Natural Disaster Reduction: Safer Sustainable
Communities, Making Better Decisions about Risk. International Council for
Science, ICSU Position Paper.
Below, R., A. Wirtz and D. Guha-Sapir, 2009: Disaster Category Classification and
Peril Terminology for Operational Purposes. Common Accord Centre for
Research on the Epidemiology of Disasters (CRED) and Munich Re, Brussels,
Belgium and Munich, Germany, cred.be/sites/default/files/DisCatClass_264.pdf.
Berkes, F., 2007: Understanding uncertainty and reducing vulnerability: Lessons from
resilience thinking. Natural Hazards, 41, 283-295.
Berkes, F., J. Colding, and C. Folke (eds.), 2004: Navigating Social-Ecological Systems:
Building Resilience for Complexity and Change. Cambridge University Press,
Cambridge, UK.
Birkmann, J., and K. von Teichman, 2010: Integrating disaster risk reduction and
climate change adaptation: key challenges – scales, knowledge, and norms.
Sustainability Science, 5(2), 171-184.
Blaikie, P., T. Cannon, I. Davis, and B. Wisner, 1994: At Risk: Natural Hazards, People’s
Vulnerability and Disasters. 1st edition. Routledge, London.
Borger, G.J. and W.A. Ligtendag, 1998: The role of water in the development of the
Netherlands – a historical perspective. Journal of Coastal Conservation, 4, 109-
114.
Bosher, L.S. (ed.), 2008: Hazards and the Built Environment: Attaining Built-in
Resilience. Taylor and Francis, London.
Bouwer, L.M. and Vellinga P., 2007: On the flood risk in the Netherlands. Advances
in Natural and Technological Hazard Research, 25, 469-484.
Brand, F.S., and K. Jax, 2007: Focusing the meaning(s) of resilience: resilience as a
descriptive concept and a boundary object. Ecology and Society, 12(1), 23,
www.ecologyandsociety.org/vol12/iss21/art23/.
Brick, T., J. Kightlinger, and D. Mann, 2010: Integrated Water Resources Plan 2010
Update. Report No. 1373, Metropolitan Water District of Southern California,
Los Angeles, CA.
Brody, S., H. Grover, E. Lindquist and A. Vedlitz, 2010: Examining climate change
mitigation and adaptation behaviours among public sector organisations in the
USA. Local Environment: The International Journal of Justice and Sustainability,
15(6), 591-603.
Brown, C., 2010: The end of reliability. Journal of Water Resources Planning and
Management. March/April, 143-145.
Brown, C., 2011: Decision-scaling for Robust Planning and Policy under Climate
Uncertainty. World Resources Report, World Resources Institute, Washington,
DC, www.worldresourcesreport.org/files/wrr/papers/wrr_brown_uncertainty.pdf.
Brown, C. and U. Lall, 2006: Water and economic development: The role of
variability and a framework for resilience. Natural Resources Forum, 30(4),
306-317.
Brown, O., A. Crawford, and A. Hammill, 2006: Natural Disaster and Resource Rights:
Building Resilience, Rebuilding Lives. International Institute for Sustainable
Development, Winnipeg, Canada.
Burby, R., 2006: Hurricane Katrina and the paradoxes of government disaster policy:
Bringing about wise governmental decisions for hazardous areas. Annals of the
American Academy of Political and Social Science, 604, 171-191.
Burby, R.J. and A.C. Nelson, 1991: Local government and public adaptation to sea-
level rise. Journal of Urban Planning and Development, 117(4), 140-163.
Burton, I., R. Kates, and G. White (eds.), 1978: The Environment as Hazard. Guildford,
New York, NY.
Butts, R.A., R.A. Schaalje, and G. Bruce, 1997: Impact of subzero temperatures on
survival, longevity, and natality of adult Russian wheat aphid (Homoptera:
Aphididae). Environmental Entomology, 26(3), 661-667.
Callaway, J. and M. Hellmuth, 2007: Assessing the incremental benefits and costs of
coping with development pressure and climate change: A South African case
study. Submissions from admitted non-governmental organizations on socio-
economic information under the Nairobi work programme. United Nations,
Geneva, Switzerland.
Canham, C.D., M.J. Papaik, and E.F. Latty, 2001: Interspecific variation in susceptibility
to windthrow as a function of tree size and storm severity for northern temperate
tree species. Canadian Journal of Forest Research,31(1), 1-10.
Cannon, S.H, R.M. Kirkham, and M. Parise, 2001: Wildfire-related debris-flow
initiation processes, Storm King Mountain, CO. Geomorphology, 39, 171-188.
Cannon, T., 2006: Vulnerability analysis, livelihoods and disasters. In: Risk21. Coping
with Risks due to Natural Hazards in the 21st Century [Amman, W., S.
Dannenmann, and L. Vulliet (eds.)]. Taylor and Francis Group, London, UK, pp.
41-50.
Cardona, O.D., 1986: Estudios de vulnerabilidad y evaluación del riesgo sísmico:
Planificación física y urbana en áreas propensas. Boletín Técnico de la
Asociación Colombiana de Ingeniería Sismica, 33(2), 32-65.
Cardona, O.D. 1996: Manejo ambiental y prevención de desastres: dos temas
asociados. In: Ciudades en Riesgo [Fernandez, M.A. (ed.)]. La RED-USAID, Lima,
Peru, pp. 79-101.
Cardona, O.D., 2001: Estimación Holística del Riesgo Sísmico utilizando Sistemas
Dinámicos Complejos. Doctoral dissertation, Department of Terrain
Engineering, Technical University of Catalonia, Spain,
www.desenredando.org/public/varios/2001/ehrisusd/index.html.
Cardona, O.D., 2004: The Need for rethinking the concepts of vulnerability and risk
from a holistic perspective: A necessary review and criticism for effective risk
management. In: Mapping Vulnerability: Disa sters, Development and People
[Bankoff, G., G. Frerks, and D. Hilhorst (eds.)]. Earthscan Publishers, London, UK,
pp. 37-51.
Cardona, O.D., 2005. Indicators of Disaster Risk and Risk Management: Program for
Latin America and the Caribbean – Summary Report. English and Spanish
edition, Inter-American Development Bank, Washington, DC.
Cardona, O.D., 2008. Indicators of Disaster Risk and Risk Management: Program for
Latin America and the Caribbean – Summary Report – Second Edition. English
and Spanish edition, Inter-American Development Bank, Washington, DC.
Cardona, O.D., 2011: Disaster risk and vulnerability: Notions and measurement of
human and environmental insecurity. In: Coping with Global Environmental
Change, Disasters and Security – Threats, Challenges, Vulnerabilities and Risks
[Brauch, H.G., U. Oswald Spring, C. Mesjasz, J. Grin, P. Kameri-Mbote, B.
Chourou, P. Dunay, and J. Birkmann (eds.)]. Hexagon Series on Human and
Environmental Security and Peace, Vol. 5, Springer Verlag, Heidelberg, Berlin
and New York, pp. 107-121.
Carter, T.R., R.N. Jones, X. Lu, S. Bhadwal, C. Conde, L.O. Mearns, B.C. O’Neill, M.D.A.
Rounsevell, and M.B. Zurek, 2007: New Assessment Methods and the
Characterisation of Future Conditions. In: Climate Change 2007: Impacts,
Adaptation and Vulnerability. Contribution of Working Group II to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change [Parry,
M.L., O.F. Canziani, J.P. Palutikof, P.J. van der Linden, and C.E. Hanson, (eds.)].
Cambridge University Press, Cambridge, UK, pp. 133-171.
Cash, D.W., W.C. Clark, F. Alcock, N.M. Dickson, N. Eckley, D.H. Guston, J. Jager, and
R.B. Mitchell, 2003: Knowledge systems for sustainable development.
Proceedings of the National Academy of Sciences, 100, 8086-8091.
Cash, D.W., J.C. Borck, and A.G. Patt, 2006: Countering the loading-dock approach to
linking science and decision making: Comparative analysis of El Niño/Southern
Oscillation (ENSO) forecasting systems. Science, Technology and Human Values,
31, 465-494.
Cavazos, T., C. Turrent, and D. P. Lettenmaier, 2008: Extreme precipitation trends
associated with tropical cyclones in the core of the North American monsoon.
Geophysical Research Letters, 35, L21703, doi:10.1029/2008GL035832.
CCSP, 2008: Preliminary Review of Adaptation Options for Climate-Sensitive
Ecosystems and Resources [Julius, S.H. and J.M. West (eds.)]. A Report by the
U.S. Climate Change Science Program and the Subcommittee on Global Change
Research, U.S. Environmental Protection Agency, Washington, DC, 873 pp.
Chen, M. (ed.), 1991: Coping with Seasonality and Drought. Sage Publications, New
Delhi, India.
Chongfu, H., 1996: Fuzzy risk assessment of urban natural hazards. Fuzzy Sets and
Systems, 83, 271-282.
Coburn, A. and R. Spence, 2002: Earthquake Protection. 2nd edition. John Wiley and
Sons, Chichester, UK.
Cochrane, H., 2004: Economic loss: Myth and measurement. Disaster Prevention
and Management, 13, 290-296.
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
58
Collins, M., 2007: Ensembles and probabilities: A new era in the prediction of climate
change. Philosophical Transactions of the Royal Society A, 365, 1957-1970.
Colls, A., N. Ash, and N. Ikkala, 2009: Ecosystem-based Adaptation: a natural
response to climate change. IUCN, Gland, Switzerland, 16 pp.
Comfort, L., 2006: Cities at risk: Hurricane Katrina and the drowning of New
Orleans. Urban Affairs Review, 41(4), 501-516.
Corfee-Morlot, J., I. Cochran, S. Hallegatte, and P.-J. Teasdale, 2011: Multilevel risk
governance and urban adaptation policy. Climatic Change, 104(1) (Special
Issue: Understanding Climate Change Impacts, Vulnerability and Adaptation at
City Scale), 169-197.
Corporación OSSO, La Red de Estudios Sociales en Prevención de Desastres en
América Latina, 2010: .DesInventar online 8.1.9-2: sistema de inventario de
efectos de desastres. OSSO/La Red de Estudios Sociales en Prevención de
Desastres en América Latina, Cali, Columbia, online.desinventar.org/.
Cox, P. and D. Stephenson, 2007: A changing climate for prediction. Science, 317,
207-208.
CRED, 2010: EM-DAT: The OFDA/CRED International Disaster Data Base. Centre for
Research on the Epidemiology of Disasters, Université Catholique de Louvain,
Louvain, Belgium, www.emdat.be/Database/.
Cronin, M., C. Gonzalez, and J.D. Sterman, 2009: Why don’t well-educated adults
understand accumulation? A challenge to researchers, educators, and citizens.
Organizational Behavior and Human Decision Processes, 108(1), 116-130.
Cuny, F.C., 1983: Disaster and Development. Oxford University Press, New York, NY.
Cutter, S.L., 1996: Vulnerability to environmental hazards. Progress in Human
Geography, 20, 529-239.
Cutter, S. L., L. Barners, M. Berry, C. Burton, E. Evans, E. Tate, and J. Webb, 2008: A
place-based model for understanding community resilience to natural disasters,
Global Environmental Change, 18, 598-606.
Davies, S., 1993: Are coping strategies a cop out? Institute of Development Studies
Bulletin, 24(4), 60-72.
Davies, S., 1996: Adaptable Livelihoods: Coping with Food Insecurity in the Malian
Sahel. Macmillan Press, London, UK.
de Bruin, K., R. Dellink, A. Ruijs, L. Bolwidt, A. van Buuren, J. Graveland, R. de Groot,
P. Kuikman, S. Reinhard, R. Roetter, V. Tassone, A. Verhagen, and E. van Ierland,
2009: Adapting to climate change in The Netherlands: an inventory of climate
adaptation options and ranking of alternatives. Climatic Change, 95(1),
23-45.
de Freitas, C.R., 1989: The hazard potential of drought for the population of the
Sahel. In: Population and Disaster [Clarke, J.I., P. Curson, S.L. Kayastha, and P.
Nag (eds.)]. Blackwell, Oxford, UK, pp. 98-113.
de Jong, C., D. Collins, and R. Ranzi (eds.), 2005:
Climate and Hydrology in Mountain
Areas. John Wiley and Sons, New York, NY.
de Vries, J., 1985: Analysis of historical climate-society interaction. In: Climate
Impact Assessment [Kates, R., J. Ausubel, and M. Berberian (eds.)]. John Wiley
and Sons, New York, NY.
Degg, M.R. and D.K. Chester, 2005: Seismic and volcanic hazards in Peru: changing
attitudes to disaster mitigation. The Geographical Journal, 171, 125-145.
del Ninno, C., P.A. Dorosh, and N. Islam, 2002: Reducing vulnerability to natural
disasters – Lessons from 1998 floods in Bangladesh. Ids Bulletin–Institute of
Development Studies, 33, 98-107.
del Ninno, C., P.A. Dorosh, and L.C. Smith, 2003: Public policy, markets and
household coping strategies in Bangladesh: Avoiding a food security crisis
following the 1998 floods. World Development, 31, 1221-1238.
Dessai, S., and M. Hulme, 2004: Does climate adaptation policy need probabilities?
Climate Policy, 4, 107-128.
Dessai, S., and M. Hulme, 2007: Assessing the robustness of adaptation decisions to
climate change uncertainties: A case study on water resources management in
the East of England. Global Environmental Change, 17(1), 59-72.
Dessai, S. and R. Wilby, 2011: How Can Developing Country Decision Makers
Incorporate Uncertainty about Climate Risks into Existing Planning and
Policymaking Processes? World Resources Report, World Resources Institute,
Washington, DC, http://www.worldresourcesreport.org/files/wrr/papers/
wrr_dessai_and_wilby_uncertainty.pdf.
Dessai, S., M. Hulme, R. Lempert, and R. Pielke Jr., 2009: Climate prediction: a limit
to adaptation? In: Adapting to Climate Change: Thresholds, Values, Governance
[Adger, W.N., I. Lorenzoni, and K.L. O’Brien (eds.)]. Cambridge University Press,
Cambridge, UK.
DFID, 2004: Adaptation to Climate Change: Making Development Disaster Proof. Key
Sheet 06, UK Department for International Development, London, UK.
DFID, 2005: Disaster Risk Reduction: A Development Concern. Policy Briefing
Document, DFID-ODG, UK Department for International Development, London,
UK.
Dhakal, A.S. and R.C. Sidle, 2004: Distributed simulations of landslides for different
rainfall conditions, Hydrological Processes, 18(4), 757-776.
Doherty, S.J., et al., 2009: Lessons learned from IPCC AR4: Scientific developments
needed to understand, predict, and respond to climate change. Bulletin of the
American Meteorological Society, 90, 497-513.
Douglas, M., 1992: Risk and Blame: Essays in Cultural Theory. Routledge, London,
UK and New York, NY.
Douglas, M. and A. Wildavsky, 1982: Risk and Culture: An Essay on the Selection of
Technical and Environmental Dangers. University of California Press, Berkeley,
CA.
Dow, K., 1992: Exploring differences in our common future(s): the meaning of
vulnerability to global environmental change. Geoforum, 23, 417-436.
Dulal, H.B., K.U. Shah, and N. Ahmad, 2009: Social equity considerations in the
implementation of Caribbean climate change adaptation policies.
Sustainability, 1, 363-383.
Dunlap, R.E. and A.M. McCright, 2008: A widening gap: republican and democratic
views on climate change. Environment, 50, 26-35.
Eakin, H., and A. L. Luers, 2006. Assessing the vulnerability of social-environmental
systems. Annual Review of Environment and Resources, 31, 365-394.
Easterling, D., G. Meehl, C. Parmesan, S. Changnon, T. Karl, and L. Mearns, 2000:
Climate extremes, observations, modeling and impacts. Science, 289, 2068-
2074.
El-Baroudy, I. and S.P. Simonovic, 2004: Fuzzy criteria for the evaluation of water
resources systems performance. Water Resources Research, 40(10), W10503,
doi:10.1029/2003WR002828.
Emanuel, K, 2001: Contribution of tropical cyclones to meridional heat transport by
the oceans. Journal of Geophysical Research, 106(D14), 14771-14781.
Emanuel, K., 2003: Tropical cyclones. Annual Review of Earth and Planetary
Sciences, 31, 75-104.
Epstein, P., H. Diaz, S. Elias, G. Grabherr, N. Graham, W. Martens, E.M. Thompson, and
J. Susskind, 1998: Biological and physical signs of climate change: focus on
mosquito borne diseases. Bulletin of the American Meteorological Society, 79,
409-417.
ERN-AL, 2011: Probabilistic Modelling of Natural Risk at Global Level: The Hybrid
Loss Exceedance Curve – Development of a Methodology and Implementation
of Case Studies, Phase 1A: Colombia, Mexico, Nepal. Report for the GAR 2011,
Consortium Evaluación de Riesgos Naturales – América Latina, Bogotá,
Colombia.
Fink, A.H., U. Ulbrich, and H. Engel, 1996: Aspects of the January 1995 flood in
Germany. Weather, 51(2), 34-39.
Fischhoff, B., 1994: What forecasts (seem to) mean. International Journal of
Forecasting, 10, 387-403.
Fischhoff, B., P. Slovic, and S. Lichtenstein, 1983: The “public” vs. the “experts”:
Perceived vs. actual disagreement about the risks of nuclear power. In: Analysis
of Actual vs. Perceived Risks [Covello,V., G. Flamm, J. Rodericks, and R. Tardiff
(eds.)]. Plenum, New York, NY, pp. 235-249.
Flood, R. and N. Romm, 1996: Diversity Management: Triple Loop Learning. Wiley,
Chichester, UK.
Folke, C., 2006: Resilience: The emergence of a perspective for social–ecological
systems analyses. Global Environmental Change, 16, 253-267.
Folke, C., F.S. Chapin III, and P. Olsson, 2009: Transformations in ecosystem
stewardship. In: Principles of Ecosystem Stewardship: Resilience-Based Natural
Resource Management in a Changing World [Chapin III, F.S., G.P. Kofinas, and
C. Folke (eds.)]. Springer, New York, NY, pp. 103-125.
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
59
Fouillet, A., G. Rey, V. Wagner, K. Laaidi, P. Empereur-Bissonnet, A. Le Tertre, P.
Frayssinet, P. Bessemoulin, F. Laurent, P. De Crouy-Chanel, E. Jougla, and D.
Hemon, 2008: Has the impact of heat waves on mortality changed in France
since the European heat wave of summer 2003?A study of the 2006 heat wave.
International Journal of Epidemiology, 37(2), 309-317.
Füssel, H., 2007: Vulnerability: A generally applicable conceptual framework for
climate change research. Global Environmental Change, 17(2), 155-167.
Gaillard, J.C., 2007: Resilience of traditional societies in facing natural hazard.
Disaster Prevention and Management, 16, 522-544.
Gaillard, J.C., 2010: Vulnerability, capacity, and resilience: Perspectives for climate
and development policy. Journal of International Development. 22, 218-232.
Gallopín, G.C., 2003: Box 1. A systemic synthesis of the relations between
vulnerability, hazard, exposure and impact, aimed at policy identification. In:
Handbook for Estimating the Socio-Economic and Environmental Effects of
Disasters. Economic Commission for Latin American and the Caribbean,
LC/MEX/G.S., Mexico, pp. 2-5.
Gallopín, G.C., 2006: Linkages between vulnerability, resilience, and adaptive capacity.
Global Environmental Change, 16, 293-303.
Gasper, D., 2010: The idea of human security. In: Climate Change, Ethics and Human
Security [O’Brien, K. A.L. St. Clair, and B. Kristoffersen (eds.)]. Cambridge
University Press, Cambridge, UK, pp. 23-46.
Gershunov, A., D. R. Cayan, and S. F. Iacobellis, 2009: The great 2006 heat wave over
California and Nevada: Signal of an increasing trend. Journal of Climate, 22,
6181-6203.
Goddard, L., W. Baethgen, B. Kirtman, and G. Meehl, 2009: The urgent need for
improved climate models and predictions. EOS, 90(39), 343.
Gordon, J.E., 1978: Structures. Penguin Books, Harmondsworth, UK.
Gottschalk, M.K.E., 1971: Stormvloeden en rivieroverstromingen in Nederland Deel
I (Storm surges and river floods in the Netherlands, part I). Van Gorcum, Assen,
The Netherlands.
Gottschalk, M.K.E., 1975: Stormvloeden en rivieroverstromingen in Nederland Deel
II (Storm surges and river floods in the Netherlands, part II). Van Gorcum, Assen,
The Netherlands.
Gottschalk, M.K.E., 1977: Stormvloeden en rivieroverstromingen in Nederland Deel
III (Storm surges and river floods in the Netherlands, part III). Van Gorcum,
Assen, The Netherlands.
Groves, D.G. and R.J. Lempert, 2007: A new analytic method for finding policy-
relevant scenarios. Global Environmental Change, 17, 73-85.
Hagman, G., 1984: Prevention Better than Cure. Swedish Red Cross, Stockholm,
Sweden.
Hallegatte, S., 2009: Strategies to adapt to an uncertain climate change. Global
Environmental Change, 19(2), 240-247.
Hargrove, R., 2002.
Masterful Coaching. Revised Edition. Jossey-Bass/Pfeiffer, Wiley,
San Francisco, CA, USA.
Hawkins, E. and R. Sutton, 2009: The potential to narrow uncertainty in regional climate
predictions. Bulletin of the American Meteorological Society, 90, 1095-1107.
Hegerl, G.C., F.W. Zwiers, P. Braconnot, N.P. Gillett, Y. Luo, J.A. Marengo Orsini, N.
Nicholls, J.E. Penner, and P.A. Stott, 2007: Understanding and attributing climate
change. In: Climate Change 2007: The Physical Science Basis. Contribution of
Working Group I to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M.
Marquis, K.B. Averyt, M. Tignor, and H.L. Miller (eds.)]. Cambridge University
Press, Cambridge, UK and New York, NY, pp. 663-745.
Heltberg, R., P.B. Siegel, and S.L. Jorgensen, 2009: Addressing human vulnerability
to climate change: Toward a “no-regrets” approach. Global Environmental
Change, 19(1), 89-99.
Hertwig, R., G. Barron, E.U. Weber, and I. Erev, 2004: Decisions from experience and
the effect of rare events. Psychological Science, 15, 534-539.
Hewitt, K., 1983: The idea of calamity in a technocratic age. In: Interpretations of
Calamity from the Viewpoint of Human Ecology [Hewitt, K. (ed.)]. Allen and
Unwin, Boston, MA and London, UK, pp 3-32.
Hewitt, K., 1997: Regions of Risk: A Geographical Introduction to Disasters.
Longman, Harlow, Essex, UK.
Hewitt, K., 2007: Preventable disasters: addressing social vulnerability, institutional
risks and civil ethics. Geographisches Bundscahu. International Edition, 3(1),
43-52.
Hewitt, K. and I. Burton, 1971: The Hazardousness of a Place: A Regional Ecology of
Damaging Events. University of Toronto, Toronto, Canada.
Hine, D. and J.W. Hall, 2010: Information gap analysis of flood model uncertainties
and regional frequency analysis. Water Resources Research, 46(19), W01514,
doi:10.1029/2008WR007620.
Hoeksma, R.J., 2006: Designed for Dry feet: Food Protection and Land Reclamation
in the Netherlands. American Society of Civil Engineers, Reson, VA.
Hohenemser, C., R.E. Kasperson, and R.W. Kates, 1984: Causal structure. In: Perilous
Progress: Managing the Hazards of Technology [Kates, R.W., C. Hohenemser
and J.X. Kasperson (eds.)]. Westview Press, Boulder, CO, pp. 25-42.
Holland, M.M., C.M. Bitz, and B. Tremblay, 2006: Future abrupt reductions in the
summer Arctic sea ice. Geophysical Research Letters, 33, L23503,
doi:10.1029/2006GL028024.
Holling, C.S., 1973: Resilience and stability of ecological systems. Annual Review of
Ecology and Systematics, 4, 1-23.
Horsburgh, K.J. and M. Horritt, 2006: The Bristol Channel Floods of 1607 –
reconstruction and analysis. Weather, 61(10), 272-277.
Huang, C., P. Vaneckova, X. Wang, G. FitzGerald, Y. Guo, and S. Tong, 2011:
Constraints and barriers to public health adaptation to climate change.
American Journal of Preventive Medicine, 40(2), 183-190.
HVRI, 2010: The Spatial Hazard Events and Losses Database for the United States.
Version 8.0 (online database). Hazards and Vulnerability Research Institute,
University of South Carolina, Columbia, SC, www.sheldus.org.
ICSU, 2002: Resilience and Sustainable Development: Building Adaptive Capacity in
a World of Transformations. Series on Science for Sustainable Development:
Resilience and Sustainable Development No. 3, International Council for
Science, Paris, France, 37 pp.
ICSU, 2008: A Science Plan for Integrated Research on Disaster Risk: addressing the
challenge of natural and human induced environmental hazards. International
Council for Science, Paris, France.
ICSU-LAC, 2009: Understanding and Managing Risk Associated with Natural
Hazards: An Integrated Scientific Approach in Latin America and the Caribbean.
International Council for Science Regional Office for Latin America and the
Caribbean, Rio de Janeiro, Brazil and Mexico City, Mexico.
IDB, 2007: Disaster Risk Management Policy. GN-2354-5, Inter-American
Development Bank, Washington, DC.
IFRC, 2007. How to Do a VCA: A Practical Step-by-Step Guide for Red Cross Red
Crescent Staff and Volunteers. International Federation of Red Cross and Red
Crescent Societies.
IIED, 2009: Natural Resilience: Healthy ecosystems as climate shock insurance. IIED
Policy Brief, International Institute for Environment and Development, London,
UK.
Ingram, K.T., M.C. Roncoli, and P.H. Kirshen, 2002: Opportunities and constraints for
farmers of West Africa to use seasonal precipitation forecasts with Burkina Faso
as a case study. Agricultural Systems, 74(3), 331-349.
International Building Codes, 2003: International Code Council (paperback), 656
pp. ISBN 1892395568.
IPCC, 2007a: Climate Change 2007: Synthesis Report. Contribution of Working
Groups I, II and III to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change [Core Writing Team, R.K. Pachauri, and A. Reisinger
(eds.)]. IPCC, Geneva, Switzerland, 104 pp.
IPCC, 2007b: Impacts, Adaptation and Vulnerability. Contribution of Working Group
II to the Fourth Assessment Report of the Intergovernmental Panel on Climate
Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. van der Linden, and C.E.
Hanson (eds.)]. Cambridge University Press, Cambridge, UK and New York, NY.
IPCC, 2007c: Appendix I: Glossary. In: Climate Change 2007: Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani,
J.P. Palutikof, P.J. van der Linden, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, and New York, NY, pp. 869-883.
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
60
IPCC, 2009. Scoping Meeting for an IPCC Special Report on Extreme Events and
Disasters: Managing the Risks. Proceedings [Barros, V., et al. (eds.)]. 23-26
March 2009, Oslo, Norway.
ISO, 2009: Guide 73: 2009. Risk Management – Vocabulary. International
Organization for Standardization, Geneva, Switzerland.
Jansen, V., N. Stollenwerk, H.J. Jensen, M.E. Ramsay, W.J. Edmunds, and C.J. Rhodes,
2003: Measles outbreaks in a population with declining vaccine uptake.
Science, 301, 804.
Jasanoff, S., 2004: States of Knowledge: The Co-Production of Science and Social
Order. Routledge, London, UK.
Jentsch, A., J. Kreyling, and C. Beierkuhnlein, 2007: A new generation of climate
change experiments: events not trends. Frontiers in Ecology and the
Environment, 6(6), 315-324.
Jones, R.N. and B.L. Preston, 2011: Adaptation and Risk Management. WIREs Climate
Change, 2, 296-308, doi:10.1002/wcc.97.
Kahan, D.M. and D. Braman, 2006: Cultural cognition and public policy. Yale Law and
Policy Review, 24(1), 149-172.
Kahan, D.M., D. Braman, J. Gastil, P. Slovic, and C.K. Mertz, 2007: Culture and identity
– protective cognition: Explaining the white male effect in risk perception.
Journal of Empirical Legal Studies, 4(3), 465-505.
Kahneman, D. and A. Tversky, 1979: Prospect theory: An analysis of decisions under
risk. Econometrica, 47, 313-327.
Kahneman, D., P. Slovic, and A. Tversky (eds.), 1982: Judgment under Uncertainty:
Heuristics and Biases. Cambridge University Press, New York, NY.
Karimi, I. and E. Hullermeier, 2007: Risk assessment system of natural hazards: A
new approach based on fuzzy probability. Fuzzy Sets and Systems, 158, 987-999.
Kasperson, R.E., O. Renn, P. Slovic, H.S. Brown, J.Emel, R. Goble, J.X. Kasperson, and
S. Ratick, 1988: The social amplification of risk: A conceptual framework. Risk
Analysis, 8,177-187.
Kates, R.W., et al., 2001: Sustainability science. Science, 292(5517), 641-642.
Keen, M., V.A. Brown, and R. Dyball, 2005: Social Learning in Environmental
Management: Towards a Sustainable Future. Earthscan, London, UK.
Kelman, I., 2008. Relocalising disaster risk reduction for urban resilience. Urban
Design and Planning, 161(DP4), 197-204.
Khan, T.A., 2011: Floods increase wheat yield in KP. Inpaper Magazine, Pakistan, 16
May 2011.
Klein
, R.J.T, R.J. Nicholls, and F. Thomalla, 2003: Resilience to natural hazards: how
useful is this concept? Global Environmental Change, 5, 35-45.
Knutti, R., 2010: The end of model democracy? Climatic Change, 102, 395-404.
Kolb, D.A., 1984: Experiential Learning: Experience as the Source of Learning and
Development. Prentice-Hall, Englewood Cliffs, NJ.
Kolb, D.A. and R. Fry, 1975: Toward an applied theory of experiential learning. In:
Theories of Group Process [Cooper, C. (ed.)]. John Wiley, London, UK, pp. 33-57.
Konikow, L.F. and E. Kendy, 2005: Groundwater depletion: a global problem.
Hydrogeology Journal, 13, 317-320.
Kunreuther, H., 2006: Disaster mitigation and insurance: Learning from Katrina. The
Annals of the American Academy of Political and Social Science, 604(1), 208-
227.
Kysar, D., 2004: Climate change, cultural transformation, and comprehensive
rationality. Boston College Environmental Affairs Law Review, 31(3), 555-590.
Lamond, J.E., D.G. Proverbs, and F.N. Hammond, 2009: Accessibility of flood risk
insurance in the UK: confusion, competition and complacency. Journal of Risk
Research, 12, 825-841.
Lau, W.K.M. and K.-M. Kim, 2011: The 2010 Pakistan flood and Russian heat wave:
Teleconnection of hydrometeorologic extremes. Journal of Hydrometeorology,
doi:10.1175/JHM-D-11-016.1.
Lavell, A., 1996: Degradación ambiental, riesgo y desastre urbano. Problemas y
conceptos: hacia la definición una agenda de investigación. In: Ciudades en
Riesgo [Fernandez, M.A. (ed.)]. La RED-USAID, Lima, Peru, pp. 21-59.
Lavell, A., 1999: Environmental degradation, risks and urban disasters. Issues and
concepts: Towards the definition of a research agenda. In: Cities at Risk:
Environmental Degradation, Urban Risks and Disasters in Latin America
[Fernandez, M.A. (ed.)]. La RED, US AID, Quito, Ecuador, pp. 19-58.
Lavell, A., 2002: Riesgo y Territorio: los niveles de intervención en la Gestión del
Riesgo (Risk and Territory: Levels of Intervention and Risk Management).
Anuario Social y Político de América Latina y el Caribe. FLACSO-Nueva
Sociedad, Latin American School of Social Sciences.
Lavell, A., 2003: Local level risk management: Concept and practices. CEPREDENAC-
UNDP. Quito, Ecuador.
Lavell, A., 2009: Technical Study in Integrating Climate Change Adaptation and
Disaster Risk Management in Development Planning and Policy. Study
undertaken for the Inter-American Development Bank, Washington, DC.
Lavell, A., 2010: Unpacking Climate Change Adaptation and Disaster Risk
Management: Searching for the Links and the Differences: A Conceptual and
Epistemological Critique and Proposal. IUCN-FLACSO, International Union for
Conservation of Nature - Latin American School of Social Sciences.
Lavell A. and E. Franco (eds.), 1996: Estado, Sociedad y Gestión de los Desastres en
América Latina. Red de Estudios Sociales en Prevención de Desastres en
América Latina, La RED, Tercer Mundo Editores, Bogotá, Colombia.
Lavell, A. and C. Lavell, 2009: Local Disaster Risk Reduction Lessons from the Andes.
Series: Significant Local Development Initiatives in the Face of Disaster Risk.
General Secretariat of the Andean Community, Lima, Peru.
Leckebusch, G.C., D. Renggli, and U. Ulbrich, 2008: Development and application of
an objective storm severity measure for the Northeast Atlantic region.
Meteorologische Zeitschrift, 17, 575-587.
Lemos, M.C., 2003: A tale of two policies: the politics of seasonal climate forecast
use in Ceará, Brazil. Policy Sciences, 32(2), 101-123.
Lemos, M.C., E. Boyd, E.L. Tompkins, H. Osbahr, and D. Liverman, 2007: Developing
adaptation and adapting development. Ecology and Society, 12(2), 26,
www.ecologyandsociety.org/vol12/iss2/art26/.
Lempert, R.J. and M. Collins, 2007: Managing the risk of uncertain threshold
responses: Comparison of robust, optimum, and precautionary approaches. Risk
Analysis, 27(4), 1009-1026.
Lempert, R.J. and D.G. Groves, 2010: Identifying and evaluating robust adaptive
policy responses to climate change for water agencies in the American west.
Technological Forecasting and Social Change, 77, 960-974.
Lempert, R.J. and M. Schlesinger, 2001: Climate-change strategy needs to be robust.
Nature, 412, 375.
Lewcowicz, A.G. and C. Harris, 2005: Frequency and magnitude of active-layer
detachment failures in discontinuous and continuous permafrost, northern
Canada. Permafrost and Periglacial Processes, 16,115-130.
Lewis, J., 1979: The vulnerable state: An alternative view. In: Disaster Assistance:
Appraisal, Reform and New Approaches [Stephens, L. and S.J. Green (eds.)].
New York University Press, New York, NY, pp. 104-129.
Lewis, J., 1984: Environmental interpretations of natural disaster mitigation: The
crucial need. The Environmentalist, 4(3), pp. 177-180.
Lewis, J., 1999: Development in Disaster-prone Places: Studies in Vulnerability. IT
Publications (Practical Action), London, UK.
Lewis, J., 2009: An island characteristic: Derivative vulnerabilities to indigenous and
exogenous hazards. Shima: The International Journal of Research into Island
Cultures. 3(1), 3-15.
Lewis, J. and I. Kelman, 2009: Housing, flooding and risk-ecology: Thames Estuary
South-Shoreland and North Kent. Journal of Architectural and Planning
Research, 26(1), 14-29.
Lindell, M. and C. Prater, 2003: Assessing community impacts of natural disasters.
Natural Hazards Review, 4(4), 176-185.
Lipton, M. and M. Ravallion, 1995: Poverty and policy. In: Handbook of Development
Economics, Vol. 3B [Behrman, J.S. and T.N. Srinivasan (eds.)]. Elsevier Science,
Amsterdam, New York, and Oxford, pp. 2551-2657.
List, D., 2006: Action research cycles for multiple futures perspectives. Futures, 38,
673-684.
Loewenstein, G.F., E.U. Weber, C.K. Hsee, and E. Welch, 2001: Risk as feelings.
Psychological Bulletin, 127, 267-286.
Malka, A. and J.A. Krosnick, 2009: The association of knowledge with concern about
global warming: trusted information sources shape public thinking. Risk
Analysis, 29, 633-647.
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
61
Manyena, S.B., 2006: The concept of resilience revisited. Disasters, 30(4),
433-450.
Maskrey, A., 1989: Disaster Mitigation: A Community Based Approach. Oxfam,
Oxford, UK.
Maskrey, A., 2011: Revisiting community based risk management. Environmental
Hazards, 10, 1-11.
Mastrandrea, M.D., C.B. Field, T.F. Stocker, O. Edenhofer, K.L. Ebi, D.J. Frame, H. Held,
E. Kriegler, K.J. Mach, P.R. Matschoss, G.-K. Plattner, G.W. Yohe, and F.W. Zwiers,
2010: Guidance Note for Lead Authors of the IPCC Fifth Assessment Report on
Consistent Treatment of Uncertainties. IPCC, www.ipcc-wg2.gov/meetings/
CGCs/Uncertainties-GN_IPCCbrochure_lo.pdf.
McBean, G., 2006: An integrated approach to air pollution, climate and weather
hazards. Policy Options, October 2006, pp. 18-24.
McGray, H., A. Hammill, and R. Bradley, 2007: Weathering the Storm: Options for
Framing Adaptation and Development. World Resources Institute, Washington,
DC.
McKee, T.B., N.J. Doesken, and J. Kleist, 1993: The duration of drought frequency and
duration to time scales. In: Eighth Conference on Applied Climatology, 17-22
January 1993, Anaheim, California, ccc.atmos.colostate.edu/relationshipof-
droughtfrequency.pdf.
McLaughlin, P. and T. Dietz, 2007: Structure, agency and environment: Toward an
integrated perspective on vulnerability. Global Environmental Change, 39(4),
99-111.
Means, E., M. Laugier, J. Daw, L. Kaatz, and M. Waage, 2010: Decision Support
Planning Methods: Incorporating Climate Change Uncertainties into Water
Planning. Water Utility Climate Alliance, San Francisco, CA,
www.wucaonline.org/assets/pdf/actions_whitepaper_012110.pdf.
Meehl, G.A., T.F. Stocker, W.D. Collins, P. Friedlingstein, A.T. Gaye, J.M. Gregory, A.
Kitoh, R. Knutti, J.M. Murphy, A. Noda, S.C.B. Raper, Watterson, I.G., A.J. Weaver
and Z.-C. Zhao, 2007: Global climate projections. In: Climate Change 2007: The
Physical Science Basis. Contribution of Working Group I to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change
[Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor,
and H.L. Miller (eds.)]. Cambridge University Press, Cambridge, UK and New
York, NY, pp. 747-843.
Mercer, J., 2010: Disaster risk reduction or climate change adaptation: are we
reinventing the wheel? Journal of International Development, 22, 247-264.
Mezirow, J., 1995: Transformation theory in adult learning. In: In Defense of the Life
World [Welton, M.R. (ed.)]. State University of New York Press, Albany, NY, pp.
39-70.
Mileti, D., 1999: Disasters by Design: A Reassessment of Natural Hazards in the
United States. Joseph Henry Press, Washington, DC.
Miller, F., H. Osbahr, E. Boyd, F. Thomalla, S. Bharwani, G. Ziervogel, B. Walker, J.
Birkmann, S. Van der Leeuw, J. Rockström, J. Hinkel, T. Downing, C. Folke, and D.
Nelson, 2010: Resilience and vulnerability: complementary or conflicting
concepts? Ecology and Society, 15(3), 11.
Milly, P.C.D., R.T. Wetherald, K.A. Dunne, and T.L. Delworth, 2002: Increasing risk of
great floods in a change climate. Nature, 415, 514-517.
Milly, P.C.D., J. Betancourt, M. Falkenmark, R. Hirsch, Z. Kundzewicz, D. Lettenmaier,
and R. Stouffer, 2008: Stationarity is dead: Whither water management?
Science, 319(5863), 573-574.
Minor, J.E., 1983: Construction tradition for housing determines disaster potential
from severe tropical cyclones. Journal of Wind Engineering and Industrial
Aerodynamics, 14, 55-66.
Mitchell, T. and M. van Aalst, 2008: Convergence of Disaster Risk Reduction and
Climate Change Adaptation. Technical Paper, October 2008, UK Department of
International Development, London, UK.
Moench, M. and A. Dixit (eds.), 2004: Adaptive Capacity and Livelihood Resilience:
Adaptive Capacities for Responding to Floods and Droughts in South Asia.
Institute for Social and Environmental Transition, Boulder, CO and Nepal.
Morgan, M.G. and M. Henrion, 1990: Uncertainty: A Guide to Dealing with
Uncertainty in Quantitative Risk and Policy Analysis. Cambridge University
Press, Cambridge, UK.
Morgan, M.G., H. Dowlatabadi, M. Henrion, D. Keith, R.J. Lempert, S. McBride, M.
Small, and T. Wilbanks, 2009: Best Practice Approaches for Characterizing,
Communicating, and Incorporating Scientific Uncertainty in Decisionmaking.
Synthesis and Assessment Product 5.2 of the U.S. Climate Change Science
Program, National Oceanic and Atmospheric Administration, Washington, DC.
Moser, S.C., 2009. Whether our levers are long enough and the fulcrum strong?
Exploring the soft underbelly of adaptation decisions and actions. In: Adapting
to Climate Change: Thresholds, Values, Governance [Adger, W.N., I. Loronzoni,
and K. O’Brien (eds.)]. Cambridge University Press, Cambridge, UK, pp. 313-
343.
Moser, S.C., 2010: Now more than ever: The need for more socially relevant research
on vulnerability and adaptation to climate change. Applied Geography, 30(4),
464-474.
Moser, S.C. and J.A. Ekstrom, 2010: Barriers to climate change adaptation: A
diagnostic framework. Proceedings of the National Academy of Sciences,
107(51), 22026-22031.
Moser, S.C. and J.A. Ekstrom, 2011: Taking ownership of climate change:
Participatory adaptation planning in two local case studies from California.
Journal of Environmental Studies and Sciences, 1(1), 63-74.
Moser, S.C. and A.L. Luers, 2008: Managing climate risks in California: The need to
engage resource managers for successful adaptation to change. Climatic
Change, 87, S309-S322.
Nelson, D.R., 2005: The Public and Private Sides of Persistent Vulnerability to
Drought: An Applied Model for Public Planning in Ceará, Brazil. PhD Thesis,
University of Arizona, Tucson, AZ.
Nelson, D. R., N. Adger, and K. Brown, 2007: Adaptation to environmental change:
contributions of a resilience framework. The Annual Review of Environment and
Resources, 32, 395-419.
Nielsen, J. and A. Reenberg, 2010: Cultural barriers to climate change adaptation:
A case study from Northern Burkina Faso. Global Environmental Change, 20(1),
142-152.
Noji, E., 1997: The Public Health Impacts of Disasters. Oxford University Press, New
York, NY.
Norris
, F.H., S.P. Stevens, B. Pfefferbaum, K.F. Wyche, and R.L. Pfefferbaum, 2008:
Community resilience as a metaphor, theory, set of capacities, and strategy for
disaster readiness. American Journal of Community Psychology, 41, 127-150
NRC, 2006: Facing Hazards and Disasters: Understanding Human Dimensions.
National Research Council, National Academies Press, Washington DC.
NRC, 2009: Informing Decisions in a Changing Climate. Panel on Strategies and
Methods for Climate-Related Decision Support, Committee on the Human
Dimensions of Global Change, Division of Behavioral and Social Sciences and
Education, National Research Council, The National Academies Press,
Washington, DC, 188 pp.
O’Brien, K. and R. Leichenko, 2000: Double exposure: assessing the impacts of
climate change within the context of economic globalization. Global
Environmental Change, 10, 221-232.
O’Brien, K., S. Eriksen, L.P. Nygaard, and A. Schjolden, 2007: Why different
interpretations of vulnerability matter in climate change discourses. Climate
Policy, 7(1), 73-88.
O’Keefe, P., K. Westgate, and B. Wisner, 1976: Taking the naturalness out of natural
disasters. Nature, 260(5552), 566-577.
OED, 1989: The Oxford English Dictionary. Oxford University Press, Oxford, UK.
Olsson, P., L.H. Gunderson, S.R. Carpenter, P. Ryan, L. Lebel, C. Folke, and C.S. Holling,
2006: Shooting the rapids: navigating transitions to adaptive governance of
social-ecological systems. Ecology and Society, 11(1), 18.
Oppenheimer, M., B.C. O’Neill, and M. Webster, 2008: Negative learning. Climatic
Change, 89, 155-172.
Ostertag, R., W.L. Silver, and A.E. Lugo, 2005: Factors affecting mortality and
resistance to damage following hurricanes in a rehabilitated subtropical moist
forest. Biotropica, 37(1), 16-24.
Pahl-Wostl, C., 2009: A conceptual framework for analysing adaptive capacity and
multi-level learning processes in resource governance regimes. Global
Environmental Change, 19(3), 354.
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
62
Pall, P., T. Aina, D.A. Stone, P.A. Stott, T. Nozawa, A.G. Hilberts, D. Lohmann, and M.R.
Allen, 2011: Anthropogenic greenhouse gas contribution to flood risk in
England and Wales in autumn 2000. Nature, 470, 382-386
Parson, E., V. Burkett, K. Fisher-Vanden, D. Keith, L. Means, H. Pitcher, C. Rosenzweig,
and M. Webster, 2007: Global Change Scenarios: Their Development and Use.
Sub-report 2.1B of Synthesis and Assessment Product 2.1 by the U.S. Climate
Change Science Program and the Subcommittee on Global Change Research,
Department of Energy, Office of Biological and Environmental Research,
Washington, DC, 106 pp.
Pelling, M., 2003: The Vulnerability of Cities: Social Resilience and Natural Disaster.
Earthscan, London, UK.
Pelling, M., 2010: Adaptation to Climate Change: From Resilience to Transformation.
Routledge, London, UK.
Pelling, M., A. Özerdem, and S. Barakat, 2002: The Macroeconomic impact of disasters.
Progress in Development Studies, 2, 283-305.
Pelling, M., C. High, J. Dearing, and D. Smith, 2008: Shadow spaces for social learning:
a relational understanding of adaptive capacity to climate change within
organisations. Environment and Planning A, 40(4), 867-884.
Pendalla, R., K.A. Foster, and M. Cowella, 2010: Resilience and regions: building
understanding of the metaphor. Cambridge Journal of Regions, Economy and
Society, 3(1), 71-84
Peschl, M.F., 2007: Triple-loop learning as foundation for profound change, individual
cultivation, and radical innovation. Construction processes beyond scientific
and rational knowledge. Constructivist Foundations, 2(2-3), 136-145.
Peters, G., A.M. DiGioia Jr, C. Hendrickson, and J. Apt, 2006: Transmission Line
Reliability: Climate Change and Extreme Weather. Carnegie Mellon Electricity
Industry Center Working Paper CEIC-05-06, Carnegie Mellon University,
Pittsburgh, PA.
Peters, O., C. Hertlein, and K. Christensen, 2001: A complexity view of rainfall.
Physical Review Letters, 88, 018701, doi:10.1103/PhysRevLett.88.018701.
Pettengell, C., 2010: Climate Change Adaptation: Enabling People Living in Poverty
to Adapt. Oxfam International Research Report, Oxfam International, Oxford,
UK.
Piao, S., P. Ciais, Y. Huang, Z. Shen, S. Peng, J. Li, L. Zhou, H. Liu, Y. Ma, Y. Ding, P.
Friedlingstein, C. Liu, K. Tan, Y. Yu, T. Zhang, and J. Fang, 2010: The impact of climate
change on water resources and agriculture in China. Nature, 467, 43-51.
Pidgeon, N. and B. Fischhoff, 2011: The role of social and decision sciences in
communicating uncertain climate risks. Nature Climate Change, 1, 35-41.
Pielke, Jr., R.A., 2007: Future economic damage from tropical cyclones: sensitivities
to societal and climate changes. Philosophical Transactions of the Royal Society
A, 365(1860), 2717-2729.
Pohl, C., S. Rist, A. Zimmermann, P. Fry, G.S. Gurung, F. Schneider, C. Ifejika Speranza,
B. Kiteme, S. Boillat, E. Serrano, G. Hirsch Hadorn, and U. Wiesmann, 2010:
Researchers’ roles in knowledge co-production: experience from sustainability
research in Kenya, Switzerland, Bolivia and Nepal. Science and Public Policy,
37(4), 267-281.
Quarantelli, E.L, 1986: Disaster Crisis Management. Preliminary Papers 113,
University of Delaware, Disaster Research Center. Newark, DE,
http://dspace.udel.edu:8080/dspace/handle/19716/487.
Ramírez, F. and O.D. y Cardona, 1996: Sistema nacional para la prevención y
atención de desastres de Colombia, Estado, sociedad y gestión de los desastres
en América Latina. In: Red de Estudios Sociales en Prevención de Desastres en
América Latina [Lavell, A. and E. y Franco (eds.)]. Tercer Mundo Editores, Bogotá,
Colombia.
Ranger, N. and S.-L. Garbett-Shiels, 2011: How can Decision-Makers in Developing
Countries Incorporate Uncertainty about Future Climate Risks into Existing
Planning and Policymaking Processes? Policy Paper, Centre for Climate Change
Economics and Policy and Grantham Research Institute on Climate Change and
the Environment in collaboration with the World Resources Report, World
Resources Institute, Washington, DC, www.worldresourcesreport.org/files/
wrr/papers/wrr_ranger_uncertainty.pdf.
Rayner, S., D. Lach, and H. Ingram, 2005: Weather forecasts are for wimps: Why water
resource managers do not use climate forecasts. Climatic Change, 69, 197-227.
Reeder, T. and N. Ranger, 2011: How Do You Adapt In an Uncertain World? Lessons
from the Thames Estuary 2100 Project. World Resources Report, World
Resources Institute, Washington, DC, www.worldresourcesreport.org/files/
wrr/papers/wrr_reeder_and_ranger_uncertainty.pdf.
Rees, H.G. and D.N. Collins, 2006: Regional differences in response of flow in glacier-
fed Himalayan rivers. Hydrological Processes, 20(10), 2157-2169.
Renn, O., 2008: Risk Governance. Coping with Uncertainty in a Complex World.
Earthscan, London, UK.
Rich, P.M., D.D. Breshears, and A.B. White, 2008: Phenology of mixed woody-
herbaceous ecosystems. Ecology, 89(2), 342-352.
Risbey, J., M. Kandlikar, H. Dowlatabadi, and D. Graetz, 1999: Scale, context, and
decision making in agricultural adaptation to climate variability and change.
Mitigation and Adaptation Strategies for Global Change, 4(2), 137-165.
Rose, A., 2004: Economic principles, issues, and research priorities in hazard loss
estimation. In: Modeling Spatial and Economic Impacts of Disasters [Okuyama,
Y. and S. Chang (eds.)]. Springer, Berlin, Germany, pp. 14–36.
Sarewitz, D., R.A. Pielke, Jr., and R. Byerly, Jr. (eds.), 2000: Prediction: Science,
Decision Making, and the Future of Nature. Island Press, Washington, DC, 405 pp.
Savage, L.J., 1972: Foundations of Statistics. Dover Publications, New York, NY.
Schipper, L., 2009: Meeting at the crossroads? Exploring the linkages between
climate change adaptation and disaster risk reduction. Climate and
Development, 1, 16-30.
Schipper, L. and I. Burton (eds.), 2009: The Earthscan Reader on Adaptation to
Climate Change. Earthscan, London, UK.
Schipper, L. and M. Pelling, 2006: Disaster risk, climate change and international
development: scope for, and challenges to, integration. Disasters,
30(1), 19-38.
Schneider, S.H., S. Semenov, A. Patwardhan, I. Burton, C.H.D. Magadza, M.
Oppenheimer, A.B. Pittock, A. Rahman, J.B. Smith, A. Suarez, and F. Yamin, 2007:
Assessing key vulnerabilities and the risk from climate change. In: Climate
Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working
Group II to the Fourth Assessment Report of the Intergovernmental Panel on
Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. van der Linden, and
C.E. Hanson, (eds.)]. Cambridge University Press, Cambridge, UK, pp. 779-810.
Sen, A.K., 1983: Poverty and famines: an essay on entitlement and deprivation.
Oxford University Press, Oxford, UK.
Seneviratne, S.I., D. Lüthi, M. Litschi, and C. Schär, 2006: Land-atmosphere coupling
and climate change in Europe. Nature, 443, 205-209.
Shapiro, M., J. Shukla, G. Brunet, C. Nobre, M. Beland, R. Dole, K. Trenberth, R.
Anthes, G. Asrar, L. Barrie, P. Bougeault, G. Brasseur, D. Burridge, A. Busalacchi,
J. Caughey, D. Chen, J. Church, T. Enomoto, B. Hoskins, O. Hov, A. Laing, H. Le
Treut, J. Marotzke, G. McBean, G. Meehl, M. Miller, B. Mills, J. Mitchell, M.
Moncrieff, T. Nakazawa, H. Olafsson, T. Palmer, D. Parsons, D. Rogers, A.
Simmons, A. Troccoli, Z. Toth, L. Uccellini, C. Velden, and J.M. Wallace, 2010: An
earth-system prediction initiative for the twenty-first century. Bulletin of the
American Meteorological Society, 91, 1377-1388.
Shukla, J., R. Hagedorn, B. Hoskins, J. Kinter, J. Marotzke, M. Miller, T.N. Palmer, and J.
Slingo, 2009: Revolution in climate prediction is both necessary and possible: A
declaration at the World Modelling Summit for Climate Prediction. Bulletin of
the American Meteorological Society, 90, 175-178.
Simonovic, S.P., 2010: Systems Approach to Management of Disasters: Methods and
Applications. John Wiley and Sons, New York, NY, 348 pp.
Sjöberg, L., 1999a: Risk perception by the public and by experts: A dilemma in risk
management. Human Ecology Review, 6(2).
Sjöberg, L., 1999b: Consequences of perceived risk: Demand for mitigation. Journal
of Risk Research, (2), 129-149.
Slovic, P., 1993: Perceived risk, trust, and democracy. Risk Analysis, 13(6), 675-676.
Slovic, P. (ed.), 2000: The Perception of Risk. Earthscan, Sterling, VA.
Slovic, P. (ed.), 2010: The Feeling of Risk: New Perspectives on Risk Perception.
Earthscan, London, UK.
Slovic, P., B. Fischhoff, and S. Lichtenstein, 1982. Facts versus fears: Understanding
perceived risk. In: Judgment Under Uncertainty: Heuristics and Biases
[Kahneman, D., P. Slovic, and A. Tversky (eds.)]. Cambridge University Press,
New York, NY.
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
63
Smit, B. and O. Pilifosova, 2003. From adaptation to adaptive capacity and
vulnerability reduction. In: Climate Change, Adaptive Capacity and
Development [Smith, J.B., R.J.T. Klein, and S. Huq (eds.)]. Imperial College Press,
London, UK.
Smit, B. and J. Wandel, 2006: Adaptation, adaptive capacity and vulnerability. Global
Environmental Change, 16, 282-292.
Smith, K., 1996: Environmental Hazards: Assessing Risk and Reducing Disaster.
Second Edition. Routledge, London, UK.
Sperling, F. and F. Szekely, 2005: Disaster Risk Management in a Changing Climate.
Discussion Paper prepared for the World Conference on Disaster Reduction on
behalf of the Vulnerability and Adaptation Resource Group, Washington, DC.
Stainforth, D.A., M.R. Allen, E.R. Tredger, and L.A. Smith, 2007: Confidence,
uncertainty, and decision-support relevance in climate predictions.
Philosophical Transactions of the Royal Society A, 365, 2145-2161.
Sterman, J.D., 2000: Business Dynamics: Systems Thinking and Modeling for a
Complex World. Irwin/McGraw-Hill, Boston, MA.
Sterman, J.D., 2006: Learning from evidence in a complex world. American Journal
of Public Health, 96(3), 505-514.
Sterman, J.D., 2008: Risk communication on climate: mental models and mass
balance. Science, 322(5901), 532-533.
Stott, P.A., D.A. Stone, and M.R. Allen, 2004: Human contribution to the European
heatwave of 2003. Nature, 432, 610-614.
Sui, J. and G. Koehler, 2001: Rain-on-snow induced flood events in Southern
Germany. Journal of Hydrology, 252(1-4), 205-220.
Thomalla, F., T. Downing, E. Spanger-Siegfried, G. Han and J. Rockström, 2006:
Reducing hazard vulnerability: towards a common approach between disaster
risk reduction and climate adaptation. Disasters, 30(1), 39-48.
Timmerman, P., 1981: Vulnerability, Resilience and Collapse in Society.
Environmental Monograph No. 1, Institute for Environmental Studies, University
of Toronto, Toronto, Canada.
Tobin, G.A. and B.E. Montz, 1997: Natural Hazards: Explanation and Integration. The
Guildford Press, London, UK.
Tompkins, E.L., M.C. Lemos, and E. Boyd, 2008: A less disastrous disaster: Managing
response to climate-driven hazards in the Cayman Islands and NE Brazil. Global
Environmental Change – Human and Policy Dimensions, 18, 736-745.
Torry, W., 1979: Intelligence, resilience and change in complex social systems:
famine management in India. Mass Emergencies, 2, 71-85.
TRB, 2008: The Potential Impacts of Climate Change on U.S. Transportation. Special
Report 290, Transportation Research Board, National Academy of Sciences,
Washington, DC. (See Appendix C: Commissioned paper: Climate Variability and
Change with Implications for U.S. Transportation, Dec. 2006).
Tversky, A., and Kahneman, D., 1974: Judgment under uncertainty: Heuristics and
biases. Science, 185, 1124-1131.
Twigg, J., 2004: Disaster Risk Reduction: Mitigation and Preparedness in Development
and Emergency Programming. Overseas Development Institute, London, UK.
Ulbrich, U., A.H. Fink, M. Klawa, and J.G. Pinto, 2001: Three extreme storms over
Europe in December 1999. Weather, 56, 70-80.
UNDP, 2004: Reducing Disaster Risk: A Challenge for Development. A Global Report.
United Nations Development Programme, New York, NY.
UNEP, 1972: Declaration of the United Nations Conference on the Human
Environment. United Nations Environment Programme, Nairobi, Kenya.
UNISDR, 2005: Hyogo Framework for Action 2005-2015. Building the Resilience of
Nations and Communities to Disasters. United Nations International Strategy
for Disaster Reduction, Geneva, Switzerland.
UNISDR, 2008a: Climate Change and Disaster Risk Reduction. Briefing Note 01,
United Nations International Strategy for Disaster Reduction, Geneva,
Switzerland.
UNISDR, 2008b: Indigenous Knowledge for Disaster Risk Reduction - Good Practices
and Lessons Learned from Experiences in the Asia-Pacific Region. United
Nations International Strategy for Disaster Reduction, Bangkok, Thailand.
UNISDR, 2008c: Linking Disaster Risk Reduction and Poverty Reduction, Good
Practices and Lessons Learned. United Nations International Strategy for
Disaster Reduction, UNDP, Geneva, Switzerland.
UNISDR, 2009a: Adaptation to Climate Change by Reducing Disaster Risk: Country
Practices and Lessons. Briefing Note 02, United Nations International Strategy
for Disaster Reduction, Geneva, Switzerland.
UNISDR, 2009b: Reducing Disaster Risk through Science: Issues and Actions. Full
Report of the Scientific and Technical Committee 2009, United Nations
International Strategy for Disaster Reduction, Geneva, Switzerland.
UNISDR, 2009c: After Action Review: Second Session Global Platform for Disaster
Risk Reduction. United Nations International Strategy for Disaster Reduction,
Geneva, Switzerland.
UNISDR (United Nations International Strategy for Disaster Reduction), 2009d:
Terminology: Basic Terms of Disaster Risk Reduction. UNISDR, Geneva,
Switzerland, www.unisdr.org/eng/library/lib-terminology-eng%20home.htm.
UNISDR, 2009e: Global Assessment Report on Disaster Risk Reduction: Risk and
Poverty in a Changing Climate – Invest Today for a Safer Tomorrow. United
Nations International Strategy for Disaster Reduction, Geneva, Switzerland, 207
pp.
UNISDR, 2011: Global Assessment Report on Disaster Risk Reduction. Revealing
Risk, Redefining Development. United Nations International Strategy for
Disaster Reduction, Geneva, Switzerland.
Vale, L.J. and T.J. Campanella, 2005: The Resilient City: How Modern Cities Recover
from Disaster. Oxford University Press, Oxford, UK.
van Baars, S. and I.M. Van Kempen, 2009: The causes and mechanisms of historical
dike failures in the Netherlands. E-Water Official Publication of the European
Water Association, Hennef, Germany, 14 pp., www.dwa.de/portale/ewa/
ewa.nsf/C125723B0047EC38/8428F628AB57BECFC125766C003024B6/$FILE/
Historical%20Dike%20Failures.pdf.
van Kerkhoff, L. and L. Lebel, 2006: Linking knowledge and action for sustainable
development. Annual Review of Environment and Resources, 31,445-477.
van Niekerk, D., 2007: Disaster risk reduction, disaster risk management and disaster
management: Academic rhetoric or practical reality? Disaster Management
South Africa, 4(1), 6.
Vandentorren, S., F. Suzan, S. Medina, M. Pascal, A. Maulpoix, J.C. Cohen, and M.
Ledrans, 2004: Mortality in 13 French cities during the August 2003 heat wave.
American Journal of Public Health, 94(9), 1518-1520.
Venton, P. and S. La Trobe, 2008: Linking Climate Change Adaptation and Disaster
Risk Reduction. Technical Paper, Tearfund, Teddington, UK.
Walker, B.H., J.M. Anderies, A.P. Kinzig, and P. Ryan (eds), 2006: Exploring Resilience
in Social-Ecological Systems: Comparative Studies and Theory Development.
CSIRO Publishing, Collingwood, Australia.
Walker, W.E., A.W. Vincent, J. Marchau, and D. Swanson, 2010: Addressing deep
uncertainty using adaptive policies. Technological Forecasting and Social
Change, 77, 917-923.
Walshaw, D., 2000: Modelling extreme wind speeds in regions prone to hurricanes.
Journal of the Royal Statistical Society, Series C (Applied Statistics), 49, 51-62.
Watts, M., 1983: On the poverty of theory: natural hazards research in context. In:
Interpretations of Calamity [Hewitt, K. (ed.)]. Allen and Unwin, Boston, MA, pp.
231-262.
Weber, E.U., 2006: Experience-based and description-based perceptions of long-term
risk: why global warming does not scare us (yet). Climatic Change, 77, 103-120.
Weber, E.U., 2010: What shapes perceptions of climate change? Wiley
Interdisciplinary Reviews: Climate Change, 1, 332-342.
Weber, E.U. and P.C. Stern, 2011: Public understanding of climate change in the
United States. American Psychologist, 66(4), 315-328.
Weber, E.U., S. Shafir, and A.R. Blais, 2004: Predicting risk-sensitivity in humans and
lower animals: Risk as variance or coefficient of variation. Psychological
Review, 111, 430-445.
Weichselgartner, J., 2001: Disaster mitigation: the concept of vulnerability revisited.
Disaster Prevention and Management, 10(2), 85-94.
Werner, E.E., J.M. Bierman, and F.E. French, 1971: The Children of Kauai Honolulu.
University of Hawaii Press, Honolulu, Hawaii.
Westerling, A.L. and T.W. Swetman, 2003: Interannual to decadal drought and
wildfire in the Western United States. EOS, Transactions American Geophysical
Union,
84(49), 545-555, doi:10.1029/2003EO490001.
Chapter 1 Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
64
Wheater, H.S., 2002: Progress in and prospects for fluvial flood modeling.
Philosophical Transactions of the Royal Society A, 360, 1409-1431.
Wijkman, A. and L. Timberlake, 1988: Natural Disasters: Acts of God or Acts of Man?
New Society Publishers, Philadelphia, PA.
Wilby, R.L. and S. Dessai, 2010: Robust adaptation to climate change. Weather,
65(7), 180-185.
Wisner, B., P. O’Keefe, and K. Westgate, 1977: Global systems and local disasters: the
untapped power of peoples’ science. Disasters, 1(1), 47-57.
Wisner, B., P. Blaikie, T. Cannon, and I. Davis, 2004: At Risk: Natural Hazards, People’s
Vulnerability, and Disasters, 2nd edition. Routledge, London, UK.
Wisner, B., J.C. Gaillard, and I. Kellman (eds.), 2011: Handbook of Hazards and
Disaster Risk Reduction. Routledge, London, UK.
WMO, 2010: Understanding Climate. World Meteorological Organization, Geneva,
Switzerland, www.wmo.int/pages/themes/climate/understanding_climate.php.
World Bank, 2001: World Development Report 2000-2001. World Bank,
Washington, DC.
World Bank, 2009: World Development Report 2010: Development and Climate
Change. World Bank, Washington, DC.
WRI (World Resources Institute), United Nations Development Programme, United
Nations Environment Programme, and World Bank, 2008: World Resources
2008: Roots of Resilience: Growing the Wealth of the Poor. WRI, Washington,
DC.
Xie, L., L.J. Pietrafesa, and M. Peng, 2004: Incorporation of a mass-conserving
inundation scheme into a three dimensional storm surge model. Journal of
Coastal Research, 20, 1209-1233.
Yohe, G. and R.S.J. Tol, 2002: Indicators for social and economic coping capacity –
moving toward a working definition of adaptive capacity. Global Environmental
Change, 12(1), 25-40.
Young, P.C. 2002: Advances in real-time flood forecasting. Philosophical Transactions
of the Royal Society, A360, 1433-1450.
Zipser, E.J., D.J. Cecil, C. Liu, S.W. Nesbitt and D.P. Yorty, 2006: Where are the most
intense thunderstorms on Earth? Bulletin of the American Meteorological
Society, 87, 1057-1071.
Chapter 1Climate Change: New Dimensions in Disaster Risk, Exposure, Vulnerability, and Resilience
65
Coordinating Lead Authors:
Omar-Dario Cardona (Colombia), Maarten K. van Aalst (Netherlands)
Lead Authors:
Jörn Birkmann (Germany), Maureen Fordham (UK), Glenn McGregor (New Zealand), Rosa Perez
(Philippines), Roger S. Pulwarty (USA), E. Lisa F. Schipper (Sweden), Bach Tan Sinh (Vietnam)
Review Editors:
Henri Décamps (France), Mark Keim (USA)
Contributing Authors:
Ian Davis (UK), Kristie L. Ebi (USA), Allan Lavell (Costa Rica), Reinhard Mechler (Germany),
Virginia Murray (UK), Mark Pelling (UK), Jürgen Pohl (Germany), Anthony-Oliver Smith (USA),
Frank Thomalla (Australia)
This chapter should be cited as:
Cardona, O.D., M.K. van Aalst, J. Birkmann, M. Fordham, G. McGregor, R. Perez, R.S. Pulwarty, E.L.F. Schipper, and B.T. Sinh,
2012: Determinants of risk: exposure and vulnerability. In: Managing the Risks of Extreme Events and Disasters to Advance
Climate Change Adaptation [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach,
G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups I and II of the
Intergovernmental Panel on Climate Change (IPCC). Cambridge University Press, Cambridge, UK, and New York, NY, USA,
pp. 65-108.
2
Determinants of Risk:
Exposure and Vulnerability
Determinants of Risk: Exposure and Vulnerability
66
Executive Summary ...................................................................................................................................67
2.1. Introduction and Scope..............................................................................................................69
2.2. Defining Determinants of Risk: Hazard, Exposure, and Vulnerability ........................................69
2.2.1. Disaster Risk and Disaster .................................................................................................................................................69
2.2.2. The Factors of Risk.............................................................................................................................................................69
2.3. The Drivers of Vulnerability .......................................................................................................70
2.4. Coping and Adaptive Capacities ................................................................................................72
2.4.1. Capacity and Vulnerability .................................................................................................................................................72
2.4.2. Different Capacity Needs...................................................................................................................................................74
2.4.2.1. Capacity to Anticipate Risk.....................................................
..........................................................................................
..................74
2.4.2.2. Capacity to Respond...........................................................................................................................................................................74
2.4.2.3. Capacity to Recover and Change........................................................................................................................................................75
2.4.3. Factors of Capacity: Drivers and Barriers...
....
....................................................................................................................76
2.5. Dimensions and Trends of Vulnerability and Exposure ..............................................................76
2.5.1. Environmental Dimensions.................................................................................................................................................76
2.5.1.1. Physical Dimensions .................................................................................................................................................
..........................77
2.5.1.2. Geography, Location, Place ................................................................................................................................................................77
2.5.1.3. Settlement Patterns and Development Trajectories.............................................................................................................................78
2.5.2. Social Dimensions ...
....
.......................................................................................................................................................80
2.5.2.1. Demography ...................................................................................................................................................
....................................80
2.5.2.2. Education............................................................................................................................................................................................81
2.5.2.3. Health and Well-Being........................................................................................................................................................................82
2.5.2.4. Cultural Dimensions............................................................................................................................................................................84
2.5.2.5. Institutional and Governance Dimensions ..........................................................................................................................................85
2.5.3. Economic Dimensions...
....
..................................................................................................................................................86
2.5.4. Interactions, Cross-Cutting Themes, and Integrations.......................................................................................................87
2.5.4.1. Intersectionality and Other Dimensions.....................................................
........................................................................................
.88
2.5.4.2. Timing, Spatial, and Functional Scales................................................................................................................................................88
2.5.4.3. Science and Technology......................................................................................................................................................................89
2.6. Risk Identification and Assessment ...........................................................................................89
2.6.1. Risk Identification ..............................................................................................................................................................90
2.6.2. Vulnerability and Risk Assessment.....................................................................................................................................90
2.6.3. Risk Communication...........................................................................................................................................................95
2.7. Risk Accumulation and the Nature of Disasters ........................................................................95
References .................................................................................................................................................96
Chapter 2
Table of Contents
67
The severity of the impacts of extreme and non-extreme weather and climate events depends strongly on
the level of vulnerability and exposure to these events (high confidence). [2.2.1, 2.3, 2.5] Trends in vulnerability
and exposure are major drivers of changes in disaster risk, and of impacts when risk is realized (high confidence). [2.5]
Understanding the multi-faceted nature of vulnerability and exposure is a prerequisite for determining how weather
and climate events contribute to the occurrence of disasters, and for designing and implementing effective adaptation
and disaster risk management strategies. [2.2, 2.6]
Vulnerability and exposure are dynamic, varying across temporal and spatial scales, and depend on economic,
social, geographic, demographic, cultural, institutional, governance, and environmental factors (high
confidence). [2.2, 2.3, 2.5] Individuals and communities are differentially exposed and vulnerable and this is based
on factors such as wealth, education, race/ethnicity/religion, gender, age, class/caste, disability, and health status. [2.5]
Lack of resilience and capacity to anticipate, cope with, and adapt to extremes and change are important causal factors
of vulnerability. [2.4]
Extreme and non-extreme weather and climate events also affect vulnerability to future extreme events,
by modifying the resilience, coping, and adaptive capacity of communities, societies, or social-ecological
systems affected by such events (high confidence). [2.4.3] At the far end of the spectrum – low-probability, high-
intensity events – the intensity of extreme climate and weather events and exposure to them tend to be more pervasive
in explaining disaster loss than vulnerability in explaining the level of impact. But for less extreme events – higher
probability, lower intensity – the vulnerability of exposed elements plays an increasingly important role (high
confidence). [2.3] The cumulative effects of small- or medium-scale, recurrent disasters at the sub-national or local
levels can substantially affect livelihood options and resources and the capacity of societies and communities to
prepare for and respond to future disasters. [2.2.1, 2.7]
High vulnerability and exposure are generally the outcome of skewed development processes, such as
those associated with environmental mismanagement, demographic changes, rapid and unplanned
urbanization in hazardous areas, failed governance, and the scarcity of livelihood options for the poor
(high confidence). [2.2.2, 2.5]
The selection of appropriate vulnerability and risk evaluation approaches depends on the decisionmaking
context (high confidence). [2.6.1] Vulnerability and risk assessment methods range from global and national
quantitative assessments to local-scale qualitative participatory approaches. The appropriateness of a specific method
depends on the adaptation or risk management issue to be addressed, including for instance the time and geographic
scale involved, the number and type of actors, and economic and governance aspects. Indicators, indices, and
probabilistic metrics are important measures and techniques for vulnerability and risk analysis. However, quantitative
approaches for assessing vulnerability need to be complemented with qualitative approaches to capture the full
complexity and the various tangible and intangible aspects of vulnerability in its different dimensions. [2.6]
Appropriate and timely risk communication is critical for effective adaptation and disaster risk management
(high confidence). Effective risk communication is built on risk assessment, and tailored to a specific audience, which
may range from decisionmakers at various levels of government, to the private sector and the public at large, including
local communities and specific social groups. Explicit characterization of uncertainty and complexity strengthens risk
communication. Impediments to information flows and limited awareness are risk amplifiers. Beliefs, values, and
norms influence risk perceptions, risk awareness, and choice of action. [2.6.3]
Adaptation and risk management policies and practices will be more successful if they take the dynamic
nature of vulnerability and exposure into account, including the explicit characterization of uncertainty
and complexity at each stage of planning and practice (medium evidence, high agreement). However,
approaches to representing such dynamics quantitatively are currently underdeveloped. Projections of the impacts of
Chapter 2 Determinants of Risk: Exposure and Vulnerability
Executive Summary
68
climate change can be strengthened by including storylines of changing vulnerability and exposure under different
development pathways. Appropriate attention to the temporal and spatial dynamics of vulnerability and exposure is
particularly important given that the design and implementation of adaptation and risk management strategies and
policies can reduce risk in the short term, but may increase vulnerability and exposure over the longer term. For
instance, dike systems can reduce hazard exposure by offering immediate protection, but also encourage settlement
patterns that may increase risk in the long term. [2.4.2.1, 2.5.4.2, 2.6.2]
Vulnerability reduction is a core common element of adaptation and disaster risk management (high
confidence). Vulnerability reduction thus constitutes an important common ground between the two areas of policy
and practice. [2.2, 2.3]
Chapter 2Determinants of Risk: Exposure and Vulnerability
69
2.1. Introduction and Scope
Many climate change adaptation efforts aim to address the implications
of potential changes in the frequency, intensity, and duration of weather
and climate events that affect the risk of extreme impacts on human
society. That risk is determined not only by the climate and weather
events (the hazards) but also by the exposure and vulnerability to these
hazards. Therefore, effective adaptation and disaster risk management
strategies and practices also depend on a rigorous understanding of the
dimensions of exposure and vulnerability, as well as a proper assessment
of changes in those dimensions. This chapter aims to provide that
understanding and assessment, by further detailing the determinants of
risk as presented in Chapter 1.
The first sections of this chapter elucidate the concepts that are needed
to define and understand risk, and show that risk originates from a
combination of social processes and their interaction with the environment
(Sections 2.2 and 2.3), and highlight the role of coping and adaptive
capacities (Section 2.4). The following section (2.5) describes the different
dimensions of vulnerability and exposure as well as trends therein.
Given that exposure and vulnerability are highly context-specific, this
section is by definition limited to a general overview (a more quantitative
perspective on trends is provided in Chapter 4). A methodological
discussion (Section 2.6) of approaches to identify and assess risk provides
indications of how the dimensions of exposure and vulnerability can be
explored in specific contexts, such as adaptation planning, and the
central role of risk perception and risk communication. The chapter
concludes with a cross-cutting discussion of risk accumulation and the
nature of disasters (Section 2.7).
2.2. Defining Determinants of Risk:
Hazard, Exposure, and Vulnerability
2.2.1. Disaster Risk and Disaster
Disaster risk signifies the possibility of adverse effects in the future. It
derives from the interaction of social and environmental processes, from
the combination of physical hazards and the vulnerabilities of exposed
elements (see Chapter 1). The hazard event is not the sole driver of risk,
and there is high confidence that the levels of adverse effects are in
good part determined by the vulnerability and exposure of societies and
social-ecological systems (UNDRO, 1980; Cuny, 1984; Cardona, 1986,
1993, 2011; Davis and Wall, 1992; UNISDR, 2004, 2009b; Birkmann,
2006a,b; van Aalst 2006a).
Disaster risk is not fixed but is a continuum in constant evolution. A
disaster is one of its many ‘moments’ (ICSU-LAC, 2010a,b), signifying
unmanaged risks that often serve to highlight skewed development
problems (Westgate and O’Keefe, 1976; Wijkman and Timberlake, 1984).
Disasters may also be seen as the materialization of risk and signify ‘a
becoming real’ of this latent condition that is in itself a social construction
(see below; Renn, 1992; Adam and Van Loon, 2000; Beck, 2000, 2008).
Disaster risk is associated with differing levels and types of adverse
effects. The effects may assume catastrophic levels or levels commensurate
with small disasters. Some have limited financial costs but very high
human costs in terms of loss of life and numbers of people affected;
others have very high financial costs but relatively limited human costs.
Furthermore, there is high confidence that the cumulative effects of
small disasters can affect capacities of communities, societies, or social-
ecological systems to deal with future disasters at sub-national or local
levels (Alexander, 1993, 2000; Quarantelli, 1998; Birkmann, 2006b;
Marulanda et al., 2008b, 2010, 2011; UNISDR, 2009a).
2.2.2. The Factors of Risk
As detailed in Section 1.1, hazard refers to the possible, future occurrence
of natural or human-induced physical events that may have adverse
effects on vulnerable and exposed elements (White, 1973; UNDRO,
1980; Cardona, 1990; UNDHA, 1992; Birkmann, 2006b). Although, at
times, hazard has been ascribed the same meaning as risk, currently it
is widely accepted that it is a component of risk and not risk itself.
The intensity or recurrence of hazard events can be partly determined
by environmental degradation and human intervention in natural
ecosystems. Landslides or flooding regimes associated with human-
induced environmental alteration and new climate change-related
hazards are examples of such socio-natural hazards (Lavell, 1996,
1999a).
Exposure refers to the inventory of elements in an area in which hazard
events may occur (Cardona, 1990; UNISDR, 2004, 2009b). Hence, if
population and economic resources were not located in (exposed to)
potentially dangerous settings, no problem of disaster risk would exist.
While the literature and common usage often mistakenly conflate
exposure and vulnerability, they are distinct. Exposure is a necessary,
but not sufficient, determinant of risk. It is possible to be exposed
but not vulnerable (for example by living in a floodplain but having
sufficient means to modify building structure and behavior to mitigate
potential loss). However, to be vulnerable to an extreme event, it is
necessary to also be exposed.
Land use and territorial planning are key factors in risk reduction. The
environment offers resources for human development at the same
time as it represents exposure to intrinsic and fluctuating hazardous
conditions. Population dynamics, diverse demands for location, and
the gradual decrease in the availability of safer lands mean it is
almost inevitable that humans and human endeavor will be located in
potentially dangerous places (Lavell, 2003). Where exposure to events is
impossible to avoid, land use planning and location decisions can be
accompanied by other structural or non-structural methods for preventing
or mitigating risk (UNISDR, 2009a; ICSU-LAC, 2010a,b).
Vulnerability refers to the propensity of exposed elements such as
human beings, their livelihoods, and assets to suffer adverse effects
when impacted by hazard events (UNDRO, 1980; Cardona, 1986, 1990,
Chapter 2 Determinants of Risk: Exposure and Vulnerability
70
1993; Liverman, 1990; Maskrey, 1993b; Cannon, 1994, 2006; Blaikie et
al., 1996; Weichselgartner, 2001; Bogardi and Birkmann, 2004; UNISDR,
2004, 2009b; Birkmann, 2006b; Janssen et al., 2006; Thywissen, 2006).
Vulnerability is related to predisposition, susceptibilities, fragilities,
weaknesses, deficiencies, or lack of capacities that favor adverse effects
on the exposed elements. Thywissen (2006) and Manyena (2006) car-
ried out an extensive review of the terminology. The former includes a
long list of definitions used for the term vulnerability and the latter
includes definitions of vulnerability and resilience and their relationship.
An early view of vulnerability in the context of disaster risk management
was related to the physical resistance of engineering structures (UNDHA,
1992), but more recent views relate vulnerability to characteristics of
social and environmental processes. It is directly related, in the context
of climate change, to the susceptibility, sensitivity, and lack of resilience
or capacities of the exposed system to cope with and adapt to extremes
and non-extremes (Luers et al., 2003; Schröter et al., 2005; Brklacich
and Bohle, 2006; IPCC, 2001, 2007).
While vulnerability is a key concept for both disaster risk and climate
change adaptation, the term is employed in numerous other contexts,
for instance to refer to epidemiological and psychological fragilities,
ecosystem sensitivity, or the conditions, circumstances, and drivers that
make people vulnerable to natural and economic stressors (Kasperson
et al., 1988; Cutter, 1994; Wisner et al., 2004; Brklacich and Bohle, 2006;
Haines et al., 2006; Villagrán de León, 2006). It is common to find
blanket descriptions of the elderly, children, or women as ‘vulnerable,’
without any indication as to what these groups are vulnerable to
(Wisner, 1993; Enarson and Morrow, 1998; Morrow, 1999; Bankoff,
2004; Cardona, 2004, 2011).
Vulnerability can be seen as situation-specific, interacting with a hazard
event to generate risk (Lavell, 2003; Cannon, 2006; Cutter et al., 2008).
Vulnerability to financial crisis, for example, does not infer vulnerability
to climate change or natural hazards. Similarly, a population might be
vulnerable to hurricanes, but not to landslides or floods. From a climate
change perspective, basic environmental conditions change progressively
and then induce new risk conditions for societies. For example, more
frequent and intense events may introduce factors of risk into new
areas, revealing underlying vulnerability. In fact, future vulnerability is
embedded in the present conditions of the communities that may be
exposed in the future (Patt et al., 2005, 2009); that is, new hazards in
areas not previously subject to them will reveal, not necessarily create,
underlying vulnerability factors (Alwang et al., 2001; Cardona et al.,
2003a; Lopez-Calva and Ortiz, 2008; UNISDR, 2009a).
While vulnerability is in general hazard-specific, certain factors, such as
poverty, and the lack of social networks and social support mechanisms,
will aggravate or affect vulnerability levels irrespective of the type of
hazard. These types of generic factors are different from the hazard-
specific factors and assume a different position in the intervention
actions and the nature of risk management and adaptation processes
(ICSU-LAC, 2010a,b). Vulnerability of human settlements and ecosystems
is intrinsically tied to different socio-cultural and environmental
processes (Kasperson et al., 1988; Cutter, 1994; Adger, 2006; Cutter and
Finch, 2008; Cutter et al., 2008; Williams et al., 2008; Décamps, 2010;
Dawson et al., 2011). Vulnerability is linked also to deficits in risk
communication, especially the lack of appropriate information that can
lead to false risk perceptions (Birkmann and Fernando, 2008), which
have an important influence on the motivation and perceived ability to
act or to adapt to climate change and environmental stressors
(Grothmann and Patt, 2005). Additionally, processes of maladaptation
or unsustainable adaptation can increase vulnerability and risks
(Birkmann, 2011a).
Vulnerability in the context of disaster risk management is the most
palpable manifestation of the social construction of risk (Aysan, 1993;
Blaikie et al., 1996; Wisner et al., 2004; ICSU-LAC 2010a,b). This notion
underscores that society, in its interaction with the changing physical
world, constructs disaster risk by transforming physical events into
hazards of different intensities or magnitudes through social processes
that increase the exposure and vulnerability of population groups, their
livelihoods, production, support infrastructure, and services (Chambers,
1989; Wilches-Chaux, 1989; Cannon, 1994; Wisner et al., 2004; Wisner,
2006a; Carreño et al., 2007a; ICSU-LAC, 2010a,b). This includes:
How human action influences the levels of exposure and
vulnerability in the face of different physical events
How human intervention in the environment leads to the creation
of new hazards or an increase in the levels or damage potential of
existing ones
How human perception, understanding, and assimilation of the
factors of risk influence societal reactions, prioritization, and
decisionmaking processes.
There is high agreement and robust evidence that high vulnerability and
exposure are mainly an outcome of skewed development processes,
including those associated with environmental mismanagement,
demographic changes, rapid and unplanned urbanization, and the scarcity
of livelihood options for the poor (Maskrey, 1993a,b, 1994, 1998; Mansilla,
1996; Lavell, 2003; Cannon, 2006; ICSU-LAC, 2010a,b; Cardona, 2011).
Increases in disaster risk and the occurrence of disasters have been in
evidence over the last five decades (Munich Re, 2011) (see Section 1.1.1).
This trend may continue and may be enhanced in the future as a result
of projected climate change, further demographic and socioeconomic
changes, and trends in governance, unless concerted actions are enacted
to reduce vulnerability and to adapt to climate change, including
interventions to address disaster risks (Lavell, 1996, 1999a, 2003; ICSU-
LAC, 2010a,b; UNISDR, 2011).
2.3. The Drivers of Vulnerability
In order to effectively manage risk, it is essential to understand how
vulnerability is generated, how it increases, and how it builds up
(Maskrey, 1989; Cardona, 1996a, 2004, 2011; Lavell, 1996, 1999a;
Chapter 2Determinants of Risk: Exposure and Vulnerability
71
O’Brien et al., 2004b). Vulnerability describes a set of conditions of
people that derive from the historical and prevailing cultural, social,
environmental, political, and economic contexts. In this sense, vulnerable
groups are not only at risk because they are exposed to a hazard but as
a result of marginality, of everyday patterns of social interaction and
organization, and access to resources (Watts and Bohle, 1993; Morrow,
1999; Bankoff, 2004). Thus, the effects of a disaster on any particular
household result from a complex set of drivers and interacting conditions.
It is important to keep in mind that people and communities are not
only or mainly victims, but also active managers of vulnerability (Ribot,
1996; Pelling, 1997, 2003). Therefore, integrated and multidimensional
approaches are highly important to understanding causes of vulnerability.
Some global processes are significant drivers of risk and are particularly
related to vulnerability creation. There is high confidence that these
include population growth, rapid and inappropriate urban development,
international financial pressures, increases in socioeconomic inequalities,
trends and failures in governance (e.g., corruption, mismanagement),
and environmental degradation (Maskrey, 1993a,b, 1994, 1998; Mansilla,
1996; Cannon, 2006). Vulnerability profiles can be constructed that take
into consideration sources of environmental, social, and economic
marginality (Wisner, 2003). This also includes the consideration of the
links between communities and specific environmental services, and the
vulnerability of ecosystem components (Renaud, 2006; Williams et al.,
2008; Décamps, 2010; Dawson et al., 2011). In climate change-related
impact assessments, integration of underlying ‘causes of vulnerability’
and adaptive capacity is needed rather than focusing on technical
aspects only (Ribot, 1995; O’Brien et al., 2004b).
Due to different conceptual frameworks and definitions, as well as
disciplinary views, approaches to address the causes of vulnerability
also differ (Burton et al., 1983; Blaikie et al., 1994; Harding et al., 2001;
Twigg, 2001; Adger and Brooks, 2003, 2006; Turner et al., 2003a,b;
Cardona, 2004; Schröter et al., 2005; Adger 2006; Füssel and Klein, 2006;
Villagrán de León, 2006; Cutter and Finch, 2008; Cutter et al., 2008).
Thomalla et al. (2006), Mitchell and van Aalst (2008), and Mitchell et al.
(2010) examine commonalities and differences between the adaptation
to climate change and disaster risk management communities, and
identify key areas of difference and convergence. The two communities
tend to perceive the nature and timescale of the threat differently:
impacts due to climate change and return periods for extreme events
frequently use the language of uncertainty; but considerable knowledge
and certainty has been expressed regarding event characteristics and
exposures related to extreme historical environmental conditions.
Four approaches to understanding vulnerability and its causes can be
distinguished, rooted in political economy, social-ecology, vulnerability,
and disaster risk assessment, as well as adaptation to climate change:
1) The pressure and release (PAR) model (Blaikie et al., 1994, 1996;
Wisner et al., 2004) is common to social science-related vulnerability
research and emphasizes the social conditions and root causes of
exposure more than the hazard as generating unsafe conditions.
This approach links vulnerability to unsafe conditions in a continuum
that connects local vulnerability to wider national and global shifts
in the political economy of resources and political power.
2) The social ecology perspective emphasizes the need to focus on
coupled human-environmental systems (Hewitt and Burton, 1971;
Turner et al., 2003a,b). This perspective stresses the ability of
societies to transform nature and also the implications of changes
in the environment for social and economic systems. It argues that
the exposure and susceptibility of a system can only be adequately
understood if these coupling processes and interactions are
addressed.
3) Holistic perspectives on vulnerability aim to go beyond technical
modeling to embrace a wider and comprehensive explanation of
vulnerability. These approaches differentiate exposure, susceptibility
and societal response capacities as causes or factors of vulnerability
(see Cardona, 1999a, 2001, 2011; Cardona and Barbat, 2000;
Cardona and Hurtado, 2000a,b; IDEA, 2005; Birkmann, 2006b;
Carreño, 2006; Carreño et al.,2007a,b, 2009; Birkmann and Fernando,
2008). A core element of these approaches is the feedback loop
which underlines that vulnerability is dynamic and is the main
driver and determinant of current or future risk.
4) In the context of climate change adaptation, different vulnerability
definitions and concepts have been developed and discussed. One
of the most prominent definitions is the one reflected in the IPCC
Fourth Assessment Report, which describes vulnerability as a
function of exposure, sensitivity, and adaptive capacity, as also
reflected by, for instance, McCarthy et al. (2001), Brooks (2003),
K. O’Brien et al. (2004a), Füssel and Klein (2006), Füssel (2007),
and G. O’Brien et al. (2008). This approach differs from the
understanding of vulnerability in the disaster risk management
perspective, as the rate and magnitude of climate change is
considered. The concept of vulnerability here includes external
environmental factors of shock or stress. Therefore, in this view, the
magnitude and frequency of potential hazard events is to be
considered in the vulnerability to climate change. This view also
differs in its focus upon long-term trends and stresses rather than
on current shock forecasting, something not explicitly excluded but
rather rarely considered within the disaster risk management
approaches.
The lack of a comprehensive conceptual framework that facilitates a
common multidisciplinary risk evaluation impedes the effectiveness of
disaster risk management and adaptation to climate change (Cardona,
2004). The option for anticipatory disaster risk reduction and adaptation
exists precisely because risk is a latent condition, which announces
potential future adverse effects (Lavell, 1996, 1999a). Understanding
disaster risk management as a social process allows for a shift in focus
from responding to the disaster event toward an understanding of
disaster risk (Cardona and Barbat, 2000; Cardona et al., 2003a). This
requires knowledge about how human interactions with the natural
environment lead to the creation of new hazards, and how persons,
property, infrastructure, goods, and the environment are exposed to
potentially damaging events. Furthermore, it requires an understanding
of the vulnerability of people and their livelihoods, including the
Chapter 2 Determinants of Risk: Exposure and Vulnerability
72
allocation and distribution of social and economic resources that can work
for or against the achievement of resistance, resilience, and security (ICSU-
LAC, 2010a,b). Overall, there is high confidence that although hazard
events are usually considered the cause of disaster risk, vulnerability
and exposure are its key determining factors. Furthermore, contrary to
the hazard, vulnerability and exposure can often be influenced by policy
and practice, including in the short to medium term. Therefore disaster
risk management and adaptation strategies have to address mainly
these same risk factors (Cardona 1999a, 2011; Vogel and O’Brien, 2004;
Birkmann, 2006a; Leichenko and O’Brien, 2008).
Despite various frameworks developed for defining and assessing
vulnerability, it is interesting to note that at least some common causal
factors of vulnerability have been identified, in both the disaster risk
management and climate change adaptation communities (see
Cardona, 1999b, 2001, 2011; Cardona and Barbat, 2000; Cardona and
Hurtado, 2000a,b; McCarthy et al., 2001; Gallopin, 2006; Manyena,
2006; Carreño et al., 2007a, 2009; IPCC, 2007; ICSU-LAC 2010a,b;
MOVE, 2010):
Susceptibility/fragility (in disaster risk management) or sensitivity
(in climate change adaptation): physical predisposition of human
beings, infrastructure, and environment to be affected by a dangerous
phenomenon due to lack of resistance and predisposition of society
and ecosystems to suffer harm as a consequence of intrinsic and
context conditions making it plausible that such systems once
impacted will collapse or experience major harm and damage due
to the influence of a hazard event.
Lack of resilience (in disaster risk management) or lack of coping
and adaptive capacities (in climate change adaptation): limitations
in access to and mobilization of the resources of the human beings
and their institutions, and incapacity to anticipate, adapt, and
respond in absorbing the socio-ecological and economic impact.
There is high confidence that at the extreme end of the spectrum, the
intensity of extreme climate and weather events – low-probability,
high-intensity – and exposure to them tend to be more pervasive in
explaining disaster loss than vulnerability itself. But as the events get
less extreme – higher-probability, lower-intensity – the vulnerability of
exposed elements plays an increasingly important role in explaining the
level of impact. Vulnerability is a major cause of the increasing adverse
effects of non-extreme events, that is, small recurrent disasters that
many times are not visible at the national or sub-national level
(Marulanda et al., 2008b, 2010, 2011; UNISDR, 2009a; Cardona, 2011;
UNISDR, 2011).
Overall, the promotion of resilient and adaptive societies requires a
paradigm shift away from the primary focus on natural hazards and
extreme weather events toward the identification, assessment, and
ranking of vulnerability (Maskrey, 1993a; Lavell, 2003; Birkmann,
2006a,b). Therefore, understanding vulnerability is a prerequisite for
understanding risk and the development of risk reduction and adaptation
strategies to extreme events in the light of climate change (ICSU-LAC,
2010a,b; MOVE, 2010; Cardona, 2011; UNISDR, 2011).
2.4. Coping and Adaptive Capacities
Capacity is an important element in most conceptual frameworks of
vulnerability and risk. It refers to the positive features of people’s
characteristics that may reduce the risk posed by a certain hazard.
Improving capacity is often identified as the target of policies and projects,
based on the notion that strengthening capacity will eventually lead to
reduced risk. Capacity clearly also matters for reducing the impact of
climate change (e.g., Sharma and Patwardhan, 2008).
As presented in Chapter 1, coping is typically used to refer to ex post
actions, while adaptation is normally associated with ex ante actions.
This implies that coping capacity also refers to the ability to react to and
reduce the adverse effects of experienced hazards, whereas adaptive
capacity refers to the ability to anticipate and transform structure,
functioning, or organization to better survive hazards (Saldaña-Zorrilla,
2007). Presence of capacity suggests that impacts will be less extreme
and/or the recovery time will be shorter, but high capacity to recover
does not guarantee equal levels of capacity to anticipate. In other
words, the capacity to cope does not infer the capacity to adapt
(Birkmann, 2011a), although coping capacity is often considered to
be part of adaptive capacity (Levina and Tirpak, 2006).
2.4.1. Capacity and Vulnerability
Most risk studies prior to the 1990s focused mainly on hazards,
whereas the more recent reversal of this paradigm has placed equal
focus on the vulnerability side of the equation. Emphasizing that risk
can be reduced through vulnerability is an acknowledgement of the
power of social, political, environmental, and economic factors in driving
risk. While these factors drive risk on one hand, they can on the other
hand be the source of capacity to reduce it (Carreño et al., 2007a;
Gaillard, 2010).
Many approaches for assessing vulnerability rely on an assessment of
capacity as a baseline for understanding how vulnerable people are to
a specific hazard. The relationship between capacity and vulnerability
is described differently among different schools of thought, stemming
from different uses in the fields of development, disaster risk
management, and climate change adaptation. Gaillard (2010) notes
that the concept of capacity “played a pivotal role in the progressive
emergence of the vulnerability paradigm within the scientific realm.
On the whole, the literature describes the relationship between
vulnerability and capacity in two ways, which are not mutually exclusive
(Bohle, 2001; IPCC, 2001; Moss et al., 2001; Yodmani, 2001; Downing
and Patwardhan, 2004; Brooks et al., 2005; Smit and Wandel, 2006;
Gaillard, 2010):
1) Vulnerability is, among other things, the result of a lack of
capacity.
2) Vulnerability is the opposite of capacity, so that increasing
capacity means reducing vulnerability, and high vulnerability
means low capacity.
Chapter 2Determinants of Risk: Exposure and Vulnerability
73
The relationship between capacity and vulnerability is interpreted
differently in the climate change community of practice and the
disaster risk management community of practice. Throughout the
1980s, vulnerability became a central focus of much work on disasters,
in some circles overshadowing the role played by hazards in driving risk.
Some have noted that the emphasis on vulnerability tended to ignore
capacity, focusing too much on the negative aspects of vulnerability
(Davis et al., 2004). Recognizing the role of capacity in reducing risk also
indicates an acknowledgement that people are not ‘helpless victims’
(Bohle, 2001; Gaillard, 2010).
In many climate change-related studies, capacity was initially subsumed
under vulnerability. The first handbooks and guidelines for adaptation
emphasized impacts and vulnerability assessment as the necessary steps
for determining adaptation options (Kate, 1985; Carter et al., 1994; Benioff
et al., 1996; Feenstra et al., 1998). Climate change vulnerability was often
placed in direct opposition to capacity. Vulnerability that was measured
was seen as the remainder after capacity had been taken into account.
However, Davis et al. (2004), IDEA (2005), Carreño et al. (2007a,b), and
Gaillard (2010) note that capacity and vulnerability are not necessarily
Chapter 2 Determinants of Risk: Exposure and Vulnerability
Box 2-1 | Coping and Adaptive Capacity: Different Origins and Uses
As set out in Section 1.4, there is a difference in understanding and use of the terms coping and adapting. Although coping capacity is
often used interchangeably with adaptive capacity in the climate change literature, Cutter et al. (2008) point out that adaptive capacity
features more frequently in global environmental change perspectives and is less prevalent in the hazards discourse.
Adaptive capacity refers to the ability of a system or individual to adapt to climate change, but it can also be used in the context of
disaster risk. Because adaptive capacity is considered to determine “the ability of an individual, family, community, or other social group
to adjust to changes in the environment guaranteeing survival and sustainability” (Lavell, 1999b), many believe that in the context of
uncertain environmental changes, adaptive capacity will be of key significance. Dayton-Johnson (2004) defines adaptive capacity as the
“vulnerability of a society before disaster strikes and its resilience after the fact. Some ways of classifying adaptive capacity include
‘baseline adaptive capacity’ (Dore and Etkin, 2003), which refers to the capacity that allows countries to adapt to existing climate
variability, and ‘socially optimal adaptive capacity,’ which is determined by the norms and rules in individual locations. Another definition
of adaptive capacity is the “property of a system to adjust its characteristics or behavior, in order to expand its coping range under
existing climate variability, or future climate conditions” (Brooks and Adger, 2004). This links adaptive capacity to coping capacity,
because coping range is synonymous with coping capacity, referring to the boundaries of systems’ ability to cope (Yohe and Tol, 2002).
In simple terms, coping capacity refers to the “ability of people, organizations, and systems, using available skills and resources, to face
and manage adverse conditions, emergencies, or disasters” (UNISDR, 2009b). Coping capacity is typically used in humanitarian discourse
to indicate the extent to which a system can survive the impacts of an extreme event. It suggests that people can deal with some
degree of destabilization, and acknowledges that at a certain point this capacity may be exceeded. Eriksen et al. (2005) link coping
capacity to entitlements – the set of commodity bundles that can be commanded – during an adverse event. The ability to mobilize this
capacity in an emergency is the manifestation of coping strategies (Gaillard, 2010). Furthermore, Birkmann (2011b) underscores that
differences between coping and adaptation are also linked to the quality of the response process. While coping aims to maintain the
system and its functions in the face of adverse conditions, adaptation involves changes and requires reorganization processes.
The capacity described by the disasters community in the past decades does not frequently distinguish between ‘coping’ or ‘adaptive’
capacities, and instead the term is used to indicate positive characteristics or circumstances that could be seen to offset vulnerability
(Anderson and Woodrow, 1989). Because the approach is focused on disasters, it has been associated with the immediate-term coping
needs, and contrasts from the long-term perspective generally discussed in the context of climate change, where the aim is to adapt to
changes rather than to just overcome them. There has been considerable discussion throughout the vulnerability and poverty and climate
change scholarly communities about whether coping strategies are a stepping stone toward adaptation, or may lead to maladaptation
(Yohe and Tol, 2002; Eriksen et al., 2005) (see Chapter 1). Useful alternative terminology is to talk about ‘capacity to change and adjust’
(Nelson and Finan, 2009) for adaptive capacity, and ‘capacity to absorb’ instead of coping capacity (Cutter et al., 2008).
In the climate change community of practice, adaptive capacity has been at the forefront of thinking regarding how to respond to the
impacts of climate change, but it was initially seen as a characteristic to build interventions on, and only later has been recognized as
the target of interventions (Adger et al., 2004). The United Nations Framework Convention on Climate Change, for instance, states in its
ultimate objective that action to reduce greenhouse gas emissions be guided by the time needed for ecosystems to adapt naturally to
the impacts of climate change.
74
opposites, because communities that are highly vulnerable may in fact
display high capacity in certain aspects. This reflects the many elements
of risk reduction and the multiple capacity needs across them. Alwang
et al. (2001) also underscore that vulnerability is dynamic and determined
by numerous factors, thus high capacity in the ability to respond to an
extreme event does not accurately reflect low vulnerability.
2.4.2. Different Capacity Needs
The capacity necessary to anticipate and avoid being affected by an
extreme event requires different assets, opportunities, social networks,
and local and external institutions from capacity to deal with impacts
and recover from them (Lavell, 1994; Lavell and Franco, 1996; Cardona,
2001, 2010; Carreño et al., 2007a,b; ICSU-LAC, 2010a,b; MOVE, 2010).
Capacity to change relies on yet another set of factors. Importantly,
however, these dimensions of capacity are not unrelated to each other:
the ability to change is also necessary for risk reduction and response
capacities.
Just like vulnerability, capacity is dynamic and will change depending
on circumstances. The discussion in Box 2-1 indicates that there are
differing perspectives on how coping and adaptive capacity relate.
When coping and adapting are viewed as different, it follows that the
capacity needs for each are also different (Cooper et al., 2008). This
suggests that work done to understand the drivers of adaptive ex ante
capacity (Leichenko and O’Brien, 2002; Yohe and Tol, 2002; Brenkert and
Malone, 2005; Brooks et al., 2005; Haddad, 2005; Vincent, 2007; Sharma
and Patwardhan, 2008; Magnan, 2010) may not be similar with the
identified drivers of capacities that helped in the past (ex post) and are
associated more closely with experienced coping processes. Many of
these elements are reflected in local, national, and international
contexts in Chapters 5, 6, and 7 of this Special Report.
2.4.2.1. Capacity to Anticipate Risk
Having the capacity to reduce the risk posed by hazards and changes
implies that people’s ability to manage is not engulfed, so they are not
left significantly worse off. Reducing risk means that people do not have
to devote substantial resources to dealing with a hazard as it occurs, but
instead have the capacity to anticipate this sort of event. This is the type
of capacity that is necessary in order to adapt to climate change, and
involves conscious, planned efforts to reduce risk. The capacity to reduce
risk also depends on ex post actions, which involve making choices after
one event that reduce the impact of future events.
Capacity for risk prevention and reduction may be understood as a
series of elements, measures, and tools directed toward intervention in
hazards and vulnerabilities with the objective of reducing existing or
controlling future possible risks (Cardona et al., 2003a). This can range
from guaranteeing survival to the ability to secure future livelihoods
(Batterbury, 2001; Eriksen and Silva, 2009).
Development planning, including land use and urban planning, river basin
and land management, hazard-resistant building codes, and landscape
design are all activities that can reduce exposure and vulnerability to
hazards and change (Cardona, 2001, 2010). The ability to carry these out
in an effective way is part of the capacity to reduce risk. Other activities
include diversifying income sources, maintaining social networks, and
collective action to avoid development that puts people at higher risk
(Maskrey, 1989, 1994; Lavell, 1994, 1999b, 2003).
Up to the early 1990s, disaster preparedness and humanitarian response
dominated disaster practice, and focus on capacity was limited to
understanding inherent response capacity. Thus, emphasizing capacity to
reduce risk was not a priority. However, in the face of growing evidence
as to significant increases in disaster losses and the inevitable increase
in financial and human resources dedicated to disaster response and
recovery, there is an increasing recognition of the need to promote the
capacity for prevention and risk reduction over time (Lavell, 1994, 1999b,
2003). Notwithstanding, different actors, stakeholders, and interests
influence the capacity to anticipate a disaster. Actions to reduce exposure
and vulnerability of one group of people may come at the cost of
increasing it for another, for example when flood risks are shifted from
upstream communities to downstream communities through large-
scale upstream dike construction (Birkmann, 2011a). Consequently, it
is not sufficient to evaluate the success of adaptation or capacities
to reduce risk by focusing on the objectives of one group only. The
evaluation of success of adaptation strategies depends on the spatial
and temporal scale used (Adger et al., 2005).
2.4.2.2. Capacity to Respond
Capacity to respond is relevant both ex post and ex ante, since it
encompasses everything necessary to be able to react once an extreme
event takes place. Response capacity is mostly used to refer to the
ability of institutions to react following a natural hazard, in particular
ex post during emergency response. However, effective response
requires substantial ex ante planning and investments in disaster
preparedness and early warning (not only in terms of financial cost but
particularly in terms of awareness raising and capacity building; IFRC,
2009). Furthermore, there are also response phases for gradual changes
in ecosystems or temperature regimes caused by climate change.
Responding spans everything from people’s own initial reactions to a
hazard upon its impact to actions to try to reduce secondary damage. It
is worth noting that in climate change literature, anticipatory actions
are often referred to as responses, which differs from the way this term
is used in the context of disaster risk, where it only implies the actions
taken once there has been an impact.
Capacity to respond is not sufficient to reduce risk. Humanitarian aid
and relief interventions have been discussed in the context of their role
in reinforcing or even amplifying existing vulnerabilities (Anderson and
Woodrow, 1991; Wisner, 2001a; Schipper and Pelling, 2006). This does
not only have implications for the capacity to respond, but also for other
Chapter 2Determinants of Risk: Exposure and Vulnerability
75
aspects of capacity. Wisner (2001a) shows how poorly constructed
shelters, where people were placed temporarily in El Salvador following
Hurricane Mitch in 1998, turned into ‘permanent’ housing when
nongovernmental organization (NGO) support ran out. When two
strong earthquakes hit in January and February 2001, the shelters
collapsed, leaving the people homeless again. This example illustrates
the perils associated with emergency measures that focus only on
responding, rather than on the capacity to reduce risk and change.
Response capacity is also differential (Chatterjee, 2010). The most
effective ex ante risk management strategies will often include a
combination of risk reduction and enhanced capacity to respond to
impacts (including smarter response by better preparedness and early
warning, as well risk transfer such as insurance).
2.4.2.3. Capacity to Recover and Change
Having the capacity to change is a requirement in order to adapt to
climate change. Viewing adaptation as requiring transformation implies
that it cannot be understood as only a set of actions that physically
protect people from natural hazards (Pelling, 2010). In the context of
natural hazards, the opportunity for changing is often greatest during the
recovery phase, when physical infrastructure has to be rebuilt and can be
improved, and behavioral patterns and habits can be contemplated
(Susman et al., 1983; Renn, 1992; Comfort et al., 1999; Vogel and O’Brien,
2004; Birkmann et al., 2010a). This is an opportunity to rethink whether
the crops planted are the most suited to the climate and whether it is
worthwhile rebuilding hotels near the coast, taking into account what
other sorts of environmental changes may occur in the area.
Capacity to recover is not only dependent on the extent of a physical
impact, but also on the extent to which society has been affected,
including the ability to resume livelihood activities (Hutton and Haque,
2003). This capacity is driven by numerous factors, including mental and
physical ability to recover, financial and environmental viability, and
political will. Because reconstruction processes often do not take
people’s livelihoods into account, instead focusing on their safety, new
settlements are often located where people do not want to be, which
brings change – but not necessarily change that leads to sustainable
development. Innumerable examples indicate how people who have been
resettled return back to their original location, moving into dilapidated
houses or setting up new housing, even if more solid housing is
available elsewhere (e.g., El Salvador after Hurricane Mitch), simply
because the new location does not allow them easy access to their
fields, to markets or roads, or to the sea (e.g., South and Southeast Asia
after the 2004 tsunami).
Recovering to return to the conditions before a natural hazard occurs
not only implies that the risk may be the same or greater, but also does
not question whether the previous conditions were desirable. In fact,
recovery processes are often out of sync with the evolving process of
development. The recovery and reconstruction phases after a disaster
provide an opportunity to rethink previous conditions and address the
root causes of risk, looking to avoid reconstructing the vulnerability
(IDB, 2007), but often the process is too rushed to enable effective
reflection, discussion, and consensus building (Christoplos, 2006).
Pushing the recovery toward transformation and change requires taking
a new approach rather than returning to ‘normalcy.’ Several examples
have shown that capacity to recover is severely limited by poverty
(Chambers, 1983; Ingham, 1993; Hutton and Haque, 2003), where
people are driven further down the poverty spiral, never returning to
their previous conditions, however undesirable.
The various capacities to respond and to survive hazard events and
changes have also been discussed within the context of the concept of
resilience. While originally, the concept of resilience was strongly linked
to an environmental perspective on ecosystems and their ability to
maintain major functions even in times of adverse conditions and crises
(Holling, 1973), the concept has undergone major shifts and has been
enhanced and applied also in the field of social-ecological systems and
disaster risk (Gunderson, 2000; Walker et al., 2004; UN, 2005; Abel et al.,
2006). Folke (2006) differentiates three different resilience concepts
that encompass an engineering resilience perspective that focuses on
recovery and constancy issues, while the ecological and social resilience
focus on persistence and robustness and, finally, the integrated social-
ecological resilience perspective deals with adaptive capacity, trans-
formability, learning, and innovation (Folke, 2006). In disaster risk
reduction the terms resilience building and the lack of resilience have
achieved a high recognition. These terms are linked to capacities of
communities or societies to deal with the impact of a hazard event
or crises and the ability to learn and create resilience through these
experiences. Recent papers, however, also criticize the unconsidered use
or the simply transfer of the concept of resilience into the wider context
of adaptation (see, e.g., Cannon and Müller-Mahn, 2010). Additionally,
the lack of resilience has also been used as an umbrella to examine
deficiencies in capacities that communities encompass in order to deal
with hazard events. Describing the lack of resilience, Cardona and
Barbat (2000) identify various capacities that are often insufficient in
societies that suffer heavily during disasters, such as the deficiencies
regarding the capacity to anticipate, to cope with, and to adapt to
changing environmental conditions and natural hazards.
Other work has argued a different view on resilience, because the very
occurrence of a disaster shows that there are gaps in the development
process (UNDP, 2004). Lessons learned from studying the impacts of the
2004 Indian Ocean tsunami (Thomalla et al., 2009; Thomalla and Larsen,
2010) are informative for climate-related hazards. They suggest that:
Social vulnerability to multiple hazards, particularly rare extreme
events, tends to be poorly understood.
There is an increasing focus away from vulnerability assessment
toward resilience building; however, resilience is poorly understood
and a lot needs to be done to go from theory to practice.
One of the key issues in sub-national risk reduction initiatives is a
need to better define the roles and responsibilities of government
and NGO actors and to improve coordination between them. Without
mechanisms for joint target setting, coordination, monitoring, and
Chapter 2 Determinants of Risk: Exposure and Vulnerability
76
evaluation, there is much duplication of effort, competition, and
tension between actors.
Risk reduction is only meaningful and prioritized by local government
authorities if it is perceived to be relevant in the context of other,
more pressing day-to-day issues, such as poverty reduction,
livelihood improvement, natural resource management, and
community development.
2.4.3. Factors of Capacity: Drivers and Barriers
There is high confidence that extreme and non-extreme weather and
climate events also affect vulnerability to future extreme events, by
modifying the resilience, coping, and adaptive capacity of communities,
societies, or social-ecological systems affected by such events. When
people repeatedly have to respond to natural hazards and changes, the
capitals that sustain capacity are broken down, increasing vulnerability
to hazards (Wisner and Adams, 2002; Marulanda et al., 2008b, 2010,
2011; UNISDR, 2009a). Much work has gone into identifying what these
factors of capacity are, to understand both what drives capacity as
well as what acts as a barrier to it (Adger et al., 2004; Sharma and
Padwardhan, 2008).
Drivers of capacity include: an integrated economy; urbanization;
information technology; attention to human rights; agricultural capacity;
strong international institutions; access to insurance; class structure; life
expectancy, health, and well-being; degree of urbanization; access to
public health facilities; community organizations; existing planning
regulations at national and local levels; institutional and decisionmaking
frameworks; existing warning and protection from natural hazards; and
good governance (Cannon, 1994; Handmer et al., 1999; Klein, 2001;
Barnett, 2005; Brooks et al., 2005; Bettencourt et al., 2006).
2.5. Dimensions and Trends of
Vulnerability and Exposure
This section presents multiple dimensions of exposure and vulnerability
to hazards, disasters, climate change, and extreme events. Some
frameworks consider exposure to be a component of vulnerability (Turner
et al., 2003a), and the largest body of knowledge on dimensions refers
to vulnerability rather than exposure, but the distinction between them
is often not made explicit. Vulnerability is: multi-dimensional and
differential – that is, it varies across physical space and among and
within social groups; scale-dependent with regard to space and units of
analysis such as individual, household, region, or system; and dynamic
– characteristics and driving forces of vulnerability change over time
(Vogel and O’Brien, 2004). As vulnerability and exposure are not fixed,
understanding the trends in vulnerability and exposure is therefore an
important aspect of the discussion.
There is high confidence that for several hazards, changes in exposure
and in some cases vulnerability are the main drivers behind observed
trends in disaster losses, rather than a change in hazard character, and
will continue to be essential drivers of changes in risk patterns over the
coming decades (Bouwer et al., 2007; Pielke Jr. and Landsea, 1998;
UNISDR, 2009a). In addition, there is high confidence that climate change
will affect disaster risk not only through changes in the frequency,
intensity, and duration of some events (see Chapter 3), but also through
indirect effects on vulnerability and exposure. In most cases, it will do
so not in isolation but as one of many sources of possible stress, for
instance through impacts on the number of people in poverty or suffering
from food and water insecurity, changing disease patterns and general
health levels, and where people live. In some cases, these changes may
be positive, but in many cases, they will be negative, especially for many
groups and areas that are already among the most vulnerable.
Although trends in some of the determinants of risk and vulnerability are
apparent (for example, accelerated urbanization), the extent to which
these are altering levels of risk and vulnerability at a range of geographical
and time scales is not always clear. While there is high confidence that
these connections exist, current knowledge often does not allow us to
provide specific quantifications with regional or global significance.
The multidimensional nature of vulnerability and exposure makes any
organizing framework arbitrary, overlapping, and contentious to a
degree. The following text is organized under three very broad headings:
environmental, social, and economic dimensions. Each of these has a
number of subcategories, which map out the major elements of interest.
2.5.1. Environmental Dimensions
Environmental dimensions include:
Potentially vulnerable natural systems (such as low-lying islands,
coastal zones, mountain regions, drylands, and Small Island
Developing States (Dow, 1992; UNCED, 1992; Pelling and Uitto,
2001; Nicholls, 2004; UNISDR, 2004; Chapter 3)
Impacts on systems (e.g., flooding of coastal cities and agricultural
lands, or forced migration)
The mechanisms causing impacts (e.g., disintegration of particular
ice sheets) (Füssel and Klein, 2006; Schneider et al., 2007)
Responses or adaptations to environmental conditions (UNEP/
UNISDR, 2008).
There are important links between development, environmental
management, disaster reduction, and climate adaptation (e.g., van Aalst
and Burton, 2002), also including social and legal aspects such as
property rights (Adger, 2000). For the purposes of vulnerability analysis
in the context of climate change, it is important to acknowledge that
the environment and human beings that form the socio-ecological
system (Gallopin et al., 2001) behave in nonlinear ways, and are
strongly coupled, complex, and evolving (Folke et al., 2002).
There are many examples of the interactions between society and
environment that make people vulnerable to extreme events (Bohle et
Chapter 2Determinants of Risk: Exposure and Vulnerability
77
al., 1994) and highlight the vulnerability of ecosystem services (Metzger
et al., 2006). As an example, vulnerabilities arising from floodplain
encroachment and increased hazard exposure are typical of the intricate
and finely balanced relationships within human-environment systems
(Kates, 1971; White, 1974) of which we have been aware for several
decades. Increasing human occupancy of floodplains increases exposure
to flood hazards. It can put not only the lives and property of human
beings at risk but can damage floodplain ecology and associated
ecosystem services. Increased exposure of human beings comes about
even in the face of actions designed to reduce the hazard. Structural
responses and alleviation measures (e.g., provision of embankments,
channel modification, and other physical alterations of the floodplain
environment), designed ostensibly to reduce flood risk, can have the
reverse result. This is variously known as the levee effect (Kates, 1971;
White, 1974), the escalator effect (Parker, 1995), or the ‘safe development
paradox’ (Burby, 2006) in which floodplain encroachment leads to
increased flood risk and, ultimately, flood damages. A maladaptive
policy response to such exposure provides structural flood defenses,
which encourage the belief that the flood risk has been removed. This
in turn encourages more floodplain encroachment and a reiteration of
the cycle as the flood defenses (built to a lower design specification) are
exceeded. This is typical of many maladaptive policy responses, which
focus on the symptoms rather than the causes of poor environmental
management.
Floodplains, even in low-lying coastal zones, have the potential to
provide benefits and/or risks and it is the form of the social interaction
(see next subsection) that determines which, and to whom. Climate
variability shifts previous risk-based decisionmaking into conditions of
greater uncertainty where we can be less certain of the probabilities of
occurrence of any extreme event.
The environmental dimension of vulnerability also deals with the role of
regulating ecosystem services and ecosystem functions, which directly
impact human well-being, particularly for those social groups that
heavily depend on these services and functions due to their livelihood
profiles. Especially in developing countries and countries in transition,
poorer rural communities often entirely depend on ecosystem services
and functions to meet their livelihood needs. The importance of these
ecosystem services and ecosystem functions for communities in the
context of environmental vulnerability and disaster risk has been
recognized by the 2009 and 2011 Global Assessment Reports on
Disaster Risk Reduction (UNISDR, 2009a, 2011) as well as by the
Millennium Ecosystem Assessment (MEA, 2005). The degradation of
ecosystem services and functions can contribute to an exacerbation of
both the natural hazard context and the vulnerability of people. The
erosion of ecosystem services and functions can contribute to the decrease
of coping and adaptive capacities in terms of reduced alternatives for
livelihoods and income-generating activities due to the degradation of
natural resources. Additionally, a worsening of environmental services
and functions might also increase the costs of accessing these services,
for example, in terms of the increased time and travel needed to access
drinking water in rural communities affected by droughts or salinization.
Furthermore, environmental vulnerability can also mean that in the case
of a hazardous event occurring, the community may lose access to the
only available water resource or face a major reduction in productivity
of the soil, which then also increases the risk of crop failure. For
instance, Renaud (2006) underscored that the salinization of wells after
the 2004 Indian Ocean tsunami had a highly negative consequence for
those communities that had no alternative access to freshwater
resources.
2.5.1.1. Physical Dimensions
Within the environmental dimension, physical aspects refer to a location-
specific context for human-environment interaction (Smithers and Smit,
1997) and to the material world (e.g., built structures).
The physical exposure of human beings to hazards has been partly
shaped by patterns of settlement of hazard-prone landscapes for the
countervailing benefits they offer (UNISDR, 2004). Furthermore, in the
context of climate change, physical exposure is in many regions also
increasing due to spatial extension of natural hazards, such as floods,
areas affected by droughts, or delta regions affected by salinization.
This does not make the inhabitants of such locations vulnerable per se
because they may have capacities to resist the impacts of extreme events;
this is the essential difference between exposure and vulnerability. The
physical dimension of vulnerability begins with the recognition of a link
between an extreme physical or natural phenomenon and a vulnerable
human group (Westgate and O’Keefe, 1976). Physical vulnerability
comprises aspects of geography, location, and place (Wilbanks, 2003);
settlement patterns; and physical structures (Shah, 1995; UNISDR,
2004) including infrastructure located in hazard-prone areas or with
deficiencies in resistance or susceptibility to damage (Wilches-Chaux,
1989). Further, Cutter’s (1996) ‘hazards of place’ model of vulnerability
expressly refers to the temporal dimension (see Section 2.5.4.2), which,
in recognizing the dynamic nature of place vulnerability, argues for a
more nuanced approach.
2.5.1.2. Geography, Location, Place
Aggregate trends in the environmental dimensions of exposure and
vulnerability as they relate to geography, location, and place are given
in Chapters 3 and 4, while this section deals with the more conceptual
aspects.
There is a significant difference in exposure and vulnerability between
developing and developed countries. While a similar (average) number of
people in low and high human development countries may be exposed
to hazards each year (11 and 15% respectively), the average numbers
killed is very different (53 and 1% respectively) (Peduzzi, 2006).
Developing countries are recognized as facing the greater impacts and
having the most vulnerable populations, in the greatest number, who
Chapter 2 Determinants of Risk: Exposure and Vulnerability
78
are least able to easily adapt to changes in inter alia temperature, water
resources, agricultural production, human health, and biodiversity
(IPCC, 2001; McCarthy et al., 2001; Beg et al., 2002). Small Island
Developing States, a number of which are also Least Developed
Countries, are recognized as being highly vulnerable to external shocks
including climate extremes (UN/DESA, 2010; Chapter 3). While efforts in
climate change adaptation have been undertaken, progress has been
limited, focusing on public awareness, research, and policy development
rather than implementation (UN/DESA, 2010).
Developed countries are also vulnerable and have geographically
distinct levels of vulnerability, which are masked by a predominant focus
on direct impacts on biophysical systems and broad economic sectors.
However, indirect and synergistic effects, differential vulnerabilities, and
assumptions of relative ease of adaptation within apparently robust
developed countries may lead to unforeseen vulnerabilities (O’Brien et al.,
2006). Thus, development per se is not a guarantee of ‘invulnerability.’
Development can undermine ecosystem resilience on the one hand but
create wealth that may enhance societal resilience overall if equitable
(Barnett, 2001).
The importance of geography has been highlighted in an analysis of
‘disaster hotspots’ by Dilley et al. (2005). Hazard exposure (event
incidence) is combined with historical vulnerability (measured by
mortality and economic loss) in order to identify geographic regions
that are at risk from a range of geophysical hazards. While flood risk is
widespread across a number of regions, drought and especially cyclone
risk demonstrate distinct spatial patterns with the latter closely related
to the climatological pattern of cyclone tracks and landfall.
2.5.1.3. Settlement Patterns and Development Trajectories
There are specific exposure/vulnerability dimensions associated with
urbanization (Hardoy and Pandiella, 2009) and rurality (Scoones, 1998;
Nelson et al., 2010a,b). The major focus below is on the urban because
of the increasing global trend toward urbanization and its potential for
increasing exposure and vulnerability of large numbers of people.
2.5.1.3.1. The urban environment
Accelerated urbanization is an important trend in human settlement,
which has implications for the consideration of exposure and vulnerability
to extreme events. There has been almost a quintupling of the global
urban population between 1950 and 2011 with the majority of that
increase being in less developed regions (UN-HABITAT, 2011).
There is high confidence that rapid and unplanned urbanization
processes in hazardous areas exacerbate vulnerability to disaster risk
(Sánchez-Rodríguez et al., 2005). The development of megacities with
high population densities (Mitchell, 1999a,b; Guha-Sapir et al., 2004)
has led to greater numbers being exposed and increased vulnerability
through, inter alia, poor infrastructural development (Uitto, 1998) and
the synergistic effects of intersecting natural, technological, and social
risks (Mitchell, 1999a). Lavell (1996) identified eight contexts of cities
that increase or contribute to disaster risk and vulnerability and are
relevant in the context of climate change:
1) The synergic nature of the city and the interdependency of its parts
2) The lack of redundancy in its transport, energy, and drainage systems
3) Territorial concentration of key functions and density of building
and population
4) Mislocation
5) Social-spatial segregation
6) Environmental degradation
7) Lack of institutional coordination
8) The contrast between the city as a unified functioning system and
its administrative boundaries that many times impede coordination
of actions.
The fact that urban areas are complex systems poses potential
management challenges in terms of the interplay between people,
infrastructure, institutions, and environmental processes (Ruth and
Coelho, 2007). Alterations or trends in any of these, or additional
components of the urban system such as environmental governance
(Freudenberg et al., 2008) or the uptake of insurance (McLemand and
Smit 2006; Lamond et al., 2009), have the potential to increase exposure
and vulnerability to extreme climate events substantially.
The increasing polarization and spatial segregation of groups with
different degrees of vulnerability to disaster have been identified as an
emerging problem (Mitchell, 1999b). For the United States, where there is
considerable regional variability, the components found to consistently
increase social vulnerability (as expressed by a Social Vulnerability Index)
are density (urbanization), race/ethnicity (see below), and socioeconomic
status, with the level of development of the built environment, age,
race/ethnicity, and gender accounting for nearly half of the variability in
social vulnerability among US counties (Cutter and Finch, 2008). Social
isolation, especially as it intersects with individual characteristics (see
Case Study 9.2.1) and other social processes of marginalization
(Duneier, 2004) also play a significant role in vulnerability creation (or,
conversely, reduction).
Rapidly growing urban populations may affect the capacity of developing
countries to cope with the effects of extreme events because of the
inability of governments to provide the requisite urban infrastructure or
for citizens to pay for essential services (UN-HABITAT, 2009). However,
there is a more general concern that there has been insufficient attention
to both existing needs for infrastructure maintenance and appropriate
ongoing adaptation of infrastructure to meet potential climate extremes
(Auld and MacIver, 2007). Further, while megacities have been associated
with increasing hazard for some time (Mitchell, 1999a), small cities and
rural communities are potentially more vulnerable to disasters than big
cities or megacities, since megacities have considerable resources for
dealing with hazards and disasters (Cross, 2001) and smaller settlements
are often of lower priority for government spending.
Chapter 2Determinants of Risk: Exposure and Vulnerability
79
The built environment can be both protective of, and subject to, climate
extremes. Inadequate structures make victims of their occupants and,
conversely, adequate structures can reduce human vulnerability. The
continuing toll of deaths and injuries in unsafe schools (UNISDR,
2009a), hospitals and health facilities (PAHO/World Bank, 2004),
domestic structures (Hewitt, 1997), and infrastructure more broadly
(Freeman and Warner 2001) are indicative of the vulnerability of many
parts of the built environment. In a changing climate, more variable
and with potentially more extreme events, old certainties about the
protective ability of built structures are undermined.
The increase in the number and extent of informal settlements or slums
(UN-HABITAT, 2003; Utzinger and Keiser, 2006) is important because they
are often located on marginal land within cities or on the periphery
because of the lack of alternative locations or the fact that areas close
to river systems or areas at the coast are sometimes state land that can
be more easily accessed than private land. Because of their location,
slums are often exposed to hydrometeorological-related hazards such
as landslides (Nathan, 2008) and floods (Bertoni, 2006; Colten, 2006;
Aragon-Durand, 2007; Douglas et al., 2008; Zahran et al., 2008).
Vulnerability in informal settlements can also be elevated because of
poor health (Sclar et al., 2005), livelihood insecurity (Kantor and Nair,
2005), lack of access to service provision and basic needs (such as
clean water and good governance), and a reduction in the capacity of
formal players to steer developments and adaptation initiatives in a
comprehensive, preventive, and inclusive way (Birkmann et al., 2010b).
Lagos, Nigeria (
Adelekan, 2010), and Chittagong, Bangladesh (Rahman
et al., 2010), serve as clear examples of where an upward trend in the
area of slums has resulted in an increase in the exposure of slum
dwellers to flooding. Despite the fact that rapidly growing informal
and poor urban areas are often hotspots of hazard exposure, for a
number of locations the urban poor have developed more or less
successful coping and adaptation strategies to reduce their vulnerability
in dealing with changing environmental conditions (e.g., Birkmann et al.,
2010b).
Globally, the pressure for urban areas to expand onto flood plains and
coastal strips has resulted in an increase in exposure of populations to
riverine and coastal flood risk (McGranahan et al., 2007; Nicholls et al.,
2011). For example, intensive and unplanned human settlements in
flood-prone areas appear to have played a major role in increasing
flood risk in Africa over the last few decades (Di Baldassarre et al.,
2010). As urban areas have expanded, urban heat has become a
management and health issue (for more on this see Section 2.5.2.3 and
Chapters 3, 5, and 9). For some cities there is clear evidence of a recent
trend in loss of green space (Boentje and Blinnikov, 2007; Sanli et al.,
2008; Rafiee et al., 2009) due to a variety of reasons including planned
and unplanned urbanization with the latter driven by internal and external
migration resulting in the expansion of informal settlements. Such
changes in green space may increase exposure to extreme climate
events in urban areas through decreasing runoff amelioration, urban
heat island mitigation effects, and alterations in biodiversity (Wilby
and Perry, 2006).
While megacities have been associated with increasing hazard for some
time (Mitchell, 1999a), small cities and rural communities (see next
section) are potentially more vulnerable to disasters than big cities or
megacities, since megacities have considerable resources for dealing
with hazards and disasters (Cross, 2001) and smaller settlements are
often of lower priority for government spending.
Urbanization itself is not always a driver for increased vulnerability.
Instead, the type of urbanization and the context in which urbanization
is embedded defines whether these processes contribute to an increase
or decrease in people’s vulnerability.
2.5.1.3.2. The rural environment
Many rural livelihoods are reliant to a considerable degree on the
environment and natural resource base (Scoones, 1998), and extreme
climate events can impact severely on the agricultural sector (Saldaña-
Zorrilla, 2007). However, despite the separation here, the urban and
the rural are inextricably linked. Inhabitants of rural areas are often
dependent on cities for employment, as a migratory destination of last
resort, and for health care and emergency services. Cities depend on
rural areas for food, water, labor, ecosystem services, and other
resources. All of these (and more) can be impacted by climate-related
variability and extremes including changes in these associated with
climate change. In either case, it is necessary to identify the many
exogenous factors that affect a household’s livelihood security.
Eakin’s (2005) examination of rural Mexico presents empirical findings
of the interactions (e.g., between neoliberalism and the opening up of
agricultural markets, and the agricultural impacts of climatic extremes),
which amplify or mitigate risky outcomes. The findings point to economic
uncertainty over environmental risk, which most influences agricultural
households’ decisionmaking. However, there is not a direct and inevitable
link between disaster impact and increased impoverishment of a rural
population. In Nicaragua, Jakobsen (2009) found that a household’s
probability of being poor in the years following Hurricane Mitch was not
affected by whether it was living in an area struck but by factors such as
off-farm income, household size, and access to credit. Successful coping
post-Hurricane Mitch resulted in poor households regaining most of
their assets and resisting a decline into a state of extreme poverty.
However, longer-term adaptation strategies, which might have lifted them
out of the poverty category, eluded the majority and were independent
of having experienced Hurricane Mitch. Thus, while poor (rural) households
may cope with the impacts of a disaster in the relatively short term,
their level of vulnerability, arising from a complex of environmental,
social, economic, and political factors, is such that they cannot escape
the poverty trap or fully reinstate development gains.
In assessing the material on exposure and vulnerability to climate
extremes in urban and rural environments it is clear that there is no
simple, deterministic relationship; it is not possible to show that either
rural or urban environments are more vulnerable (or resilient). In
Chapter 2 Determinants of Risk: Exposure and Vulnerability
80
either context there is the potential that climate risks can be either
ameliorated or exacerbated by positive or negative adaptation processes
and outcomes.
2.5.2. Social Dimensions
The social dimension is multi-faceted and cross-cutting. It focuses
primarily on aspects of societal organization and collective aspects
rather than individuals. However, some assessments also use the
‘individual’ descriptor to clarify issues of scale and units of analysis
(Adger and Kelly, 1999; K. O’Brien et al., 2008). Notions of the individual
are also useful when considering psychological trauma in and after
disasters (e.g., Few, 2007), including that related to family breakdown
and loss. The social dimension includes demography, migration, and
displacement, social groups, education, health and well-being, culture,
institutions, and governance aspects.
2.5.2.1. Demography
Certain population groups may be more vulnerable than others to climate
variability and extremes. For example, the very young and old are more
vulnerable to heat extremes than other population groups (Staffogia et
al., 2006; Gosling et al., 2009). A rapidly aging population at the
community to country scale bears implications for health, social isolation,
economic growth, family composition, and mobility, all of which are
social determinants of vulnerability. However, as discussed further
below (Social Groups section), static checklists of vulnerable groups do
not reflect the diversity or dynamics of people’s changing conditions.
2.5.2.1.1. Migration and displacement
Trends in migration, as a component of changing population dynamics,
have the potential to rise because of alterations in extreme climate
event frequency. The United Nations Office for the Coordination of
Humanitarian Affairs and the Internal Displacement Monitoring Centre
have estimated that around 20 million people were displaced or
evacuated in 2008 because of rapid onset climate-related disasters
(OCHA/IDMC, 2009). Further, over the last 30 years, twice as many
people have been affected by droughts (slow onset events not included
in the previous point) as by storms (1.6 billion compared with
approximately 718 million) (IOM, 2009). However, because of the multi-
causal nature of migration, the relationship between climatic variability
and change in migration is contested (Black, 2001) as are the terms
environmental and climate refugees (Myers, 1993; Castles 2002; IOM,
2009). Despite an increase in the number of hydrometeorological
disasters between 1990 and 2009, the International Organization on
Migration reports no major impact on international migratory flows
because displacement is temporary and often confined within a region,
and displaced individuals do not possess the financial resources to
migrate (IOM, 2009).
Although there is also a lack of clear evidence for a systematic trend in
extreme climate events and migration, there are clear instances of the
impact of extreme hydrometeorological events on displacement. For
example, floods in Mozambique displaced 200,000 people in 2001,
163,000 people in 2007, and 102,000 more in 2008 (INGC, 2009; IOM,
2009); in Niger, large internal movements of people are due to
pervasive changes related to drought and desertification trends (Afifi,
2011); in the Mekong River Delta region, changing flood patterns appear
to be associated with migratory movements (White, 2002; IOM, 2009);
and Hurricane Katrina, for which social vulnerability, race, and class
played an important role in outward and returning migration (Elliott
and Pais, 2006; Landry et al., 2007; Myers et al., 2008), resulted in the
displacement of over one million people. As well as the displacement
effect, there is evidence for increased vulnerability to extreme events
among migrant groups because of an inability to understand extreme
event-related information due to language problems, prioritization of
finding employment and housing, and distrust of authorities (Enarson
and Morrow, 2000; Donner and Rodriguez, 2008).
Migration can be both a condition of, and a response to, vulnerability –
especially political vulnerability created through conflict, which can drive
people from their homelands. Increasingly it relates to economically and
environmentally displaced persons but can also refer to those who do
not cross international borders but become internally displaced persons
as a result of extreme events in both developed and developing countries
(e.g., Myers et al., 2008).
Although data on climate change-forced displacement is incomplete, it
is clear that the many outcomes of climate change processes will be
seen and felt as disasters by the affected populations (Oliver-Smith,
2009). For people affected by disasters, subsequent displacement and
resettlement often constitute a second disaster in their lives. As part
of the Impoverishment Risks and Reconstruction approach, Cernea
(1996) outlines the eight basic risks to which people are subjected by
displacement: landlessness, joblessness, homelessness, marginalization,
food insecurity, increased morbidity, loss of access to common property
resources, and social disarticulation. When people are forced from their
known environments, they become separated from the material and
cultural resource base upon which they have depended for life as
individuals and as communities (Altman and Low, 1992). The material
losses most often associated with displacement and resettlement are
losses of access to customary housing and resources. Displaced people
are often distanced from their sources of livelihood, whether land,
common property (water, forests, etc.), or urban markets and clientele
(Koenig, 2009). Disasters and displacement may sever the identification
with an environment that may once have been one of the principle
features of cultural identity (Oliver-Smith, 2006). Displacement for any
group can be distressing, but for indigenous peoples it can result in
particularly severe impacts. The environment and ties to land are
considered to be essential elements in the survival of indigenous societies
and distinctive cultural identities (Colchester, 2000). The displacement
and resettlement process has been consistently shown to disrupt and
destroy those networks of social relationships on which the poor
Chapter 2Determinants of Risk: Exposure and Vulnerability
81
depend for resource access, particularly in times of stress (Cernea, 1996;
Scudder, 2005).
Migration is an ancient coping mechanism in response to environmental
(and other) change and does not inevitably result in negative outcomes,
either for the migrants themselves or for receiving communities (Barnett
and Webber, 2009). Climate variability will result in some movement of
stressed people but there is low confidence in ability to assign direct
causality to climatic impacts or to the numbers of people affected.
2.5.2.1.2. Social groups
Research evidence of the differential vulnerability of social groups is
extensive and raises concerns about the disproportionate effects of
climate change on identifiable, marginalized populations (Bohle et al.,
1994; Kasperson and Kasperson, 2001; Thomalla et al., 2006). Particular
groups and conditions have been identified as having differential
exposure or vulnerability to extreme events, for example race/ethnicity
(Fothergill et al., 1999; Elliott and Pais, 2006; Cutter and Finch, 2008),
socioeconomic class and caste (O’Keefe et al., 1976; Peacock et al.,
1997; Ray-Bennett, 2009), gender (Sen, 1981), age (both the elderly and
children; Jabry, 2003; Wisner, 2006b; Bartlett, 2008), migration, and
housing tenure (whether renter or owner), as among the most common
social vulnerability characteristics (Cutter and Finch, 2008). Morrow
(1999) extends and refines this list to include residents of group living
facilities; ethnic minorities (by language); recent migrants (including
immigrants); tourists and transients; physically or mentally disabled (see
also McGuire et al., 2007; Peek and Stough, 2010); large households;
renters; large concentrations of children and youth; poor households;
the homeless (see also Wisner, 1998); and women-headed households.
Generally, the state of vulnerability is defined by a specific population
at a particular scale; aggregations (and generalizations) are often less
meaningful and require careful interpretation (Adger and Kelly, 1999).
One of the largest bodies of research evidence, and one which can be an
exemplar for the way many other marginalized groups are differentially
impacted or affected by extreme events, has been on gender and disaster,
and on women in particular (e.g., Neal and Phillips, 1990; Enarson and
Morrow, 1998; Neumayer and Plümper, 2007). This body of literature is
relatively recent, particularly in a developed world context, given the
longer recognition of gender concerns in the development field
(Fordham, 1998). The specific gender and climate change link including
self-defined gender groups has been even more recent (e.g., Masika,
2002; Pincha and Krishna, 2009). The research evidence emphasizes the
social construction of gendered vulnerability in which women and girls
are often (although not always) at greater risk of dying in disasters,
typically marginalized from decisionmaking fora, and discriminated and
acted against in post-disaster recovery and reconstruction efforts
(Houghton, 2009; Sultana, 2010).
Women or other socially marginalized or excluded groups are not
vulnerable through biology (except in very particular circumstances) but
are made so by societal structures and roles. For example, in the Indian
Ocean tsunami of 2004, many males were out to sea in boats, fulfilling
their roles as fishermen, and were thus less exposed than were many
women who were on the seashore, fulfilling their roles as preparers and
marketers of the fish catch. However, the women were made vulnerable
not simply by their location and role but by societal norms which did not
encourage survival training for girls (e.g., to swim or climb trees) and
which placed the majority of the burden of child and elder care with
women. Thus, escape was made more difficult for women carrying
children and responsible for others (Doocy et al., 2007).
The gender and disaster/climate change literature has also recognized
resilience/capacity/capability alongside vulnerability. This elaboration of the
vulnerability approach makes clear that vulnerability in these identified
groups is not an immutable or totalizing condition. The vulnerability
‘label’ can reinforce notions of passivity and helplessness, which obscure
the very significant, active contributions that socially marginalized
groups make in coping with and adapting to extremes. An example is
provided in Box 2-2.
2.5.2.2. Education
The education dimension ranges across the vulnerability of educational
building structures; issues related to access to education; and also
sharing and access to disaster risk reduction and climate adaptation
information and knowledge (Wisner, 2006b). Priority 3 of the Hyogo
Framework for Action 2005-2015 recommends the use of knowledge,
innovation, and education to build a ‘culture of safety and resilience’ at
all levels (UNISDR, 2007a). A well-informed and motivated population
can lead to disaster risk reduction but it requires the collection and
dissemination of knowledge and information on hazards, vulnerabilities,
and capacities. However, “It is not information per se that determines
action, but how people interpret it in the context of their experience,
beliefs and expectations. Perceptions of risks and hazards are culturally
and socially constructed, and social groups construct different meanings
for potentially hazardous situations” (McIvor and Paton, 2007). In addition
to knowledge and information, explicit environmental education programs
among children and adults may have benefits for public understanding
of risk, vulnerability, and exposure to extreme events (UNISDR, 2004;
Kobori, 2009; Nomura, 2009; Patterson et al., 2009; Kuhar et al., 2010),
because they promote resilience building in socio-ecological systems
through their role in stewardship of biological diversity and ecosystem
services, provide the opportunity to integrate diverse forms of knowledge
and participatory processes in resource management (Krasny and Tidball,
2009), and help promote action towards sustainable development
(Waktola, 2009; Breiting and Wikenberg, 2010).
Many lives have been lost through the inability of education infrastructure
to withstand extreme events. Where flooding is a recurrent phenomenon
schools can be exposed or vulnerable to floods. For example, a survey
of primary schools’ flood vulnerability in the Nyando River catchment
of western Kenya revealed that 40% were vulnerable, 48% were
Chapter 2 Determinants of Risk: Exposure and Vulnerability
82
marginally vulnerable, and 12% were not vulnerable; the vulnerability
status was attributed to a lack of funds, poor building standards, local
topography, soil types and inadequate drainage (Ochola et al., 2010).
Improving education infrastructure safety can have multiple benefits.
For example, the Malagasy Government initiated the Development
Intervention Fund IV project to reduce cyclone risk, including safer school
construction and retrofitting. In doing so, awareness and understanding
of disaster issues were increased within the community (Madagascar
Development Intervention Fund, 2007).
The impact of extreme events can limit the ability of parents to afford
to educate their children or require them (especially girl children, whose
access to education is typically prioritized less than that of boy children)
to work to meet basic needs (UNDP, 2004; UNICEF, 2009).
Access to information related to early warnings, response strategies,
coping and adaptation mechanisms, science and technology, and human,
social, and financial capital is critical for reduction of vulnerability and
increasing resilience. A range of factors may control or influence the
access to information, including economic status, race (Spence et al.,
2007), trust (Longstaff and Yang, 2008), and belonging to a social
network (Peguero, 2006). However, the mode of information transfer or
exchange must be considered because there is emerging evidence of a
growing digital inequality (Rideout, 2003) that may influence trends in
vulnerability as an increasing amount of information about extreme event
preparedness and response is often made available via the internet (see
Chapter 9). Evidence has existed for some time that people who have
experienced natural hazards (and thus may have information and
knowledge gained directly through that experience) are, in general,
better prepared than those who have not (Kates, 1971). However, this
does not necessarily translate into protective behavior because of what
has been called the ‘prison of experience’ (Kates, 1962), in which people’s
response behavior is determined by the previous experience and is not
based on an objective assessment of current risk. In the uncertain
context of climate-related extremes, this may mean people are not
appropriately educated regarding the risk.
2.5.2.3. Health and Well-Being
The health dimension of vulnerability includes differential physical,
physiological, and mental health effects of extreme events in different
regions and on different social groups (McMichael et al., 2003; van
Lieshout et al., 2004; Haines et al., 2006; Few, 2007; Costello et al.,
Chapter 2Determinants of Risk: Exposure and Vulnerability
Box 2-2 | Integrating Disaster Risk Reduction, Climate Adaptation, and Resilience-Building:
the Garifuna Women of Honduras
The Garifuna women of Honduras could be said to show multiple vulnerability characteristics (Brondo, 2007). They are women, the gender
often made vulnerable by patriarchal structures worldwide; they come from Honduras, a developing country exposed to many hazards;
they belong to an ethnic group descended from African slaves, which is socially, economically, and politically marginalized; and they
depend largely upon a subsistence economy, with a lack of education, health, and other resources. However, despite these markers of
vulnerability, the Garifuna women have organized to reduce their communities’ exposure to hazards and vulnerability to disasters
through the protection and development of their livelihood opportunities (Fordham et al., 2011).
The women lead the Comité de Emergencia Garifuna de Honduras, which is a grassroots, community-based group of the Afro-Indigenous
Garifuna that was developed in the wake of Hurricane Mitch in 1998. After Mitch, there was a lack of external support and so the Comité
women organized themselves and repaired hundreds of houses, businesses, and public buildings, in the process of which women were
empowered and trained in non-traditional work. They campaigned to buy land for relocating housing to safer areas, in which the poorest
families participated in the reconstruction process. Since being trained themselves in vulnerability and capacity mapping by grassroots
women in Jamaica, they have in turn trained 60 trainers in five Garifuna communities to carry out mapping exercises in their communities.
The Garifuna women have focused on livelihood-based activities to ensure food security by reviving and improving the production of
traditional root crops, building up traditional methods of soil conservation, carrying out training in organic composting and pesticide use,
and creating the first Garifuna farmers’ market. In collaborative efforts, 16 towns now have established tool banks, and five have seed
banks. Through reforestation, the cultivation of medicinal and artisanal plants, and the planting of wild fruit trees along the coast, they
are helping to prevent erosion and reducing community vulnerability to hazards and the vagaries of climate.
The Garifuna women’s approach, which combines livelihood-based recovery, disaster risk reduction, and climate change adaptation, has
had wide-ranging benefits. They have built up their asset base (human, social, physical, natural, financial, and political), and improved
their communities’ nutrition, incomes, natural resources, and risk management. They continue to partner with local, regional, and
international networks for advocacy and knowledge exchange. The women and communities are still at risk (Drusine, 2005) but these
strategies help reduce their socioeconomic vulnerability and dependence on external aid (Fordham et al., 2011).
83
2009). It also includes, in a link to the institutional dimension, health
service provision (e.g., environmental health and public health issues,
infrastructure and conditions; Street et al., 2005), which may be impacted
by extreme events (e.g., failures in hospital/health center building
structures; inability to access health services because of storms and
floods). Vulnerability can also be understood in terms of functionality
related to communication, medical care, maintaining independence,
supervision, and transportation. In addition individuals including children,
senior citizens, and pregnant women and those who may need additional
response assistance including the disabled, those living in institutionalized
settings, those from diverse cultures, people with limited English
proficiency or are non-English speaking, those with no access to transport,
have chronic medical disorders, and have pharmacological dependency
can also be considered vulnerable in a health context.
Unfortunately, the health dimensions of disasters are difficult to measure
because of difficulties in attributing the health condition (including
mortality) directly to the extreme event because of secondary effects; in
addition, some of the effects are delayed in time, which again makes
attribution difficult (Bennet, 1970; Hales et al., 2003). The difficulty of
collection of epidemiological data in crisis situations is also a factor,
especially in low-income countries. Further understanding the post-
traumatic stress disorder dimensions of extreme climate events and the
psychological aspects of climate change presents a number of challenges
(Amstadter et al., 2009; Kar, 2009; Mohay and Forbes, 2009; Furr et al.,
2010; Doherty and Clayton, 2011).
Health vulnerability is the sum of all the risk and protective factors
that determine the degree to which individuals or communities could
experience adverse impacts from extreme weather events (Balbus and
Malina, 2009). Vulnerabilities can arise from a wide range of institutional,
geographic, environmental, socioeconomic, biological sensitivity, and other
factors, which can vary spatially and temporally. Biological sensitivity
can be associated with developmental stage (e.g., children are at
increased mortality risk from diarrheal diseases); pre-existing medical
conditions (e.g., diabetics are at increased risk during heat waves);
acquired conditions (e.g., malaria immunity); and genetic factors
(Balbus and Malina, 2009). Vulnerability can be viewed both from the
perspective of the population groups more likely to experience adverse
health outcomes and from the perspective of the public health and
health care interventions required to prevent adverse health impacts
during and following an extreme event.
For some extreme weather events the vulnerable population groups
depend on the adverse health outcome considered. For example, in the
case of heat waves socially isolated elderly people with pre-existing
medical conditions are vulnerable to heat-related health effects (see
Chapter 9). For floods, children are at greater risk for transmission of
fecal-oral diseases, and those with mobility and cognitive constraints
can be at increased risk of injuries and deaths (Ahern et al., 2005), while
people on low incomes are less likely to be able to afford insurance
against risks associated with flooding, such as storm and flood damage
(Marmot, 2010). Flooding has been found to increase the risk of mental
health problems, pre- and post-event, in both adults and children
(Ginexi et al., 2000; Reacher et al., 2004; Ahern et al., 2005; Carroll et
al., 2006; Tunstall et al., 2006; UK Department of Health, 2009). A UK
study of over 1,200 households affected by flooding suggested that
there were greater impacts on physical and mental health among more
vulnerable groups and poorer households and communities (Werritty et
al., 2007). However, while there is evidence for impacts on particular
social groups in identified disaster types, there are some social groups
that are more likely to be vulnerable whatever the hazard type; these
include those at the extremes of the age range, those with underlying
medical conditions, and those otherwise stressed by low socioeconomic
status. The role of socioeconomic factors supports the necessity of a
social, and not just a medical, model of response and adaptation.
A number of public health impacts are expected to worsen in climate-
related disasters such as storms, floods, landslides, heat, drought, and
wildfire. These are highly context-specific but range from worsening of
existing chronic illnesses (which could be widespread), through possible
toxic exposures (in air, water or food), to deaths (expected to be few to
moderate but may be many in low-income countries) (Keim, 2008).
Public health and health care services required for preventing adverse
health impacts from an extreme weather event include surveillance and
control activities for infectious diseases, access to safe water and improved
sanitation, food security, maintenance of solid waste management and
other critical infrastructure, maintenance of hospitals and other health
care infrastructure, provision of mental health services, sufficient and
safe shelter to prevent or mitigate displacement, and effective warning
and informing systems (Keim, 2008). Further, it is important to consider
the synergistic effects of NaTech disasters (Natural Hazard Triggering a
Technological Disaster) where impacts can be considerable if only single,
simple hazard events are planned for. In an increasingly urbanized world,
interactions between natural disasters and simultaneous technological
accidents must be given attention (Cruz et al., 2004); the combination
of an earthquake, tsunami, and radiation release at the Japanese
Fukushima Nuclear Power plant in March 2011 is the most recent
example. Lack of provision of these services increases population
vulnerability, particularly in individuals with greater biological sensitivity
to an adverse health outcome. Although there is little evidence for
trends in the exposure or vulnerability of public health infrastructure,
the imperative for a resilient health infrastructure is widely recognized
in the context of extreme climate events (Burkle and Greenough, 2008;
Keim, 2008).
Deteriorating environmental conditions as a result of extremes (including
land clearing, salinization, dust generation, altered ecology; Renaud,
2006; Middleton et al., 2008; Ellis and Wilcox, 2009; Hong et al., 2009;
Ljung et al., 2009; Johnson et al., 2010; Tong et al., 2010) can impact key
ecosystem services and exacerbate climate sensitive disease incidence
(e.g., diarrheal disease; Clasen et al., 2007), particularly via deteriorating
water quality and quantity.
For some health outcomes, which have direct or indirect implications for
vulnerability to extreme climate events, there is evidence of trends. For
Chapter 2 Determinants of Risk: Exposure and Vulnerability
84
example, obesity, a risk factor for cardiovascular disease, which in turn
is a heat risk factor (Bouchama et al., 2007) has been noted to be on the
increase in a number of developed countries (Skelton et al., 2009;
Stamatakis et al., 2010). Observed trends in major public health threats
such as the infectious or communicable diseases HIV/AIDS, tuberculosis,
and malaria, although not directly linked to the diminution of long-term
resilience of some populations, have been identified as having the
potential to do so (IFRC, 2008). In addition to the diseases themselves,
persistent and increasing obstacles to expanding or strengthening
health systems such as inadequate human resources and poor hospital
and laboratory infrastructure as observed in some countries (Vitoria et
al., 2009) may also contribute indirectly to increasing vulnerability and
exposure where, for example, malaria and HIV/Aids occasionally reach
epidemic proportions.
However, trends in well-being and health are difficult to assess.
Indicators that characterize a lack of well-being and a high degree of
susceptibility are, for example, indicators of undernourishment and
malnutrition. The database for the Millennium Development Goals and
respective statistics of the Food and Agriculture Organization (FAO)
underscore that trends in undernourishment are spatially and temporally
differentiated. While, as but one example, the trend in undernourished
people in Burundi shows a significant increase from 1991 to 2005, an
opposite trend of a reduction in the percentage of undernourished
people can be observed in Angola (see UN Statistics Division, 2011;
FAOSTAT, 2011). Thus, evidence exists that trends in vulnerability, e.g.,
in terms of well-being and undernourishment change over time and are
highly differentiated in terms of spatial patterns.
In considering health-related exposure and vulnerability to extreme events,
evidence from past climate/weather-related disaster events (across a
range of hazard types for which lack of space precludes coverage)
makes clear the links to a range of negative outcomes for physical and
mental health and health infrastructure. Furthermore, there is clear
evidence (Haines et al., 2006; Confalonieri et al., 2007) that current and
projected health impacts from climate change are multifarious and will
affect low-income groups and low-income countries the most severely,
although high-income countries are not immune.
2.5.2.4. Cultural Dimensions
The broad term ‘culture’ embraces a complexity of elements that can
relate to a way of life, behavior, taste, ethnicity, ethics, values, beliefs,
customs, ideas, institutions, art, and intellectual achievements that
affect, are produced, or are shared by a particular society. In essence, all
these characteristics can be summarized to describe culture as ‘the
expression of humankind within society’ (Aysan and Oliver, 1987).
Culture is variously used to describe many aspects of extreme risks from
natural disasters or climate change, including:
Cultural aspects of risk perception
Negative culture of danger/ vulnerability/ fear
Culture of humanitarian concern
Culture of organizations / institutions and their responses
Culture of preventive actions to reduce risks, including the creation
of buildings to resist extreme climatic forces
Ways to create and maintain a ‘Risk Management Culture,’ a
‘Safety Culture,’ or an Adaptation Culture.
In relation to our understanding of risk, certain cultural issues need to
be noted. Typical examples are cited below:
Ethnicity and Culture. Deeply rooted cultural values are a dominant
factor in whether or not communities adapt to climate change. For
example, recent research in Northern Burkina Faso indicates that
two ethnic groups have adopted very different strategies due to
cultural values and historical relations, despite their presence in
the same physical environment and their shared experience of
climate change (Nielsen and Reenberg, 2010).
Locally Based Risk Management Culture. Wisner (2003) has argued
that the point in developing a ‘culture of prevention’ is to build
networks at the neighborhood level capable of ongoing hazard
assessment and mitigation at the micro level. He has noted that while
community based NGOs emerged to support recovery after the
Mexico City and Northridge earthquakes, these were not sustained
over time to promote risk reduction activities. This evidence
confirms other widespread experience indicating that ways still
need to be found to extend the agenda of Community-Based
Organizations into effective action to reduce climate risks and
promote adaptation to climate change.
Conflicting Cultures: Who Benefits, and Who Loses when Risks are
Reduced? A critical cultural conflict can arise when private actions
to reduce disaster risks and adapting to climate change by one
party have negative consequences on another. This regularly applies
in river flood hazard management where upstream measures to
reduce risks can significantly increase downstream threats to
persons and property. Adger has argued that if appropriate risk
reduction actions are to occur, the key players must bear all the
costs and receive all the benefits from their actions (Adger, 2009).
However, this can be problematic if adaptation is limited to specific
local interests only.
Traditional behaviors tied to local (and wider) tradition and cultural
practices can increase vulnerability – for example, unequal gender norms
that put women and girls at greater risk, or traditional uses of the
environment that have not adapted (or cannot adapt) to changed
environmental circumstances. On the other hand, local or indigenous
knowledge can reduce vulnerabilities too (Gaillard et al., 2007, 2010).
Furthermore, cultural practices are often subtle and may be opaque to
outsiders. The early hazards paradigm literature (White, 1974; Burton et
al., 1978) referred often to culturally embedded fatalistic attitudes,
which resulted in inaction in the face of disaster risk. However,
Schmuck-Widmann (2000), in her social anthropological studies of char
dwellers in Bangladesh, revealed how a belief that disaster occurrence
and outcomes were in the hands of God did not preclude preparatory
activities. Perceptions of risk (and their interpretation by others) depend
Chapter 2Determinants of Risk: Exposure and Vulnerability
85
on the cultural and social context (Slovic, 2000; Oppenheimer and
Todorov, 2006; Schneider et al., 2007).
Research findings emphasize the importance of considering the role –
and cultures – of religion and faith in the context of disaster. This
includes the role of faith in the recovery process following a disaster
(e.g., Davis and Wall, 1992; Massey and Sutton, 2007); religious
explanations of nature (e.g., Orr, 2003; Peterson, 2005); the role of
religion in influencing positions on environment and climate change
policy (e.g., Kintisch, 2006; Hulme, 2009); and religion and vulnerability
(Guth et al., 1995; Chester, 2005; Elliott et al., 2006; Schipper, 2010).
The cultural dimension also includes the potential vulnerability of
aboriginal and native peoples in the context of climate extremes.
Globally, indigenous populations are frequently dependent on primary
production and the natural resource base while being subject to
(relatively) poor socioeconomic conditions (including poor health, high
unemployment, low levels of education, and greater poverty). This
applies to groups from Canada (Turner and Clifton, 2009), to Australia
(Campbell et al., 2008), to the Pacific (Mimura et al., 2007). Small island
states, often with distinct cultures, typically show high vulnerability and
low adaptive capacity to climate change (Nurse and Sem, 2001).
However, historically, indigenous groups have had to contend with many
hazards and, as a consequence, have developed capacities to cope
(Campbell, 2006) such as the use of traditional knowledge systems,
locally appropriate building construction with indigenous materials, and
a range of other customary practices (Campbell, 2006).
Given the degree of cultural diversity identified, the importance of
understanding differential risk perceptions in a cultural context is
reinforced (Marris et al., 1998). Cultural Theory has contributed to an
understanding of how people interpret their world and define risk
according to their worldviews: hierarchical, fatalistic, individualistic,
and egalitarian (Douglas and Wildavsky, 1982). Too often policies and
studies focus on ‘the public’ in the aggregate and too little on the needs,
interests, and attitudes of different social and cultural groups (see also
Sections 2.5.2.1.2 and 2.5.4).
2.5.2.5. Institutional and Governance Dimensions
The institutional dimension is a key determinant of vulnerability to
extreme events (Adger, 1999). Institutions have been defined in a broad
sense to include “habitualized behavior and rules and norms that govern
society” (Adger, 2000) and not just the more typically understood
formal institutions. This view allows for a discussion of institutional
structures such as property rights and land tenure issues (Toni and
Holanda, 2008) that govern natural resource use and management. It
forms a bridge between the social and the environmental/ecological
dimensions and can induce sustainable or unsustainable exploitation
(Adger, 2000). Expanding the institutional domain to include political
economy (Adger, 1999) and different modes of production – feudal,
capitalist, socialist (Wisner, 1978) – raises questions about the
vulnerability of institutions and the vulnerability caused by institutions
(including government). Institutional factors play a critical role in
adaptation (Adger, 2000) as they influence the social distribution of
vulnerability and shape adaptation capacity (Næss et al., 2005).
This broader understanding of the institutional dimension also takes us
into a recognition of the role of social networks, community bonds and
organizing structures, and processes that can buffer the impacts of
extreme events (Nakagawa and Shaw, 2004) partly through increasing
social cohesion but also recognizing ambiguous or negative forms
(UNISDR, 2004). For example, social capital/assets (Portes, 1998;
Putnam, 2000) – “the norms and networks that enable people to act
collectively” (Woolcock and Narayan, 2000) – have a role in vulnerability
reduction (Pelling, 1998). Social capital (or its lack) is both a cause and
effect of vulnerability and thus can result in either positive benefit or
negative impact; to be a part of a social group and accrue social assets
is often to indicate others’ exclusion. It also includes attempts to
reframe climate debates by acknowledging the possibility of diverse
impacts on human security, which opens up human rights discourses
and rights-based approaches to disaster risk reduction (Kuwali, 2008;
Mearns and Norton, 2010).
The institutional dimension includes the relationship between policy
setting and policy implementation in risk and disaster management. Top-
down approaches assume policies are directly translated into action on
the ground; bottom-up approaches recognize the importance of other
actors in shaping policy implementation (Urwin and Jordan, 2008). Twigg’s
categorization of the characteristics of the ideal disaster resilient
community (Twigg, 2007) adopts the latter approach. This guideline
document, which has been field tested by NGOs, identifies the important
relations between the community and the enabling environment of
governance at various scales in creating resilience, and by inference,
reducing vulnerability. This set of 167 characteristics (organized under five
thematic areas) also refers to institutional forms for (and processes of)
engagement with risk assessment, risk management, and hazard and
vulnerability mapping. These have been championed by institutions
working across scales to create the Hyogo Framework for Action (UNISDR,
2007a) and associated tools (Davis et al., 2004; UNISDR, 2007b) with
the goal to reduce disaster risk and vulnerability. However, linkages
across scales and the inclusion of local knowledge systems are still not
integrated well in formal institutions (Næss et al., 2005).
A lack of institutional interaction and integration between disaster risk
reduction, climate change, and development may mean policy responses
are redundant or conflicting (Schipper and Pelling, 2006; Mitchell and
van Aalst, 2008; Mitchell et al., 2010). Thus, the institutional model
operational in a given place and time (more or less participatory,
deliberative, and democratic; integrated; or disjointed) could be an
important factor in either vulnerability creation or reduction (Comfort et
al., 1999). Furthermore, risk-specific policies must also be integrated
(see the slippage between UK heat and cold wave policies, Wolf et al.,
2010a). However, further study of the role of institutions in influencing
vulnerability is called for (O’Brien et al., 2004b).
Chapter 2 Determinants of Risk: Exposure and Vulnerability
86
Governance is also a key topic for vulnerability and exposure.
Governance is broader than governmental actions; governance can be
understood as the structures of common governance arrangements and
processes of steering and coordination – including markets, hierarchies,
networks, and communities (Pierre and Peters, 2000). Institutionalized
rule systems and habitualized behavior and norms that govern society
and guide actors are representing governance structures (Adger, 2000;
Biermann et al., 2009). These formal and informal governance structures
also determine vulnerability, since they influence power relations, risk
perceptions, and constitute the context in which vulnerability, risk
reduction, and adaptation are managed.
Conflicts between formal and informal governance or governmental
and nongovernmental strategies and norms can generate additional
vulnerabilities for communities exposed to environmental change. An
example of these conflicts of formal and informal strategies is linked to
flood protection measures. While local people might expend resources
to deal with increasing flood events (e.g., adapting their livelihoods and
production patterns to changing flood regimes), formal adaptation
strategies, particularly in developing countries, prioritize structural
measures (e.g., dike systems or relocation strategies) that have severe
consequences for the vulnerability of communities dependent on local
ecosystem services, such as fishing and farming systems (see Birkmann,
2011a,b). These conflicts between formal and informal or governmental
and nongovernmental management systems and norms are an important
factor that increase vulnerability and reduce adaptive capacity of the
overall system (Birkmann et al., 2010b). Countries with institutional and
governance fragilities often lack the capacity to identify and reduce
risks and to deal with emergencies and disasters effectively. The recent
disaster and problems in coping and recovery in the aftermath of the
earthquake in Haiti or the problems in terms of managing recovery and
emergency management after the Pakistan floods are examples that
illustrate the importance of governance as a subject of resilience and
vulnerability.
In some developed countries, the last 30 years have witnessed a shift in
environmental governance practices toward more integrated approaches.
With the turn of the century, there has been recognition of the need to
move beyond technical solutions and to deal with the patterns and
drivers of unsustainable demand and consumption. This has resulted in
the emergence of a more integrated approach to environmental
management, a focus on prevention (UNEP, 2007), the incorporation of
knowledge from the local to the global in environment policies
(Karlsson, 2007), and co-management and involvement of stakeholders
from all sectors in the management of natural resources (Plummer, 2006;
McConnell, 2008), although some have also questioned the efficacy of
this new paradigm (Armitage et al., 2007; Sandstrom, 2009).
2.5.3. Economic Dimensions
Economic vulnerability can be understood as the susceptibility of an
economic system, including public and private sectors, to potential
(direct) disaster damage and loss (Rose, 2004; Mechler et al., 2010) and
refers to the inability of affected individuals, communities, businesses,
and governments to absorb or cushion the damage (Rose, 2004).
The degree of economic vulnerability is exhibited post-event by the
magnitude and duration of the indirect follow-on effects. These effects
can comprise business interruption costs to firms unable to access
inputs from their suppliers or service their customers, income losses of
households unable to get to work, or the deterioration of the fiscal
stance post-disasters as less taxes are collected and significant public
relief and reconstruction expenditure is required. At a macroeconomic
level, adverse impacts include effects on gross domestic product (GDP),
consumption, and the fiscal position (Mechler et al., 2010). Key drivers
of economic vulnerability are low levels of income and GDP, constrained
tax revenue, low domestic savings, shallow financial markets, and high
indebtedness with little access to external finance (OAS, 1991; Benson
and Clay, 2000; Mechler, 2004).
Economic vulnerability to external shocks, including natural hazards,
has been inexactly defined in the literature and conceptualizations
often have overlapped with risk, resilience, or exposure. One line of
research focusing on financial vulnerability, as a subset of economic
vulnerability, framed the problem in terms of risk preference and
aversion, a conceptualization more common to economists. Risk
aversion, in this context, denotes the ability of economic agents to
absorb risk financially (Arrow and Lind, 1970). There are many ways to
absorb the financial burdens of disasters, with market-based insurance
being one, albeit prominent, option, although more particularly in a
developed country context. Households as economic agents often use
informal mechanisms relying on family and relatives abroad or outside
a disaster area; governments may simply rely on their tax base or
international assistance. Yet, in the face of large and covariate risks,
such ad hoc mechanisms often break down, particularly in developing
countries (see Linnerooth-Bayer and Mechler, 2007).
Research on financial vulnerability to disasters has hitherto focused
on developing countries’ financial vulnerability describing financial
vulnerability as a country’s ability to access domestic and foreign
savings for financing post-disaster relief and reconstruction needs in
order to quickly recover and avoid substantial adverse ripple effects
(Mechler et al., 2006; Marulanda et al., 2008a; Cardona, 2009; Cummins
and Mahul, 2009). Reported and estimated substantial financial
vulnerability and risk aversion in many exposed countries, as well as the
emergence of novel public-private partnership instruments for pricing
and transferring catastrophe risks globally, has motivated developing
country governments, as well as development institutions, NGOs, and
other donor organizations, to consider pre-disaster financial instruments
as an important component of disaster risk management (Linnerooth-
Bayer et al., 2005).
There is a distinct scale aspect to the economic dimension of exposure
and vulnerability. While evidence of the economic costs of known
disasters indicate impacts may be under 10% of GDP (Wilbanks et al.,
Chapter 2Determinants of Risk: Exposure and Vulnerability
87
2007), at smaller and more local scales the costs can be significantly
greater. A lack of good data makes it difficult to provide meaningful and
specific assessments other than to acknowledge that, without investment
in adaptation and resilience building measures, the intensification or
increased frequency of extreme weather events is bound to impact GDP
growth in the future (Wilbanks et al., 2007).
Work and Livelihoods
At the individual and community levels, work and livelihoods are an
important facet of the economic dimension. These are often impacted
by extreme events and by the responses to extreme events.
Humanitarian/disaster relief in response to extreme events can induce
dependency and weaken local economic and social systems (Dudasik,
1982) but livelihood-based relief is of growing importance (Pantuliano
and Wekesa, 2008). Further, there is increasing recognition that
disasters and extreme events are stresses and shocks within livelihood
development processes (Cannon et al., 2003; see Kelman and Mather,
2008, for a discussion of cases applying to volcanic events).
Paavola’s (2008) analysis of livelihoods, vulnerability, and adaptation to
climate change in Morogoro, Tanzania, is indicative of the way extreme
events impact livelihoods in specific ways. Here, rural households are
found to be more vulnerable to climate variability and climate change
than are those in urban environments (see also Section 2.5.1.3). This is
because rural incomes and consumption levels are significantly lower,
there are greater levels of poverty, and more limited access to markets
and other services. More specifically, women are made more vulnerable
than men because they lack access to livelihoods other than climate-
sensitive agriculture. Local people have employed a range of strategies
(extensification, intensification, diversification, and migration) to
manage climate variability but these have sometimes had undesirable
environmental outcomes, which have increased their vulnerability. In
the absence of opportunities to fundamentally change their livelihood
options, we see here an example of short-term coping rather than long-
term climate adaptation (Paavola, 2008).
Human vulnerability to natural hazards and income poverty are largely
codependent (Adger, 1999; UNISDR, 2004) but poverty does not equal
vulnerability in a simple way (e.g., Blaikie et al., 1994); the determinants
and dimensions of poverty are complex as well as its association with
climate change (Khandlhela and May, 2006; Demetriades and Esplen,
2008; Hope, 2009). It is important to recognize that adaptation
measures need to specifically target climate extremes-poverty linkages
as not all poverty reduction measures reduce vulnerability to climate
extremes and vice versa. Further, measures are required across scales
because the drivers of poverty, although felt at a local level, may
necessitate tackling political and economic issues at a larger scale
(Eriksen and O’Brien, 2007; K. O’Brien et al., 2008).
Given the relationship between poverty and vulnerability, it can be
argued (Tol et al., 2004) that economic growth could reduce vulnerability
(with caveats). However, increasing economic growth would not
necessarily decrease climate impacts because it has the potential to
simultaneously increase greenhouse gas emissions. Furthermore,
growth is often reliant on critical infrastructure which itself may be
affected by extreme events. There are many questions still to be
answered by research about the impacts of varying economic policy
changes including the pursuit of narrow development trajectories
and how this might shape vulnerability (Tol et al., 2004; UNDP, 2004;
UNISDR, 2004)
2.5.4. Interactions, Cross-Cutting Themes, and Integrations
This section began by breaking down the vulnerability concept into its
constitutive dimensions, with evidence derived from a number of
discrete research and policy communities (e.g., disaster risk reduction;
climate change adaptation; environmental management; and poverty
reduction) that have largely worked independently (Thomalla et al.,
2006). Increasingly it is recognized that collaboration and integration is
necessary both to set appropriate policy agendas and to better
understand the topic of interest (K. O’Brien et al., 2008), although
McLaughlin and Dietz (2008) have made a critical analysis of the
absence of an integrated perspective on the interrelated dynamics of
social structure, human agency, and the environment.
Reviewing singular dimensions of vulnerability cannot provide an
appropriate level of synthesis. Considerable conceptual advances arose
from the early recognition that so-called natural disasters were not
‘natural’ at all (O’Keefe et al., 1976) but were the result of structural
inequalities rooted in political economy. This critique required analysis
of more than the hazard component (Blaikie et al., 1994). Further, it
demonstrated how crossing disciplinary and other boundaries (e.g.,
those separating disaster and development, or developed and developing
countries) can be fruitful in better understanding extremes of various
kinds (see Hewitt, 1983). If we consider food security/vulnerability (as
just one example), an inclusive analysis of the vulnerability of food
systems (to put it broadly), must take account of aspects related to, inter
alia: physical location in susceptible areas; political economy (Watts and
Bohle, 1993); entitlements in access to resources (Sen, 1981); social
capital and networks (Eriksen et al., 2005); landscape ecology (Fraser,
2006); human ecology (Bohle et al., 1994); and political ecology (Pulwarty
and Riebsame, 1997; Holling, 2001; see Chapter 4 for further discussion
of food systems and food security). More generally, in relation to hazards,
disaster risk reduction, and climate extremes, productive advances have
been made in research adopting a coupled human/social-environment
systems approach (Holling, 2001; Turner et al., 2003b) which recognizes
the importance of integrating often separate domains. For example, in
analyzing climate change impacts, vulnerability, and adaptation in
Norway, O’Brien et al. (2006) argue that a simple examination of direct
climate change impacts underestimates the, perhaps more serious and
larger, synergistic impacts. They use an example of projected climate
change effects in the Barents Sea, which may directly impact keystone
fish species. However, important as this finding is, climate change may
Chapter 2 Determinants of Risk: Exposure and Vulnerability
88
also influence the transport sector (through reduction in ice cover);
increase numbers of pollution events (through increased maritime
transport of oil and other goods); may risk ecological and other damages
as a result of competition from introduced species in ballast water;
which, in turn, are aggravated by increases in ocean temperatures.
Neither the potential level of impact nor the processes of adaptation are
best represented by a singular focus on a particular sector but must
consider interactions between sectors and institutional, economic,
social, and cultural conditions (O’Brien et al., 2006).
2.5.4.1. Intersectionality and Other Dimensions
The dimensions discussed above generate differential effects but it is
important to consider not just differences between single categories
(e.g., between women and men) but the differences within a given
category (e.g., ‘women’). This refers to intersectionality, where, for
example, gender may be a significant variable but only when allied with
race/ethnicity or some other variable. In Hurricane Katrina, it mattered
(it still matters) whether you were black or white, upper class or work-
ing class, home owner or renter, old or young, woman or man in terms
of relative exposure and vulnerability factors (Cutter et al., 2006; Elliott
and Pais, 2006).
Certain factors are identified as cross-cutting themes of particular
importance for understanding the dynamic changes within exposure,
vulnerability, and risk. In the Sphere Project’s minimum standards in
humanitarian response, children, older people, persons with disabilities,
gender, psychosocial issues, HIV and AIDS, and environment, climate
change, and disaster risk reduction are identified as cross-cutting
themes and must be considered, not as separate sectors, which people
may or may not select for attention, but must be integrated within each
sector (Sphere Project, 2011). Exactly which topics are selected as cross-
cutting themes, to be incorporated throughout an activity, is context-
specific. Below, we consider just two: different timing (diachronic
aspects within a single day or across longer time periods) and different
spatial and functional scales.
2.5.4.2. Timing, Spatial, and Functional Scales
Cross-cutting themes of particular importance for understanding the
dynamic changes within exposure, vulnerability, and risk are different
timing (diachronic aspects within a single day or across longer time
periods) and different spatial and functional scales.
2.5.4.2.1. Timing and timescales
Timing and timescales are important cross-cutting themes that need
more attention when dealing with the identification and management
of extreme climate and weather events, disasters, and adaptation
strategies. The first key issue when dealing with timing and timescales
is the fact that different hazards and their recurrence intervals might
fundamentally change in terms of the time dimension. This implies that
the identification and assessment of risk, exposure, and vulnerability
needs also to deal with different time scales and in some cases might
need to consider different time scales. At present most of the climate
change scenarios focus on climatic change within the next 100 or
200 years, while often the projections of vulnerability just use present
socioeconomic data. However, a key challenge for enhancing knowledge
of exposure and vulnerability as key determinants of risk requires
improved data and methods to project and identify directions and
different development pathways in demographic, socioeconomic, and
political trends that can adequately illustrate potential increases or
decreases in vulnerability with the same time horizon as the changes in
the climate system related to physical-biogeochemical projections (see
Birkmann et al., 2010b).
Furthermore, the time dependency of risk analysis, particularly if the
analysis is conducted at a specific point in time, has been shown to be
critical. Newer research underlines that exposure – especially the
exposure of different social groups – is a highly dynamic element that
changes not only seasonally, but also during the day and over different
days of the week (e.g., Setiadi, 2011). Disasters also exacerbate pre-
disaster trends in vulnerability (Colten et al., 2008).
Consequently, time scales and dynamic changes over time have to be
considered carefully when conducting risk and vulnerability assessments
for extreme events and creeping changes in the context of climate change.
Additionally, changes in the hazard frequency and timing of hazard
occurrence during the year will have a strong impact on the ability of
societies and ecosystems to cope and adapt to these changes.
The timing of events may also create ‘windows of vulnerability,’ periods in
which the hazards are greater because of the conjunction of circumstances
(Dow, 1992). Time is a cross-cutting dimension that always needs to be
considered but particularly so in the case of anthropogenic climate change,
which may be projected some years into the future (Füssel, 2005). In
fact, this time dimension is regarded (Thomalla et al., 2006) as a key
difference between the disaster management and climate change
communities. To generalize somewhat, the former group typically
(with obvious exceptions like slow-onset hazards such as drought or
desertification) deals with fast-onset events, in discrete, even if extensive,
locations, requiring immediate action. The latter group typically focuses on
conditions that occur in a dispersed form over lengthy time periods and
which are much more challenging in their identification and measurement
(Thomalla et al., 2006). Risk perception may be reduced (Leiserowitz,
2006) for such events remote in time and/or space, such as some climate
change impacts are perceived to be. Thus, in this conceptualization,
different time scales are an important constraint when dealing with the
link between disaster risk reduction and climate change adaptation (see
Thomalla et al., 2006; Birkmann and von Teichman, 2010).
However, it is important to also acknowledge that disaster risk reduction
considers risk reduction within different time frames; it encompasses
Chapter 2Determinants of Risk: Exposure and Vulnerability
89
short-term emergency management/response strategies and long-term
risk reduction strategies, for example, building structures to resist
10,000-year earthquakes or flood barriers to resist 1,000-year storm
surges. Modern prospective risk management debates involve security
considerations decades ahead for production, infrastructure, houses,
hospitals, etc.
2.5.4.2.2. Spatial and functional scales
Spatial and functional scales are another cross-cutting theme that is of
particular relevance when dealing with the identification of exposure and
vulnerability to extreme events and climate change. Leichenko and O’Brien
(2002) conclude that in many areas of climate change and natural hazards
societies are confronted with dynamic vulnerability, meaning that
processes and factors that cause vulnerability operate simultaneously at
multiple scales making traditional indicators insufficient. Leichenko and
O’Brien (2002) analyze a complex mix of influences (both positive and
negative) on the vulnerability, and coping and adaptive capacity of
southern African farmers in dealing with climate variability. These
include the impacts of globalization on national-level policies and local-
level experiences (e.g., structural adjustment programs reducing local-
level agricultural subsidies on the one hand, and on the other, trade
liberalization measures opening up new opportunities through
diversification of production in response to drought). Also Turner et al.
(2003a,b) stress that vulnerability and resilience assessments need to
consider the influences on vulnerability from different scales, however,
the practical application and analysis of these interacting influences on
vulnerability from different spatial scales is a major challenge and in most
cases not sufficiently understood. Furthermore, vulnerability analysis
particularly linked to the identification of institutional vulnerability has
also to take into account the various functional scales that climate
change, natural hazards, and vulnerability as well as administrative
systems operate on. In most cases, current disaster management
instruments and measures of urban or spatial planning as well as water
management tools (specific plans, zoning, norms) operate on different
functional scales compared to climate change. Even the various hazards
that climate change may modify encompass different functional scales
that cannot be sufficiently captured with one approach. For example,
policy setting and management of climate change and of disaster risk
reduction are usually the responsibility of different institutions or
departments, thus it is a challenge to develop a coherent and integrated
strategy (Birkmann and von Teichman, 2010). Consequently, functional
and spatial scale mismatches might even be part of institutional
vulnerabilities that limit the ability of governance system to adequately
respond to hazards and changes induced by climate change.
2.5.4.3. Science and Technology
Science and technology possess the potential to assist with adaptation
to extreme climate events, however there are a number of factors that
determine the ultimate utility of technology for adaptation. These
include an understanding of the range of technologies available, the
identification of the appropriate role for technology, the process of
technology transfer, and the criteria applied in selection of the technology
(Klein et al., 2006). For major sectors such as water, agriculture, and
health a range of possible so-called ‘hard’ and ‘soft’ technologies exist
such as irrigation and crop rotation pattern (Klein et al., 2006) or the
development of drought-resistant crops (IAASTD, 2009) in the case of
the agricultural sector.
Although approaches alternative to pure science- and technology-
based ones have been suggested for decreasing vulnerability (Haque
and Etkin, 2007; Marshall and Picou, 2008), such as blending western
science and technology with indigenous knowledge (Mercer et al., 2010)
and ecological cautiousness and the creation of eco-technologies with
a pro-nature, pro-poor, and pro-women orientation (Kesavan and
Swaminathan, 2006), their efficacy in the context of risk and vulnerability
reduction remain undetermined.
The increasing integration of a range of emerging weather and climate
forecasting products into early warning systems (Glantz, 2003) has
helped reduce exposure to extreme climate events because of an
increasing improvement of forecast skill over a range of time scales
(Goddard et al., 2009; Stockdale et al., 2009; van Aalst, 2009; Barnston
et al., 2010; Hellmuth et al., 2011). Moreover, there is an increasing use
of weather and climate information for planning and climate risk
management in business (Changnon and Changnon, 2010), food
security (Verdin et al., 2005), and health (Ceccato et al., 2007; Degallier
et al., 2010) as well as the use of technology for the development of a
range of decision support tools for climate-related disaster management
(van de Walle and Turoff, 2007).
2.6. Risk Identification and Assessment
Risk accumulation, dynamic changes in vulnerabilities, and the different
phases of crises and disaster situations constitute a complex environment
for identifying and assessing risks and vulnerabilities, risk reduction
measures, and adaptation strategies. Understanding of extreme events
and disasters is a pre-requisite for the development of adaptation
strategies in the context of climate change and risk reduction in the
context of disaster risk management.
Current approaches to disaster risk management typically involve four
distinct public policies or components (objectives) (IDEA, 2005; Carreño,
2006; IDB, 2007; Carreño et al., 2007b):
1) Risk identification (involving individual perception, evaluation of
risk, and social interpretation)
2) Risk reduction (involving prevention and mitigation of hazard or
vulnerability)
3) Risk transfer (related to financial protection and in public
investment)
4) Disaster management (across the phases of preparedness, warnings,
response, rehabilitation, and reconstruction after disasters).
Chapter 2 Determinants of Risk: Exposure and Vulnerability
90
The first three actions are mainly ex ante – that is, they take place in
advance of disaster – and the fourth refers mainly to ex post actions,
although preparedness and early warning do require ex ante planning
(Cardona, 2004; IDB, 2007). Risk identification, through vulnerability and
risk assessment can produce common understanding by the stakeholders
and actors. It is the first step for risk reduction, prevention, and transfer,
as well as climate adaptation in the context of extremes.
2.6.1. Risk Identification
Understanding risk factors and communicating risks due to climate
change to decisionmakers and the general public are key challenges.
These challenges include developing an improved understanding of
underlying vulnerabilities, and societal coping and response capacities.
There is high confidence that the selection of appropriate vulnerability
and risk evaluation approaches depends on the decisionmaking context.
The promotion of a higher level of risk awareness regarding climate
change-induced hazards and changes requires an improved understanding
of the specific risk perceptions of different social groups and individuals,
including those factors that influence and determine these perceptions,
such as beliefs, values, and norms. This also requires attention for
appropriate formats of communication that characterize uncertainty
and complexity (see, e.g., Patt et al., 2005; Bohle and Glade, 2008; Renn,
2008, pp. 289; Birkmann et al., 2009; ICSU-LAC, 2011a,b, p. 15).
Appropriate information and knowledge are essential prerequisites for
risk-aware behavior and decisions. Specific information and knowledge
on the dynamic interactions of exposed and vulnerable elements
include livelihoods and critical infrastructures, and potentially damaging
events, such as extreme weather events or potential irreversible
changes such as sea level rise. Based on the expertise of disaster risk
research and findings in the climate change and climate change
adaptation community, requirements for risk understanding related to
climate change and extreme events particularly encompass knowledge
of various elements (Kasperson et al., 2005; Patt et al., 2005; Renn and
Graham, 2006; Biermann, 2007; Füssel, 2007; Bohle and Glade, 2008;
Cutter and Finch, 2008; Renn, 2008; Biermann et al., 2009, Birkmann et
al., 2009, 2010b; Cardona, 2010; Birkmann, 2011a; ICSU-LAC, 2011a,b),
including:
Processes by which persons, property, infrastructure, goods, and
the environment itself are exposed to potentially damaging events,
for example, understanding exposure in its spatial and temporal
dimensions.
Factors and processes that determine or contribute to the
vulnerability of persons and their livelihoods or of socio-ecological
systems. This includes an understanding of increases or decreases
in susceptibility and response capacity, including the distribution of
socio- and economic resources that make people more vulnerable
or that increase their level of resilience.
How climate change affects hazards, particularly regarding
processes by which human activities in the natural environment or
changes in socio-ecological systems lead to the creation of new
hazards (e.g., NaTech hazards), irreversible changes, or increasing
probabilities of hazard events occurrence.
Different tools, methodologies, and sources of knowledge (e.g.,
expert/scientific knowledge, local or indigenous knowledge) that
allow capturing new hazards, risk, and vulnerability profiles, as well
as risk perceptions. In this context, new tools and methodologies
are also needed that allow for the evaluation, for example, of new
risks (sea level rise) and of current adaptation strategies.
How risks and vulnerabilities can be modified and reconfigured
through forms of governance, particularly risk governance –
encompassing formal and informal rule systems and actor
networks at various levels. Furthermore, it is essential to improve
knowledge on how to promote adaptive governance within the
framework of risk assessment and risk management.
Adaptive capacity status and limits of adaptation. This includes the
need to assess potential capacities for future hazards and for
dealing with uncertainty. Additionally, more knowledge is needed
on the various and socially differentiated limits of adaptation.
These issues also imply an improved understanding on how different
adaptation measures influence resilience and adaptive capacities.
2.6.2. Vulnerability and Risk Assessment
The development of modern risk analysis and assessments were closely
linked to the establishment of scientific methodologies for identifying
causal links between adverse health effects and different types of
hazardous events and the mathematical theories of probability (Covello
and Mumpower, 1985). Today, risk and vulnerability assessments
encompass a broad and multidisciplinary research field. In this regard,
vulnerability and risk assessments can have different functions and
goals.
Risk and vulnerability assessment depend on the underlying
understanding of the terms. In this context, two main schools of
thought can be differentiated. The first school of thought defines risk
as a decision by an individual or a group to act in such a way that the
outcome of these decisions can be harmful (Luhmann, 2003; Dikau and
Pohl, 2007). In contrast, the disaster risk research community views risk
as the product of the interaction of a potentially damaging event and
the vulnerable conditions of a society or element exposed (UNISDR,
2004; IPCC, 2007).
Vulnerability and risk assessment encompass various approaches and
techniques ranging from indicator-based global or national assessments
to qualitative participatory approaches of vulnerability and risk assessment
at the local level. They serve different functions and goals (see IDEA,
2005; Birkmann, 2006a; Cardona, 2006; Dilley, 2006; Wisner, 2006a;
IFRC, 2008; Peduzzi et al., 2009).
Risk assessment at the local level presents specific challenges related to
a lack of data (including climate data at sufficient resolution, but also
Chapter 2Determinants of Risk: Exposure and Vulnerability
91
socioeconomic data at the lowest levels of aggregation) but also the
highly complex and dynamic interplay between the capacities of the
communities (and the way they are distributed among community
members, including their power relationships) and the challenges they
face (including both persistent and acute aspects of vulnerability).
To inform risk management, it is desirable that risk assessments are
locally based and result in increased awareness and a sense of local
ownership of the process and the options that may be employed to
address the risks. Several participatory risk assessment methods, often
based on participatory rural appraisal methods, have been adjusted to
explicitly address changing risks in a changing climate. Examples of
guidance on how to assess climate vulnerability at the community level
are available from several sources (see Willows and Connell, 2003;
Moench and Dixit, 2007; van Aalst et al., 2007; CARE, 2009; IISD et al.,
2009; Tearfund, 2009). In integrating climate change, a balance needs to
be struck between the desire for a sophisticated assessment that includes
high-quality scientific inputs and rigorous analysis of the participatory
findings, and the need to keep the process simple, participatory, and
implementable at scale. Chapter 5 provides further details on the
implementation of risk management at local levels.
The International Standards Organization defines risk assessment as a
process to comprehend the nature of risk and to determine the level of
risk (ISO, 2009a,b). Additionally, communication within risk assessment
and management are seen as key elements of the process (Renn, 2008).
More specifically, vulnerability and risk assessment deal with the
identification of different facets and factors of vulnerability and risk, by
means of gathering and systematizing data and information, in order to
be able to identify and evaluate different levels of vulnerability and risk
of societies – social groups and infrastructures – or coupled socio-
ecological systems at risk. A common goal of vulnerability and risk
assessment approaches is to provide information about profiles, patterns
of, and changes in risk and vulnerability (see, e.g., IDEA, 2005; Birkmann,
2006a; Cardona, 2008; IFRC, 2008), in order to define priorities, select
alternative strategies, or formulate new response strategies. In this
context, the Hyogo Framework for Action stresses “that the starting
point for reducing disaster risk and for promoting a culture of disaster
resilience lies in the knowledge of the hazards and the physical, social,
economic, and environmental vulnerabilities to disasters that most
societies face, and of the ways in which hazards and vulnerabilities are
changing in the short and long term, followed by action taken on the
basis of that knowledge” (UN, 2005).
Vulnerability and risk assessments are key strategic activities that
inform both disaster risk management and climate change adaptation.
These require the use of reliable methodologies that allow an adequate
estimation and quantification of potential losses and consequences to
the human systems in a given exposure time.
Risk estimates are thus intended to be prospective, anticipating
scientifically possible hazard events that may occur in the future.
Usually technical risk analyses have been associated with probabilities.
Taking into account epistemic and aleatory uncertainties the probabilistic
estimations of risk attempt to forecast damage or losses even where
insufficient data are available on the hazards and the system being
analyzed (UNDRO, 1980; Fournier d’Albe, 1985; Spence and Coburn,
1987; Blockley, 1992; Coburn and Spence, 1992; Sheldon and Golding,
1992; Woo, 1999; Grossi and Kunreuther, 2005; Cardona et al., 2008a,b;
Cardona 2011). In most cases, approaches and criteria for simplification
and for aggregation of different information types and sources are used,
due to a lack of data or the inherent low resolution of the information. This
can result in some scientific or technical and econometric characteristics,
accuracy, and completeness that are desirable features when the risk
evaluation is the goal of the process (Cardona et al., 2003b). Measures
such as loss exceedance curves and probable maximum loss for different
event return periods are of particular importance for the stratification of
risk and the design of disaster risk intervention strategy considering risk
reduction, prevention, and transfer (Woo, 1999; Grossi and Kunreuther,
2005; Cardona et al., 2008a,b; ERN-AL, 2011; UNISDR, 2011). However,
it is also evident that more qualitative-oriented risk assessment
approaches are focusing on deterministic approaches and the profiling
of vulnerability using participatory methodologies (Garret, 1999).
Vulnerability and risk indicators or indices are feasible techniques for
risk monitoring and may take into account both the harder aspects of
risk as well as its softer aspects. The usefulness of indicators depends
on how they are employed to make decisions on risk management
objectives and goals (Cardona et al., 2003a; IDEA, 2005; Cardona, 2006,
2008, 2010; Carreño et al., 2007b).
However, quantitative approaches for assessing vulnerability need to be
complemented with qualitative approaches to capture the full complexity
and the various tangible and intangible aspects of vulnerability in its
different dimensions. It is important to recognize that complex systems
involve multiple variables (physical, social, cultural, economic, and
environmental) that cannot be measured using the same methodology.
Physical or material reality have a harder topology that allows the use
of quantitative measure, while collective and historical reality have a
softer topology in which the majority of the attributes are described in
qualitative terms (Munda, 2000). These aspects indicate that a weighing
or measurement of risk involves the integration of diverse disciplinary
perspectives. An integrated and interdisciplinary focus can more
consistently take into account the nonlinear relations of the parameters,
the context, complexity, and dynamics of social and environmental systems,
and contribute to more effective risk management by the different
stakeholders involved in risk reduction or adaptation decisionmaking.
Results can be verified and risk management/adaptation priorities can
be established (Carreño et al., 2007a, 2009).
To ensure that risk and vulnerability assessments are also understood,
the key challenges for future vulnerability and risk assessments, in the
context of climate change, are, in particular, the promotion of more
integrative and holistic approaches; the improvement of assessment
methodologies that also account for dynamic changes in vulnerability,
exposure, and risk; and the need to address the requirements of
Chapter 2 Determinants of Risk: Exposure and Vulnerability
92
Chapter 2Determinants of Risk: Exposure and Vulnerability
Box 2-3 | Developing a Regional Common Operating Picture of Vulnerability
in the Americas for Various Kinds of Decisionmakers
The Program of Indicators of Disaster Risk and Risk Management for the Americas of the Inter-American Development Bank (IDEA, 2005;
Cardona, 2008, 2010) provides a holistic approach to relative vulnerability assessment using social, economic, and environmental indicators
and a metric for sovereign fiscal vulnerability assessment taking into account that extreme impacts can generate financial deficit due to
a sudden elevated need for resources to restore affected inventories or capital stock.
The Prevalent Vulnerability Index (PVI) depicts predominant vulnerability conditions of the countries over time to identify progresses and
regressions. It provides a measure of direct effects (as result of exposure and susceptibility) as well as indirect and intangible effects of
hazard events (as result of socioeconomic fragilities and lack of resilience). The indicators used are made up of a set of demographic,
socioeconomic, and environmental national indicators that reflect situations, causes, susceptibilities, weaknesses, or relative absences of
development affecting the country under study. The indicators are selected based on existing indices, figures, or rates available from
reliable worldwide databases or data provided by each country. These vulnerability conditions underscore the relationship between risk
and development. Figure 2-1 shows the aggregated PVI (Exposure, Social Fragility, Lack of Resilience) for 2007.
Vulnerability and therefore risk are also the result of unsustainable economic growth and deficiencies that may be corrected by means of
adequate development processes, reducing susceptibility of exposed assets, socioeconomic fragilities, and improving capacities and
resilience of society (IDB, 2007). The information provided by an index such as the PVI can prove useful to ministries of housing and
urban development, environment, agriculture, health and social welfare, economy, and planning. The main advantage of PVI lies in its
ability to disaggregate results and identify factors that may take priority in risk management actions as corrective and prospective measures
or interventions of vulnerability from a development point of view. The PVI can be used at different territorial levels, however often the
indicators used by the PVI are only available at the national level; this is a limitation for its application at other sub-national scales.
On the other hand, future disasters have been identified as contingency liabilities and could be included in the balance of each nation.
As pension liabilities or guaranties that the government has to assume for the credit of territorial entities or due to grants, disaster
34.2
50.4
46.6
50.3
60.8
56.0
55.7
45.8
38.4
51.1
52.0
64.9
65.9
60.2
61.4
73.9
65.4
68.0
25.0
17.4
25.7
27.0
28.7
23.3
22.5
27.7
32.6
33.0
46.2
20.5
34.1
39.7
38.3
43.5
35.1
40.4
15.1
15.5
23.5
18.5
16.7
26.8
25.7
33.6
54.6
34.6
30.3
45.1
37.1
42.0
42.4
30.2
53.6
47.5
Prevalent Vulnerability Index (PVI) Evaluated for 2007
Exposure and Susceptibility
Socioeconomic Fragilities
Lack of Resilience
Chile
Argentina
Mexico
Colombia
Peru
Ecuador
Bolivia
Panama
Barbados
Costa Rica
Belize
Trinidad & Tobago
Dominican Republic
Honduras
El Salvador
Guatemala
Jamaica
Nicaragua
0 20 40 60 80 100 120 140 160
Figure 2-1 | Aggregate Prevalent Vulnerability Index (PVI) for 19 countries of the Americas for 2007. Source: Cardona, 2010.
Continued next page
93
decisionmakers and the general public. Many assessments still focus
solely on one dimension, such as economic risk and vulnerability. Thus,
they consider a very limited set of vulnerability factors and dimensions.
Some approaches, e.g., at the global level, view vulnerability primarily with
regard to the degree of experienced loss of life and economic damage (see
Dilley et al., 2005; Dilley 2006). A more integrative and holistic perspective
captures a greater range of dimensions and factors of vulnerability and
disaster risk. Successful adaptation to climate change has been based
on a multi-dimensional perspective, encompassing, for example, social,
economic, environmental, and institutional aspects. Hence, risk and
vulnerability assessments – that intend to inform these adaptation
strategies – require also a multi-dimensional perspective.
Chapter 2 Determinants of Risk: Exposure and Vulnerability
reposition costs are liabilities that become materialized when the hazard events occur. The Disaster Deficit Index (DDI) provides an
estimation of the extreme impact (due to hurricane, floods, tsunami, earthquake, etc.) during a given exposure time and the financial
ability to cope with such a situation. The DDI captures the relationship between the loss that the country could experience when an
extreme impact occurs (demand for contingent resources) and the public sector’s economic resilience – that is, the availability of funds
to address the situation (restoring affected inventories). This macroeconomic risk metric underscores the relationship between extreme
impacts and the capacity to cope of the government. Figure 2-2 shows the DDI for 2008.
A DDI greater than 1.0 reflects the country’s inability to cope with extreme disasters, even when it would go into as much debt as
possible: the greater the DDI, the greater the gap between the potential losses and the country’s ability to face them. This disaster risk
figure is interesting and useful for a Ministry of Finance and Economics. It is related to the potential financial sustainability problem of
the country regarding the potential disasters. On the other hand, the DDI gives a compressed picture of the fiscal vulnerability of the
country due to extreme impacts. The DDI has been a guide for economic risk management; the results at national and sub-national levels
can be studied by economic, financial, and planning analysts, who can evaluate the potential budget problem and the need to take into
account these figures in the financial planning.
0123456780 10,000 20,000 30,000
Chile
Mexico
Argentina
Trinidad & Tobago
Costa Rica
Ecuador
Bolivia
Colombia
Jamaica
Peru
Guatemala
Panama
El Salvador
Nicaragua
Belize
Dominican Republic
Barbados
Honduras
1.47
1.46
0.96
0.80
0.46
0.37
0.42
6.96
5.75
5.41
4.59
4.55
3.42
2.84
2.73
2.47
1.85
2.40
6,942
17,544
5,664
1,197
1,616
4,043
2,887
3,540
3,103
426
7,818
1,420
3,984
2,346
2,270
8,223
26,289
23,256
Disaster Deficit Index (DDI) for 500-year Return Period
Evaluated for 2008
Probable Maximum Loss (PML) for 500-year Return Period
Evaluated for 2008
$US millions
Figure 2-2 | Disaster deficit index (DDI) and probable maximum loss (PML) in 500 years for 19 countries of the Americas for 2008. Source: Cardona, 2010.
94
Assessment frameworks with integrative and holistic perspectives have
been developed by Turner et al. (2003a), Birkmann (2006b), and Cardona
(2001). Key elements of these holistic views are the identification of
causal linkages between factors of vulnerability and risk and the
interventions (structural, non-structural) that nations, societies, and
communities or individuals make to reduce their vulnerability or exposure
to hazards. Turner et al. (2003a) underline the need to focus on different
scales simultaneously, in order to capture the linkages between different
scales (local, national, regional, etc.). The influences and linkages
between different scales can be difficult to capture, especially due to
their dynamic nature during and after disasters, for example, through
inputs of external disaster aid (Cardona, 1999a,b; Cardona and Barbat,
2000; Turner et al., 2003a; Carreño et al., 2005, 2007a, 2009; IDEA,
2005; Birkmann, 2006b; ICSU-LAC, 2011a,b).
Several methods have been proposed to measure vulnerability from a
comprehensive and multidisciplinary perspective. In some cases composite
indices or indicators intend to capture favorable conditions for direct
physical impacts – such as exposure and susceptibility – as well as indirect
or intangible impacts of hazard events – such as socio-ecological fragilities
or lack of resilience (IDEA, 2005; Cardona, 2006; Carreño et al., 2007a).
In these holistic approaches, exposure and physical susceptibility are
representing the ‘hard’ and hazard-dependent conditions of vulnerability.
On the other hand, the propensity to suffer negative impacts as a result
of the socio-ecological fragilities and not being able to adequately cope
and anticipate future disasters can be considered ‘soft’ and usually
non-hazard dependent conditions, that aggravate the impact. Box 2-3
describes two of these approaches, based on relative indicators, useful
for monitoring vulnerability of countries over time and to communicate
it to country’s development and financial authorities in their own
language.
To enhance disaster risk management and climate change adaptation,
risk identification and vulnerability assessment may be undertaken in
different phases, that is, before, during, and even after disasters occur.
This includes, for instance, the evaluation of the continued viability of
measures taken and the need for further or different adaptation/risk
management measures. Although risk and vulnerability reduction are the
primary actions to be conducted before disasters occur, it is important
to acknowledge that ex post and forensic studies of disasters provide a
laboratory in which to study risk and disasters as well as vulnerabilities
revealed (see Birkmann and Fernando, 2008; ICSU-LAC, 2011a,b).
Disasters draw attention to how societies and socio-ecological processes
are changing and acting in crises and catastrophic situations, particularly
regarding the reconfiguration of access to different assets or the role
of social networks and formal organizations (see Bohle, 2008). It is
noteworthy that, until today, many post-disaster processes and strategies
have failed to integrate aspects of climate change adaptation and long-
term risk reduction (see Birkmann et al., 2009, 2010a).
In the broader context of the assessments and evaluations, it is also
crucial to improve the different methodologies to measure and evaluate
hazards, vulnerability, and risks. The disaster risk research has paid more
attention to sudden-onset hazards and disasters such as floods, storms,
tsunamis, etc., and less on the measurement of creeping changes and
integrating the issue of tipping points into these assessments (see also
Section 3.1.7). Therefore, the issue of measuring vulnerability and risk, in
terms of quantitative and qualitative measures also remains a challenge.
Lastly, the development of appropriate assessment indicators and
evaluation criteria would also be strengthened if respective integrative
and consistent goals for vulnerability reduction and climate change
adaptation could be defined for specific regions, such as coastal,
mountain, or arid environments. Most assessments to date have based
their judgment and evaluation on a relative comparison of vulnerability
levels between different social groups or regions.
There is medium evidence (given the generally limited amount of
long-term evaluations of impacts of adaptation and risk management
interventions and complications associated with such assessments), but
high agreement that adaptation and risk management policies and
practices will be more successful if they take the dynamic nature of
vulnerability and exposure into account, including the explicit
characterization of uncertainty and complexity (Cardona 2001, 2011;
Hilhorst, 2004, ICSU-LAC, 2010, Pelling, 2010). Projections of the impacts
of climate change can be strengthened by including storylines of changing
vulnerability and exposure under different development pathways.
Appropriate attention to the dynamics of vulnerability and exposure is
particularly important given that the design and implementation of
adaptation and risk management strategies and policies can reduce risk
in the short term, but may increase vulnerability and exposure over the
longer term. For instance, dike systems can reduce hazard exposure by
offering immediate protection, but also encourage settlement patterns
that may increase risk in the long term. For instance, in the 40-year span
between Hurricanes Betsy and Katrina, protective works – new and
improved levees, drainage pumps, and canals – successfully protected
New Orleans and surrounding parishes against three hurricanes in 1985,
1997, and 1998. These works were the basis for the catastrophe of
Katrina, having enabled massive development of previously unprotected
areas and the flooding of these areas that resulted when the works
themselves were shown to be inadequate (Colten et al., 2008). For other
examples, see Décamps (2010).
The design of public policy on disaster risk management is related to the
method of evaluation used to orient policy formulation. If the diagnosis
invites action it is much more effective than where the results are limited
to identifying the simple existence of weaknesses or failures. The main
quality attributes of a risk model are represented by its applicability,
transparency, presentation, and legitimacy (Corral, 2000). For more
details see Cardona (2004, 2011).
Several portfolio-level climate risk assessment methods for development
agencies have paid specific attention to the risk of variability and
extremes (see, e.g., Burton and van Aalst, 1999, 2004; Klein, 2001; van
Aalst, 2006b; Klein et al., 2007; Agrawala and van Aalst, 2008; Tanner,
2009). Given the planning horizons of most development projects
(typically up to about 20 years), even if the physical lifetime of the
Chapter 2Determinants of Risk: Exposure and Vulnerability
95
investment may be much longer, and need to combine attention to current
and future risks, these tools provide linkages between adaptation to
climate change and enhanced disaster risk management even in light of
current hazards. For more details on the implementation of risk
management at the national level, see Chapter 6.
2.6.3. Risk Communication
How people perceive a specific risk is a key issue for risk management
and climate change adaptation effectiveness (e.g., Burton et al., 1993;
Alexander, 2000; Kasperson and Palmlund, 2005; van Sluis and van Aalst,
2006; ICSU-LAC, 2011a,b) since responses are shaped by perception of
risk (Grothmann and Patt, 2005; Wolf et al., 2010b; Morton et al., 2011).
Risk communication is a complex cross-disciplinary field that involves
reaching different audiences to make risk comprehensible, understanding
and respecting audience values, predicting the audience’s response to
the communication, and improving awareness and collective and
individual decisionmaking (e.g., Cardona, 1996c; Mileti, 1996; Greiving,
2002; Renn, 2008). Risk communication failures have been revealed in
past disasters, such as Hurricane Katrina in 2005 or the Pakistan floods
in 2010 (DKKV, 2011). Particularly, the loss of trust in official institutions
responsible for early warning and disaster management were a key
factor that contributed to the increasing disaster risk. Effective and
people-centered risk communication is therefore a key to improve
vulnerability and risk reduction in the context of extreme events,
particularly in the context of people-centered early warning (DKKV,
2011). Weak and insufficient risk communication as well as the loss of
trust in government institutions in the context of early warning or climate
change adaptation can be seen as a core component of institutional
vulnerability.
Risk assessments and risk identification have to be linked to different types
and strategies of risk communication. Risk communication or the failure
of effective and people-centered risk communication can contribute to
an increasing vulnerability and disaster risk. Knowledge on factors that
determine how people perceive and respond to a specific risk or a set
of multi-hazard risks is key for risk management and climate change
adaptation (see Grothmann and Patt 2005; van Aalst et al., 2008).
Understanding the ways in which disasters are framed requires more
information and communication about vulnerability factors, dynamic
temporal and spatial changes of vulnerability, and the coping and
response capacities of societies or social-ecological systems at risk (see
Turner et al., 2003a; Birkmann, 2006a,b,c; Cardona, 2008; Cutter and
Finch, 2008; ICSU-LAC, 2011a,b). ‘Framing’ refers to the way a particular
problem is presented or viewed. Frames are shaped by knowledge of
and underlying views of the world (Schon and Rein, 1994). It is related
to the organization of knowledge that people have about their world in
the light of their underlying attitudes toward key social values (e.g.,
nature, peace, freedom), their notions of agency and responsibility (e.g.,
individual autonomy, corporate responsibility), and their judgments about
reliability, relevance, and weight of competing knowledge claims (Jasanoff
and Wynne, 1997). ‘Early warning’ implies information interventions
into an environment in which much about vulnerability is assumed. In
this regard, risk communication is not solely linked to a top-down
communication process, rather effective risk communication requires
recognition of communication as a social process meaning that risk
communication also deals with local risk perceptions and local framing
of risk. Risk communication thus functions also as a tool to upscale local
knowledge and needs (bottom-up approach). Therefore, effective risk
communication achieves both informing people at risk about the key
determinants of their particular risks and of impending disaster risk (early
warning), and also engages different stakeholders in the definition of a
problem and the identification of respective solutions (see van Aalst et al.,
2008).
Climate change adaptation strategies as well as disaster risk reduction
approaches need public interest, leadership, and acceptance. The
generation and receipt of risk information occurs through a diverse array
of channels. Chapter 5 and others discuss the important role of mass
media and other sources (see, e.g., the case of Japan provided in Sampei
and Aoyagi-Usui, 2009). Within the context of risk communication,
particularly in terms of climate change and disasters, decisionmakers,
scientists, and NGOs have to act in accordance with media requirements
concerning news production, public discourse, and media consumption
(see Carvalho and Burgess, 2005). Carvalho (2005) and Olausson (2009)
underline that mass media is often closely linked to political awareness
and is framed by its own journalistic norms and priorities; that means
also that mass media provides little space for alternative frames of
communicating climate change (Carvalho, 2005; Olausson, 2009).
Boykoff and Boykoff (2007) conclude that this process might also lead
to an informational bias, especially toward the presentation of events
instead of a comprehensive analysis of the problem. Thus, an important
aspect of improving risk communication and the respective knowledge
base is the acceptance and admission of the limits of knowledge about
the future (see Birkmann and von Teichman, 2010).
2.7. Risk Accumulation and
the Nature of Disasters
The concept of risk accumulation describes a gradual build-up of disaster
risk in specific locations, often due to a combination of processes, some
persistent and/or gradual, others more erratic, often in a combination of
exacerbation of inequality, marginalization, and disaster risk over time
(Maskrey, 1993b; Lavell, 1994). It also reflects that the impacts of one
hazard – and the response to it – can have implications for how the
next hazard plays out. This is well illustrated by the example of El
Salvador, where people living in temporary shelters after the 1998
Hurricane Mitch were at greater risk during the 2001 earthquakes due
to the poor construction of the shelters (Wisner, 2001b). The concept of
risk accumulation acknowledges the multiple causal factors of risk by
the connecting development patterns and risk, as well as the links
between one disaster and the next.
Chapter 2 Determinants of Risk: Exposure and Vulnerability
96
Risk accumulation can be driven by underlying factors such as a decline
in the regulatory services provided by ecosystems, inadequate water
management, land use changes, rural-urban migration, unplanned
urban growth, the expansion of informal settlements in low-lying areas,
and an underinvestment in drainage infrastructure. Development and
governance processes that increase the marginalization of specific
groups, for example, through the reduction of access to health services
or the exclusion from information and power – to name just a few – can
also severely increase the susceptibility of these groups and at the same
time erode societal response capacities. The classic example is disaster
risk in urban areas in many rapidly growing cities in developing countries
(Pelling and Wisner, 2009b). In these areas, disaster risk is often very
unequally distributed, with the poor facing the highest risk, for instance
because they live in the most hazard-prone parts of the city, often in
unplanned dense settlements with a lack of public services; where lack
of waste disposal may lead to blocking of drains and increases the risk
of disease outbreaks when floods occur; with limited political influence
to ensure government interventions to reduce risk. The accumulation of
disaster risk over time may be partly caused by a string of smaller
disasters due to continued exposure to small day-to-day risks in urban
areas (e.g., Pelling and Wisner, 2009a), aggravated by limited resources
to cope and recover from disasters when they occur – creating a vicious
cycle of poverty and disaster risk. Analysis of disaster loss data suggests
that frequent low-intensity losses often highlight an accumulation of risks,
which is then realized when an extreme hazard event occurs (UNISDR,
2009a). Similar accumulation of risk may occur at larger scales in hazard-
prone states, especially in the context of conflict and displacement (e.g.,
UNDP, 2004).
A context-based understanding of these risks is essential to identify
appropriate risk management strategies. This may include better collection
of sub-national disaster data that allows visualization of complex patterns
of local risk (UNDP, 2004), as well as locally owned processes of risk
identification and reduction. Bull-Kamanga et al. (2003) suggest that
one of the most effective methods to address urban disaster risk in
Africa is to support community processes among the most vulnerable
groups so they can identify risks and set priorities – both for community
action and for action by external agencies (including local governments).
Such local risk assessment processes also avoid the pitfalls of planning
based on dated maps used to plan and develop large physical construction
and facilities.
Disaster risk is not an autonomous or externally generated circumstance
to which society reacts, adapts, or responds (as is the case with natural
phenomena or events per se), but rather the result of the interaction of
society and the natural or built environment. Thus disasters are often
the product of parallel developments that sometimes reach a tipping
point, where the cumulative effect of these parallel processes results in
disaster (Dikau and Pohl, 2007; Birkmann, 2011b). After that point,
recovery may be slowed by conflict between processes and goals of
reconstruction (Colten et al., 2008). In addition, there is often strong
pressure to restore the status quo as soon as possible after a disaster
has happened, even if that status quo means continued high levels of
disaster risk. Sometimes, however, disasters themselves can be a
window of opportunity for addressing the determinants of disaster risk.
With proactive risk assessment and reconstruction planning, more
appropriate solutions can be realized while restoring essential assets
and services during and after disasters (Susman et al., 1983, Renn,
1992; Comfort et al., 1999; Vogel and O’Brien, 2004).
Chapter 2Determinants of Risk: Exposure and Vulnerability
References
A digital library of non-journal-based literature cited in this chapter that
may not be readily available to the public has been compiled as part of
the IPCC review and drafting process, and can be accessed via either the
IPCC Secretariat or IPCC Working Group II web sites.
Abel, N., D. Cumming, and J. Anderies, 2006: Collapse and reorganization in social-
ecological systems: Questions, some ideas, and policy implications. Ecology and
Society, 11(1), 17-42.
Adam, B. and J. van Loon, 2000: Repositioning risk; the challenge for social theory.
In: The Risk Society and Beyond [Adam, B., U. Beck, and J. van Loon (eds.)].
SAGE Publications, London, UK, pp. 1-31.
Adelekan, I.O., 2010: Vulnerability of poor urban coastal communities to flooding in
Lagos, Nigeria. Environment and Urbanization, 22, 433, doi:10.1177/
0956247810380141.
Adger, W.N., 1999: Social vulnerability to climate change and extremes in coastal
Vietnam. World Development, 27(3), 249-269.
Adger, W.N., 2000: Social and ecological resilience: are they related? Progress in
Human Geography, 24(3), 347-364.
Adger, W.N., 2003: Social capital, collective action, and adaptation to climate
change. Economic Geography, 79(4), 387-404.
Adger, W.N., 2006: Vulnerability. Global Environmental Change, 16, 268-281.
Adger, W.N. (ed.), 2009: Adapting to Climate Change, Thresholds, Values,
Governance. Cambridge University Press, Cambridge, UK.
Adger, W.N. and N. Brooks, 2003: Does global environmental change cause
vulnerability to disaster? In: Natural Disasters and Development in a
Globalizing World [Pelling, M. (ed.)]. Routledge, London, UK, pp. 19-42.
Adger, W.N., and P.M. Kelly, 1999: Social vulnerability to climate change and the
architecture of entitlements. Mitigation and Adaptation Strategies for Global
Change, 4, 253-266.
Adger, W.N., N. Brooks, M. Kelly, S. Bentham, and S. Eriksen, 2004: New Indicators of
Vulnerability and Adaptive Capacity. Tyndall Centre for Climate Change
Research, Technical Report 7, University of East Anglia, Norwich, UK.
Adger, W.N., N.W. Arnell, and E.L. Tompkins, 2005: Successful adaptation to climate
change across scales. Global Environmental Change, 15(2), 77-86.
Afifi, T., 2011: Economic or environmental migration? The push factors in Niger.
International Migration, 49(S1), e95-e124.
Agrawala, S. and M.K. van Aalst, 2008: Adapting development co-operation to
adapt to climate change. Climate Policy, 8, 183-193.
Ahern, M., R.S. Kovats, P. Wilkinson, R. Few, and F. Matthies, 2005: Global health
impacts of floods: epidemiologic evidence. Epidemiologic Reviews,
27,
doi:10.1093/epirev/mxi004.
Alexander, D.E., 1993: Natural Disasters, UCL Press Limited, London, UK.
Alexander, D.E., 2000: Confronting Catastrophe. Terra Publishing, Harpenden, UK.
Altman, I. and S. Low, 1992: Place Attachment. Plenum Press, New York, NY.
Alwang, J., P.B. Siegel, and S.L. Jorgensen, 2001: Vulnerability: A View From Different
Disciplines. Social Protection Discussion Paper Series, No. 115, World Bank,
Washington, DC.
97
Amstadter, A.B., R. Acierno, L.K. Richardson, D.G. Kilpatrick, D.F. Gros, R.J. Johnson,
M.T. Gaboury, L.T. Trinh, T.T. Lam, T.T. Nguyen, T. Tran, T.B. La, T.H. Tran, D.C. Tran,
and G. Sandro, 2009: Post typhoon prevalence of posttraumatic stress disorder,
major depressive disorder, panic disorder, and generalized anxiety disorder in a
Vietnamese sample. Journal of Traumatic Stress, 22(3), 180-188.
Anderson, M.B. and P.J. Woodrow, 1989: Rising from the Ashes, Development
Strategies in Times of Disaster. 1998 ed. Lynne Rienner, London, UK, 338 pp.
Anderson, M.B. and P.J. Woodrow, 1991: Reducing vulnerability to drought and
famine: Developmental approaches to relief. Disasters, 15(1), 43-54.
Aragon-Durand, F., 2007: Urbanisation and flood vulnerability in the peri-urban
interface of Mexico City. Disasters, 31(4), 477-494.
Armitage, D., B. Fikret, and N. Doubleday (eds.), 2007: Adaptive Co-Management:
Collaboration, Learning, and Multi-Level Governance. UBC Press, Vancouver,
Canada, 160 pp.
Arrow, K. and R. Lind, 1970: Uncertainty and the evaluation of public investment
decisions. The American Economic Review, 60, 364-378.
Auld, H. and D. MacIver, 2007: Changing Weather Patterns, Uncertainty And
Infrastructure Risks: Emerging Adaptation Requirements. Occasional Paper 9,
Environment Canada, Toronto, Canada.
Aysan, Y., 1993: Vulnerability assessment. In: Natural Disasters: Protecting Vulnerable
Communities [Merriman, P.A. and C.W.A. Browitt (eds.)]. IDNDR-Thomas
Telford, London, UK.
Aysan, Y. and P. Oliver, 1987: Housing and Culture after Earthquakes. A guide for
future policy making on housing in seismic areas. Oxford Polytechnic for the
Overseas Development Administration (ODA) of the UK Government, Oxford,
UK.
Balbus, J.M. and C. Malina, 2009: Identifying vulnerable subpopulations for climate
change health effects in the United States. Journal of Occupational and
Environmental Medicine, 51, 33-37.
Bankoff, G., 2004: The Historical Geography of Disaster: ‘Vulnerability’ and ‘Local
Knowledge’ in Western Discourse. Earthscan, London, UK.
Barnett, J., 2001: Adapting to climate change in Pacific Island countries: The
problem of uncertainty. World Development, 29(6), 977-993.
Barnett, J., 2005: Titanic states? Impacts and responses to climate change in the
Pacific Islands. Journal of International Affairs, 59, 203-219.
Barnett, J. and M. Webber, 2009: Accommodating Migration to Promote Adaptation
to Climate Change. A policy brief prepared for the Secretariat of the Swedish
Commission on Climate Change and Development and the World Bank World
Development Report 2010 team, Commission on Climate Change and
Development, Stockholm, Sweden.
Barnston, A.G., S. Li, S.L. Mason, D.G. DeWitt, L. Goddard, and X. Gong, 2010:
Verification of the first 11 years of iri’s seasonal climate forecasts. Journal of
Applied Meteorology and Climatology, 49, 493-520.
Bartlett, S., 2008: Climate change and urban children: impacts and implications for
adaptation in low- and middle income countries. Environment & Urbanization,
20(2), 501-519.
Batterbury, S., 2001: Landscapes of diversity: A local political ecology of livelihood
diversification in south-western Niger. Ecumene, 8(4), 437-464.
Beck, U., 2000: Risk society revisited: Theory, politics and research programmes. In:
The Risk Society and Beyond [Adam, B., U. Beck, and J. van Loon (eds.)]. SAGE
Publications, London, UK, 211-229.
Beck, U., 2008: La Sociedad del Riesgo Mundial. Paidós, Barcelona, Spain.
Beg, N., J.C. Morlot, O. Davidson, Y. Afrane-Okesse, L. Tyani, F. Denton, Y. Sokona, J.P.
Thomas, E.L. La Rovere, J.K. Parikh, K. Parikh, and A.A. Rahman, 2002: Linkages
between climate change and sustainable development. Climate Policy, 2, 129-144.
Benioff, R., S. Guill, and J. Lee (eds.), 1996: Vulnerability and Adaptation
Assessments: An International Handbook. Kluwer Academic Publishers,
Dordrecht, The Netherlands.
Bennet, G., 1970: Bristol floods 1968: controlled survey of effects on health of local
community disaster. British Medical Journal, 3, 454-458.
Benson, C. and E. Clay, 2000: Developing countries and the economic impacts of
natural disasters. In Managing disaster risk in emerging economies [Kreimer, A.
and M. Arnold (eds.)]. World Bank, Washington, DC, pp. 11-21.
Bertoni, J.C., 2006: Urban floods in Latin America: reflections on the role of risk
factors. In: Frontiers in Flood Research [Tchiguirinskaia, I., K.N.N. Thein, and P.
Hubert (eds.)]. IAHS publication 305, International Association of Hydrological
Sciences, Wallingford, UK, pp. 123-141.
Bettencourt, S., R. Croad, P. Freeman, J. Hay, R. Jones, P. King, P. Lal, A. Mearns, G.
Miller, I. Pswarayi-Riddihough, A. Simpson, N. Teuatabo, U. Trotz, and M.K. van
Aalst, 2006: Not If, But When. Adapting To Natural Hazards in the Pacific Islands
Region. World Bank, Washington, DC.
Biermann, F., 2007: ‘Earth system governance’ as a crosscutting theme of global
change research. Global Environmental Change, 17(3-4), 326-337.
Biermann, F., M. Betsill, J. Gupta, N. Kanie, L. Lebel, D. Livermann, H. Schroeder, and
B. Siebenhüner, 2009: Earth System Governance – People, Places, and the
Planet. Report No. 1, Science and Implementation Plan of the Earth System
Governance Project, Bonn, Germany.
Birkmann, J., 2006a: Measuring Vulnerability to Natural Hazards – Towards Disaster
Resilient Societies. United Nations University Press, Tokyo, Japan, 450 pp.
Birkmann, J., 2006b: Measuring vulnerability to promote disaster-resilient societies:
conceptual frameworks and definitions. In: Measuring Vulnerability to Natural
Hazards: Towards Disaster Resilient Societies [Birkmann, J. (ed.)]. United
Nations University Press, Tokyo, Japan, pp. 9-54.
Birkmann, J., 2006c: Conclusions and recommendations. In: Measuring Vulnerability
to Natural Hazards: Towards Disaster Resilient Societies [Birkmann, J. (ed.)].
United Nations University Press, Tokyo, Japan, pp. 432-447.
Birkmann, J., 2011a: First and second-order adaptation to natural hazards and
extreme events in the context of climate change. Natural Hazards, 58(2), 811-
840, doi:10.1007/s11069-011-9806-8.
Birkmann, J., 2011b: Regulation and coupling of society and nature in the context
of natural hazards. In: Coping with Global Environmental Change, Disasters and
Security [Brauch, H.G., U. Oswald Spring, C. Mesjasz, J. Grin, P. Kameri-Mbote, B.
Chourou, P. Dunay, J. Birkmann (eds.)]. Springer, Berlin, Germany, pp. 1103-1127.
Birkmann, J. and N. Fernando, 2008: Measuring revealed and emergent vulnerabili-
ties of coastal communities to tsunami in Sri Lanka. Disasters, 32(1), 82-104.
Birkmann, J. and K. von Teichman, 2010: Integrating disaster risk reduction and
climate change adaptation: key challenges – scales, knowledge, and norms.
Sustainability Science, 5(2), 171-184.
Birkmann, J., K. von Teichman, P. Aldunce, C. Bach, N.T. Binh, M. Garschagen, S.
Kanwar, N. Setiadi, and L.N. Thach, 2009: Addressing the Challenge:
Recommendations and Quality Criteria for Linking Disaster Risk Reduction and
Adaptation to Climate Change [Birkmann, J., G. Tetzlaff, and K-O. Zentel (eds.)].
DKKV Publication Series No. 38, German Committee for Disaster Reduction,
Bonn, Germany.
Birkmann, J., P. Buckle, J. Jaeger, M. Pelling, N. Setiadi, M. Garschagen, N. Fernando,
and J. Kropp, 2010a: Extreme events and disasters: A window of opportunity for
change? Analysis of changes, formal and informal responses after mega-
disasters. Natural Hazards, 55(3), 637-669.
Birkmann, J., M. Garschagen, F. Kraas, and N. Quang, 2010b: Adaptive urban
governance: new challenges for the second generation of urban adaptation
strategies to climate change. Sustainability Science, 5(2), 185-206.
Black, R., 2001: Environmental Refugees: Myth or Reality? UNHCR Working Paper
34, UN High Commissioner for Refugees, Geneva, Switzerland, 19 pp.
Blaikie, P., T. Cannon, I. Davis, and B. Wisner, 1994: At Risk: Natural Hazards, People,
Vulnerability, and Disasters. Routledge, London, UK.
Blaikie, P., T. Cannon, I. Davis, and B. Wisner, 1996: Vulnerabilidad, el entorno social
de los desastres. La RED-ITDG, Bogota, Colombia.
Blockley, D. (ed.), 1992: Engineering Safety. McGraw-Hill International Series in Civil
Engineering, London, UK.
Boentje, J.P. and M.S. Blinnikov, 2007: Post-Soviet forest fragmentation and loss in the
Green Belt around Moscow, Russia (1991-2001): a remote sensing perspective.
Landscape and Urban Planning, 82(4), 208-221.
Bogardi, J. and J. Birkmann, 2004: Vulnerability assessment: the first step towards
sustainable risk reduction. In: Disasters and Society – From Hazard Assessment
to Risk Reduction [Malzahn, D. and T. Plapp (eds.)]. Logos Verlag, Berlin,
Germany, pp. 75-82.
Chapter 2 Determinants of Risk: Exposure and Vulnerability
98
Bohle, H-G., 2001: Vulnerability and criticality: Perspectives from social geography.
IHDP Update, 2/2001, 1-7.
Bohle, H-G., 2008: Krisen, katastrophen, kollaps – Geographien von verwundbarkeit
in der risikogesellschaft. In: Umgang mit Risiken. Katastrophen –
Destabilisierung – Sicherheit [Kulke, E. and H. Popp (eds.)]. Deutscher
Geographentag 2007, Bautzen, Lausitzer Druck- und Verlagshaus GmbH,
Bayreuth, Germany, pp. 69-82.
Bohle, H-G., and T. Glade, 2008: Vulnerabilitätskonzepte in Sozial- und
Naturwissenschaften. In: Naturrisiken und Sozialkatastrophen [Felgentreff, C.
and T. Glade (eds.)]. Spektrum Verlag, Berlin, Germany, pp. 99-119.
Bohle, H.G., T.E. Downing, and M.J. Watts, 1994: Climate change and social
vulnerability: Toward a sociology and geography of food insecurity. Global
Environmental Change, 4(1), 37-48.
Bouchama, A., M. Dehbi, G. Mohamed, F. Matthies, M. Shoukri, and B. Menne, 2007:
Prognostic factors in heat wave related deaths: a meta-analysis. Archives of
Internal Medicine, 167, 2170-2176.
Bouwer, L.M., R.P. Crompton, E. Faust, P. Hoppe, and R.A. Pielke Jr., 2007: Disaster
management: Confronting disaster losses. Science, 318 (5851), 753.
Boykoff, M.T. and J. Boykoff, 2007: Climate change and journalistic norms: a
case-study of US mass-media coverage. Geoforum, 38, 1190-1204.
Breiting, S. and P. Wickenberg, 2010: The progressive development of environmental
education in Sweden and Denmark. Environmental Education Research, 16(1),
9-37.
Brenkert, A.L. and E.L. Malone, 2005: Modeling vulnerability and resilience to
climate change: A case study of India and Indian States. Climatic Change,
72(1-2), 57-102.
Brklacich, M. and H.-G. Bohle, 2006: Assessing human vulnerability to global climatic
change. In: Earth System Science in the Anthropocene [Ehlers, E. and T. Krafft
(eds.)]. Springer, Berlin, Germany, pp. 51-61.
Brondo, K.V., 2007: Land loss and Garifuna women’s activism on Honduras’ North
Coast. Journal of International Women’s Studies, 9(1), 99-116.
Brooks, N., 2003: Vulnerability, Risk and Adaptation: A Conceptual Framework.
Tyndall Centre for Climate Change Working Paper 38, University of East Anglia,
Norwich, UK.
Brooks, N. and W.N. Adger, 2004: Assessing and enhancing adaptive capacity,
Technical Paper 7. In: Adaptation Policy Frameworks for Climate Change:
Developing Strategies, Policies and Measures [Lim, B. and E. Spanger-Siegfried
(eds.)]. United Nations Development Programme and Cambridge University
Press, New York, NY, pp. 165-182.
Brooks, N., W.N. Adger, and M. Kelly, 2005: The determinants of vulnerability and
adaptive capacity at the national level and the implications for adaptation.
Global Environmental Change Part B: Environmental Hazards, 15, 151-163.
Bull-Kamanga, L., K. Diagne, A. Lavell, E. Leon, F. Lerise, H. MacGregor, A. Maskrey,
M. Meshack, M. Pelling, H. Reid, D. Satterthwaite, J. Songsore, K. Westgate, and
A. Yitambe, 2003: From everyday hazards to disasters: the accumulation of risk
in urban areas. Environment & Urbanization
, 15(1), 193-203.
Burby, R.J., 2006: Hurricane Katrina and the paradoxes of government disaster
policy: Bringing about wise governmental decisions for hazardous areas. The
Annals of the American Academy of Political and Social Science, 604, 171-191.
Burkle, F.M. and P.G. Greenough, 2008: Impact of public health emergencies on
modern disaster taxonomy, planning, and response. Disaster Medicine and
Public Health Preparedness, 2(4), e2-e9.
Burton, I. and M.K. van Aalst, 1999: Come Hell or High Water: Integrating Climate
Change Vulnerability and Adaptation into Bank Work. World Bank Environment
Department Papers 72, World Bank, Washington, DC.
Burton, I. and M.K. van Aalst, 2004: Look Before You Leap. A Risk Management
Approach for Climate Change Adaptation in World Bank Operations. World
Bank, Washington, DC.
Burton, I., R.W. Kates, and G.F. White, 1978: The Environment as Hazard. Oxford
University Press, New York, NY.
Burton, I., J. Wilson, and R.E. Munn, 1983: Environmental impact assessment:
national approaches and international needs. Environmental Monitoring and
Assessment, 3, 133-150.
Burton, I., R.W. Kates, and G.F. White, 1993: The Environment as Hazard – Second
Edition. Oxford University Press, New York, NY.
Campbell, D., M. Stafford Smith, J. Davies, P. Kuipers, J. Wakerman, and M.J.
McGregor, 2008: Responding to health impacts of climate change in the
Australian desert. Rural and Remote Health, 8, 1008.
Campbell, J.R. 2006. Traditional Disaster Reduction in Pacific Island Communities.
GNS Science Report 2006/38. Institute of Geological and Nuclear Sciences,
Avalon, NZ, ISBN 0-478-09961-4.
Cannon, T., 1994: Vulnerability analysis and the explanation of ‘natural’ disasters. In:
Disasters, Development and Environment [Varley, A. (ed.)]. John Wiley and Sons,
Chichester, UK, pp. 13-29.
Cannon, T., 2006: Vulnerability analysis, livelihoods and disasters. In: Risk 21: Coping
with Risks Due to Natural Hazards in the 21st Century [Ammann, W.J., S.
Dannenmann, and L. Vulliet (eds.)]. Taylor and Francis Group, London, UK, pp.
41-49.
Cannon, T. and D. Müller-Mahn, 2010: Vulnerability, resilience and development
discourses in the context of climate change. Natural Hazards, 55(3), 621-635.
Cannon, T., J. Twigg, and J. Rowell, 2003: Social Vulnerability, Sustainable Livelihoods
and Disasters. Conflict and Humanitarian Assistance Department and
Sustainable Livelihoods Support Office, Department for International
Development, London, UK, 63 pp.
Cardona, O.D., 1986: Estudios de vulnerabilidad y evaluación del riesgo sísmico:
Pla nificación física y urbana en áreas propensas. Boletín Técnico de la
Asociación Colombiana de Ingeniería Sísmica, 33(2), 32-65.
Cardona, O.D., 1990: Terminología de Uso Común en Manejo de Riesgos. AGID
Reporte No. 13, Escuela de Administración, Finanzas, y Tecnología, Medellín,
Colombia.
Cardona, O.D., 1993: Evaluación de la amenaza, la vulnerabilidad y el riesgo:
Elementos para el Ordenamiento y la Planeación del Desarrollo. In:
Los
Desastres No son Naturales [Maskrey, A. (ed.)]. La RED/Tercer Mundo Editores,
Bogotá, Colombia.
Cardona, O.D., 1996a: Manejo ambiental y prevención de desastres: dos temas aso-
ciados. In: Ciudades en Riesgo [Fernandez, M.A. (ed.)]. La RED-USAID, Lima,
Peru, pp. 79-101.
Cardona, O.D., 1996b: El Manejo de riesgos y los preparativos para Desastres:
Compromiso Institucional para mejorar la calidad de vida. In: Desastres:
modelo para armar [Mansilla, E. (ed.)]. La RED, Lima, Peru, pp. 128-147.
Cardona, O.D., 1996c: Variables involucradas en el manejo de riesgos. In: Desastres
y Sociedad, Especial: Predicciones, Pronósticos, Alertas y Respuestas Sociales
[Cardona, O.D. (ed.)]. 4, La RED, Tarea Gráfica, Lima, Peru, pp. 7-35.
Cardona, O.D., 1999a: Environmental management and disaster prevention: Two
related topics: A holistic risk assessment and management approach. In:
Natural Disaster Management [Ingleton, J. (ed.)]. IDNDR-Tudor Rose, London,
UK, pp. 151-153.
Cardona, O.D., 1999b: Environmental management and disaster prevention: Two
related topics. In: Cities at Risk: Environmental Degradation, Urban Risks and
Disasters in Latin America [Fernandez, M.A. (ed.)]. A/H Editorial, La RED, US AID,
Quito, Peru, pp. 79-102.
Cardona, O.D., 2001: Estimación Holística del Riesgo Sísmico utilizando Sistemas
Dinámicos Complejos. Doctoral dissertation, Department of Terrain
Engineering, Technical University of Catalonia, Spain.
Cardona, O.D., 2004: The need for rethinking the concepts of vulnerability and risk
from a holistic perspective: A necessary review and criticism for effective risk
management. In: Mapping Vulnerability: Disa sters, Development and People
[Bankoff, G., G. Frerks, and D. Hilhorst (eds.)]. Earthscan Publishers, London, UK,
pp. 37-51.
Cardona, O.D., 2006: A system of indicators for disaster risk management in the
Americas. In: Measuring Vulnerability to Hazards of Natural Origin: Towards
Disaster Resilient Societies [Birkmann, J. (ed.)]. UNU Press, Tokyo, Japan, pp.
189-209.
Cardona, O.D., 2008: Indicators of Disaster Risk and Risk Management: Program for
Latin America and the Caribbean – Summary Report – Second Edition. INE-08-
002, Inter-American Development Bank, Washington, DC.
Chapter 2Determinants of Risk: Exposure and Vulnerability
99
Cardona, O.D., 2009: La Gestión Financiera del Riesgo de Desastres: Instrumentos
Financieros de Retención y Transferencia para la Comunidad Andina. PREDECAN,
Comunidad Andina, Lima, Peru.
Cardona, O.D., 2010: Indicators of Disaster Risk and Risk Management – Program
for Latin America and the Caribbean: Summary Report. Evaluación de Riesgos
Naturales - Latino America, ERN-AL, Inter-American Development Bank,
Washington, DC.
Cardona, O.D., 2011: Disaster risk and vulnerability: Notions and measurement of
human and environmental insecurity. In: Coping with Global Environmental
Change, Disasters and Security – Threats, Challenges, Vulnerabilities and
Risks [Brauch, H.G., U. Oswald Spring, C. Mesjasz, J. Grin, P. Kameri-Mbote,
B. Chourou, P. Dunay, J. Birkmann]. Springer Verlag, Berlin, Germany, pp. 107-
122.
Cardona, O.D. and A.H. Barbat, 2000: El Riesgo Sísmico y su Prevención. Cuaderno
Técnico 5, Calidad Siderúrgica, Madrid, Spain, 190 pp.
Cardona, O.D. and J.E. Hurtado, 2000a: Holistic seismic risk estimation of a
metropolitan center. In: Proceedings of the 12th World Conference of
Earthquake Engineering, January 30 – February 4, 2000, paper 2643. New
Zealand Society for Earthquake Engineering, Auckland, New Zealand.
Cardona, O.D. and J.E. Hurtado, 2000b: Modelación numérica para la estimación
holística del riesgo sísmico urbano, considerando variables técnicas, sociales y
económicas. In: Métodos Numéricos en Ciencias Sociales (MENCIS 2000)
[Oñate, E., F. García-Sicilia, and L. Ramallo (eds.)]. Centro Internacional de
Métodos Numéricos en Ingeniería - Universidad Politécnica de Cataluña,
Barcelona, Spain, pp. 452-466.
Cardona, O.D., J.E. Hurtado, G. Duque, A. Moreno, A.C. Chardon, L.S. Velásquez, and
S.D. Prieto, 2003a: The Notion of Disaster Risk: Conceptual Framework for
Integrated Management. National University of Colombia / Inter-American
Development Bank, Washington, DC.
Cardona, O.D., J.E. Hurtado, G. Duque, A. Moreno, A.C. Chardon, L.S. Velásquez, and
S.D. Prieto, 2003b: Indicators for Risk Measurement: Methodological
Fundamentals. National University of Colombia / Inter-American Development
Bank, Washington, DC.
Cardona, O.D., M.G. Ordaz, M.C. Marulanda, and A.H. Barbat, 2008a: Estimation of
probabilistic seismic losses and the public economic resilience – An approach
for a macroeconomic impact evaluation. Journal of Earthquake Engineering,
12(S2), 60-70.
Cardona, O.D., M.G. Ordaz, L.E. Yamín, M.C. Marulanda, and A. H. Barbat, 2008b:
Earthquake loss assessment for integrated disaster risk management. Journal of
Earthquake Engineering, 12(S2), 48-59.
CARE, 2009: Climate Vulnerability and Capacity Analysis Handbook. CARE
International, Chatelaine, Switzerland.
Carreño, M.L., 2006: Técnicas innovadoras para la evaluación del riego sísmico y su
gestión en centros urbanos: Acciones ex ante y ex post. Doctoral dissertation,
Department of Terrain Engineering, Technical University of Catalonia, Spain.
Carreño, M.L., O.D. Cardona, and A.H. Barbat, 2005: Sistema de Indicadores para la
Evaluación de Riesgos. Monografía CIMNE IS-52, Technical University of
Catalonia, Barcelona, Spain, 160 pp.
Carreño, M.L., O.D. Cardona, and A.H. Barbat, 2007a: Urban seismic risk evaluation:
A holistic approach. Journal of Natural Hazards, 40(1), 137-172.
Carreño, M.L., O.D. Cardona, and A.H. Barbat, 2007b: A disaster risk management
performance index. Journal of Natural Hazards, 41(1), 1-20.
Carreño, M.L., O.D. Cardona, M.C. Marulanda, and A.H. Barbat, 2009: Holistic urban
seismic risk evaluation of megacities: Application and robustness. In: The 1755
Lisbon Earthquake: Revisited [Mendes-Victor, L.A., C.S. Sousa Oliveira, J.
Azevedo, and A. Ribeiro (eds.)]. Springer, Berlin, Germany.
Carroll, B., H. Morbey, R. Balogh, and G. Araoz, 2006: Living in Fear: Health and
Social Impacts of the Floods in Carlisle 2005
. Research Report, Centre for Health
Research and Practice Development, St. Martins College, Carlisle, UK.
Carter, T.R., M.L. Parry, H. Harasawa, and S. Nishioka, 1994: IPCC Technical
Guidelines for Assessing Climate Change Impacts and Adaptations. Center for
Global Environmental Research, National Institute for Environmental Studies,
Tsukuba, Japan.
Carvalho, A., 2005: Representing the politics of the greenhouse effect: discursive
strategies in the British media. Critical Discourse Studies, 2(1), 1-29.
Carvalho, A. and J. Burgess, 2005: Cultural circuits of climate change in U.K.
broadsheet newspapers, 1985-2003. Society for Risk Analysis, 25(6), 1457-1469.
Castles, S., 2002: Environmental Change and Forced Migration: Making Sense of the
Debate. UNHCR Issues in Refugee Research, Working Paper No. 70, UN High
Commissioner for Refugees, Geneva, Switzerland.
Ceccato, P., T. Ghebremeskel, M. Jaiteh, P.M. Graves, M. Levy, S. Ghebreselassie, A.
Ogbamariam, A.G. Barnston, M. Bell, J. del Corral, S.J. Connor, I. Fesseha, E.P.
Brantly, and M.C. Thomson, 2007: Malaria stratification, climate, and epidemic
early warning in Eritrea. American Journal of Tropical Medicine and Hygiene,
77, 61-68.
Cernea, M., 1996: Eight Main Risks: Impoverishment and Social Justice in
Resettlement. World Bank Environment Department, Washington, DC.
Chambers, R., 1983: Rural Development – Putting the Last First. Longmans Scientific
and Technical Publishers, Essex, UK, 246 pp.
Chambers, R., 1989: Vulnerability, coping and policy. Institute of Development
Studies Bulletin, 20(2), 1-7.
Changnon, D. and S.A. Changnon, 2010: Major growth in some business related
uses of climate information. Journal of Applied Meteorology and Climatology,
49, 325-331.
Chatterjee, M., 2010: Slum dwellers response to flooding events in the megacities
of India. Adaptation and Mitigation Strategies for Global Change, 15, 337-353.
Chester, D.K., 2005: Theology and disaster studies: the need for dialogue. Journal of
Volcanology and Geothermal Research, 146 (4), 319-328.
Christoplos, I., 2006: The elusive ‘window of opportunity’ for risk reduction in post-
disaster recovery. Discussion Paper, ProVention Consortium Forum, 2-3 February
2006, Bangkok, Thailand.
Clasen, T., W.P Schmidt, T. Rabie, I. Roberts, and S. Cairncross, 2007: Interventions to
improve water quality for preventing diarrhoea: systematic review and meta-
analysis. British Medical Journal, 334(7597), 782-785.
Coburn, A. and R. Spence, 1992: Earthquake Protection. John Wiley & Sons,
Chichester, UK.
Colchester, M., 2000: Dams, Indigenous People and Vulnerable Ethnic Minorities.
Thematic Review 1.2. World Commission on Dams, Cape Town, South Africa.
Colten, C.E., 2006: Vulnerability and place: Flat land and uneven risk in New Orleans.
American Anthropologist,
108(4), 731-734.
Colten, C., R. Kates, and S. Laska, 2008: Community Resilience: Lessons for New
Orleans and Hurricane Katrina. CARRI Research Report 3, Community and
Regional Resilience Institute, Oak Ridge, TN, 47 pp.
Comfort, L., B. Wisner, S. Cutter, R. Pulwarty, K. Hewitt, A. Oliver-Smith, J. Wiener, M.
Fordham, W. Peacock, and F. Krimgold, 1999: Reframing disaster policy: the
global evolution of vulnerable communities. Environmental Hazards, 1, 39-44.
Confalonieri, U., B. Menne, R. Akhtar, K.L. Ebi, M. Hauengue, R.S. Kovats, B. Revich,
and A. Woodward, 2007: Human health. In: Climate Change 2007. Impacts,
Adaptation and Vulnerability. Contribution of Working Group II to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change [Parry,
M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)].
Cambridge University Press, Cambridge, UK, 391-431.
Cooper, P.J.M., J. Dimes, K.P.C. Rao, B. Shapiro, B. Shiferaw, and S. Twomlow, 2008:
Coping better with current climatic variability in the rain-fed farming systems
of sub-Saharan Africa: An essential first step in adapting to future climate
change? Agriculture, Ecosystems and Environment, 126, 24-35.
Corral, S., 2000: Explorando la calidad de los procesos de elaboración de políticas
ambientales. In: Métodos Numéricos en Ciencias Sociales (MENCIS 2000)
[Oñate, E., F. García-Sicilia, and L. Ramallo (eds.)]. Centro Internacional de
Métodos Numéricos en Ingeniería - Universidad Politécnica de Cataluña,
Barcelona, Spain, pp. 391-401.
Costello, A., M. Abbas, A. Allen, S. Ball, S. Bell, R. Bellamy, S. Friel, N. Groce, A. Johnson,
M. Kett, M. Lee, C. Levy, M. Maslin, D. McCoy, B. McGuire, H. Montgomery, D.
Napier, C. Pagel, J. Patel, C. Patterson, J.A. Puppim de Oliveira, N. Redclift, H.
Rees, D. Rogger, J. Scott, J. Stephenson, J. Twigg, and J. Wolff, 2009: Managing
the health effects of climate change. The Lancet, 373(16), 1693-1733.
Chapter 2 Determinants of Risk: Exposure and Vulnerability
100
Covello, V. and J. Mumpower, 1985: Risk analysis and risk management: An
historical perspective. Risk Analysis, 5(2), 103-120.
Cross, J.A., 2001: Megacities and small towns: Different perspectives on hazard
vulnerability. Environmental Hazards, 3(2), 63-80.
Cruz, A.M., L.J. Steinberg, A.L.V. Arellano, J.-P. Nordvik, and F. Pisano, 2004: State of
the Art in Natech Risk Management (NATECH: Natural Hazard Triggering a
Technological Disaster). EUR 21292, European Communities, European
Commission, Brussels, Belgium.
Cummins, J. and O. Mahul, 2009: Catastrophe Risk Financing in Developing
Countries: Principles for Public Intervention. The World Bank, Washington, DC,
268 pp.
Cuny, F.C., 1984: Disaster and Development. Oxford University Press, New York, NY.
Cutter, S.L. (ed.), 1994: Environmental Risks and Hazards. Prentice Hall, Upper
Saddle River, NJ.
Cutter, S.L., 1996: Vulnerability to environmental hazards. Progress in Human
Geography, 20, 529-539.
Cutter, S.L. and C. Finch, 2008: Temporal and spatial changes in social vulnerability
to natural hazards. Proceedings of the National Academy of Sciences, 105(7),
2301-2306.
Cutter, S.L., C.T. Emrich, J.T. Mitchell, B.J. Boruff, M. Gall, M.C. Schmidtlein, C.G.
Burton, and G. Melton, 2006: The long road home: Race, class, and recovery
from Hurricane Katrina. Environment, 48(2), 8-20.
Cutter, S.L., L. Barnes, M. Berry, C. Burton, E. Evans, E. Tate, and J. Webb, 2008: A
place-based model for understanding community resilience to natural disasters.
Global Environmental Change, 18, 598-606.
Davis, I. and M. Wall (eds.), 1992: Christian Perspectives on Disaster Management: A
Training Manual. International Relief and Development Association, Middlesex,
UK.
Davis, I., B. Haghebaert, and D. Peppiatt, 2004: Social Vulnerability and Capacity
Analysis Workshop Discussion paper and workshop report. ProVention
Constortium, Geneva, Switzerland.
Dawson, T.P., S.T. Jackson, J.I. House, C.I. Prentice, and G.M. Mace, 2011: Beyond
predictions: Biodiversity conservation in a changing climate. Science, 332, 53.
Dayton-Johnson, J., 2004: Natural Disasters and Adaptive Capacity. OECD Working
Paper No. 237, Organisation for Economic Co-operation and Development,
Paris, France.
Décamps, H. (ed.), 2010: Événements Climatiques Extrêmes: Réduire les
Vulnérabilités des Systèmes Écologiques et Sociaux. Institut de France,
Académie des Sciences, Paris, France, 194 pp.
Degallier, N., C. Favier, C. Menkes, M. Lengaigne, W.M. Ramalho, R. Souza, J. Servain,
J.-P. Boulanger, 2010: Toward an early warning system for dengue prevention:
modeling climate impact on dengue transmission. Climatic Change, 98, 581-592.
Demetriades
, J. and E. Esplen, 2008: The gender dimensions of poverty and climate
change adaptation. IDS Bulletin, 39(4), 24.
Di Baldassarre, G., A. Montanari, H. Lins, D. Koutsoyiannis, L. Brandimarte, and G.
Bloeschl, 2010: Flood fatalities in Africa: From diagnosis to mitigation.
Geophysical Research Letters, 37, L22402.
Dikau, R. and J. Pohl, 2007: “Hazards” Naturgefahren und Naturrisiken. In:
Geographie, Physische Geographie und Humangeographie [Gebhardt, H., R.
Glaser, U. Radtke, and P. Reuber (eds.)]. Spektrum Akademischer Verlag, Berlin,
Germany, pp. 1029-1076.
Dilley, M., 2006: Disaster risk hotspots: A project summary. In: Measuring
Vulnerability to Natural Hazards – Towards Disaster Resilient Societies
[Birkmann, J. (ed.)]. United Nations University Press, Tokyo, Japan, pp. 182-188.
Dilley, M., R. Chen, U. Deichmann, A. Lerner-Lam, and M. Arnold, 2005: Natural
Disaster Hotspots: A Global Risk Analysis. Columbia University and World Bank,
New York, NY and Washington, DC.
DKKV, 2011: Adaptive Disaster Risk Reduction – Enhancing Methods and Tools of
Disaster Risk Reduction in the light of Climate Change. DKKV Publication Series
no. 43, German Committee for Disaster Reduction, Bonn, Germany,
www.dkkv.org/de/publications/schriftenreihe.asp?h=5.
Doherty, T.J. and S. Clayton, 2011: The psychological impacts of global climate
change. American Psychologist, 66(4), 265-276.
Donner, W. and H. Rodriguez, 2008: Population composition, migration and inequality:
The influence of demographic changes on disaster risk and vulnerability. Social
Forces, 87(2), 1089-1114.
Doocy, S., Y. Gorokhovich, G. Burnham, D. Balk, and C. Robinson, 2007: Tsunami
mortality estimates and vulnerability mapping in Aceh, Indonesia. American
Journal of Public Health, 97(S1), 146-151.
Dore, M.H.I. and D. Etkin, 2003: Natural disasters, adaptive capacity and development
in the twenty-first century. In: Natural Disasters and Development in A
Globalizing World [Pelling, M. (ed.)]. Routledge, London, UK, pp. 75-91.
Douglas, I., K. Alam, M. Maghenda, Y. McDonnell, L. McLean, and J. Campbell, 2008:
Unjust waters: climate change, flooding and the urban poor in Africa.
Environment And Urbanization, 20(1), 187-205.
Douglas, M. and A. Wildavsky, 1982: Risk and culture: An essay on the selection of
technological and environmental dangers. University of California Press,
Berkeley, CA.
Dow, K., 1992: Exploring differences in our common future(s): the meaning of
vulnerability to global environmental change. Geoforum, 23 (3), 417-436.
Downing, T.E. and A. Patwardhan, 2004: Assessing vulnerability for climate adaptation,
Technical Paper 3. In: Adaptation Policy Frameworks for Climate Change:
Developing Strategies, Policies and Measures [Lim, B. and E. Spanger-Siegfried
(eds.)]. United Nations Development Programme and Cambridge University
Press, New York, NY, pp. 67-89.
Drusine, H., 2005: The Garifuna fight back. Third Text, 19(2), 197-202.
Dudasik, S., 1982: Unanticipated repercussions of international disaster relief.
Disasters, 6,31-37.
Duneier, M., 2004: Scrutinizing the heat: On ethnic myths and the importance of
shoe leather.
Contemporary Sociology, 33 (2), 139-150.
Eakin, H., 2005: Institutional change, climate risk, and rural vulnerability: Cases from
Central Mexico. World Development, 33(11), 1923-1938.
Elliot, L., M. Beeson, S. Akbarzadeh, G. Fealy, and S. Harris, 2006: Religion, Faith and
Global Politics. Department of International Relations, Canberra, Australia.
Elliott, J.R. and Pais, J. 2006: Race, class, and Hurricane Katrina: Social differences in
human responses to disaster. Social Science Research, 35, 295-321.
Ellis, B.R. and B.A. Wilcox, 2009: The ecological dimensions of vector-borne disease
research and control. Cadernos De Saude Publica, 25, S155-S167.
Enarson, E. and B.H. Morrow (eds.), 1998: The Gendered Terrain of Disaster: Through
Women’s Eyes. Praeger Publishers, Westport, CT.
Enarson, E. and B.H. Morrow, 2000: A gendered perspective: The voices of women.
In: Hurricane Andrew: Ethnicity, Gender and the Sociology of Disasters
[Peacock, W.G., B.H. Morrow, and H. Gladwin (eds)]. International Hurricane
Centre, Laboratory for Social and Behavioural Research, Miami, FL, pp. 116-137.
Eriksen, S.H., and K. O’Brien, 2007: Vulnerability, poverty and the need for
sustainable adaptation measures. Climate Policy, 7(4), 337-352.
Eriksen, S. and J.A. Silva, 2009: The vulnerability context of a savanna area in
Mozambique: household drought strategies and responses to economic
change. Environmental Science and Policy, 12, 33-52.
Eriksen, S., H.K. Brown, and P.M. Kelly, 2005: The dynamics of vulnerability: locating
coping strategies in Kenya and Tanzania. Geographical Journal, 171 (4), 287-305.
ERN-AL, 2011: Probabilistic Modelling of Natural Risk at Global Level: The Hybrid
Loss Exceedance Curve – Development of a Methodology and Implementation
of Case Studies, Phase 1A: Colombia, Mexico, Nepal. Report for the GAR 2011,
Consortium Evaluación de Riesgos Naturales – América Latina, Bogotá.
FAOSTAT, 2011: FAOSTAT Statistical database on agricultural employment. Food and
Agriculture Organization, Rome, Italy, faostat.fao.org.
Feenstra, J.F., I. Burton, J.B. Smith, and R.S.J. Tol, 1998: Handbook on Methods for
Climate Change Impact Assessment and Adaptation Strategies. Version 2.0,
United Nations Environment Programme and Dutch National Institute for Public
Health and the Environment (UNEP/RIVM), Nairobi, Kenya and Amsterdam, The
Netherlands.
Few, R., 2007: Health and climatic hazards: framing social research on vulnerability,
response and adaptation. Global Environmental Change, 17, 281-295.
Folke, C., 2006: Resilience: The emergence of a perspective for social-ecological
systems analyses. Global Environmental Change, 16(3), 253-267.
Chapter 2Determinants of Risk: Exposure and Vulnerability
101
Folke, C., S. Carpenter, T. Elmqvist, L. Gunderson, C.S. Holling, and B. Walker, 2002:
Resilience and sustainable development: Building adaptive capacity in a
world of transformations. AMBIO: A Journal of the Human Environment,
31(5), 437-440.
Fordham, M., 1998: Making women visible in disasters: problematising the private
domain. Disasters, 22(2), 126-143.
Fordham, M. and S. Gupta with S. Akerkar and M. Scharf, 2011: Leading Resilient
Development: Grassroots Women’s Priorities, Practices and Innovations.
Grassroots Organizations Operating Together in Sisterhood and the UN
Development Programme, New York, NY.
Fothergill, A., E.G.M. Maestas, and J.D. Darlington, 1999: Race, ethnicity and disasters
in the United States: A review of the literature. Disasters, 23(2), 156-173.
Fournier d’Albe, M., 1985: The quantification of seismic hazard for the purposes of
risk assessment. In: International Conference on Reconstruction, Restoration
and Urban Planning of Towns and Regions in Seismic Prone Areas, 5-9
November 1985. Skopje, Yugoslavia, pp. 77-84.
Fraser, E.D.G., 2006: Food system vulnerability: using past famines to help understand
how food systems may adapt to climate change. Ecological Complexity, 3, 328-335.
Freeman, P. and K. Warner, 2001: Vulnerability of Infrastructure to Climate
Variability: How Does This Affect Infrastructure Lending Policies? Disaster
Management Facility of The World Bank and the ProVention Consortium,
Washington, DC, 40 pp.
Freudenberg, W., R. Gramling, S. Laska, and K. Erickson, 2008: Organising hazards,
engineering disasters? Improving the recognition of politico-economic factors
in the creation of disasters. Social Factors, 78(2), 1015-1038.
Furr, J.M., J.S. Comer, J.M. Edmunds, and P.C. Kendall, 2010: Disasters and youth: A
meta-analytic examination of posttraumatic stress. Journal of Consulting and
Clinical Psychology, 78, 765-780.
Füssel, H.-M., 2005: Vulnerability to climate change: a comprehensive conceptual
framework. University of California International and Area Studies Breslauer
Symposium Paper 6, Berkeley, CA, 36 pp.
Füssel, H.-M., 2007: Vulnerability: A generally applicable conceptual framework for
climate change research. Global Environmental Change, 17, 155-167.
Füssel, H.-M. and R.J.T. Klein, 2006: Climate change vulnerability assessments: an
evolution of conceptual thinking. Climatic Change, 75, 301-329.
Gaillard, J.C., 2007: Resilience of traditional societies in facing natural hazard.
Disaster Prevention and Management, 16, 522-544.
Gaillard, J.C., 2010: Vulnerability, capacity and resilience: Perspectives for climate
and development policy, Journal of International Development, 22, 218-232,
doi:10.1002/jid.1675.
Gallopin, G.C., 2006: Linkages between vulnerability, resilience, and adaptive
capacity. Global Environmental Change,
16(3), 293-303.
Gallopin, G.C., S. Funtowicz, M. O’Connor, and J. Ravetz, 2001: Science for the twenty-
first century: from social contract to the scientific core. International Social
Science Journal, 53(168), 219-229
Garret, M.J., 1999: Health Futures: A Handbook for Health Professionals. World
Health Organization, Geneva. Switzerland, ISBN 92-4-154521-6.
Ginexi, E.M., K. Weihs, S.J. Simmens, and D. R. Hoyt, 2000: Natural disaster and
depression: a prospective investigation of reactions to the 1993 midwest
floods. American Journal of Community Psychology, 28, 495-518.
Glantz, M., 2003: Early warning systems: Dos and don’ts. Report of Workshop on
Early Warning Systems, Shanghai, China, 20-24 October 2003, ISBN 978-
0756744953.
Goddard, L., Y. Aitchellouche, W. Baethgen, M. Dettinger, R. Graham, P. Hayman, M.
Kadi, R. Martínez, and H. Meinke, 2009: Providing Seasonal-to-Interannual
Climate Information for Risk Management and Decision Making. White Paper
presented at the World Climate Conference 3 in Report of the World Climate
Conference 3, Report No. 1048, World Meteorological Organization, Geneva,
Switzerland.
Gosling, S., G. McGregor, and J. Lowe, 2009: Climate change and heat-related
mortality in six cities – part 2: climate model evaluation and projected impacts
from changes in the mean and variability of temperature with climate change.
International Journal of Biometeorology, 53(1), 31-51.
Greiving, S., 2002: Räumliche Planung und Risiko. Gerling Akademischer Verlag,
München, Germany, 320 pp.
Grossi, P. and H. Kunreuther (eds.), 2005: Catastrophe Modeling: A new approach to
managing risk. Springer, New York, NY, 245 pp.
Grothmann, T. and A. Patt, 2005: Adaptive capacity and human cognition: the
process of individual adaptation to climate change. Global Environmental
Change, 15, 199-213
Guha-Sapir, D., D. Hargitt, and P. Hoyois, 2004: Thirty Years Of Natural Disasters
1974-2003: The Numbers. Centre for Research on the Epidemiology of
Disasters, Presses Universitaires de Louvain, Brussels, Belgium, ISBN
2930344717.
Gunderson, L.H., 2000: Resilience in theory and practice. Annual Review of Ecology
and Systematics, 31, 425-439.
Guth, J.L., J.C. Green, L.A. Kellstedt, and C.E. Smith, 1995: Faith and the environment:
Religious beliefs and attitudes on environmental policy. American Journal of
Political Science, 39(2), 364-382.
Haddad, B.M., 2005: Ranking the adaptive capacity of nations to climate change when
socio-political goals are explicit. Global Environmental Change, 15, 165-176.
Haines, A., R.S. Kovats, D. Campbell-Lendrum, and C. Corvalan, 2006: Climate change
and human health: Impacts, vulnerability and public health. Public Health, 120,
585-596
Hales, S., S. Edwards, and R.S. Kovats, 2003: Impacts on health of climate extremes.
In: Climate Change and Human Health: Risks and Responses [McMichael, A.J.,
D. Campbell-Lendrum, K. Ebi, and C. Corvalan (eds.)]. World Health
Organisation, Geneva, Switzerland, pp. 79-96.
Handmer, J.W., S. Dovers, and T.E. Downing, 1999: Societal vulnerability to climate
change and variability. Mitigation and Adaptation Strategies for Global
Change, 4, 267-281.
Haque, C.E. and D. Etkin, 2007: People and community as constituent parts of
hazards: the significance of societal dimensions in hazards analysis. Natural
Hazards, 41(2), 271-282.
Harding, T.W., F. Romerio, J. Rossiaud, J.J. Wagner, S. Bertrand, C. Frischknecht, and
J.D. Laporte, 2001 : Management des risques majeurs: des disciplines à
l’interdisciplinarité. Document de travail No 1 du Groupe de recherche
Management des risques majeurs, Université de Genève, Programme
plurifacultaire du Rectorat, Geneva, Switzerland.
Hardoy, J. and G. Pandiella, 2009: Urban poverty and vulnerability to climate change
in Latin America. Environment & Urbanization, 21(1), 203-224.
Hellmuth, M.E., S.J. Mason, C. Vaughan, M.K. van Aalst, and R. Choularton (eds.),
2011: A Better Climate for Disaster Risk Management. International Research
Institute for Climate and Society, Columbia University, New York, NY.
Hewitt, K., 1997: Regions of Risk: A Geographical Introduction to Disasters. Addison
Wesley Longman, Harlow, UK, 408 pp.
Hewitt, K. (ed.), 1983: Interpretations of Calamity, Allen & Unwin, London, UK.
Hewitt, K. and I. Burton, 1971: The Hazardousness of a Place; a Regional Ecology of
Damaging Events. University of Toronto Press, Toronto, Canada.
Hilhorst, D. 2004: Complexity and diversity: Unlocking social domains of disaster
response. In: Mapping Vulnerability: Disa sters, Development and People
[Bankoff, G., G. Frerks, and D. Hilhorst (eds.)]. Earthscan Publishers, London, UK,
pp. 52-66.
Holling, C.S., 1973: Resilience and stability of ecological systems. Annual Reviews of
Ecology and Systematics, 4, 1-23.
Holling, C.S., 2001: Understanding the complexity of economic, ecological, and
social systems. Ecosystems, 4, 390-405.
Hong, Y.C., X.C. Pan , S.Y. Kim, K. Park, E.J. Park, X. Jin, S.M. Yi, K.S. Kim, Y.H. Kim, C.H.
Park, S. Song, and H. Kim, 2009: Asian dust storms and pulmonary function of
school children. Epidemiology, 20(6), S121-S121.
Hope, K.R. Sr., 2009: Climate change and poverty in Africa. International Journal of
Sustainable Development and World Ecology, 16(6), 451-461.
Houghton, R., 2009: Domestic violence reporting and disasters in New Zealand.
Regional Development Dialogue, 30(1), 79-90.
Hulme, M., 2009: Why We Disagree About Climate Change. Cambridge University
Press, Cambridge, UK.
Chapter 2 Determinants of Risk: Exposure and Vulnerability
102
Hutton, D. and C.E. Haque, 2003: Patterns of coping and adaptation among erosion-
induced displacees in Bangladesh: Implications for hazard analysis and mitigation.
Natural Hazards, 29, 405-421.
IAASTD, 2009: Agriculture at a Crossroads: International Assessment of Agricultural
Knowledge, Science and Technology for Development. International Assessment
of Agricultural Knowledge, Science and Technology for Development, Island
Press, Washington, DC.
ICSU-LAC, 2010a: Science for a better life: Developing regional scientific programs
in priority areas for Latin America and the Caribbean. Vol 2, Understanding and
Managing Risk Associated with Natural Hazards: An Integrated Scientific
Approach in Latin America and the Caribbean [Cardona, O.D., J.C. Bertoni, A.
Gibbs, M. Hermelin, and A. Lavell (eds.)]. ICSU Regional Office for Latin America
and the Caribbean, Rio de Janeiro, Brazil.
ICSU-LAC, 2010b: Entendimiento y gestión del riesgo asociado a las amenazas
naturales: Un enfoque científico integral para América Latina y el Caribe.
Ciencia para una vida mejor: Desarrollando programas científicos regionales en
áreas prioritarias para América Latina y el Caribe. Vol 2 [Cardona, O.D., J.C.
Bertoni, A. Gibbs, M. Hermelin, and A. Lavell (eds.)]. ICSU Regional Office for
Latin America and the Caribbean, Rio de Janeiro, Brazil.
IDB, 2007: Disaster Risk Management Policy. GN-2354-5, Inter-American
Development Bank, Washington, DC.
IDEA, 2005: Indicators of Disaster Risk and Risk Management – Main Technical
Report. English and Spanish edition, National University of Colombia/
Manizales, Institute of Environmental Studies/IDEA, Inter-American Development
Bank, Washington, DC, 223 pp.
IFRC, 2008: Vulnerability and Capacity Assessment – Guidelines. International
Federation of Red Cross and Red Crescent Societies, Geneva, Switzerland.
IFRC, 2009: World Disasters Report 2009 – Focus on Early Warning, Early Action.
International Federation of Red Cross and Red Crescent Societies, Geneva,
Switzerland.
IISD, Intercooperation, IUCN, and SEI, 2009: CRiSTAL: Community-based Risk Screening
Tool – Adaptation & Livelihoods, User’s Manual, v 4.0. IISD, Geneva, Switzerland.
INGC, 2009. Synthesis report. INGC Climate Change Report: Study on the impact of
climate change on disaster risk in Mozambique [van Logchem, B. and R. Brito R
(eds.)]. National Institute for Disaster Management, Mozambique.
Ingham, B., 1993: The meaning of development: interactions between ‘new’ and
‘old’ ideas. World Development, 21, 1803-1821.
IOM, 2009: Migration, Environment and Climate Change: Assessing the Evidence.
International Organization for Migration, Geneva, Switzerland, 441 pp.
IPCC, 2001: IPCC Third Assessment Report. Synthesis Report, Cambridge University
Press, Cambridge, UK.
IPCC, 2007: Climate Change 2007. Impacts, Adaptation and Vulnerability.
Contribution of Working Group II to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P.
Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 7-22.
ISO, 2009a: Risk Management – Principles and guidelines. ISO 31000, International
Organization for Standardization, Geneva, Switzerland.
ISO, 2009b: Risk Management – Vocabulary. ISO Guide 73, International Organization
for Standardization, Geneva, Switzerland.
Jabry, A., 2003: Children in Disasters: After the Cameras Have Gone. Plan UK, Plan
International, Woking, UK, 54 pp.
Jakobsen, K.T., 2009, Views on vulnerability following hurricane Mitch in Nicaragua,
IOP Conference Series: Earth and Environmental Science, 6, pp. 1-2.
Janssen, M.A., M.L. Schoon, W. Ke, and K. Börner, 2006: Scholarly networks on
resilience, vulnerability and adaptation within the human dimensions of global
environmental change. Global Environmental Change, 16(3), 240-252.
Jasanoff, S. and B. Wynne, 1997: Handbook of Science and Technology Studies. Sage
Publications, London, UK, 828 pp., ISBN 0803940211.
Johnson, P.T.J., A.R. Townsend, C.C. Cleveland, P.M. Glibert, R.W. Howarth, V.J.
McKenzie, E. Rejmankova, and M.H. Ward, 2010: Linking environmental nutrient
enrichment and disease emergence in humans and wildlife. Ecological
Applications, 20(1), 16-29.
Kantor, P. and P. Nair, 2005: Vulnerability among slum dwellers in Lucknow, India –
Implications for urban livelihood security. International Development Planning
Review, 27(3), 333-358.
Kar, N. 2009: Psychological impact of disasters on children: review of assessment and
interventions. World Journal of Pediatrics, 5(1), 5-11.
Karlsson, S., 2007: Allocating responsibilities in multi-level governance for
sustainable development. Journal of Social Economics, 34(1-2), 103-126.
Kasperson, R.E. and J.X. Kasperson, 2001: Climate Change, Vulnerability, and Social
Justice. Risk and Vulnerability Programme, Stockholm Environment Institute,
Stockholm, Sweden, pp. 18.
Kasperson, R. and I. Palmund, 2005: Evaluating risk communication. In: The Social
Contours of Risk. Volume I: Publics, Risk Communication & the Social
Amplification of Risk [Kasperson, J. and R. Kasperson (eds.)]. Earthscan, London,
UK, pp. 51-67.
Kasperson, R.E., O. Renn, P. Slovic, H.S. Brown, J. Emel, R. Goble, J.X. Kasperson, and
S. Ratick, 1988: The social amplification of risk: a conceptual framework. Risk
Analysis, 8(2), 177-187.
Kasperson, J., R. Kasperson, B.L. Turner, W. Hsieh, and A. Schiller, 2005: Vulnerability
to global environmental change. In: The Social Contours of Risk. Volume II: Risk
Analysis, Corporations & the Globalization of Risk [Kasperson, J. and R.
Kasperson (eds.)]. Earthscan, London, UK, pp. 245-285.
Kates, R.W., 1962: Hazard and Choice Perception in Flood Plain Management. Dept.
of Geography Research Paper No. 78, University of Chicago, Chicago, IL.
Kates, R.W., 1971: Natural hazard in human ecological perspective: Hypotheses and
models. Economic Geography, 47(3), pp. 438-451.
Kates, R.W., 1985: The interaction of climate and society. In: Climate Impact
Assessment [Kates, R.W., J.H. Ausubel, and M. Berberian (eds.)]. SCOPE 27,
Wiley, Chichester, UK.
Keim, M.E., 2008: Building human resilience: the role of public health preparedness
and response as an adaptation to climate change. American Journal of
Preventive Medicine, 35, 508-516.
Kelman, I. and T.A. Mather, 2008: Living with volcanoes: the sustainable livelihoods
approach for volcano-related opportunities. Journal of Volcanology and
Geothermal Research, 172(3-4), 189-198.
Kesavan, P.C. and M.S. Swaminathan, 2006: Managing extreme natural disasters in
coastal areas. Philosophical Transactions of The Royal Society A, 364 (1845),
2191-2216.
Khandlhela, M. and J. May, 2006: Poverty, vulnerability and the impact of flooding
in the Limpopo Province, South Africa. Natural Hazards, 39(2), 275-287.
Kintisch, E., 2006: Evangelicals, scientists reach common ground on climate change.
Science, 311(5764), 1082-1083.
Klein, R.J.T., 2001: Adaptation to Climate Change in German Official Development
Assistance: An Inventory of Activities and Opportunities, with a Special Focus on
Africa. GTZ, Eschborn, Germany.
Klein, R.J.T., M. Alam, I. Burton, W.W. Dougherty, K.L. Ebi, M. Fernandes, A. Huber-Lee,
A.A. Rahman, and C. Swartz, 2006: Application of Environmentally Sound
Technologies For Adaptation To Climate Change. Technical Paper
FCCC/TP/2006/2, United Nations Framework Convention on Climate Change
Secretariat, Bonn, Germany.
Klein, R., S. Eriksen, L.O. Naess, A. Hammill, C. Robledo, T.M. Tanner, and K. O’Brien,
2007: Portfolio screening to support the mainstreaming of adaptation to
climate change into development assistance. Climatic Change, 84, 23-44.
Kobori, H., 2009: Current trends in conservation education in Japan. Biological
Conservation, 142(9), 1950-1957.
Koenig, D., 2009: Urban relocation and resettlement: Distinctive problems, distinctive
opportunities. In: Development and Dispossession: The Crisis of Development
Forced Displacement and Resettlement [Oliver-Smith, A. (ed.)]. SAR Press, Santa
Fe, NM, pp. 119-139.
Krasny, M.E. and K.G. Tidball, 2009: Applying a resilience systems framework to urban
environmental education. Environmental Education Research, 15(4), 465-482.
Kuhar, C.W., T.L. Bettinger, K. Lehnhardt, O. Tracy, and D. Cox, 2010: Evaluating for
long-term impact of an environmental education program at the Kalinzu Forest
Reserve, Uganda. American Journal Of Primatology, 72(5), 407-413.
Chapter 2Determinants of Risk: Exposure and Vulnerability
103
Kuwali, D., 2008, From the West to the rest: Climate change as a challenge to human
security in Africa. African Security Review, 17 (3), 18-38.
Lamond, J.E., D.G. Proverbs, and F.N. Hammond, 2009: Accessibility of flood risk
insurance in the UK: confusion, competition and complacency. Journal of Risk
Research, 12(6), 825-841.
Landry, C.E., O. Bin, P. Hindsley, J.C. Whitehead, and K. Wilson, 2007: Going Home:
Evacuation-Migration decisions of Hurricane Katrina Survivors. Center for
Natural Hazards Research Working Paper, Natural Hazards Center, University of
Colorado at Boulder, Boulder, CO.
Lavell, A. (ed.), 1994: Viviendo en riesgo: comunidades vulnerables y prevención de
desastres en América Latina. LA RED, Tercer Mundo Editores, Bogotá, Colombia.
Lavell, A., 1996: Degradación ambiental, riesgo y desastre urbano. Problemas y
conceptos: hacia la definición de una agenda de investigación. In: Ciudades en
Riesgo [Fernandez, M.A. (ed.)]. La RED-USAID, Lima, Peru, pp. 21-59.
Lavell, A., 1999a: Environmental degradation, risks and urban disasters. issues and
concepts: Towards the definition of a research agenda. In: Cities at Risk:
Environmental Degradation, Urban Risks and Disasters in Latin America
[Fernandez, M.A. (ed.)]. A/H Editorial, La RED, US AID, Quito, Ecuador, pp. 19-58.
Lavell, A., 1999b: Natural and Technological Disasters: Capacity Building and Human
Resource Development for Disaster Management. Concept Paper commissioned
by Emergency Response Division, United Nations Development Program,
Geneva Switzerland.
Lavell, A., 2003: Local Level Risk Management: Concept and Practices. CEPRE-
DENAC-UNDP, Quito, Ecuador.
Lavell A. and E. Franco (eds.), 1996: Estado, Sociedad y Gestión de los Desastres en
América Latina. Red de Estudios Sociales en Prevención de Desastres en
América Latina, La RED, Tercer Mundo Editores, Bogotá, Colombia.
Leichenko, R.M. and K.L. O’Brien, 2002: The dynamics of rural vulnerability to global
change: the case of southern Africa. Mitigation and Adaptation Strategies for
Global Change, 7, 1-18.
Leichenko, R.M. and K.L. O’Brien, 2008: Environmental Change and Globalization,
Double Exposures. Oxford University Press, New York, NY.
Leiserowitz, A., 2006: Climate change risk perception and policy preferences: the
role of affect, imagery, and values. Climatic Change, 77, 45-72
Levina, E. and D. Tirpak, 2006: Adaptation to Climate Change: Key Terms.
COM/ENV/EPOC/IEA/SLT(2006)1, OECD, Paris, France.
Linnerooth-Bayer, J. and R. Mechler, 2007: Disaster safety nets for developing
countries: Extending public-private partnerships. Environmental Hazards, 7, 54-61.
Linnerooth-Bayer, J., R, Mechler, and G. Pflug, 2005: Refocusing disaster aid.
Science, 309, 1044-1046.
Liverman, D.M., 1990: Vulnerability to global environmental change. In:
Understanding Global Environmental Change: The Contributions of Risk
Analysis and Management [Kasperson, R.E., K. Dow, D. Golding, and J.X.
Kasperson (eds.)]. Clark University, Worcester, MA, pp. 27-44.
Longstaff, P.H. and S.-U. Yang, 2008: Communication management and trust: Their
role in building resilience to “surprises” such as natural disasters, pandemic flu,
and terrorism. Ecology and Society, 13 (1), 3.
Lopez-Calva, L.F. and E. Ortiz, 2008: Evidence and Policy Lessons on the Link
between Disaster Risk and Poverty in Latin America: Summary of Regional
Studies. RPP LAC – MDGs and Poverty – 10/2008, RBLAC-UNDP, New York, NY,
USA.
Luers, A.L., D.B. Lobell, L.S. Sklar, C.L. Addams, and P.A. Matson, 2003: A method for
quantifying vulnerability, applied to the Yaqui Valley, Mexico. Global
Environmental Change, 13, 255-267.
Luhmann, N., 2003: Soziologie des Risikos. Walter De Gruyterl, Berlin, Germany.
Ljung, K., F. Maley, A. Cook, and P. Weinstein, 2009: Acid sulfate soils and human
health – A Millennium Ecosystem Assessment. Environment International,
35(8), 1234-1242.
Madagascar Development Intervention Fund, 2007: Disaster-Resistant
Schools, a Tool for Universal Primary Education. UNISDR, Geneva, Switzerland.
Magnan, A., 2010: For a better understanding of adaptive capacity to climate
change: a research framework. IDDRI Analysis, 2/10, IDDRI, Paris, France.
Mansilla, E. (ed.), 1996: Desastres: modelo para armar. La RED, Lima, Peru.
Manyena, S.B. 2006: The concept of resilience revisited. Disasters, 30(4), 433-450.
Marmot, M., 2010: Marmot Review final report, Fair Society, Healthy Lives.
University College London, London, UK.
Marris, C., I. Langford, T. Saunderson, and T. O’Riordan, 1998, A quantitative test of
the cultural theory of risk perceptions: Comparison with the psychometric
paradigm. Risk Analysis, 18(5), 635-647.
Marshall B.K. and J.S. Picou, 2008: Postnormal science, precautionary principle, and
worst cases: The challenge of twenty-first century catastrophes. Sociological
Enquiry, 78(2), 230-247.
Marulanda, M.C., O.D. Cardona, M.G. Ordaz, and A.H. Barbat, 2008a: La gestión
financiera del riesgo desde la perspectiva de los desastres: Evaluación de la
exposición fiscal de los Estados y alternativas de instrumentos financieros de
retención y transferencia del riesgo. Monografía CIMNE IS-61, Universidad
Politécnica de Cataluña, Barcelona, Spain.
Marulanda, M.C., O.D. Cardona, and A.H. Barbat, 2008b: The economic and social
effects of small disasters: Revision of the Local Disaster Index and the case
study of Colombia. In: Megacities: Resilience and Social Vulnerability [Bohle,
H.G. and K. Warner (eds.)]. SOURCE No. 10, United Nations University (EHS),
Munich Re Foundation, Bonn, Germany.
Marulanda, M.C., O.D. Cardona, and A.H. Barbat, 2010: Revealing the socioeconomic
impact of small disasters in Colombia using the DesInventar database.
Disasters, 34(2), 552-570.
Marulanda, M.C., O.D. Cardona, and A.H. Barbat, 2011: Revealing the impact of
small disasters to the economic and social development. In: Coping with Global
Environmental Change, Disasters and Security - Threats, Challenges,
Vulnerabilities and Risks [Brauch, H.G., U. Oswald Spring, C. Mesjasz, J. Grin, P.
Kameri-Mbote, B. Chourou, P. Dunay, and J. Birkmann (eds.)]. Springer-Verlag,
Berlin, Germany.
Masika, R., 2002: Editorial. Gender & Development, 10(2), 2-9.
Maskrey, A., 1989: Disaster Mitigation: A Community Based Approach. Oxfam,
Oxford, UK.
Maskrey, A. (Comp.), 1993a: Los Desastres No son Naturales. Red de Estudios
Sociales en Prevención de Desastres en América Latina, LA RED, Tercer Mundo
Editores, La RED, Bogotá, Colombia.
Maskrey, A., 1993b: Vulnerability accumulation in peripheral regions in Latin
America: The challenge for disaster prevention and management. In: Natural
Disasters: Protecting Vulnerable Communities [Merriman, P.A. and C.W. Browitt
(eds.)]. IDNDR, Telford, London, UK, pp. 461-472.
Maskrey, A., 1994: Disaster mitigation as a crisis paradigm: Reconstructing after the
Alto Mayo Earthquake, Peru. In: Disaster, Development and Environment
[Varley, A. (ed.)]. John Wiley & Sons, Chichester, UK, pp. 109-123.
Maskrey, A. (ed.), 1998: Navegando entre Brumas: La Aplicación de los Sistemas de
Información Geográfica al Análisis de Riesgo en América Latina. LA RED, ITDG,
Lima, Peru.
Massey, K. and J. Sutton, 2007: Faith community’s role in responding to disasters.
Southern Medical Journal, 100(9), 944-945, doi:10.1097/
SMJ.0b013e318145a847.
McCarthy, J.J., O.F. Canziani, N.A. Leary, D.J. Dokken, and K.S. White (eds.), 2001:
Climate Change 2001: Impacts, Adaptation, and Vulnerability. Working Group II
of the Intergovernmental Panel on Climate Change, Cambridge University
Press, Cambridge, UK.
McConnell, W.J., 2008: Comanagement of natural resources: Local learning for
poverty reduction. Society & Natural Resources, 21(3), 273-275.
McGranahan, G., D. Balk, and B. Anderson, 2007: The rising tide: assessing the risks
of climate change and human settlements in low elevation coastal zones.
Environment and Urbanization, 19(1), 17-37, doi:10.1177/0956247807076960.
McGuire, L.C., E.S. Ford, and C.A. Okoro, 2007: Natural disasters and older US adults
with disabilities: implications for evacuation. Disasters, 31 (1), 49-56.
McIvor, D. and D. Paton, 2007: Preparing for natural hazards: normative and
attitudinal influences. Disaster Prevention and Management, 16(1), 79-88.
McLaughlin, P., and T. Dietz, 2008: Structure, agency and environment: toward an
integrated perspective on vulnerability. Global Environmental Change, 18, 99-
111.
Chapter 2 Determinants of Risk: Exposure and Vulnerability
104
McLeman, R. and B. Smit, 2006: Vulnerability to climate change hazards and risks:
crop and flood insurance. Canadian Geographer, 50(2), 217-226.
McMichael, D.H., C.F. Campbell-Lendrum, K. Corvalan, L. Ebi, A. Githeki,
J.D.Scheraga, and A. Woodward, 2003: Climate Change and Human Health:
Risks And Responses. WMO, Geneva, Switzerland.
MEA, 2005: Millennium Ecosystem Assessment, Ecosystems and Human Well-being:
Synthesis. Island Press, Washington, DC.
Mearns, R. and A. Norton, 2010: The Social Dimensions of Climate Change: Equity
and Vulnerability in a Warming World. New Frontiers of Social Policy Series,
World Bank, Washington, DC, ISBN 978-0-8213-7887-8.
Mechler, R., 2004: Natural Disaster Risk Management and Financing Disaster Losses
in Developing Countries. Verlag für Versicherungswirtschaft, Karlsruhe, Germany.
Mechler, R., J. Linnerooth-Bayer, S. Hochrainer, and G. Pflug, 2006: Assessing financial
vulnerability and coping capacity: The IIASA CATSIM Model. In: Measuring
Vulnerability and Coping Capacity to Hazards of Natural Origin. Concepts and
Methods [Birkmann, J. (ed.)]. United Nations University Press, Tokyo, Japan.
Mechler, R., S. Hochrainer, A. Aaheim, H. Salen and A. Wreford, 2010: Modelling
economic impacts and adaptation to extreme events: Insights from European
case studies. Mitigation and Adaptation Strategies for Global Change, 15(7),
737-762.
Mercer J., I. Kelman, L. Taranis, and S. Suchet-Pearson, 2010: Framework for integrating
indigenous and scientific knowledge for disaster risk reduction. Disasters,
34(1), 214-239.
Metzger, M.J., M.D.A. Rounsevell, L. Acosta-Michlik, R. Leemans, and D. Schröter,
2006: The vulnerability of ecosystem services to land use change. Agriculture,
Ecosystems and Environment, 114, 69-85.
Mileti, D.S., 1996: Psicología social de las alertas públicas efectivas de desastres. In:
Desastres y Sociedad, Especial: Predicciones, Pronósticos, Alertas y Respuestas
Sociales [Cardona, O.D. (ed.)], 4, La RED, Tarea Gráfica, Lima, Peru.
Mimura, N.L., L. Nurse, R.F. McLean, J. Agard, L. Briguglio, P. Lefale, R. Payet, and G.
Sem, 2007: Small islands. In: Climate Change 2007. Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani,
J.P. Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 687-716.
Mitchell, J.K. (ed.), 1999a: Crucibles of Hazards: Megacities and Disasters in
Transition. United Nations University Press, Tokyo, Japan.
Mitchell, J.K., 1999b: Megacities and natural disasters: A comparative analysis.
GeoJournal, 49(2), 137-142.
Mitchell, T., and M.K. van Aalst, 2008: Convergence of Disaster Risk Reduction and
Climate Change Adaptation: A Review for DFID. Department for International
Development (DFID), London, UK.
Mitchell, T., M. van Aalst, and P.S. Villanueva, 2010: Assessing Progress on Integrating
Disaster Risk Reduction and Climate Change Adaptation in Development
Processes. Strengthening Climate Resilience Discussion Papers, 2, Strengthening
Climate Resilience, Institute of Development Studies, Brighton, UK.
Moench, M. and A. Dixit (eds.), 2007: Working with the Winds of Change: Towards
Strategies for Responding to the Risks Associated with Climate Change and
Other Hazards. ISET, Kathmandu, Nepal, 285 pp.
Mohay, H. and N. Forbes, 2009: Reducing the risk of posttraumatic stress disorder in
children following natural disasters. Australian Journal of Guidance and
Counselling, 19(2), 179-195.
Morrow, B.H., 1999: Identifying and mapping community vulnerability. Disasters,
23(1), 1-18.
Morton, T.A., A. Rabinovich, D. Marshall, and P. Bretschneider, 2011: The future that
may (or may not) come: How framing changes responses to uncertainty in
climate change communications. Global Environmental Change, 21, 103-109.
Moss, R.H., A.L. Brenkert, and E.L. Malone, 2001: Vulnerability to Climate Change: A
Quantitative Approach. Technical Report PNNL-SA-33642, Pacific Northwest
National Laboratories, Richland, WA.
MOVE, 2010: Generic Conceptual Framework for Vulnerability Measurement. Seven
Framework Programme, Methods for the Improvement of Vulnerability
Assessment in Europe, European Commission, Brussels, Belgium.
Munda, G., 2000: Multicriteria Methods and Process for Integrated Environmental
Assessment. In: Métodos Numéricos en Ciencias Sociales (MENCIS 2000)
[Oñate, E., F. García-Sicilia, and L. Ramallo (eds.)]. Centro Internacional de
Métodos Numéricos en Ingeniería - Universidad Politécnica de Cataluña,
Barcelona, Spain, pp. 364-375, ISBN 84-89925-71-2.
Munich Re, 2011: TOPICS GEO, Natural Catastrophes 2010, Analyses, Assessments,
Positions. Munich Re, Munich, Germany.
Myers, C.A., T. Slack, and J. Singelmann, 2008: Social vulnerability and migration in
the wake of disaster: the case of Hurricanes Katrina and Rita. Population &
Environment, 29(6), 271-291.
Myers, N., 1993: Environmental refugees in a globally warmed world. Bioscience,
43(11), 752-761.
Næss, L., O. Bang, S. Eriksen, and J. Vevatne, 2005, Institutional adaptation to climate
change: Flood responses at the municipal level in Norway. Global
Environmental Change, 15, 125-138, doi:10.1016/j.gloenvcha.2004.10.003.
Nakagawa, Y. and R. Shaw, 2004: Social capital: A missing link to disaster recovery.
International Journal of Mass Emergencies and Disasters, 22, 5-34.
Nathan, F., 2008: Risk perception, risk management and vulnerability to landslides
in the hill slopes in the city of La Paz, Bolivia. Disasters, 32 (3), 337-357.
Neal, D.M. and B.D. Phillips, 1990: Female-dominated local movement organizations
in disaster-threat situations. In: Women and Social Protest [West, G. and R.L.
Blumberg (eds.)]. Oxford University Press, New York, NY, pp. 243-255.
Nelson, D.R. and T.J. Finan, 2009: Praying for drought: Persistent vulnerability and the
politics of patronage in Ceara, Northeast Brazil. American Anthropologist, 111,
302-316, doi:10.1111/j.1548-1433.2009.01134.x.
Nelson, R., P. Kokic, S. Crimp, H. Meinke, and S.M. Howden, 2010a: The vulnerability of
Australian rural communities to climate variability and change, Pt. I: Conceptualising
and measuring vulnerability. Environmental Science & Policy,
13, 8-17.
Nelson, R., P. Kokic, S. Crimp, P. Martin, H. Meinke, S.M. Howden, P. de Voil, and U.
Nidumolu, 2010b: The vulnerability of Australian rural communities to climate
variability and change, Part II: Integrating impacts with adaptive capacity.
Environmental Science & Policy, 13, 18-27.
Neumayer, E. and T. Pluemper, 2007: The gendered nature of natural disasters: The
impact of catastrophic events on the gender gap in life expectancy, 1981–2002.
Annals of the Association of American Geographers, 97(3), 551-566
Nicholls, R.J., 2004: Coastal flooding and wetland loss in the 21st century: changes
under the SRES climate and socio-economic scenarios. Global Environmental
Change, 14(1), 69-86.
Nicholls, R.J., N. Marinova, J. Lowe, S. Brown, P. Vellinga, D. De Gusmao, J. Hinkel, and
R.S. Tol, 2011: Sea-level rise and its possible impacts given a ‘beyond 4 degrees
C world’ in the twenty-first century. Philosophical Transactions of The Royal
Society A, 369(1934), 161-181.
Nielsen. J.O. and A. Reenberg, 2010: Cultural barriers to climate change adaptation:
A case study from Northern Burkina Faso. Global Environmental Change, 20,
142-152.
Nomura, K., 2009: A perspective on education for sustainable development:
Historical development of environmental education in Indonesia. International
Journal of Educational Development, 29(6), 621-627.
Nurse, L. and G. Sem, 2001: Small island states. In: Climate Change 2001: Impacts,
Adaptation & Vulnerability [McCarthy, J., O. Canziani, N. Leary, D. Dokken, and
K. White (eds.)]. Cambridge University Press, Cambridge, UK, pp. 842-875.
O’Brien, G., P. O’Keefe, H. Meena, J. Rose, and L. Wilson, 2008: Climate adaptation
from a poverty perspective. Climate Policy, 8(2), 194-201.
O’Brien, K., S. Eriksen, A. Schjolen, and L. Nygaard, 2004a: What’s in a word?
Conflicting interpretations of vulnerability in climate change research.
CICEROWorking Paper 2004:04, CICERO, Oslo University, Oslo, Norway.
O’Brien, K, R. Leichenko, U. Kelkar, H. Venema, G. Aandahl, H. Tompkins, A. Javed, S.
Bhadwal, S. Barg, L. Nygaard, and J. West, 2004b: Mapping vulnerability to multiple
stressors: climate change and globalization in India. Global Environmental
Change, 14, 303-313.
O’Brien, K., S. Eriksen, L. Sygna, and L.O. Naess, 2006: Questioning complacency:
Climate change impacts, vulnerability, and adaptation in Norway. Ambio,
35(2), 50-56.
Chapter 2Determinants of Risk: Exposure and Vulnerability
105
O’Brien, K., L. Sygna, R. Leinchenko, W.N. Adger, J. Barnett, T. Mitchell, L. Schipper, T.
Tanner, C. Vogel, and C. Mortreux, 2008: Disaster Risk Reduction, Climate
Change Adaptation and Human Security. GECHS Report 2008:3, Global
Environmental Change and Human Security, Oslo, Norway.
O’Keefe, P., K. Westgate, and B. Wisner, 1976: Taking the naturalness out of natural
disasters. Nature, 260, 566-567.
OAS, 1991: Primer on Natural Hazard Management in Integrated Regional
Development Planning. Organization of American States, Washington, DC.
OCHA/IDMC, 2009: Monitoring Disaster Displacement in the Context of Climate
Change. Findings of a study by the United Nations Office for the Coordination
of Humanitarian Affairs and the Internal Displacement Monitoring Centre,
Geneva, Switzerland.
Ochola, S.O., B. Eitel, and D.O. Olago, 2010: Vulnerability of schools to floods in
Nyando River catchment, Kenya. Disasters, 34(3), 732-754.
Olausson, U., 2009: Global warming – global responsibility? Media frames of
collective action and scientific certainty. Public Understanding of Science, 18,
421-436
Oliver-Smith, A., 2006: Communities after catastrophe: Reconstructing the material,
reconstituting the social. In: Community Building in the 21st Century [Hyland,
S. (ed.)]. School of American Research Press, Santa Fe, NM, pp. 45-70.
Oliver-Smith, A., 2009: Disasters and diasporas: Global climate change and
population displacement in the 21st century. In: Anthropology and Climate
Change: From Encounters to Actions [Crate, S.A. and M. Nuttall (eds.)]. Left
Coast Press, Walnut Creek, CA, pp. 116-138.
Oppenheimer, M. and A. Todorov (eds.), 2006: Global warming: the psychology of
long term risk. Climatic Change, 77(1-6).
Orr, M., 2003: Environmental decline and the rise of religion. Zygon, 38(4), 895-910.
Paavola, J., 2008: Livelihoods, vulnerability and adaptation to climate change in
Morogoro, Tanzania. Environmental Science & Policy, 11, 642-654.
PAHO/World Bank, 2004: Guidelines for Vulnerability Reduction in the Design of
New Health Facilities. Prepared by Rubén Boroschek Krauskopf and Rodrigo
Retamales Saavedra, PAHO/World Bank in collaboration with the World Health
Organization, Washington, DC, 106 pp.
Pantuliano, S. and M. Wekesa, 2008: Improving drought response in pastoral areas
of Ethiopia: Somali and Afar Regions and Borena Zone of Oromiya Region.
Prepared for the CORE group (CARE, FAO, Save the Children UK and Save the
Children US), Overseas Development Institute, Humanitarian Policy Group,
London, UK.
Parker, D.J., 1995: Floodplain development policy in England and Wales. Applied
Geography, 15(4), 341-363.
Patt, A., R. Klein, and A. Vega-Leinert, 2005: Taking the uncertainty in climate-change
vulnerability seriously. Geoscience,
337, 411-424.
Patt, A.G., D. Schröter, R.L.T. Klein, and A.C. Vega-Leinert, 2009: Assessing
Vulnerability to Global Environmental Change. Earthscan Publications, London,
UK.
Patterson, J., E. Linden, J.K.P. Edward, D. Wilhelmsson, and I. Lofgren, 2009:
Community-based environmental education in the fishing villages of Tuticorin
and its role in conservation of the environment. Australian Journal of Adult
Learning, 49(2), 382-393.
Peacock, W.G., B.H. Morrow, and H. Gladwin (eds.), 1997: Hurricane Andrew:
Ethnicity, Gender and the Sociology of Disasters. Routledge, London, UK.
Peduzzi, P., 2006: The Disaster Risk Index: Overview of a quantitative approach. In:
Measuring Vulnerability to Natural Hazards - Towards Disaster Resilient
Societies [Birkmann, J. (ed.)]. United Nations University Press, Tokyo, Japan,
pp.171-181.
Peduzzi, P., H. Dao, C. Herold, and F. Mouton, 2009: Assessing global exposure and
vulnerability towards natural hazards: the Disaster Risk Index. Natural Hazards
and the Earth System Science, 9, 1149-1159.
Peek, L. and L.M. Stough, 2010: Children with disabilities in the context of disaster:
A social vulnerability perspective. Child Development, 81(4), 1260-1270.
Peguero, A.A., 2006: Latino disaster vulnerability – The dissemination of hurricane
mitigation information among Florida’s homeowners. Hispanic Journal of
Behavioral Sciences, 28 (1), 5-22.
Pelling, M., 1997: What determines vulnerability to floods: a case study in
Georgetown, Guyana. Environment and Urbanization, 9, 203-226.
Pelling, M., 1998: Participation, social capital and vulnerability to urban flooding in
Guyana. Journal of International Development, 10, 469-486.
Pelling, M., 2003: The Vulnerability of Cities: Natural Disasters and Social Resilience.
Earthscan Publications, London, UK.
Pelling, M., 2010: Adaptation to Climate Change: From Resilience to Transformation.
Routledge, London, UK.
Pelling, M. and J.I. Uitto, 2001: Small island developing states: natural disaster
vulnerability and global change. Environmental Hazards, 3, 49-62.
Pelling, M. and B. Wisner (eds.), 2009a: Disaster Risk Reduction, Cases From Urban
Africa. Earthscan, London, UK.
Pelling, M. and B. Wisner, 2009b: Introduction: Urbanization, human security and
disaster risk in Africa. In: Disaster Risk Reduction, cases from urban Africa
[Pelling, M. and B. Wisner (eds.)]. Earthscan, London, UK, pp. 1-16.
Pielke, Jr., R.A. and C.W. Landsea, 1998: Normalized hurricane damages in the
United States: 1925-1995. Weather and Forecasting, 13, 621-631.
Pierre, J. and B.G. Peters, 2000:
Governance, Politics, and the State. St. Martin’s Press,
New York, NY.
Pincha, C. and N.H. Krishna, 2009: Post-disaster death ex gratia payments and their
gendered impact. Regional Development Dialogue, 30(1), 95-105.
Plummer, R., 2006: Sharing the management of a river corridor: A case study of the
comanagement process. Society & Natural Resources, 19(8), 709-721.
Portes, A., 1998: Social capital: its origins and applications in modern sociology.
Annual Review of Sociology, 24, 1-24.
Pulwarty, R.S. and W.E. Riebsame, 1997: The political ecology of vulnerability to
hurricane-related hazards. In: Hurricanes: Climate and Socio-Economic Impacts
[Diaz, H.F. and R.S. Pulwarty (eds.)]. Springer, Heidelberg, Germany, pp. 185-214.
Putnam, R.D., 2000: Bowling Alone: The Collapse and Revival of American
Community. Simon & Schuster, New York, NY.
Quarantelli, E.L., 1998: What is a Disaster? Routledge, New York, NY.
Rafiee, R., A.S. Mahiny, and N. Khorasani, 2009: Assessment of changes in urban
green spaces of Mashad city using satellite data. International Journal of
Applied Earth Observation and Geoinformation, 11(6), 431-438.
Rahman, M.M., G. Graham Haughton, and A.E.G. Jonas, 2010: The challenges of
local environmental problems facing the urban poor in Chittagong,
Bangladesh: a scale-sensitive analysis. Environment and Urbanization, 22(2),
561-578, doi:10.1177/0956247810377560.
Ray-Bennett, N.S., 2009: The influence of caste, class and gender in surviving
multiple disasters: A case study from Orissa, India. Environmental Hazards-
Human and Policy Dimensions, 8(1), 5-22.
Reacher, M., K. McKenzie, C. Lane, T. Nichols, I. Kedge, A. Iversen, P. Hepple, T.
Walter, C. Laxton, and J. Simpson, on behalf of the Lewes Flood Action Recovery
Team, 2004: Health impacts of flooding in Lewes: a comparison of reported
gastrointestinal and other illness and mental health in flooded and non-flooded
households. Communicable Disease and Public Health, 7(1), 39-46.
Renaud, F.G., 2006: Environmental components of vulnerability. In: Measuring
Vulnerability to Natural Hazards. Towards Disaster Resilient societies
[Birkmann, J. (ed.)]. United Nations University Press, Tokyo, Japan, pp. 117-127.
Renn, O., 1992: Concepts of risk: A classification. In: Social Theories of Risk [Krimsky,
S. and D. Golding (eds.)]. Praeger, Westport, CT, pp. 53-79.
Renn, O., 2008: Risk Governance – Coping with Uncertainty in a Complex World.
Earthscan, London, UK.
Renn, O. and P. Graham, 2006: Risk Governance – Towards an integrative approach.
White paper no. 1, International Risk Governance Council, Geneva, Switzerland.
Ribot, J., 1995: The causal structure of vulnerability and its application to climate
impact analysis. GeoJournal, 35(2), 119-122.
Ribot, J.C., 1996: Introduction: Climate variability, climate change and vulnerability:
Moving forward by looking back. In: Climate Variability: Climate Change and
Social Vulnerability in the Semi-Arid Tropics [Ribot, J.C., A.R. Magalhaes, and S.S.
Panagides (eds.)]. Cambridge University Press, Cambridge, UK.
Rideout
V., 2003: Digital inequalities in eastern Canada. Canadian Journal of
Information and Library Science, 27(2), 3-31.
Chapter 2 Determinants of Risk: Exposure and Vulnerability
106
Rose, A., 2004: Economic principles, issues, and research priorities in hazard loss
estimation. In: Modeling Spatial and Economic Impacts of Disasters [Okuyama,
Y. and S.E. Chang (eds.)]. Springer, New York, NY.
Ruth, M. and D. Coelho, 2007: Understanding and managing the complexity of urban
systems under climate change. Climate Policy, 7(4), 317-336.
Saldaña-Zorrilla, S.R., 2007: Socioeconomic vulnerability to natural disasters in
Mexico: rural poor, trade and public response. CEPAL Report 92, UN-ECLAC,
Disaster Evaluation Unit, Mexico, ISBN 978-92-1-121661-5.
Sampei, Y. and M. Aoyagi-Usui, 2008: Mass-media coverage, its influence on public
awareness of climate-change issues, and implications for Japan’s national
campaign to reduce greenhouse gas emissions. Global Environmental Change,
19, 203-212.
Sánchez-Rodríguez, R., K.C. Seto, D. Simon, W.D. Solecki, F. Kraas, and G. Laumann,
2005: Science Plan Urbanization and Global Environmental Change. IHDP
Report 15, International Human Dimensions Programme on Global
Environmental Change, Bonn, Germany.
Sandstrom, C., 2009: institutional dimensions of comanagement: Participation,
power, and process. Society & Natural Resources, 22(3), 230-244.
Sanli, F.B., F.B. Balcik, and C. Goksel, 2008: Defining temporal spatial patterns of
mega city Istanbul to see the impacts of increasing population. Environmental
Monitoring and Assessment, 146(1-3), 267-275.
Schipper, E.L.F, 2010: Religion as an integral part of determining and reducing
climate change and disaster risk: an agenda for research. In: Climate Change:
The Social Science Perspective [Voss, M. (ed.)]. VS-Verlag, Wiesbaden, Germany,
pp. 377-393.
Schipper, L. and M. Pelling, 2006: Disaster risk, climate change and international
development: Scope for, and challenges to, integration. Disasters, 30(1), 19-38.
Schmuck-Widmann, H., 2000: Wissenskulturen im Vergleich. Bäuerliche und
ingenieurwissenschaftliche Wahrnehmungen und Strategien zur Bewältigung
der Flut in Bangladesh.PhD Thesis, Free University Berlin, Germany. (also published
in English as Schmuck-Widmann, H., 2001: Facing the Jamuna River. Indigenous
and engineering knowledge in Bangladesh. Bangladesh Resource Centre for
Indigenous Knowledge in Dhaka, Bangladesh, 242 pp.).
Schneider, S.H., S. Semenov, A. Patwardhan, I. Burton, C.H.D. Magadza, M.
Oppenheimer, A.B. Pittock, A. Rahman, J.B. Smith, A. Suarez, and F. Yamin, 2007:
Assessing key vulnerabilities and the risk from climate change. In: Climate
Change 2007. Impacts, Adaptation and Vulnerability. Contribution of Working
Group II to the Fourth Assessment Report of the Intergovernmental Panel on
Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der Linde, and
C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK, pp. 779-810.
Schon, D.A. and M. Rein, 1994: Frame Reflection: Toward the Resolution of
Intractable Policy Controversies. BasicBooks, New York, NY, USA, 247 pp. ISBN
0465025064.
Schröter, D., C. Polsky, and A.G. Patt, 2005: Assessing vulnerabilities to the effects
of global change: an eight step approach. Mitigation and Adaptation Strategies
for Global Change, 10, 573-595.
Sclar, E.D., P. Garau, and G. Carolini, 2005: The 21st Century health challenge of
slums and cities. Lancet, 365, 901-903.
Scoones, I., 1998: Sustainable Rural Livelihoods: A Framework For Analysis
. IDS
Working Paper 72, Brighton, UK, ISBN 1-85964-224-8.
Scudder, T., 2005: The Future of Large Dams: Dealing with Social, Environmental,
Institutional and Political Costs. Earthscan, London, UK, 408 pp.
Sen, A., 1981: Poverty and Famines: An Essay on Entitlements and Deprivation.
Clarendon Press, Oxford, UK.
Setiadi, N., 2011: Daily mobility – excursus Padang, Indonesia. In: Early Warning in
the Context of Environmental Shocks: Demographic Change, Dynamic Exposure
to Hazards, and the Role of EWS in Migration Flows and Human Displacement
[Chang Seng, D. and J. Birkmann (eds.)]. Migration and Global Environmental
Change, Foresight SR4b, Vulnerability Assessment, Risk Management and
Adaptive Planning Section, United Nations University for Environment and
Human Security, Government Office for Science, pp. 41-45.
Shah, H.C., 1995: The increasing nature of global earthquake risk. Global
Environmental Change, 5(l), 65-67.
Sharma, U. and A. Patwardhan, 2008: An empirical approach to assessing generic
adaptive capacity to tropical cyclone risk in coastal districts of India. Mitigation
and Adaptation Strategies to Global Change, 13, 819-831.
Sheldon, K. and D. Golding (eds.), 1992: Social Theories of Risk. Praeger, Westport,
CT.
Skelton, J.A., S.R. Cook, P. Auinger, J.D. Klein, and S.E. Barlow, 2009: Prevalence and
trends of severe obesity among U.S. children and adolescents. Academic
Pediatrics, 9(5), 322-329.
Slovic, P., 2000: The Perception of Risk. Earthscan, London, UK.
Smit, B. and J. Wandel, 2006: Adaptation, adaptive capacity and vulnerability. Global
Environmental Change, 16, 282-292.
Smithers, J. and B. Smit, 1997: Human adaptation to climatic variability and change.
Global Environmental Change, 7, 129-146.
Spence, P.R., K.A. Lachlan, and D.R. Griffin, 2007: Crisis communication, race, and
natural disasters. Journal of Black Studies, 37(4), 539-554.
Spence, R.J.S. and A.W. Coburn, 1987: Earthquake Protection: An Task for the 1990s.
The Structural Engineer, 65A(8), 290-296
Sphere Project, 2011: The Sphere Project, Humanitarian Charter and Minimum
Standards in Humanitarian Response. Distributed for the Sphere Project by
Practical Action Publishing, Lippincott Williams & Wilkins, Philadelphia, PA,
USA, ISBN 978-1-908176-00-4.
Stafoggia M., F. Forastiere, P. Michelozzi, and C.A. Perucci, 2006: Vulnerability to
heat-related mortality: a multicity, population-based, case-crossover analysis.
Epidemiology, 17, 315-323.
Stamatakis, E., J. Wardle, and T.J. Cole, 2010: Childhood obesity and overweight
prevalence trends in England: evidence for growing socioeconomic disparities.
International Journal of Obesity, 34(1), 41-47.
Stockdale, T.N., O. Alves, G. Boer, M. Deque, Y. Ding, K. Kumar, W. Landman, S.J.
Mason, P. Nobre, A. Scaife, O. Tomoaki, and W.-T. Yun, 2009: Understanding and
predicting seasonal to interannual climate variability – the producer perspective.
White Paper presented at the World Climate Conference 3 in Report of the
World Climate Conference 3, World Meteorological Organization Report No.
1048, Geneva, Switzerland.
Street
, R., A. Maarouf, and H. Jones-Otazo, 2005: Extreme weather and climate
events: implications for public health. In: Integration of Public Health with
Adaptation to Climate Change: Lessons Learned and New Directions [Ebi, K.L.,
J.B. Smith, and I. Burton (eds.)]. Taylor & Francis, London, UK, pp. 161-190.
Sultana, F. 2010: Living in hazardous waterscapes: Gendered vulnerabilities and
experiences of floods and disasters. Environmental Hazards-Human and Policy
Dimensions, 9(1), 43-53.
Susman, P., P. O’Keefe, and B. Wisner, 1983: Global disasters: A radical interpretation.
In: Interpretations of Calamity [Hewitt, K. (ed.)]. Allen & Unwin, Winchester, MA,
pp. 264-283.
Tanner, T., 2009: Screening climate risks to development cooperation. Focus, 2.5,
IDS, Brighton, UK.
Tearfund, 2009: Climate change and Environmental Degradation Risk and
Adaptation assessment. Tearfund, London, UK.
Thomalla, F. and R.K. Larsen, 2010: Resilience in the context of tsunami early
warning systems and community disaster preparedness in the Indian Ocean
region. Environmental Hazards, 9(3), 249-265.
Thomalla, F., T. Downing, E. Spanger-Siegfried, G. Han, and J. Rockström, 2006:
Reducing hazard vulnerability: towards a common approach between disaster
risk reduction and climate adaptation. Disasters, 30(1), 39-48.
Thomalla, F., R.K. Larsen, F. Kanji, S. Naruchaikusol, C. Tepa, B. Ravesloot, and A.K.
Ahmed, 2009: From Knowledge to Action: Learning to Go the Last Mile. A
Participatory Assessment of the Conditions for Strengthening the Technology –
community Linkages of Tsunami Early Warning Systems in the Indian Ocean,
Project Report, Stockholm Environment Institute, Macquarie University, Asian
Disaster Preparedness Centre, and Raks Thai Foundation, Bangkok and Sydney,
Australia.
Thywissen, K., 2006: Core terminology of disaster risk reduction: A comparative
glossary. In: Measuring Vulnerability to Natural Hazards [Birkmann, J. (ed.)].
UNU Press, Tokyo, Japan, pp. 448-496.
Chapter 2Determinants of Risk: Exposure and Vulnerability
107
Tol, R.S.J., T.E. Downing, O.J. Kuik, and J.B. Smith, 2004: Distributional aspects of
climate change impacts. Global Environmental Change, 14, 259-272.
Tong, S.M., P. Mather, G. Fitzgerald, D. McRae, K. Verrall, and D. Walker, 2010:
Assessing the vulnerability of eco-environmental health to climate change.
International Journal Of Environmental Research And Public Health, 7(2),
546-564.
Toni, F. and E. Holanda, 2008: The effects of land tenure on vulnerability to droughts
in Northeastern Brazil. Global Environmental Change-Human and Policy
Dimensions, 18(4), 575-582.
Tunstall, S., S. Tapsell, C. Green, and P. Floyd, 2006: The health effects of flooding:
social research results from England and Wales. Journal of Water Health, 4,
365-380.
Turner, B.L., R.E. Kasperson, P.A. Matson, J.J. McCarthy, R.W. Corell, L. Christensen,
N. Eckley, J.X. Kasperson, A. Luers, M.L. Martello, C. Polsky, A. Pulsipher, and A.
Schiller, 2003a: A framework for vulnerability analysis in sustainability science.
Proceedings of the National Academy of Sciences, 100(14), 8074-8079.
Turner, B.L. II, P.A. Matson, J.J. McCarthy, R.W. Corell, L. Christensen, N. Eckley, G. K.
Hovelsrud-Broda, J.X. Kasperson, R.E. Kasperson, A. Luers, M.L. Martello, S.
Mathiesen, R. Naylor, C. Polsky, A. Pulsipher, A. Schiller, H. Selin, and N. Tyler,
2003b: Illustrating the coupled human-environment system for vulnerability
analysis: Three case studies. Proceedings of the National Academy of Sciences,
100(14), 8080-8085.
Turner, N.J. and H. Clifton, 2009: “It’s so different today”: Climate change and
indigenous lifeways in British Columbia, Canada. Global Environmental
Change, 19, 180-190.
Twigg, J., 2001: Sustainable Livelihoods and Vulnerability to Disasters. Benfield Greig
Hazard Research Centre, Disaster Management Working Paper 2/2001, BGHRC,
UCL, London UK.
Twigg, J., 2007: Characteristics of a Disaster-resilient Community. A Guidance Note.
Version 1 (for field testing). DFID, Disaster Risk Reduction Interagency
Coordination Group, Benfield, UK.
Uitto, J.I., 1998: The geography of disaster vulnerability in megacities: a theoretical
framework. Applied Geography, 18(1), 7-16.
UK Department of Health, 2009: NHS Emergency Planning Guidance: Planning for
the psychosocial and mental health care of people affected by major incidents
and disasters: Interim national strategic guidance. UK DoH, London, UK.
UN, 2005: Hyogo Framework for Action 2005-2015: Building the Resilience of
Nations and Communities to Disasters, World Conference on Disaster
Reduction, 18-22 January 2005, Kobe, Japan, www.unisdr.org/wcdr/intergover/
official-doc/L-docs/Hyogo-framework-for-action-english.pdf.
UN Statistics Division, 2011, Millennium Development Goals Data Base. UN
Statistics Division, New York, NY, mdgs.un.org/unsd/mdg/Metadata.aspx.
UN/DESA, 2010: Trends in Sustainable Development in Small Island Development
States. United Nations, New York, NY, USA, 40 pp. ISBN 978-92-1-104610-6.
UN-HABITAT, 2003: Slums of the World: The Face of Urban Poverty in the New
Millennium? Monitoring the Millennium Development Goal, Target 11 – World-
Wide Slum Dweller Estimation. UN-Habitat, Nairobi, Kenya.
UN-HABITAT, 2009: Global Report on Human Settlements 2011: Planning
Sustainable Cities, United Nations Human Settlements Program, Earthscan,
London, UK, ISBN 978-1-84407-898-1.
UN-HABITAT
, 2011: Global Report on Human Settlements 2011: Cities and Climate
Change. United Nations Human Settlements Program, Earthscan, London, UK.
UNCED, 1992: The Global Partnership for Environment and Development: A Guide to
Agenda 21. United Nations Commission on Environment and Development
(UNCED), Geneva, Switzerland.
UNDHA, 1992: Internationally agreed glossary of basic terms relating to disaster
management. UNDHA, Geneva, Switzerland.
UNDP, 2004: Reducing Disaster Risk: A Challenge for Development, A Global Report.
UNDP, New York, NY.
UNDRO, 1980: Natural Disasters and Vulnerability Analysis. Report of Experts Group
Meeting of 9-12 July 1979, UNDRO, Geneva, Switzerland.
UNEP, 2007: Global Environment Outlook 4. United Nations Environment
Programme, Nairobi, Kenya.
UNEP/UNISDR, 2008: Environment and Disaster Risk Emerging Perspectives. UNISDR/
UNEP, Geneva, Switzerland.
UNICEF, 2009: Education in Emergencies and Post-Crisis Transition Consolidated
2009. Progress Report to the Government of The Netherlands and The European
Commission, UNICEF, New York, NY.
UNISDR, 2004: Living With Risk. United Nations International Strategy for Disaster
Reduction, Geneva, Switzerland.
UNISDR, 2007a: Words Into Action: A Guide for Implementing the Hyogo
Framework. United Nations International Strategy for Disaster Reduction,
Geneva, Switzerland.
UNISDR, 2007b: Towards a Culture of Prevention: Disaster Risk Reduction Begins at
School. Good Practices and Lessons Learned. United Nations International
Strategy for Disaster Reduction, Geneva, Switzerland.
UNISDR, 2009a: Global Assessment Report on Disaster Risk Reduction - Risk and
Poverty in a Changing Climate: Invest Today for a Safer Tomorrow. United
Nations International Strategy for Disaster Reduction, Geneva, Switzerland, 207
pp.
UNISDR, 2009b: Terminology on Disaster Risk Reduction. United Nations
International Strategy for Disaster Reduction, Geneva, Switzerland.
unisdr.org/eng/library/lib-terminology-eng.htm.
UNISDR, 2011: Global Assessment Report on Disaster Risk Reduction: Revealing
Risk, Redefining Development. United Nations International Strategy for
Disaster Reduction, Geneva, 178 pp., www.preventionweb.net/gar.
Urwin, K. and A. Jordan, 2008: Does public policy support or undermine climate change
adaptation? Exploring policy interplay across different scales of governance.
Global Environmental Change, 18, 180-191.
Utzinger, J. and J. Keiser, 2006: Urbanization and tropical health – then and now.
Annals of Tropical Medicine and Parasitology, 100(5-6), 517-533.
van Aalst, M.K., 2006a: The impacts of climate change on the risk of natural
disasters. Disasters, 30, 5-18.
van Aalst, M.K., 2006b: Managing Climate Risk: Integrating Climate Change
Adaptation into World Bank Operations. World Bank, Washington, DC,
USA.
van Aalst
, M.K., 2009: Bridging timescales. In: World Disasters Report 2009. Focus
on Early Warning, Early Action. IFRC, Geneva, Switzerland, pp. 68-93.
van Aalst, M.K. and I. Burton, 2002: The Last Straw. Integrating Natural Disaster
Mitigation with Environmental Management. World Bank Disaster Risk
Management Working Paper Series, 5, World Bank, Washington, DC.
van Aalst, M.K., M. Helmer, C. de Jong, F. Monasso, E. van Sluis, and P. Suarez, 2007:
Red Cross/Red Crescent Climate Guide. Red Cross/Red Crescent Climate Centre,
The Hague, The Netherlands, 144 pp.
van Aalst, M.K., T. Cannon, and I. Burton, 2008: Community level adaptation to climate
change: The potential role of participatory community risk assessment. Global
Environmental Change, 18, 165-179.
Van de Walle, B. and M. Turoff, 2007: Emergency response information systems:
Emerging trends and technologies. Communications of the ACM, 50(3), 29-31.
Van Lieshout, M., R.S. Kovats, M.T.J. Livermore, and P. Martens, 2004: Climate
change and malaria: analysis of the SRES climate and socio-economic scenarios.
Global Environmental Change, 14, 87-99.
Van Sluis, E. and M.K. van Aalst, 2006: Climate change and disaster risk in urban
environments. Humanitarian Exchange, 35.
Verdin, J., C. Funk, G. Senay, and R. Choularton, 2005: Climate science and famine
early warning. Philosophical Transactions of the Royal Society B, 360(1463),
2155-2168.
Villagrán de León, J.C., 2006: Vulnerability: A conceptual and methodological
review. SOURCE Publication Series of UNU-EHS, 4, United Nations University
Institute for Environment and Human Security (UNU-EHS), Bonn, Germany.
Vincent, K., 2007: Uncertainty in adaptive capacity and the importance of scale.
Global Environmental Change, 17(1), 12-24.
Vitoria, M., R. Granich, C.F. Gilks, C. Gunneberg, M. Hosseini, W. Were, M. Raviglione,
and K.M. Cock, 2009: The global fight against HIV/AIDS, tuberculosis, and
malaria: Current status and future perspectives. American Journal of Clinical
Pathology, 131(6), 844-848.
Chapter 2 Determinants of Risk: Exposure and Vulnerability
108
Vogel, C. and K. O’Brien, 2004: Vulnerability and Global Environmental Change:
Rhetoric and Reality. AVISO 13, Global Environmental Change and Human
Security Project, Ottawa, Canada.
Waktola, D.K., 2009: Challenges and opportunities in mainstreaming environmental
education into the curricula of teachers’ colleges in Ethiopia. Environmental
Education Research, 15(5), 589-605.
Walker, B., C.S. Holling, S.R. Carpenter, and A. Kinzig, 2004: Resilience, adaptability
and transformability in social–ecological systems. Ecology and Society, 9(2), 5.
Watts, M.J. and H.G. Bohle, 1993: The space of vulnerability: the causal structure of
hunger and famine. Progress in Human Geography, 17(1), 43-67.
Weichselgartner, J., 2001: Disaster mitigation: the concept of vulnerability revisited.
Disaster Prevention and Management, 10(2), 85-94.
Werritty, A., D. Houston, T. Ball, A. Tavendale, and A. Black, 2007: Exploring The
Social Impacts Of Flood Risk And Flooding In Scotland. Scottish Executive Social
Research, Edinburgh, Scotland.
Westgate, K.N., P. O’Keefe, 1976: Some Definitions of Disaster. Disaster Research
Unit Occasional Paper 4, Department of Geography, University of Bradford, UK.
White, G.F., 1973: Natural hazards research. In: Directions in Geography [Chorley, R.J.
(ed.)]. Methuen and Co., London, UK, pp. 193-216.
White, G.F. (ed.), 1974: Natural Hazards: Local, National, Global. Oxford University
Press, New York, NY.
White, I., 2002: Water Management in the Mekong Delta: Changes, Conflicts and
Opportunities. IHP-VI Technical Papers in Hydrology No. 61, UNESCO, Paris, France.
Wijkman, A. and L. Timberlake, 1984: Natural Disasters: Act of God or Acts of Man.
Earthscan, Washington, DC.
Wilbanks, T.J., 2003: Integrating climate change and sustainable development in a
place-based context. Climate Policy, 3(S1), S147-S154.
Wilbanks, T.J., P. Romero Lankao, M. Bao, F. Berkhout, S. Cairncross, J.-P. Ceron, M.
Kapshe, R. Muir-Wood, and R. Zapata-Marti, 2007: Industry, settlement and
society. In: Climate Change 2007. Impacts, Adaptation and Vulnerability.
Contribution of Working Group II to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P.
Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, 357-390.
Wilby R.L. and G. L.W. Perry, 2006: Climate change, biodiversity and the urban
environment: a critical review based on London, UK. Progress in Physical
Geography, 30(1), 73-98.
Wilches-Chaux, G., 1989: Desastres, ecologismo y formación profesional. SENA,
Popayán, Co lom bia.
Williams, S.E, L.P. Shoo, J.L. Isaac, A.A. Hoffmann, and G. Langham, 2008: Towards
an integrated framework for assessing the vulnerability of species to climate
change. PLoS Biology, 6(12), e325, doi:10.1371/journal. pbio.0060325.
Willows, R.I. and R.K. Connell (eds.), 2003: Climate Adaptation: Risk, Uncertainty
and Decision-Making. UKCIP Technical Report. UKCIP, Oxford, UK.
Wisner, B.G., 1978: An appeal for a significantly comparative method in disaster
research. Disasters, 2, 80-82.
Wisner, B., 1993: Disaster vulnerability: Scale, power and daily life. GeoJournal, 30,
127-140.
Wisner, B., 1998: Marginality and vulnerability: why the homeless of Tokyo don’t
‘count’ in disaster preparations. Applied Geography, 18(1), 25-33.
Wisner, B., 2001a: Capitalism and the shifting spatial and social distribution of
hazard and vulnerability. Australian Journal of Emergency Management,
Winter 2001, 44-50.
Wisner, B., 2001b: Risk and the Neoliberal State: Why post-Mitch lessons didn’t
reduce El Salvador’s earthquake losses. Disasters, 25 (3), 251-268.
Wisner, B., 2003: Disaster risk reduction in megacities – Making the most of human
and social capital. In: Building Safer Cities- The Future of Disaster Risk [Kreimer,
A., M. Arnold, and C. Carlin (eds.)]. Disaster Management Facility, World Bank,
Washington, DC, pp. 181-196.
Wisner, B., 2006a: Self-assessment of coping capacity: Participatory, proactive, and
qualitative engagement of communities in their own risk management. In:
Measuring Vulnerability to Natural Hazards - Towards Disaster Resilient
Societies [Birkmann, J. (ed.)]. United Nations University Press, Tokyo, Japan, pp.
316-328.
Wisner, B., 2006b: Let Our Children Teach Us! A Review of the Role of Education and
Knowledge in Disaster Risk Reduction. UNISDR System Thematic
Cluster/Platform on Knowledge and Education, Geneva, Switzerland, 135 pp.
Wisner, B. and J. Adams (eds.), 2002: Environment and Health in Emergencies and
Disasters: A Practical Guide. World Health Organization, Geneva, Switzerland.
Wisner, B., P. Blaikie, T. Cannon, and I. Davis, 2004: At Risk, Natural Hazards, People’s
Vulnerability and Disasters. Routledge, London, UK.
Wolf, J., N.W. Adger, and I. Lorenzoni, 2010a: Heat waves and cold spells: an analysis
of policy response and perceptions of vulnerable populations in the UK.
Environment and Planning A, 42, 2721-2734, doi:10.1068/a42503.
Wolf, J., W.N. Adger, I. Lorenzoni, V. Abrahamson, and R. Raine, 2010b: Social capital,
individual responses to heat waves and climate change adaptation: An empirical
study of two UK cities. Global Environmental Change, 20, 44-52.
Woo, G., 1999: The Mathematics of Natural Catastrophes. Imperial College Press,
London, UK.
Woolcock, M. and D. Narayan, 2000: Social capital: Implications for development
theory, research and policy. World Bank Research Observer, 15, 225-249.
Yodmani, S. 2001; Disaster preparedness and management. In: Social Protection in
Asia and the Pacific [Ortiz, I.D. (ed.)]. Asian Development Bank, Manila, pp. 481-
502.
Yohe, G. and R.S.J. Tol, 2002: Indicators for social and economic coping capacity:
moving toward a working definition of adaptive capacity,
Global Environmental
Change, 12, 25-40.
Chapter 2Determinants of Risk: Exposure and Vulnerability
109
Coordinating Lead Authors:
Sonia I. Seneviratne (Switzerland), Neville Nicholls (Australia)
Lead Authors:
David Easterling (USA), Clare M. Goodess (United Kingdom), Shinjiro Kanae (Japan), James Kossin
(USA), Yali Luo (China), Jose Marengo (Brazil), Kathleen McInnes (Australia), Mohammad Rahimi (Iran),
Markus Reichstein (Germany), Asgeir Sorteberg (Norway), Carolina Vera (Argentina), Xuebin Zhang
(Canada)
Review Editors:
Matilde Rusticucci (Argentina), Vladimir Semenov (Russia)
Contributing Authors:
Lisa V. Alexander (Australia), Simon Allen (Switzerland), Gerardo Benito (Spain), Tereza Cavazos
(Mexico), John Clague (Canada), Declan Conway (United Kingdom), Paul M. Della-Marta (Switzerland),
Markus Gerber (Switzerland), Sunling Gong (Canada), B. N. Goswami (India), Mark Hemer (Australia),
Christian Huggel (Switzerland), Bart van den Hurk (Netherlands), Viatcheslav V. Kharin (Canada),
Akio Kitoh (Japan), Albert M.G. Klein Tank (Netherlands), Guilong Li (Canada), Simon Mason (USA),
William McGuire (United Kingdom), Geert Jan van Oldenborgh (Netherlands), Boris Orlowsky
(Switzerland), Sharon Smith (Canada), Wassila Thiaw (USA), Adonis Velegrakis (Greece), Pascal Yiou
(France), Tingjun Zhang (USA), Tianjun Zhou (China), Francis W. Zwiers (Canada)
This chapter should be cited as:
Seneviratne, S.I., N. Nicholls, D. Easterling, C.M. Goodess, S. Kanae, J. Kossin, Y. Luo, J. Marengo, K. McInnes, M. Rahimi,
M. Reichstein, A. Sorteberg, C. Vera, and X. Zhang, 2012: Changes in climate extremes and their impacts on the natural
physical environment. In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation
[Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen,
M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate
Change (IPCC). Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 109-230.
3
Changes in Climate Extremes
and their Impacts on the
Natural Physical Environment
Changes in Climate Extremes and their Impacts on the Natural Physical Environment
110
Executive Summary .................................................................................................................................111
3.1. Weather and Climate Events Related to Disasters ..................................................................115
3.1.1. Categories of Weather and Climate Events Discussed in this Chapter ...........................................................................115
3.1.2. Characteristics of Weather and Climate Events Relevant to Disasters ...........................................................................115
3.1.3. Compound (Multiple) Events............................................................................................................................................118
3.1.4. Feedbacks.........................................................................................................................................................................118
3.1.5. Confidence and Likelihood of Assessed Changes in Extremes ........................................................................................120
3.1.6. Changes in Extremes and Their Relationship to Changes in Regional and Global Mean Climate..................................121
3.1.7. Surprises / Abrupt Climate Change..................................................................................................................................122
3.2. Requirements and Methods for Analyzing Changes in Extremes............................................122
3.2.1. Observed Changes............................................................................................................................................................122
3.2.2. The Causes behind the Changes ......................................................................................................................................125
3.2.3. Projected Long-Term Changes and Uncertainties ............................................................................................................128
3.3. Observed and Projected Changes in Weather and Climate Extremes .....................................133
3.3.1. Temperature .....................................................................................................................................................................133
3.3.2. Precipitation.....................................................................................................................................................................141
3.3.3. Wind .................................................................................................................................................................................149
3.4. Observed and Projected Changes in
Phenomena Related to Weather and Climate Extremes..........................................................152
3.4.1. Monsoons .........................................................................................................................................................................152
3.4.2. El Niño-Southern Oscillation............................................................................................................................................155
3.4.3. Other Modes of Variability...............................................................................................................................................157
3.4.4. Tropical Cyclones..............................................................................................................................................................158
3.4.5. Extratropical Cyclones......................................................................................................................................................163
3.5. Observed and Projected Impacts on the Natural Physical Environment.................................167
3.5.1. Droughts...........................................................................................................................................................................167
3.5.2. Floods...............................................................................................................................................................................175
3.5.3. Extreme Sea Levels ..........................................................................................................................................................178
3.5.4. Waves ...............................................................................................................................................................................180
3.5.5. Coastal Impacts................................................................................................................................................................182
3.5.6. Glacier, Geomorphological, and Geological Impacts .......................................................................................................186
3.5.7. High-latitude Changes Including Permafrost...................................................................................................................189
3.5.8. Sand and Dust Storms......................................................................................................................................................190
References ...............................................................................................................................................203
Boxes and Frequently Asked Questions
Box 3-1. Definition and Analysis of Climate Extremes in the Scientific Literature .......................................................................116
FAQ 3.1. Is the Climate Becoming More Extreme? ........................................................................................................................124
FAQ 3.2. Has Climate Change Affected Individual Extreme Events? .............................................................................................127
Box 3-2. Variations in Confidence in Projections of Climate Change: Mean versus Extremes, Variables, Scale ...........................132
Box 3-3. The Definition of Drought................................................................................................................................................167
Box 3-4. Small Island States...........................................................................................................................................................184
Supplementary Material
Appendix 3.A: Notes and Technical Details on Chapter 3 Figures................................................................................Available On-Line
Chapter 3
Table of Contents
111
This chapter addresses changes in weather and climate events relevant to extreme impacts and disasters.
An extreme (weather or climate) event is generally defined as the occurrence of a value of a weather or climate
variable above (or below) a threshold value near the upper (or lower) ends (‘tails’) of the range of observed values of
the variable. Some climate extremes (e.g., droughts, floods) may be the result of an accumulation of weather or climate
events that are, individually, not extreme themselves (though their accumulation is extreme). As well, weather or
climate events, even if not extreme in a statistical sense, can still lead to extreme conditions or impacts, either by
crossing a critical threshold in a social, ecological, or physical system, or by occurring simultaneously with other
events. A weather system such as a tropical cyclone can have an extreme impact, depending on where and when it
approaches landfall, even if the specific cyclone is not extreme relative to other tropical cyclones. Conversely, not all
extremes necessarily lead to serious impacts. [3.1]
Many weather and climate extremes are the result of natural climate variability (including phenomena
such as El Niño), and natural decadal or multi-decadal variations in the climate provide the backdrop for
anthropogenic climate changes. Even if there were no anthropogenic changes in climate, a wide variety of natural
weather and climate extremes would still occur. [3.1]
A changing climate leads to changes in the frequency, intensity, spatial extent, duration, and timing of
weather and climate extremes, and can result in unprecedented extremes. Changes in extremes can also be
directly related to changes in mean climate, because mean future conditions in some variables are projected to lie
within the tails of present-day conditions. Nevertheless, changes in extremes of a climate or weather variable are not
always related in a simple way to changes in the mean of the same variable, and in some cases can be of opposite
sign to a change in the mean of the variable. Changes in phenomena such as the El Niño-Southern Oscillation or
monsoons could affect the frequency and intensity of extremes in several regions simultaneously. [3.1]
Many factors affect confidence in observed and projected changes in extremes. Our confidence in observed
changes in extremes depends on the quality and quantity of available data and the availability of studies analyzing
these data. It consequently varies between regions and for different extremes. Similarly, our confidence in projecting
changes (including the direction and magnitude of changes in extremes) varies with the type of extreme, as well as
the considered region and season, depending on the amount and quality of relevant observational data and model
projections, the level of understanding of the underlying processes, and the reliability of their simulation in models
(assessed from expert judgment, model validation, and model agreement). Global-scale trends in a specific extreme
may be either more reliable (e.g., for temperature extremes) or less reliable (e.g., for droughts) than some regional-
scale trends, depending on the geographical uniformity of the trends in the specific extreme. Low confidence’ in
observed or projected changes in a specific extreme neither implies nor excludes the possibility of changes in this
extreme. [3.1.5, 3.1.6, 3.2.3; Box 3-2; Figures 3-3, 3-4, 3-5, 3-6, 3-7, 3-8, 3-10]
There is evidence from observations gathered since 1950 of change in some extremes. It is very likely that
there has been an overall decrease in the number of cold days and nights, and an overall increase in the number of
warm days and nights, at the global scale, that is, for most land areas with sufficient data. It is likely that these changes
have also occurred at the continental scale in North America, Europe, and Australia. There is medium confidence of a
warming trend in daily temperature extremes in much of Asia. Confidence in observed trends in daily temperature
extremes in Africa and South America generally varies from low to medium depending on the region. Globally, in many
(but not all) regions with sufficient data there is medium confidence that the length or number of warm spells or heat
waves has increased since the middle of the 20th century. It is likely that there have been statistically significant
increases in the number of heavy precipitation events (e.g., 95th percentile) in more regions than there have been
statistically significant decreases, but there are strong regional and subregional variations in the trends. There is
low confidence that any observed long-term (i.e., 40 years or more) increases in tropical cyclone activity are robust,
after accounting for past changes in observing capabilities. It is likely that there has been a poleward shift in the
main Northern and Southern Hemisphere extratropical storm tracks. There is low confidence in observed trends in
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Executive Summary
112
small-scale phenomena such as tornadoes and hail because of data inhomogeneities and inadequacies in monitoring
systems. There is medium confidence that since the 1950s some regions of the world have experienced a trend to
more intense and longer droughts, in particular in southern Europe and West Africa, but in some regions droughts
have become less frequent, less intense, or shorter, for example, in central North America and northwestern Australia.
There is limited to medium evidence available to assess climate-driven observed changes in the magnitude and
frequency of floods at regional scales because the available instrumental records of floods at gauge stations are limited
in space and time, and because of confounding effects of changes in land use and engineering. Furthermore, there is
low agreement in this evidence, and thus overall low confidence at the global scale regarding even the sign of these
changes. It is likely that there has been an increase in extreme coastal high water related to increases in mean sea
level in the late 20th century. [3.2.1, 3.3.1, 3.3.2, 3.3.3, 3.4.4, 3.4.5, 3.5.1, 3.5.2, 3.5.3; Tables 3-1, 3-2]
There is evidence that some extremes have changed as a result of anthropogenic influences, including
increases in atmospheric concentrations of greenhouse gases. It is likely that anthropogenic influences have led
to warming of extreme daily minimum and maximum temperatures at the global scale. There is medium confidence
that anthropogenic influences have contributed to intensification of extreme precipitation at the global scale. It is
likely that there has been an anthropogenic influence on increasing extreme coastal high water due to an increase in
mean sea level. The uncertainties in the historical tropical cyclone records, the incomplete understanding of the physical
mechanisms linking tropical cyclone metrics to climate change, and the degree of tropical cyclone variability provide
only low confidence for the attribution of any detectable changes in tropical cyclone activity to anthropogenic
influences. Attribution of single extreme events to anthropogenic climate change is challenging. [3.2.2, 3.3.1, 3.3.2,
3.4.4, 3.5.3; Table 3-1]
The following assessments of the likelihood of and/or confidence in projections are generally for the end
of the 21st century and relative to the climate at the end of the 20th century. There are three main sources of
uncertainty in the projections: the natural variability of climate; uncertainties in climate model parameters and
structure; and projections of future emissions. Projections for differing emissions scenarios generally do not strongly
diverge in the coming two to three decades, but uncertainty in the sign of change is relatively large over this time
frame because climate change signals are expected to be relatively small compared to natural climate variability. For
certain extremes (e.g., precipitation-related extremes), the uncertainty in projected changes by the end of the 21st
century is more the result of uncertainties in climate models rather than uncertainties in future emissions. For other
extremes (in particular temperature extremes at the global scale and in most regions), the emissions uncertainties are
the main source of uncertainty in projections for the end of the 21st century. In the assessments provided in this
chapter, uncertainties in projections from the direct evaluation of multi-model ensemble projections are modified by
taking into account the past performance of models in simulating extremes (for instance, simulations of late 20th-
century changes in extreme temperatures appear to overestimate the observed warming of warm extremes and
underestimate the warming of cold extremes), the possibility that some important processes relevant to extremes may
be missing or be poorly represented in models, and the limited number of model projections and corresponding
analyses currently available of extremes. For these reasons the assessed uncertainty is generally greater than would be
assessed from the model projections alone. Low-probability, high-impact changes associated with the crossing of
poorly understood climate thresholds cannot be excluded, given the transient and complex nature of the climate
system. Feedbacks play an important role in either damping or enhancing extremes in several climate variables.
[3.1.4, 3.1.7, 3.2.3, 3.3.1, 3.3.2; Box 3-2]
Models project substantial warming in temperature extremes by the end of the 21st century. It is virtually
certain that increases in the frequency and magnitude of warm daily temperature extremes and decreases in cold
extremes will occur through the 21st century at the global scale. It is very likely that the length, frequency, and/or
intensity of warm spells or heat waves will increase over most land areas. For the Special Report on Emissions
Scenarios (SRES) A2 and A1B emission scenarios, a 1-in-20 year annual hottest day is likely to become a 1-in-2 year
annual extreme by the end of the 21st century in most regions, except in the high latitudes of the Northern
Hemisphere where it is likely to become a 1-in-5 year annual extreme. In terms of absolute values, 20-year extreme
annual daily maximum temperature (i.e., return value) will likely increase by about 1 to 3°C by mid-21st century and
by about 2 to 5°C by the late 21st century, depending on the region and emissions scenario (considering the B1, A1B,
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
113
and A2 scenarios). Regional changes in temperature extremes will often differ from the mean global temperature
change. [3.3.1; Table 3-3; Figure 3-5]
It is likely that the frequency of heavy precipitation or the proportion of total rainfall from heavy rainfalls
will increase in the 21st century over many areas of the globe. This is particularly the case in the high latitudes
and tropical regions, and in winter in the northern mid-latitudes. Heavy rainfalls associated with tropical cyclones are
likely to increase with continued warming induced by enhanced greenhouse gas concentrations. There is medium
confidence that, in some regions, increases in heavy precipitation will occur despite projected decreases in total
precipitation. For a range of emission scenarios (SRES A2, A1B, and B1), a 1-in-20 year annual maximum 24-hour
precipitation rate is likely to become a 1-in-5 to 1-in-15 year event by the end of the 21st century in many regions,
and in most regions the higher emissions scenarios (A1B and A2) lead to a greater projected decrease in return
period. Nevertheless, increases or statistically non-significant changes in return periods are projected in some regions.
[3.3.2; Table 3-3; Figure 3-7]
There is generally low confidence in projections of changes in extreme winds because of the relatively few
studies of projected extreme winds, and shortcomings in the simulation of these events. An exception is
mean tropical cyclone maximum wind speed, which is likely to increase, although increases may not occur in all ocean
basins. It is likely that the global frequency of tropical cyclones will either decrease or remain essentially unchanged.
There is low confidence in projections of small-scale phenomena such as tornadoes because competing physical
processes may affect future trends and because climate models do not simulate such phenomena. There is medium
confidence that there will be a reduction in the number of mid-latitude cyclones averaged over each hemisphere due
to future anthropogenic climate change. There is low confidence in the detailed geographical projections of mid-latitude
cyclone activity. There is medium confidence in a projected poleward shift of mid-latitude storm tracks due to future
anthropogenic forcings. [3.3.3, 3.4.4, 3.4.5]
Uncertainty in projections of changes in large-scale patterns of natural climate variability remains large.
There is low confidence in projections of changes in monsoons (rainfall, circulation), because there is little consensus
in climate models regarding the sign of future change in the monsoons. Model projections of changes in El Niño-
Southern Oscillation variability and the frequency of El Niño episodes as a consequence of increased greenhouse gas
concentrations are not consistent, and so there is low confidence in projections of changes in the phenomenon.
However, most models project an increase in the relative frequency of central equatorial Pacific events (which typically
exhibit different patterns of climate variations than do the classical East Pacific events). There is low confidence in the
ability to project changes in other natural climate modes including the North Atlantic Oscillation, the Southern Annular
Mode, and the Indian Ocean Dipole. [3.4.1, 3.4.2, 3.4.3]
It is very likely that mean sea level rise will contribute to upward trends in extreme coastal high water
levels in the future. There is high confidence that locations currently experiencing adverse impacts such as coastal
erosion and inundation will continue to do so in the future due to increasing sea levels, all other contributing factors
being equal. There is low confidence in wave height projections because of the small number of studies, the lack of
consistency of the wind projections between models, and limitations in the models’ ability to simulate extreme winds.
Future negative or positive changes in significant wave height are likely to reflect future changes in storminess and
associated patterns of wind change. [3.5.3, 3.5.4, 3.5.5]
Projected precipitation and temperature changes imply possible changes in floods, although overall there
is low confidence in projections of changes in fluvial floods. Confidence is low due to limited evidence and
because the causes of regional changes are complex, although there are exceptions to this statement. There is medium
confidence (based on physical reasoning) that projected increases in heavy rainfall would contribute to increases in
local flooding, in some catchments or regions. Earlier spring peak flows in snowmelt and glacier-fed rivers are very
likely. [3.5.2]
There is medium confidence that droughts will intensify in the 21st century in some seasons and areas,
due to reduced precipitation and/or increased evapotranspiration. This applies to regions including southern
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
114
Europe and the Mediterranean region, central Europe, central North America, Central America and Mexico, northeast
Brazil, and southern Africa. Definitional issues, lack of observational data, and the inability of models to include all the
factors that influence droughts preclude stronger confidence than medium in the projections. Elsewhere there is overall
low confidence because of inconsistent projections of drought changes (dependent both on model and dryness index).
There is low confidence in projected future changes in dust storms although an increase could be expected where
aridity increases. [3.5.1, 3.5.8; Box 3-3; Table 3-3; Figure 3-10]
There is high confidence that changes in heat waves, glacial retreat, and/or permafrost degradation will
affect high-mountain phenomena such as slope instabilities, mass movements, and glacial lake outburst
floods. There is also high confidence that changes in heavy precipitation will affect landslides in some
regions. There is low confidence regarding future locations and timing of large rock avalanches, as these depend on
local geological conditions and other non-climatic factors. There is low confidence in projections of an anthropogenic
effect on phenomena such as shallow landslides in temperate and tropical regions, because these are strongly
influenced by human activities such as land use practices, deforestation, and overgrazing. [3.5.6, 3.5.7]
The small land area and often low elevation of small island states make them particularly vulnerable to
rising sea levels and impacts such as inundation, shoreline change, and saltwater intrusion into
underground aquifers. Short record lengths and the inadequate resolution of current climate models to represent
small island states limit the assessment of changes in extremes. There is insufficient evidence to assess observed
trends and future projections in rainfall across the small island regions considered here. There is medium confidence in
projected temperature increases across the Caribbean. The very likely contribution of mean sea level rise to increased
extreme coastal high water levels, coupled with the likely increase in tropical cyclone maximum wind speed, is a
specific issue for tropical small island states. [3.4.4, 3.5.3; Box 3-4]
This chapter does not provide assessments of projected changes in extremes at spatial scales smaller than
for large regions. These large-region projections provide a wider context for national or local projections,
where these exist, and where they do not exist, a first indication of expected changes, their associated
uncertainties, and the evidence available. [3.2.3.1]
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
115
3.1. Weather and Climate Events
Related to Disasters
A changing climate leads to changes in the frequency, intensity, spatial
extent, duration, and timing of weather and climate extremes, and can
result in unprecedented extremes (Sections 3.1.7, 3.3, 3.4, and 3.5). As
well, weather or climate events, even if not extreme in a statistical sense,
can still lead to extreme conditions or impacts, either by crossing a critical
threshold in a social, ecological, or physical system, or by occurring
simultaneously with other events (Sections 3.1.2, 3.1.3, 3.1.4, 3.3, 3.4,
and 3.5). Some climate extremes (e.g., droughts, floods) may be the result
of an accumulation of weather or climate events that are, individually,
not extreme themselves (though their accumulation is extreme, e.g.,
Section 3.1.2). A weather system such as a tropical cyclone can have an
extreme impact, depending on where and when it approaches landfall,
even if the specific cyclone is not extreme relative to other tropical
cyclones. Conversely, not all extremes necessarily lead to serious impacts.
Changes in extremes can also be directly related to changes in mean
climate, because mean future conditions in some variables are projected
to lie within the tails of present-day conditions (Section 3.1.6). Hence,
the definition of extreme weather and climate events is complex
(Section 3.1.2 and Box 3-1) and the assessment of changes in climate
that are relevant to extreme impacts and disasters needs to consider
several aspects. Those related to vulnerability and exposure are
addressed in Chapters 2 and 4 of this report, while we focus here on the
physical dimension of these events.
Many weather and climate extremes are the result of natural climate
variability (including phenomena such as El Niño), and natural decadal
or multi-decadal variations in the climate provide the backdrop for
anthropogenic climate changes. Even if there were no anthropogenic
changes in climate, a wide variety of natural weather and climate
extremes would still occur.
3.1.1. Categories of Weather and Climate Events
Discussed in this Chapter
This chapter addresses changes in weather and climate events relevant
to extreme impacts and disasters grouped into the following categories:
1) Extremes of atmospheric weather and climate variables (temperature,
precipitation, wind)
2) Weather and climate phenomena that influence the occurrence of
extremes in weather or climate variables or are extremes themselves
(monsoons, El Niño and other modes of variability, tropical and
extratropical cyclones)
3) Impacts on the natural physical environment (droughts, floods,
extreme sea level, waves, and coastal impacts, as well as other
physical impacts, including cryosphere-related impacts, landslides,
and sand and dust storms).
The distinction between these three categories is somewhat arbitrary, and
the categories are also related. In the case of the third category, ‘impacts
on the natural physical environment,’ a specific distinction between
these events and those considered under ‘extremes of atmospheric
weather and climate variables’ is that they are not caused by variations
in a single atmospheric weather and climate variable, but are generally
the result of specific conditions in several variables, as well as of some
surface properties or states. For instance, both floods and droughts are
related to precipitation extremes, but are also impacted by other
atmospheric and surface conditions (and are thus often better viewed as
compound events, see Section 3.1.3). Most of the impacts on the natural
physical environment discussed in the third category are extremes
themselves, as well as often being caused or affected by atmospheric
weather or climate extremes. Another arbitrary choice made here is the
separate category for phenomena (or climate or weather systems) that
are related to weather and climate extremes, such as monsoons, El Niño,
and other modes of variability. These phenomena affect the large-scale
environment that, in turn, influences extremes. For instance, El Niño
episodes typically lead to droughts in some regions with, simultaneously,
heavy rains and floods occurring elsewhere. This means that all
occurrences of El Niño are relevant to extremes and not only extreme
El Niño episodes. A change in the frequency or nature of El Niño episodes
(or in their relationships with climate in specific regions) would affect
extremes in many locations simultaneously. Similarly, changes in monsoon
patterns could affect several countries simultaneously. This is especially
important from an international disaster perspective because coping
with disasters in several regions simultaneously may be challenging
(see also Section 3.1.3 and Chapters 7 and 8).
This section provides background material on the characterization and
definition of extreme events, the definition and analysis of compound
events, the relevance of feedbacks for extremes, the approach used for
the assignment of confidence and likelihood assessments in this chapter,
and the possibility of ‘surprises’ regarding future changes in extremes.
Requirements and methods for analyzing changes in climate extremes
are addressed in Section 3.2. Assessments regarding changes in the
climate variables, phenomena, and impacts considered in this chapter
are provided in Sections 3.3 to 3.5. Table 3-1 provides summaries of
these assessments for changes at the global scale. Tables 3-2 and 3-3
(found on pages 191-202) provide more regional detail on observed and
projected changes in temperature extremes, heavy precipitation, and
dryness (with regions as defined in Figure 3-1). Note that impacts on
ecosystems (e.g., bushfires) and human systems (e.g., urban flooding)
are addressed in Chapter 4.
3.1.2. Characteristics of Weather and Climate Events
Relevant to Disasters
The identification and definition of weather and climate events that are
relevant from a risk management perspective are complex and depend
on the stakeholders involved (Chapters 1 and 2). In this chapter, we focus
on the assessment of changes in ‘extreme climate or weather events’
(also referred to herein as ‘climate extremes’ see below and Glossary),
which generally correspond to the ‘hazards’ discussed in Chapter 1.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
116
Hence, the present chapter does not directly consider the dimensions of
vulnerability or exposure, which are critical in determining the human
and ecosystem impacts of climate extremes (Chapters 1, 2, and 4).
This report defines an ‘extreme climate or weather event’ or ‘climate
extreme’ as “the occurrence of a value of a weather or climate variable
above (or below) a threshold value near the upper (or lower) ends of the
range of observed values of the variable” (see Glossary). Several
aspects of this definition can be clarified thus:
Definitions of thresholds vary, but values with less than 10, 5, 1%, or
even lower chance of occurrence for a given time of the year (day,
month, season, whole year) during a specified reference period
(generally 1961-1990) are often used. In some circumstances,
information from sources other than observations, such as model
projections, can be used as a reference.
Absolute thresholds (rather than these relative thresholds based
on the range of observed values of a variable) can also be used to
identify extreme events (e.g., specific critical temperatures for
health impacts).
What is called an extreme weather or climate event will vary from
place to place in an absolute sense (e.g., a hot day in the tropics
will be a different temperature than a hot day in the mid-latitudes),
and possibly in time given some adaptation from society (see
Box 3-1).
Some climate extremes (e.g., droughts, floods) may be the result
of an accumulation of moderate weather or climate events (this
accumulation being itself extreme). Compound events (see Section
3.1.3), that is, two or more events occurring simultaneously, can
lead to high impacts, even if the two single events are not extreme
per se (only their combination).
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Box 3-1 | Definition and Analysis of Climate Extremes in the Scientific Literature
This box provides some details on the definition of climate extremes in the scientific literature and on common approaches employed for
their investigation.
A large amount of the available scientific literature on climate extremes is based on the use of so-called ‘extreme indices,’ which can
either be based on the probability of occurrence of given quantities or on threshold exceedances (Section 3.1.2). Typical indices that are
seen in the scientific literature include the number, percentage, or fraction of days with maximum temperature (Tmax) or minimum
temperature (Tmin), below the 1st, 5th, or 10th percentile, or above the 90th, 95th, or 99th percentile, generally defined for given time
frames (days, month, season, annual) with respect to the 1961-1990 reference time period. Commonly, indices for 10th and 90th
percentiles of Tmax/Tmin computed on daily time frames are referred to as ‘cold/warm days/nights’ (e.g., Figures 3-3 and 3-4; Tables 3-1
to 3-3, and Section 3.3.1; see also Glossary). Other definitions relate to, for example, the number of days above specific absolute
temperature or precipitation thresholds, or more complex definitions related to the length or persistence of climate extremes. Some
advantages of using predefined extreme indices are that they allow some comparability across modelling and observational studies and
across regions (although with limitations noted below). Moreover, in the case of observations, derived indices may be easier to obtain
than is the case with daily temperature and precipitation data, which are not always distributed by meteorological services. Peterson and
Manton (2008) discuss collaborative international efforts to monitor extremes by employing extreme indices. Typically, although not
exclusively, extreme indices used in the scientific literature reflect ‘moderate extremes,’ for example, events occurring as often as 5 or 10%
of the time. More extreme ‘extremes’ are often investigated using Extreme Value Theory (EVT) due to sampling issues (see below).
Extreme indices are often defined for daily temperature and precipitation characteristics, and are also sometimes applied to seasonal
characteristics of these variables, to other weather and climate variables, such as wind speed, humidity, or to physical impacts and
phenomena. Beside analyses for temperature and precipitation indices (see Sections 3.3.1 and 3.3.2; Tables 3-2 and 3-3), other studies
are, for instance, available in the literature for wind-based (Della-Marta et al., 2009) and pressure-based (Beniston, 2009a) indices, for
health-relevant indices (e.g., ‘heat index’) combining temperature and relative humidity characteristics (e.g., Diffenbaugh et al., 2007;
Fischer and Schär, 2010; Sherwood and Huber, 2010), and for a range of dryness indices (see Box 3-3).
Extreme Value Theory is an approach used for the estimation of extreme values (e.g., Coles, 2001), which aims at deriving a probability
distribution of events from the tail of a probability distribution, that is, at the far end of the upper or lower ranges of the probability
distributions (typically occurring less frequently than once per year or per period of interest, i.e., generally less than 1 to 5% of the
considered overall sample). EVT is used to derive a complete probability distribution for such low-probability events, which can also help
analyzing the probability of occurrence of events that are outside of the observed data range (with limitations). Two different approaches
can be used to estimate the parameters for such probability distributions. In the block maximum approach, the probability distribution
parameters are estimated for maximum values of consecutive blocks of a time series (e.g., years). In the second approach, instead of the
block maxima the estimation is based on events that exceed a high threshold (peaks over threshold approach). Both approaches are
used in climate research.
Continued next page
117
Not all extreme weather and climate events necessarily have
extreme impacts.
The distinction between extreme weather events and extreme climate
events is not precise, but is related to their specific time scales:
An extreme weather event is typically associated with changing
weather patterns, that is, within time frames of less than a day
to a few weeks.
An extreme climate event happens on longer time scales. It can
be the accumulation of several (extreme or non-extreme)
weather events (e.g., the accumulation of moderately below-
average rainy days over a season leading to substantially below-
average cumulated rainfall and drought conditions).
For simplicity, we collectively refer to both extreme weather events and
extreme climate events with the term ‘climate extremes’ in this chapter.
From this definition, it can be seen that climate extremes can be defined
quantitatively in two ways:
1) Related to their probability of occurrence
2) Related to a specific (possibly impact-related) threshold.
The first type of definition can either be expressed with respect to given
percentiles of the distribution functions of the variables, or with respect
to specific return frequencies (e.g., ‘100-year event’). Compound events
can be viewed as a special category of climate extremes, which result
from the combination of two or more events, and which are again
‘extreme’ either from a statistical perspective (tails of distribution functions
of climate variables) or associated with a specific threshold (Section
3.1.3.). These two definitions of climate extremes, probability-based or
threshold-based, are not necessarily antithetic. Indeed, hazards for
society and ecosystems are often extreme both from a probability and
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Recent publications have used other approaches for evaluating characteristics of extremes or changes in extremes, for instance, analyzing
trends in record events or investigating whether records in observed time series are being set more or less frequently than would be
expected in an unperturbed climate (Benestad, 2003, 2006; Zorita et al., 2008; Meehl et al., 2009c; Trewin and Vermont, 2010).
Furthermore, besides the actual magnitude of extremes (quantified in terms of probability/return frequency or absolute threshold), other
relevant aspects for the definition of climate extremes from an impact perspective include the event’s duration, the spatial area affected,
timing, frequency, onset date, continuity (i.e., whether there are ‘breaks’ within a spell), and preconditioning (e.g., rapid transition from a
slowly developing meteorological drought into an agricultural drought, see Box 3-3). These aspects, together with seasonal variations in
climate extremes, are not as frequently examined in climate models or observational analyses, and thus can only be partly assessed
within this chapter.
As noted in the discussion of ‘extreme weather or climate events’ in Section 3.1.2, thresholds, percentiles, or return values used for the
definition of climate extremes are generally defined with respect to a given reference period (generally historical, i.e., 1961-1990, but
possibly also based on climate model data). In some cases, a transient baseline can also be considered (i.e., the baseline uses data from
the period under examination and changes as the period being considered changes, rather than using a standard period such as
1961-1990). The choice of the reference period may be relevant for the magnitude of the assessed changes as highlighted, for example,
in Lorenz et al. (2010). The choice of the reference period (static or transient) could also affect the assessment of the respective role of
changes in mean versus changes in variability for changes in extremes discussed in Section 3.1.6. If extremes are based on the probability
distribution from which they are drawn, then a simple change in the mean (and keeping the same distribution) would, strictly speaking,
produce no relative change in extremes at all. The question of the choice of an appropriate reference period is tied to the notion of
adaptation. Events that are considered extreme nowadays in some regions could possibly be adapted to if the vulnerability and exposure
to these extremes is reduced (Chapters 1, 2, and 4 through 7). However, there are also some limits to adaptation as highlighted in
Chapter 8. These considerations are difficult to include in the statistical analyses of climate scenarios because of the number of (mostly
non-physical) aspects that would need to be taken into account.
To conclude, there is no precise definition of an extreme (e.g., D.B. Stephenson et al., 2008). In particular, we note limitations in the
definition of both probability-based or threshold-based climate extremes and their relations to impacts, which apply independently of
the chosen method of analysis:
An event from the extreme tails of probability distributions is not necessarily extreme in terms of impact.
Impact-related thresholds can vary in space and time, that is, single absolute thresholds (e.g., a daily rainfall exceeding 25 mm or
the number of frost days) will not reflect extremes in all locations and time periods (e.g., season, decade).
As an illustration, projected patterns (in the magnitude but not the sign) of changes in annual heat wave length were shown to be highly
dependent on the choice of index used for the assessment of heat wave or warm spell duration (using the mean and maximum Heat
Wave Duration Indices, HWDImean and HWDImax, and the Warm Spell Duration Index, WSDI; see Orlowsky and Seneviratne, 2011),
because of large geographical variations in the variability of daily temperature (Alexander et al., 2006). Similar definition issues apply to
other types of extremes, especially those characterizing dryness (see Section 3.5.1 and Box 3-3).
118
threshold perspective (e.g., a 40°C threshold for midday temperature in
the mid-latitudes).
In the scientific literature, several aspects are considered in the definition
and analysis of climate extremes (Box 3-1).
3.1.3. Compound (Multiple) Events
In climate science, compound events can be (1) two or more extreme
events occurring simultaneously or successively, (2) combinations of
extreme events with underlying conditions that amplify the impact of
the events, or (3) combinations of events that are not themselves
extremes but lead to an extreme event or impact when combined. The
contributing events can be of similar (clustered multiple events) or
different type(s). There are several varieties of clustered multiple events,
such as tropical cyclones generated a few days apart with the same
path and/or intensities, which may occur if there is a tendency for
persistence in atmospheric circulation and genesis conditions. Examples
of compound events resulting from events of different types are
varied – for instance, high sea level coinciding with tropical cyclone
landfall (Section 3.4.4), or cold and dry conditions (e.g., the Mongolian
Dzud, see Case Study 9.2.4), or the impact of hot events and droughts
on wildfire (Case Study 9.2.2), or a combined risk of flooding from sea
level surges and precipitation-induced high river discharge (Svensson
and Jones, 2002; Van den Brink et al., 2005). Compound events can even
result from ‘contrasting extremes’, for example, the projected occurrence
of both droughts and heavy precipitation events in future climate in
some regions (Table 3-3).
Impacts on the physical environment (Section 3.5) are often the result
of compound events. For instance, floods will more likely occur over
saturated soils (Section 3.5.2), which means that both soil moisture
status and precipitation intensity play a role. The wet soil may itself be
the result of a number of above-average but not necessarily extreme
precipitation events, or of enhanced snow melt associated with
temperature anomalies in a given season. Similarly, droughts are the
result of pre-existing soil moisture deficits and of the accumulation of
precipitation deficits and/or evapotranspiration excesses (Box 3-3), not
all (or none) of which are necessarily extreme for a particular drought
event when considered in isolation. Also, impacts on human systems or
ecosystems (Chapter 4) can be the results of compound events, for
example, in the case of health-related impacts associated with combined
temperature and humidity conditions (Box 3-1).
Although compound events can involve causally unrelated events, the
following causes may lead to a correlation between the occurrence of
extremes (or their impacts):
1) A common external forcing factor for changing the probability of
the two events (e.g., regional warming, change in frequency or
intensity of El Niño events)
2) Mutual reinforcement of one event by the other and vice versa due
to system feedbacks (Section 3.1.4)
3) Conditional dependence of the occurrence or impact of one event
on the occurrence of another event (e.g., extreme soil moisture levels
and precipitation conditions for floods, droughts, see above).
Changes in one or more of these factors would be required for a changing
climate to induce changes in the occurrence of compound events.
Unfortunately, investigation of possible changes in these factors has
received little attention. Also, much of the analysis of changes of
extremes has, up to now, focused on individual extremes of a single
variable. However, recent literature in climate research is starting to
consider compound events and explore appropriate methods for their
analysis (e.g., Coles, 2001; Beirlant et al., 2004; Benestad and Haugen,
2007; Renard and Lang, 2007; Schölzel and Friederichs, 2008; Beniston,
2009b; Tebaldi and Sanso, 2009; Durante and Salvadori, 2010).
3.1.4. Feedbacks
A special case of compound events is related to the presence of feedbacks
within the climate system, that is, mutual interaction between several
climate processes, which can either lead to a damping (negative feedback)
or enhancement (positive feedback) of the initial response to a given
forcing (see also ‘climate feedback’ in the Glossary). Feedbacks can play
an important role in the development of extreme events, and in some
cases two (or more) climate extremes can mutually strengthen one
another. One example of positive feedback between two extremes is the
possible mutual enhancement of droughts and heat waves in transitional
regions between dry and wet climates. This feedback has been identified
as having an influence on projected changes in temperature variability
and heat wave occurrence in Central and Eastern Europe and the
Mediterranean (Seneviratne et al., 2006a; Diffenbaugh et al., 2007), and
possibly also in Britain, Eastern North America, the Amazon, and East
Asia (Brabson et al., 2005; Clark et al., 2006). Further results also suggest
that it is a relevant factor for past heat waves and temperature
extremes in Europe and the United States (Durre et al., 2000; Fischer et
al., 2007a,b; Hirschi et al., 2011). Two main mechanisms that have been
suggested to underlie this feedback are: (1) enhanced soil drying during
heat waves due to increased evapotranspiration (as a consequence of
higher vapor pressure deficit and higher incoming radiation); and (2)
higher relative heating of the air from sensible heat flux when soil
moisture deficit starts limiting evapotranspiration/latent heat flux (e.g.,
Seneviratne et al., 2010). Additionally, there may also be indirect and/or
non-local effects of dryness on heat waves through, for example,
changes in circulation patterns or dry air advection (e.g., Fischer et al.,
2007a; Vautard et al., 2007; Haarsma et al., 2009). However, the
strength of these feedbacks is still uncertain in current climate models
(e.g., Clark et al., 2010), in particular if additional feedbacks with
precipitation (e.g., Koster et al., 2004b; Seneviratne et al., 2010) and
with land use and land cover state and changes (e.g., Lobell et al., 2008;
Pitman et al., 2009; Teuling et al., 2010) are considered. Also, feedbacks
between trends in snow cover and changes in temperature extremes have
been highlighted as being relevant for projections (e.g., Kharin et al.,
2007; Orlowsky and Seneviratne, 2011). Feedbacks with soil moisture
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
119
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Likely anthropogenic influence on
trends in warm/cold days/nights at
the global scale. No attribution of
trends at a regional scale with a
few exceptions.
Likely statistically significant increases in the number
of heavy precipitation events (e.g., 95th percentile) in
more regions than those with statistically significant
decreases, but strong regional and subregional
variations in the trends. [Regional details in Table 3-2]
Medium confidence that
anthropogenic influences have
contributed to intensification of
extreme precipitation at the global
scale.
Likely increase in frequency of heavy precipitation
events or increase in proportion of total rainfall from
heavy falls over many areas of the globe, in particular
in the high latitudes and tropical regions, and in
winter in the northern mid-latitudes. [Regional details
in Table 3-3]
Low confidence in trends due to insufficient evidence. Low confidence in the causes of
trends due to insufficient evidence.
Low confidence in projections of extreme winds (with
the exception of wind extremes associated with
tropical cyclones).
Low confidence in trends because of insufficient
evidence.
Low confidence due to insufficient
evidence.
Low confidence in projected changes in monsoons,
because of insufficient agreement between climate
models.
Medium confidence in past trends toward more
frequent central equatorial Pacific El Niño-Southern
Oscillation (ENSO) events.
Insufficient evidence for more specific statements on
ENSO trends.
Likely trends in Southern Annular Mode (SAM).
Likely anthropogenic influence on
identified trends in SAM.
1
Anthropogenic influence on trends
in North Atlantic Oscillation (NAO)
are about as likely as not. No
attribution of changes in ENSO.
Low confidence in projections of changes in behavior
of ENSO and other modes of variability because of
insufficient agreement of model projections.
Low confidence that any observed long-term (i.e., 40
years or more) increases in tropical cyclone activity are
robust, after accounting for past changes in observing
capabilities.
Low confidence in attribution of
any detectable changes in tropical
cyclone activity to anthropogenic
influences (due to uncertainties in
historical tropical cyclones record,
incomplete understanding of
physical mechanisms, and degree
of tropical cyclone variability).
Likely decrease or no change in frequency of tropical
cyclones.
Likely increase in mean maximum wind speed, but
possibly not in all basins.
Likely increase in heavy rainfall associated with
tropical cyclones.
Likely poleward shift in extratropical cyclones.
Low confidence in regional changes in intensity.
Medium confidence in an
anthropogenic influence on
poleward shift.
Likely impacts on regional cyclone activity but low
confidence in detailed regional projections due to only
partial representation of relevant processes in current
models.
Medium confidence in a reduction in the numbers of
mid-latitude storms.
Medium confidence in projected poleward shift of
mid-latitude storm tracks.
Medium confidence that some regions of the world
have experienced more intense and longer droughts,
in particular in southern Europe and West Africa, but
opposite trends also exist. [Regional details in Table
3-2]
Medium confidence that
anthropogenic influence has
contributed to some observed
changes in drought patterns.
Low confidence in attribution of
changes in drought at the level of
single regions due to inconsistent
or insufficient evidence.
Medium confidence in projected increase in duration
and intensity of droughts in some regions of the
world, including southern Europe and the
Mediterranean region, central Europe, central North
America, Central America and Mexico, northeast
Brazil, and southern Africa.
Overall low confidence elsewhere because of
insufficient agreement of projections.
[Regional details in Table 3-3]
Limited to medium evidence available to assess
climate-driven observed changes in the magnitude
and frequency of floods at regional scale.
Furthermore, there is low agreement in this evidence,
and thus overall low confidence
at the global scale
regarding even the sign of these changes.
High confidence in trend toward earlier occurrence of
spring peak river flows in snowmelt- and glacier-fed
rivers.
Low confidence that anthropogenic
warming has affected the
magnitude or frequency of floods at
a global scale.
Medium confidence to high
confidence in anthropogenic
influence on changes in some
components of the water cycle
(precipitation, snowmelt) affecting
floods.
Low confidence in global projections of changes in
flood magnitude and frequency because of insufficient
evidence.
Medium confidence (based on physical reasoning)
that projected increases in heavy precipitation would
contribute to rain-generated local flooding in some
catchments or regions.
Very likely earlier spring peak flows in snowmelt- and
glacier-fed rivers.
Virtually certain decrease in frequency and magnitude
of unusually cold days and nights at the global scale.
Virtually certain increase in frequency and magnitude
of unusually warm days and nights at the global scale.
Very likely increase in length, frequency, and/or
intensity of warm spells or heat waves over most land
areas. [Regional details in Table 3-3]
El Niño and
other Modes of
Variability
(Sections 3.4.2
and 3.4.3)
Monsoons
(Section 3.4.1)
Winds
(Section 3.3.3)
Precipitation
(Section 3.3.2)
Temperature
(Section 3.3.1)
Tropical
Cyclones
(Section 3.4.4)
Extratropical
Cyclones
(Section 3.4.5)
Droughts
(Section 3.5.1)
Floods
(Section 3.5.2)
Very likely decrease in number of unusually cold days
and nights at the global scale. Very likely increase in
number of unusually warm days and nights at the
global scale. Medium confidence in increase in length
or number of warm spells or heat waves in many (but
not all) regions. Low or medium confidence in trends in
temperature extremes in some subregions due either
to lack of observations or varying signal within
subregions. [Regional details in Table 3-2]
Continued next page
Observed Changes (since 1950)
Attribution of Observed
Changes
Projected Changes (up to 2100) with
Respect to Late 20th Century
Weather
and
Climate
Variables
Phenomena
Related to
Weather and
Climate
Extremes
Impacts on
Physical
Environment
Table 3-1 | Overview of considered extremes and summary of observed and projected changes at a global scale. Regional details on observed and projected changes in temperature
and precipitation extremes are provided in Tables 3-2 and 3-3. Extremes (e.g., cold/warm days/nights, heat waves, heavy precipitation events) are defined with respect to late 20th-
century climate (see also Box 3-1 for discussion of reference period).
120
and snow affect extremes in specific regions (hot extremes in transitional
climate regions, and cold extremes in snow-covered regions), where they
may induce significant deviations in changes in extremes versus changes
in the average climate, as also discussed in Section 3.1.6. Other relevant
feedbacks involving extreme events are those that can lead to impacts
on the global climate, such as modification of land carbon uptake due
to enhanced drought occurrence (e.g., Ciais et al., 2005; Friedlingstein et
al., 2006; Reichstein et al., 2007) or carbon release due to permafrost
degradation (see Section 3.5.7). These aspects are not, however,
specifically considered in this chapter (but see Section 3.1.7, on
projections of possible increased Amazon drought and forest dieback in
this region). Chapter 4 also addresses feedback loops between
droughts, fire, and climate change (Section 4.2.2.1).
3.1.5. Confidence and Likelihood
of Assessed Changes in Extremes
In this chapter, all assessments regarding past or projected changes in
extremes are expressed following the new IPCC Fifth Assessment Report
uncertainty guidance (Mastrandrea et al., 2010). The new uncertainty
guidance makes a clearer distinction between confidence and likelihood
(see Box SPM.2). Its use complicates comparisons between assessments
in this chapter and those in the IPCC Fourth Assessment Report (AR4), as
they are not directly equivalent in terms of nomenclature. The following
procedure was adopted in this chapter (see in particular the Executive
Summary and Tables 3-1, 3-2, and 3-3.):
For each assessment, the confidence level for the given assessment
is first assessed (low, medium, or high), as discussed in the next
paragraph.
For assessments with high confidence, likelihood assessments of a
direction of change are also provided (virtually certain for 99-100%,
very likely for 90-100%, likely for 66-100%, more likely than not
for 50-100%, about as likely as not for 33-66%, unlikely for 0-33%,
very unlikely for 0-10%, and exceptionally unlikely for 0-1%). In
a few cases for which there is high confidence (e.g., based on
physical understanding) but for which there are not sufficient
model projections to provide a more detailed likelihood assessment
(such as likely’), only the confidence assessment is provided.
For assessments with medium confidence, a direction of change is
provided, but without an assessment of likelihood.
For assessments with low confidence, no direction of change is
generally provided.
The confidence assessments are expert-based evaluations that consider
the confidence in the tools and data basis (models, data, proxies) used
to assess or project changes in a specific element, and the associated
level of understanding. Examples of cases of low confidence for model
projections are if models display poor performance in simulating the
specific extreme in the present climate (see also Box 3-2), or if insufficient
literature on model performance is available for the specific extreme, for
example, due to lack of observations. Similarly for observed changes,
the assessment may be of low confidence if the available evidence is
based only on scattered data (or publications) that are insufficient to
provide a robust assessment for a large region, or the observations may
be of poor quality, not homogeneous, or only of an indirect nature
(proxies). In cases with low confidence regarding past or projected
changes in some extremes, we indicate whether the low confidence is
due to lack of literature, lack of evidence (data, observations), or lack of
understanding. It should be noted that there are some overlaps
between these categories, as for instance a lack of evidence can be at
the root of a lack of literature and understanding. Cases of changes in
extremes for which confidence in the models and data is rated as
medium’ are those where we have some confidence in the tools and
evidence available to us, but there remain substantial doubts about
some aspects of the quality of these tools. It should be noted that an
assessment of low confidence in observed or projected changes or
trends in a specific extreme neither implies nor excludes the possibility
of changes in this extreme. Rather the assessment indicates low
confidence in the ability to detect or project any such changes.
Changes (observed or projected) in some extremes are easier to assess
than in others either due to the complexity of the underlying processes
or to the amount of evidence available for their understanding. This
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Likely increase in extreme coastal high water
worldwide related to increases in mean sea level in
the late 20th century.
Likely anthropogenic influence via
mean sea level contributions.
Very likely that mean sea level rise will contribute to
upward trends in extreme coastal high water levels.
High confidence that locations currently experiencing
coastal erosion and inundation will continue to do so
due to increasing sea level, in the absence of changes
in other contributing factors.
Low confidence in global trends in large landslides in
some regions. Likely increased thawing of permafrost
with likely resultant physical impacts.
Likely anthropogenic influence on
thawing of permafrost.
Low confidence of other
anthropogenic influences because
of insufficient evidence for trends in
other physical impacts in cold
regions.
High confidence that changes in heat waves, glacial
retreat, and/or permafrost degradation will affect high
mountain phenomena such as slope instabilities, mass
movements, and glacial lake outburst floods. High
confidence that changes in heavy precipitation will
affect landslides in some regions.
Low confidence in projected future changes in dust
activity.
Other Physical
Impacts
(Sections 3.5.6,
3.5.7, and 3.5.8)
Extreme Sea
Level and
Coastal Impacts
(Sections 3.5.3,
3.5.4, and 3.5.5)
Observed Changes (since 1950)
Attribution of Observed
Changes
Projected Changes (up to 2100) with
Respect to Late 20th Century
Impacts on
Physical
Environment
(Continued)
Notes: 1. Due to trends in stratospheric ozone concentrations.
Table 3-1 (continued)
121
results in differing levels of uncertainty in climate simulations and
projections for different extremes (Box 3-2). Because of these issues,
projections in some extremes are difficult or even impossible to provide,
although projections in some other extremes have a high level of
confidence. In addition, uncertainty in projections also varies over different
time frames for individual extremes, because of varying contributions
over time of internal climate variability, model uncertainty, and emission
scenario uncertainty to the overall uncertainty (Box 3-2 and Section 3.2).
Overall, we can infer that our confidence in past and future changes in
extremes varies with the type of extreme, the data available, and the
region, season, and time frame being considered, linked with the level
of understanding and reliability of simulation of the underlying physical
processes. These various aspects are addressed in more detail in Box 3-2,
Section 3.2, and the subsections on specific extremes in Sections 3.3-3.5.
3.1.6. Changes in Extremes and Their Relationship
to Changes in Regional and Global Mean Climate
Changes in extremes can be linked to changes in the mean, variance, or
shape of probability distributions, or all of these (see, e.g., Figure 1-2).
Thus a change in the frequency of occurrence of hot days (i.e., days
above a certain threshold) can arise from a change in the mean daily
maximum temperature, and/or from a change in the variance and/or
shape of the frequency distribution of daily maximum temperatures. If
changes in the frequency of occurrence of hot days were mainly linked
to changes in the mean daily maximum temperature, and changes in the
shape and variability of the distribution of daily maximum temperatures
were of secondary importance, then it might be reasonable to use
projected changes in mean temperature to estimate how changes in
extreme temperatures might change in the future. If, however, changes in
the shape and variability of the frequency distribution of daily maximum
temperature were important, such naive extrapolation would be less
appropriate or possibly even misleading (e.g., Ballester et al., 2010). The
results of both empirical and model studies indicate that although in
several situations changes in extremes do scale closely with changes in
the mean (e.g., Griffiths et al., 2005), there are sufficient exceptions from
this that changes in the variability and shape of probability distributions
of weather and climate variables need to be considered as well as
changes in means, if we are to project future changes in extremes (e.g.,
Hegerl et al., 2004; Schär et al., 2004; Caesar et al., 2006; Clark et al.,
2006; Della-Marta et al., 2007a; Kharin et al., 2007; Brown et al., 2008;
Ballester et al., 2010; Orlowsky and Seneviratne, 2011). This appears to be
especially the case for short-duration precipitation, and for temperatures
in mid- and high latitudes (but not all locations in these regions). In mid-
and high latitudes stronger increases (or decreases) in some extremes
are generally associated with feedbacks with soil moisture or snow
cover (Section 3.1.4). Note that the respective importance of changes in
mean versus changes in variability also depends on the choice of the
reference period used to define the extremes (Box 3-1).
An additional relevant question is the extent to which regional changes
in extremes scale with changes in global mean climate. Indeed, recent
publications and the public debate have focused, for example, on global
mean temperature targets (e.g., Allen et al., 2009; Meinshausen et al.,
2009), however, the exact implications of these mean global changes
(e.g., ‘2°C target’) for regional extremes have not been widely assessed
(e.g., Clark et al., 2010). Orlowsky and Seneviratne (2011) investigated
the scaling between projected changes in the 10th and 90th percentile
of Tmax on annual and seasonal (June-July-August: JJA, December-
January-February: DJF) time scales with globally averaged annual mean
changes in Tmax based on the whole CMIP3 ensemble (see Section 3.2.3
for discussion of the CMIP3 ensemble). The results highlight particularly
large projected changes in the 10th percentile Tmax in the northern
high-latitude regions in winter and the 90th percentile Tmax in
Southern Europe in summer with scaling factors of about 2 in both
cases (i.e., increases of about 4°C for a mean global increase of 2°C).
However, in some regions and seasons, the scaling can also be below 1
(e.g., changes in 10th percentile in JJA in the high latitudes). This is also
illustrated in Figure 3-5a, which compares analyses of changes in return
values of annual extremes of maximum daily temperatures for the overall
land and specific regions, and shows high region-to-region variability in
these changes. The changes in return values at the global scale (‘Globe
(Land only)’) for their part are almost identical to the changes in global
mean daily maximum temperature, suggesting that the scaling issues are
related to regional effects rather than overall differences in the changes
in the tails versus the means of the distributions of daily maximum
temperature. The situation is very different for precipitation (Figure 3-7a),
with clearly distinct behavior between changes in mean and extreme
precipitation at the global scale, highlighting the dependency of any
scaling on the variable being considered. The lack of consistent scaling
between regional and seasonal changes in extremes and changes in
means has also been highlighted in empirical studies (e.g., Caesar et al.,
2006). It should further be noted that not only do regional extremes not
necessarily scale with global mean changes, but also mean global
warming does not exclude the possibility of cooling in some regions and
seasons, both in the recent past and in the coming decades: it has for
instance been recently suggested that the decrease in sea ice caused by
the mean warming could induce, although not systematically, more
frequent cold winter extremes over northern continents (Petoukhov and
Semenov, 2010). Also parts of central North America and the eastern
United States present cooling trends in mean temperature and some
temperature extremes in the spring to summer season in recent decades
(Section 3.3.1). It should be noted that, independently of scaling issues
for the means and extremes of the same variable, some extremes can
be related to mean climate changes in other variables, such as links
between mean global changes in relative humidity and some regional
changes in heavy precipitation events (Sections 3.2.2.1 and 3.3.2).
Global-scale trends in a specific extreme may be either more reliable or less
reliable than some regional-scale trends, depending on the geographical
uniformity of the trends in the specific extreme. In particular, climate
projections for some variables are not consistent, even in the sign of the
projected change, everywhere across the globe (e.g., Christensen et al.,
2007; Meehl et al., 2007b). For instance, projections typically include
some regions with a tendency toward wetter conditions and others with
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
122
a tendency toward drier conditions, with some regions displaying a shift
in climate regimes (e.g., from humid to transitional or transitional to dry).
Some of these regional changes will depend on how forcing changes may
alter the regional atmospheric circulation, especially in coastal regions
and regions with substantial orography. Hence for certain extremes such
as floods and droughts, regional projections might indicate larger
changes than is the case for projections of global averages (which
would average the regional signals exhibiting changes of opposite
signs). This also means that signals at the regional scale may be more
reliable (and meaningful) in some cases than assessments at the global
scale. On the other hand, temperature extremes projections, which are
consistent across most regions, are thus more reliable at the global
scale (‘virtually certain’) than at the regional scale (at most very likely’).
3.1.7. Surprises / Abrupt Climate Change
This report focuses on the most probable changes in extremes based on
current knowledge. However, the possible future occurrence of low-
probability, high-impact scenarios associated with the crossing of poorly
understood climate thresholds cannot be excluded, given the transient
and complex nature of the climate system. Such scenarios have important
implications for society as highlighted in Section 8.5.1. So, an assessment
that we have low confidence in projections of a specific extreme, or even
lack of consideration of given climate changes under the categories
covered in this chapter (e.g., shutdown of the meridional overturning
circulation), should not be interpreted as meaning that no change is
expected in this extreme or climate element (see also Section 3.1.5).
Feedbacks play an important role in either damping or enhancing
extremes in several climate variables (Section 3.1.4), and this can also
lead to ‘surprises,’ that is, changes in extremes greater (or less) than
might be expected with a gradual warming of the climate system.
Similarly, as discussed in 3.1.3, contrasting or multiple extremes can
occur but our understanding of these is insufficient to provide credible
comprehensive projections of risks associated with such combinations.
One aspect that we do not address in this chapter is the existence of
possible tipping points in the climate system (e.g., Meehl et al., 2007b;
Lenton et al., 2008; Scheffer et al., 2009), that is, the risks of abrupt,
possibly irreversible changes in the climate system. Abrupt climate
change is defined as follows in the Glossary: “The nonlinearity of the
climate system may lead to abrupt climate change, sometimes called
rapid climate change, abrupt events, or even surprises. The term abrupt
often refers to time scales faster than the typical time scale of the
responsible forcing. However, not all abrupt climate changes need be
externally forced. Some changes may be truly unexpected, resulting
from a strong, rapidly changing forcing of a nonlinear system.
Thresholds associated with tipping points may be termed ‘critical
thresholds,’ or, in the case of the climate system, ‘climate thresholds’.
Scheffer et al. (2009) illustrate the possible equilibrium responses of a
system to forcing. In the case of a linear response, only a large forcing
can lead to a major state change in the system. However, in the presence
of a critical threshold even a small change in forcing can lead to a similar
major change in the system. For systems with critical bifurcations in the
equilibrium state function two alternative stable conditions may exist,
whereby an induced change may be irreversible. Such critical transitions
within the climate system represent typical low-probability, high-impact
scenarios, which were also noted in the AR4 (Meehl et al., 2007b).
Lenton et al. (2008) provided a recent review on potential tipping elements
within the climate system, that is, subsystems of the Earth system that
are at least subcontinental in scale and which may entail a tipping
point. Some of these would be especially relevant to certain extremes
[e.g., El Niño-Southern Oscillation (ENSO), the Indian summer monsoon,
and the Sahara/Sahel and West African monsoon for drought and heavy
precipitation, and the Greenland and West Antarctic ice sheets for sea
level extremes], or are induced by changes in extremes (e.g., Amazon
rainforest die-back induced by drought). For some of the identified
tipping elements, the existence of bistability has been suggested by
paleoclimate records, but is still debated in some cases (e.g., Brovkin et
al., 2009). There is often a lack of agreement between models regarding
these low-probability, high-impact scenarios, for instance, regarding a
possible increased drought and consequent die-back of the Amazon
rainforest (e.g., Friedlingstein et al., 2006; Poulter et al., 2010; see
Table 3-3 for dryness projections in this region), the risk of an actual
shutdown of the Atlantic thermohaline circulation (e.g., Rahmstorf et
al., 2005; Lenton et al., 2008), or the potential irreversibility of the
decrease in Arctic sea ice (Tietsche et al., 2011). For this reason,
confidence in these scenarios is assessed as low.
3.2. Requirements and Methods
for Analyzing Changes in Extremes
3.2.1. Observed Changes
Sections 3.3 to 3.5 of this chapter provide assessments of the literature
regarding changes in extremes in the observed record published mainly
since the AR4 and building on the AR4 assessment. Summaries of these
assessments are provided in Table 3-1. Overviews of observed regional
changes in temperature and precipitation extremes are provided in
Table 3-2. In this section issues are discussed related to the data and
observations used to examine observed changes in extremes.
Issues with data availability are especially critical when examining
changes in extremes of given climate variables (Nicholls, 1995). Indeed,
the more rare the event, the more difficult it is to identify long-term
changes, simply because there are fewer cases to evaluate (Frei and
Schär, 2001; Klein Tank and Können, 2003). Identification of changes in
extremes is also dependent on the analysis technique employed (X.
Zhang et al., 2004; Trömel and Schönwiese, 2005). Another important
criterion constraining data availability for the analysis of extremes is the
respective time scale on which they occur (Section 3.1.2), since this
determines the required temporal resolution for their assessment (e.g.,
heavy hourly or daily precipitation versus multi-year drought). Longer
time resolution data (e.g., monthly, seasonal, and annual values) for
temperature and precipitation are available for most parts of the world
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
123
starting late in the 19th to early 20th century, and allow analysis of
meteorological drought (see Box 3-3) and unusually wet periods of the
order of a month or longer. To examine changes in extremes occurring on
short time scales, particularly of climate variables such as temperature
and precipitation (or wind), normally requires the use of high-temporal
resolution data, such as daily or sub-daily observations, which are
generally either not available, or available only since the middle of the
20th century and in many regions only from as recently as 1970. Even
where sufficient data are available, several problems can still limit their
analysis. First, although the situation is changing (especially for the
situation with respect to ‘extreme indices,’ Box 3-1), many countries still
do not freely distribute their higher temporal resolution data. Second,
there can be issues with the quality of measurements. A third important
issue is climate data homogeneity (see below). These and other issues
are discussed in detail in the AR4 (Trenberth et al., 2007). For instance,
the temperature and precipitation stations considered in the daily data
set used in Alexander et al. (2006) are not globally uniform. Although
observations for most parts of the globe are available, measurements
are lacking in Northern South America, Africa, and part of Australia. The
other data set commonly used for extremes analyses is from Caesar et al.
(2006; used, e.g., in Brown et al., 2008), which also has data gaps in
most of South America, Africa, Eastern Europe, Mexico, the Middle East,
India, and Southeast Asia. Also the study by Vose et al. (2005) has data
gaps in South America, Africa, and India. It should be further noted
that the regions with data coverage do not all have the same density of
stations (Alexander et al., 2006; Caesar et al., 2006). While some studies
are available on a country or regional basis for areas not covered in
global studies (see, e.g., Tables 3-2 and 3-3), lack of data in many parts
of the globe leads to limitations in our ability to assess observed
changes in climate extremes for many regions.
Whether or not climate data are homogeneous is of clear relevance for
an analysis of extremes, especially at smaller spatial scales. Data are
defined as homogeneous when the variations and trends in a climate time
series are due solely to variability and changes in the climate system. Some
meteorological elements are especially vulnerable to uncertainties caused
by even small changes in the exposure of the measuring equipment. For
instance, erection of a small building or changes in vegetative cover near
the measuring equipment can produce a bias in wind measurements
(Wan et al., 2010). When a change occurs it can result in either a
discontinuity in the time series (step change) or a more gradual change
that can manifest itself as a false trend (Menne and Williams Jr., 2009),
both of which can impact on whether a particular observation exceeds
a threshold. Homogeneity detection and data adjustments have been
implemented for longer averaging periods (e.g., monthly, seasonal,
annual); however, techniques applicable to shorter observing periods
(e.g., daily) data have only recently been developed (e.g., Vincent et
al., 2002; Della-Marta and Wanner, 2006), and have not been widely
implemented. Homogeneity issues also affect the monitoring of other
meteorological and climate variables, for which further and more severe
limitations also can exist. This is in particular the case regarding
measurements of wind and relative humidity, and data required for the
analysis of weather and climate phenomena (tornadoes, extratropical
and tropical cyclones; Sections 3.3.3, 3.4.4, and 3.4.5), as well as
impacts on the physical environment (e.g., droughts, floods, cryosphere
impacts; Section 3.5).
Thunderstorms, tornadoes, and related phenomena are not well
observed in many parts of the world. Tornado occurrence since 1950 in
the United States, for instance, displays an increasing trend that mainly
reflects increased population density and increased numbers of people
in remote areas (Trenberth et al., 2007; Kunkel et al., 2008). Such trends
increase the likelihood that a tornado would be observed. A similar
problem occurs with thunderstorms. Changes in reporting practices,
increased population density, and even changes in the ambient noise
level at an observing station all have led to inconsistencies in the
observed record of thunderstorms.
Studies examining changes in extratropical cyclones, which focus on
changes in storm track location, intensities, and frequency, are limited
in time due to a lack of suitable data prior to about 1950. Most of these
studies have relied on model-based reanalyses that also incorporate
observations into a hybrid model-observational data set. However,
reanalyses can have homogeneity problems due to changes in the
amount and type of data being assimilated, such as the introduction of
satellite data in the late 1970s and other observing system changes
(Trenberth et al., 2001; Bengtsson et al., 2004). Recent reanalysis efforts
have attempted to produce more homogeneous reanalyses that show
promise for examining changes in extratropical cyclones and other climate
features (Compo et al., 2006). Results, however, are strongly dependent
on the reanalysis and cyclone tracking techniques used (Ulbrich et al.,
2009).
The robustness of analyses of observed changes in tropical cyclones has
been hampered by a number of issues with the historical record. One of
the major issues is the heterogeneity introduced by changing technology
and reporting protocols within the responsible agencies (e.g., Landsea
et al., 2004). Further heterogeneity is introduced when records from
multiple ocean basins are combined to explore global trends, because data
quality and reporting protocols vary substantially between agencies (Knapp
and Kruk, 2010). Much like other weather and climate observations,
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
2
18
24
7
17
3
6
26
22
9
15
5
1
10
23
25
14
4
11
16
13
19
8
21
12
20
Figure 3-1 | Definitions of regions used in Tables 3-2 and 3-3, and Figures 3-5 and 3-7.
Exact coordinates of the regions are provided in the on-line supplement, Appendix 3.A.
Assessments and analyses are provided for land areas only.
124
tropical cyclone observations are taken to support short-term forecasting
needs. Improvements in observing techniques are often implemented
without any overlap or calibration against existing methods to document
the impact of the changes on the climate record. Additionally, advances
in technology have enabled better and more complete observations. For
example, the introduction of aircraft reconnaissance in some basins in
the 1940s and satellite data in the 1960s had a profound effect on our
ability to accurately identify and measure tropical cyclones, particularly
those that never encountered land or a ship. While aircraft reconnaissance
programs have continued in the North Atlantic, they were terminated in
the Western Pacific in 1987. The introduction of geostationary satellite
imagery in the 1970s, and the introduction (and subsequent improvement)
of new tropical cyclone analysis methods (such as the Dvorak technique
for estimating storm intensity), further compromises the homogeneity
of historical records of tropical cyclone activity.
Regarding impacts to the physical environment, soil moisture is a key
variable for which data sets are extremely scarce (e.g., Robock et al.,
2000; Seneviratne et al., 2010). This represents a critical issue for the
validation and correct representation of soil moisture (agricultural) as
well as hydrological drought (Box 3-3) in climate, land surface, and
hydrological models, and the monitoring of ongoing changes in regional
terrestrial water storage. As a consequence, these need to be inferred
from simple climate indices or model-based approaches (Box 3-3). Such
estimates rely in large part on precipitation observations, which have,
however, inadequate spatial coverage for these applications in many
regions of the world (e.g., Oki et al., 1999; Fekete et al., 2004; Koster et
al., 2004a). Similarly, runoff observations are not globally available,
which results in significant uncertainties in the closing of the global and
some regional water budgets (Legates et al., 2005; Peel and McMahon,
2006; Dai et al., 2009; Teuling et al., 2009), as well as for the global
analysis of changes in the occurrence of floods (Section 3.5.2).
Additionally, ground observations of snow, which are lacking in several
regions, are important for the investigation of physical impacts,
particularly those related to the cryosphere and runoff generation (e.g.,
Essery et al., 2009; Rott et al., 2010).
All of the above-mentioned issues lead to uncertainties in observed
trends in extremes. In many instances, great care has been taken to
develop procedures to reduce the confounding influences of these
issues on the data, which in turn helps to reduce uncertainty, and
progress has been made in the last 15 years (e.g., Caesar et al., 2006;
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
FAQ 3.1 | Is the Climate Becoming More Extreme?
While there is evidence that increases in greenhouse gases have likely caused changes in some types of extremes, there is no simple
answer to the question of whether the climate, in general, has become more or less extreme. Both the terms ‘more extreme’ and ‘less
extreme’ can be defined in different ways, resulting in different characterizations of observed changes in extremes. Additionally, from a
physical climate science perspective it is difficult to devise a comprehensive metric that encompasses all aspects of extreme behavior in
the climate.
One approach for evaluating whether the climate is becoming more extreme would be to determine whether there have been changes
in the typical range of variation of specific climate variables. For example, if there was evidence that temperature variations in a given
region had become significantly larger than in the past, then it would be reasonable to conclude that temperatures in that region had
become more extreme. More simply, temperature variations might be considered to be becoming more extreme if the difference
between the highest and the lowest temperature observed in a year is increasing. According to this approach, daily temperature over
the globe may have become less extreme because there have generally been greater increases in mean daily minimum temperatures
globally than in mean daily maximum temperatures, over the second half of the 20th century. On the other hand, one might conclude
that daily precipitation has become more extreme because observations suggest that the magnitude of the heaviest precipitation events
has increased in many parts of the world. Another approach would be to ask whether there have been significant changes in the
frequency with which climate variables cross fixed thresholds that have been associated with human or other impacts. For example, an
increase in the mean temperature usually results in an increase in hot extremes and a decrease in cold extremes. Such a shift in the
temperature distribution would not increase the ‘extremeness’ of day-to-day variations in temperature, but would be perceived as
resulting in a more extreme warm temperature climate, and a less extreme cold temperature climate. So the answer to the question
posed here would depend on the variable of interest, and on which specific measure of the extremeness of that variable is examined. As
well, to provide a complete answer to the above question, one would also have to collate not just trends in single variables, but also
indicators of change in complex extreme events resulting from a sequence of individual events, or the simultaneous occurrence of
different types of extremes. So it would be difficult to comprehensively describe the full suite of phenomena of concern, or to find a way
to synthesize all such indicators into a single extremeness metric that could be used to comprehensively assess whether the climate as a
whole has become more extreme from a physical perspective. And to make such a metric useful to more than a specific location, one
would have to combine the results at many locations, each with a different perspective on what is ‘extreme.
Continued next page
125
Brown et al., 2008). As a consequence, more complete and homogenous
information about changes is now available for at least some variables
and regions (Nicholls and Alexander, 2007; Peterson and Manton,
2008). For instance, the development of global databases of daily
temperature and precipitation covering up to 70% of the global land
area has allowed robust analyses of extremes (see Alexander et al.,
2006). In addition, analyses of temperature and precipitation extremes
using higher temporal resolution data, such as that available in the
Global Historical Climatology Network – Daily data set (Durre et al., 2008)
have also proven robust at both a global (Alexander et al., 2006) and
regional scale (Sections 3.3.1 and 3.3.2). Nonetheless, as highlighted
above, for many extremes, data remain sparse and problematic, resulting
in lower ability to establish changes, particularly on a global basis and
for specific regions.
3.2.2. The Causes behind the Changes
This section discusses the main requirements, approaches, and
considerations for the attribution of causes for observed changes in
extremes. In Sections 3.3 to 3.5, the causes of observed changes in
specific extremes are assessed. A global summary of these assessments
is provided in Table 3-1. Climate variations and change are induced by
variability internal to the climate system, and changes in external
forcings, which include natural external forcings such as changes in solar
irradiance and volcanism, and anthropogenic forcings such as aerosol
and greenhouse gas emissions principally due to the burning of fossil
fuels, and land use and land cover changes. The mean state, extremes,
and variability are all related aspects of the climate, so external forcings
that affect the mean climate would in general result in changes in
extremes. For this reason, we provide in Section 3.2.2.1 a brief overview
of human-induced changes in the mean climate to aid the understanding
of changes in extremes as the literature directly addressing the causes
of changes in extremes is quite limited.
3.2.2.1. Human-Induced Changes in the Mean Climate
that Affect Extremes
The occurrence of extremes is usually the result of multiple factors,
which can act either on the large scale or on the regional (and local)
scale (see also Section 3.1.6). Some relevant large-scale impacts of
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Three types of metrics have been considered to avoid these problems, and thereby allow an answer to this question. One approach is to
count the number of record-breaking events in a variable and to examine such a count for any trend. However, one would still face the
problem of what to do if, for instance, hot extremes are setting new records, while cold extremes are not occurring as frequently as in
the past. In such a case, counting the number of records might not indicate whether the climate was becoming more or less extreme,
rather just whether there was a shift in the mean climate. Also, the question of how to combine the numbers of record-breaking events
in various extremes (e.g., daily precipitation and hot temperatures) would need to be considered. Another approach is to combine
indicators of a selection of important extremes into a single index, such as the Climate Extremes Index (CEI), which measures the fraction
of the area of a region or country experiencing extremes in monthly mean surface temperature, daily precipitation, and drought. The CEI,
however, omits many important extremes such as tropical cyclones and tornadoes, and could, therefore, not be considered a complete
index of ‘extremeness.’ Nor does it take into account complex or multiple extremes, nor the varying thresholds that relate extremes to
impacts in various sectors.
A third approach to solving this dilemma arises from the fact that extremes often have deleterious economic consequences. It may
therefore be possible to measure the integrated economic effects of the occurrence of different types of extremes into a common
instrument such as insurance payout to determine if there has been an increase or decrease in that instrument. This approach would
have the value that it clearly takes into account those extremes with economic consequences. But trends in such an instrument will be
dominated by changes in vulnerability and exposure and it will be difficult, if not impossible, to disentangle changes in the instrument
caused by non-climatic changes in vulnerability or exposure in order to leave a residual that reflects only changes in climate extremes.
For example, coastal development can increase the exposure of populations to hurricanes; therefore, an increase in damage in coastal
regions caused by hurricane landfalls will largely reflect changes in exposure and may not be indicative of increased hurricane activity.
Moreover, it may not always be possible to associate impacts such as the loss of human life or damage to an ecosystem due to climate
extremes to a measurable instrument.
None of the above instruments has yet been developed sufficiently as to allow us to confidently answer the question posed here. Thus
we are restricted to questions about whether specific extremes are becoming more or less common, and our confidence in the answers
to such questions, including the direction and magnitude of changes in specific extremes, depends on the type of extreme, as well as
on the region and season, linked with the level of understanding of the underlying processes and the reliability of their simulation in
models.
126
external forcings affecting extremes include net increases in temperature
induced by changes in radiation, enhanced moisture content of the
atmosphere, and increased land-sea contrast in temperatures, which can,
for example, affect circulation patterns and to some extent monsoons.
At regional and local scales, additional processes can modulate the
overall changes in extremes, including regional feedbacks, in particular
linked to land-atmosphere interactions with, for example, soil moisture
or snow (e.g., Section 3.1.4). This section briefly reviews the current
understanding of the causes (i.e., in the sense of attribution to either
external forcing or internal climate variability) of large-scale (and some
regional) changes in the mean climate that are of relevance to extreme
events, to the extent that they have been considered in detection and
attribution studies.
Regarding observed increases in global average annual mean surface
temperatures in the second half of the 20th century, we base our analysis
on the following AR4 assessment (Hegerl et al., 2007): Most of the
observed increase in global average temperatures is very likely due to
the observed increase in anthropogenic greenhouse gas concentrations.
Greenhouse gas forcing alone would likely have resulted in a greater
warming than observed if there had not been an offsetting cooling
effect from aerosol and other forcings. It is extremely unlikely (<5%)
that the global pattern of warming can be explained without external
forcing, and very unlikely that it is due to known natural external causes
alone. Anthropogenically forced warming over the second half of the
20th century has also been detected in ocean heat content and air
temperatures in all continents (Hegerl et al., 2007; Gillett et al., 2008b).
Hegerl et al. (2007) assessed literature that considered detection in
temperature trends at scales as small as approximately 500 km. Recent
work has provided more evidence of detection of an anthropogenic
influence at increasingly smaller spatial scales and for seasonal averages
(Stott et al., 2010). For instance, Min and Hense (2007) found that
estimates of response to anthropogenic forcing from the multi-model
Coupled Model Intercomparison Project 3 (CMIP3) ensemble (see
Section 3.2.3.3) provided a better explanation for observed continental-
scale seasonal temperature changes than alternative explanations such
as natural external forcing or internal variability. In another study, an
anthropogenic signal was detected in 20th-century summer temperatures
in Northern Hemisphere subcontinental regions except central North
America, although the results were more uncertain when anthropogenic
and natural signals were considered together (Jones et al., 2008). An
anthropogenic signal has also been detected in multi-decadal trends
in a US climate extreme index (Burkholder and Karoly, 2007), in the
hydrological cycle of the western United States (Barnett et al., 2008), in
New Zealand temperatures (Dean and Stott, 2009), and in European
temperatures (Christidis et al., 2011a).
Attribution has more stringent demands than those for the detection of
an external influence in observations. Overall, attribution at scales
smaller than continental has still not yet been established primarily due
to the low signal-to-noise ratio and the difficulties of separately
attributing effects of the wider range of possible driving processes
(either attributable to external forcing or internal climate variability) at
these scales (Hegerl et al., 2007). One reason is that averaging over
smaller regions reduces the internal variability less than does averaging
over large regions. In addition, the small-scale details of external forcing,
and the responses simulated by models, are less credible than large-
scale features. For instance, temperature changes are poorly simulated
by models in some regions and seasons (Dean and Stott, 2009; van
Oldenborgh et al., 2009). Also the inclusion of additional forcing factors,
such as land use change and aerosols that can be more important at
regional scales, remains a challenge (Lohmann and Feichter, 2007;
Pitman et al., 2009; Rotstayn et al., 2009).
One of the significant advances since AR4 is emerging evidence of human
influence on global atmospheric moisture content and precipitation.
According to the Clausius-Clapeyron relationship, the saturation vapor
pressure increases approximately exponentially with temperature. It is
physically plausible that relative humidity would remain roughly constant
under climate change (e.g., Hegerl et al., 2007). This means that specific
humidity increases about 7% for a one degree increase in temperature
in the current climate. Indeed, observations indicate significant increases
between 1973 and 2003 in global surface specific humidity but not in
relative humidity (Willett et al., 2008), and at the largest spatial-temporal
scales moistening is close to the Clausius-Clapeyron scaling of the
saturated specific humidity (~7% K
-1
; Willett et al., 2010), though relative
humidity over low- and mid-latitude land areas decreased over a 10-year
period prior to 2008 possibly due to a slower temperature increase in
the oceans than over the land (Simmons et al., 2010). By comparing
observations with model simulations, changes in the global surface
specific humidity for 1973-2003 (Willett et al., 2007), and in lower
tropospheric moisture content over the 1988-2006 period (Santer et al.,
2007) can be attributed to anthropogenic influence.
The increase in the atmospheric moisture content would be expected to
lead to an increase in extreme precipitation when other factors do not
change. Min et al. (2011) detected an anthropogenic influence in annual
maxima of daily precipitation over Northern Hemisphere land areas. The
influence of anthropogenic forcing has been detected in the latitudinal
pattern of land precipitation trends though the model-simulated
magnitude of changes is smaller than that observed (X. Zhang et al., 2007).
The smaller changes in model simulations may be due in part to averaging
precipitation trends from different model simulations, as spatial patterns
of trends simulated by different models are not exactly the same. The
influence of anthropogenic greenhouse gases and aerosols on changes
in precipitation over high-latitude land areas north of 55°N has also been
detected (Min et al., 2008). Detection is possible there, despite limited
data coverage, in part because the response to forcing is relatively strong,
and because internal variability in precipitation is low in this region.
3.2.2.2. How to Attribute a Change in Extremes to Causes
The good practice guidance paper on detection and attribution (Hegerl
et al., 2010) reconciles terminologies of detection and attribution used
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
127
by Working Groups I and II in the AR4. It provides detailed guidance on
the procedures that include two main approaches to attribute a change
in climate to causes. One is single-step attribution, which involves
assessments that attribute an observed change within a system to an
external forcing based on explicitly modelling the response of the
variable to the external forcings. The alternate procedure is multi-step
attribution, which combines an assessment that attributes an observed
change in a variable of interest to a change in climate, with a separate
assessment that attributes the change to external forcings. Attribution
of changes in climate extremes has some unique issues. Observed data
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
FAQ 3.2 | Has Climate Change Affected Individual Extreme Events?
A changing climate can be expected to lead to changes in climate and weather extremes. But it is challenging to associate a single
extreme event with a specific cause such as increasing greenhouse gases because a wide range of extreme events could occur even in
an unchanging climate, and because extreme events are usually caused by a combination of factors. Despite this, it may be possible to
make an attribution statement about a specific weather event by attributing the changed probability of its occurrence to a particular
cause. For example, it has been estimated that human influences have more than doubled the probability of a very hot European summer
like that of 2003.
Recent years have seen many extreme events including the extremely hot summer in parts of Europe in 2003 and 2010, and the intense
North Atlantic hurricane seasons of 2004 and 2005. Can the increased atmospheric concentrations of greenhouse gases be considered
the ‘cause’ of such extreme events? That is, could we say these events would not have occurred if CO
2
had remained at pre-industrial
concentrations? For instance, the monthly mean November temperature averaged across the state of New South Wales in Australia for
November 2009 is about 3.5 standard deviations warmer than the 1950-2008 mean, suggesting that the chance of such a temperature
occurring in the 1950-2008 climate (assuming a stationary climate) is quite low. Is this event, therefore, an indication of a changing
climate? In the CRUTEM3V global land surface temperature data set, about one in every 900 monthly mean temperatures observed
between 1900 and 1949 lies more than 3.5 standard deviations above the corresponding monthly mean temperature for 1950-2008.
1
Since global temperature was lower in the first half of the 20th century, this clearly indicates that an extreme warm event as rare as the
November 2009 temperature in any specific location could have occurred in the past, even if its occurrence in recent times is more probable.
A second complicating issue is that extreme events usually result from a combination of factors, and this will make it difficult to attribute
an extreme to a single causal factor. The hot 2003 European summer was associated with a persistent high-pressure system (which led
to clear skies and thus more solar energy received at the surface) and too-dry soil (which meant that less solar energy was used for
evaporation, leaving more energy to heat the soil). Another example is that hurricane genesis requires weak vertical wind shear, as well
as very warm sea surface temperatures. Since some factors, but not others, may be affected by a specific cause such as increasing
greenhouse gas concentrations, it is difficult to separate the human influence on a single, specific extreme event from other factors
influencing the extreme.
Nevertheless, climate models can sometimes be used to identify if specific factors are changing the likelihood of the occurrence of
extreme events. In the case of the 2003 European heat wave, a model experiment indicated that human influences more than doubled
the likelihood of having a summer in Europe as hot as that of 2003, as discussed in the AR4. The value of such a probability-based
approach – “Does human influence change the likelihood of an event?” – is that it can be used to estimate the influence of external
factors, such as increases in greenhouse gases, on the frequency of specific types of events, such as heat waves or cold extremes. The
same likelihood-based approach has been used to examine anthropogenic greenhouse gas contribution to flood probability.
The discussion above relates to an individual, specific occurrence of an extreme event (e.g., a single heat wave). For the reasons outlined
above it remains very difficult to attribute any individual event to greenhouse gas-induced warming (even if physical reasoning or model
experiments suggest such an extreme may be more likely in a changed climate). On the other hand, a long-term trend in an extreme
(e.g., heat wave occurrences) is a different matter. It is certainly feasible to test whether such a trend is likely to have resulted from
anthropogenic influences on the climate, just as a global warming trend can be assessed to determine its likely cause.
____________
1
We used the CRUTEM3V land surface temperature data. We limit our calculation to grid points with long-term observations, requiring at least 50 non-missing values
during 1950-2008 for a calendar month and a grid point to be included. A standard deviation is computed for the period 1950-2008. We then count the number of
occurrences when the temperature anomaly during 1900-1949 relative to 1950-2008 mean is greater than 3.5 standard deviations, and compare it with the total
number of observations for the grid and month in that period. The ratio of these two numbers is 0.00107.
128
are limited in both quantity and quality (Section 3.2.1), resulting in
uncertainty in the estimation of past changes; the signal-to-noise ratio
may be low for many variables and insufficient data may be available to
detect such weak signals. In addition, global climate models (GCMs)
have several issues in simulating extremes and downscaling techniques
can only partly circumvent these issues (Section 3.2.3).
Single-step attribution based on optimal detection and attribution (e.g.,
Hegerl et al., 2007) can in principle be applied to climate extremes.
However, the difference in statistical properties between mean values
and extremes needs to be carefully addressed (e.g., Zwiers et al., 2011;
see also Section 3.1.6). Post-processing of climate model simulations to
derive a quantity of interest that is not explicitly simulated by the models,
by applying empirical methods or physically based models to the outputs
from the climate models, may make it possible to directly compare
observed extremes with climate model results. For example, sea level
pressure simulated by multiple GCMs has been used to derive
geostrophic wind to represent atmospheric storminess and to derive
significant wave height on the oceans for the detection of external
influence on trends in atmospheric storminess and northern oceans
wave heights (X.L. Wang et al., 2009a). GCM-simulated precipitation
and temperature have also been downscaled as input to hydrological
and snowpack models to infer past and future changes in temperature,
timing of the peak flow, and snow water equivalent for the western
United States, and this enabled a detection and attribution analysis of
human-induced changes in these variables (Barnett et al., 2008).
If a single-step attribution of causes to effects on extremes or physical
impacts of extremes is not feasible, it might be feasible to conduct a
multiple-step attribution. The assessment would then need to be based
on evidence not directly derived from model simulations, that is, physical
understanding and expert judgment, or their combination. For instance,
in the northern high-latitude regions, spring temperature has increased,
and the timing of spring peak flows in snowmelt-fed rivers has shifted
toward earlier dates (Regonda et al., 2005; Knowles et al., 2006). A
change in streamflow may be attributable to external influence if
streamflow regime change can be attributed to a spring temperature
increase and if the spring temperature increase can be attributed to
external forcings (though these changes may not necessarily be linked to
changes in floods; Section 3.5.2). If the chain of processes is established
(e.g., in this case additionally supported by the physical understanding
that snow melts earlier as spring temperature increases), the confidence
in the overall assessment would be similar to, or weaker than, the lower
confidence in the two steps in the assessment. In cases where the
underlying physical mechanisms are less certain, such as those linking
tropical cyclones and sea surface temperature (see Section 3.4.4), the
confidence in multi-step attribution can be severely undermined. A
necessary condition for multi-step attribution is to establish the chain of
mechanisms responsible for the specific extremes being considered.
Physically based process studies and sensitivity experiments that help
the physical understanding (e.g., Findell and Delworth, 2005;
Seneviratne et al., 2006a; Haarsma et al., 2009) can possibly play a role
in developing such multi-step attributions.
Extreme events are rare, which means that there are also few data
available to make assessments regarding changes in their frequency or
intensity (Section 3.2.1). When a rare and high-impact meteorological
extreme event occurs, a question that is often posed is whether such an
event is due to anthropogenic influence. Because it is very difficult to
rule out the occurrence of low-probability events in an unchanged
climate and because the occurrence of such events usually involves
multiple factors, it is very difficult to attribute an individual event to
external forcing (Allen, 2003; Hegerl et al., 2007; Dole et al., 2011; see
also FAQ 3.2). However, in this case, it may be possible to estimate
the influence of external forcing on the likelihood of such an event
occurring (e.g., Stott et al., 2004; Pall et al., 2011; Zwiers et al., 2011).
3.2.3. Projected Long-Term Changes and Uncertainties
In this section we discuss the requirements and methods used for
preparing climate change projections, with a focus on projections of
extremes and the associated uncertainties. The discussion draws on the
AR4 (Christensen et al., 2007; Meehl et al., 2007b; Randall et al., 2007)
with consideration of some additional issues relevant to projections of
extremes in the context of risk and disaster management. More detailed
assessments of projections for specific extremes are provided in
Sections 3.3 to 3.5. Summaries of these assessments are provided in
Table 3-1. Overviews of projected regional changes in temperature
extremes, heavy precipitation, and dryness are provided in Table 3-3
(see pages 196-202).
3.2.3.1. Information Sources for Climate Change Projections
Work on the construction, assessment, and communication of climate
change projections, including regional projections and of extremes,
draws on information from four sources: (1) GCMs; (2) downscaling of
GCM simulations; (3) physical understanding of the processes governing
regional responses; and (4) recent historical climate change (Christensen
et al., 2007; Knutti et al., 2010b). At the time of the AR4, GCMs were the
main source of globally available regional information on the range of
possible future climates including extremes (Christensen et al., 2007).
This is still the case for many regions, as can be seen in Table 3-3.
The AR4 concluded that statistics of extreme events for present-day
climate, especially temperature, are generally well simulated by current
GCMs at the global scale (Randall et al., 2007). Precipitation extremes
are, however, less well simulated (Randall et al., 2007; Box 3-2). As
they continue to develop, and their spatial resolution as well as their
complexity continues to improve, GCMs could become increasingly useful
for investigating smaller-scale features, including changes in extreme
weather events. However, when we wish to project climate and weather
extremes, not all atmospheric phenomena potentially of relevance can
be realistically or explicitly simulated. GCMs include a number of
approximations, known as parameterizations, of processes (e.g., relating
to clouds) that cannot be fully resolved in climate models. Furthermore,
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
129
the assessment of climate model performance with respect to extremes
(summarized in Sections 3.3 to 3.5 for specific extremes), particularly at
the regional scale, is still limited by the rarity of extreme events that
makes evaluation of model performance less robust than is the case for
average climate. Evaluation is further hampered by incomplete data on
the historical frequency and severity of extremes, particularly for variables
other than temperature and precipitation, and for specific regions
(Section 3.2.1; Table 3-2).
The requirement for projections of extreme events has provided one of
the motivations for the development of regionalization or downscaling
techniques (Carter et al., 2007). These have been specifically developed
for the study of regional- and local-scale climate change, to simulate
weather and climate at finer spatial resolutions than is possible with
GCMs – a step that is particularly relevant for many extremes given
their spatial scale. These techniques are, nonetheless, constrained by the
reliability of large-scale information coming from GCMs. Recent
advances in downscaling for extremes are discussed below.
As indicated in the Glossary, downscaling “is a method that derives local-
to regional-scale (up to 100 km) information from larger-scale models
or data analyses. Two main methods are distinguished: dynamical
downscaling and empirical/statistical downscaling (Christensen et al.,
2007). The dynamical method uses the output of regional climate
models (RCMs), global models with variable spatial resolution, or high-
resolution global models. The empirical/statistical methods develop
statistical relationships that link the large-scale atmospheric variables
with local/regional climate variables. In all cases, the quality of the
downscaled product depends on the quality of the driving model.
Dynamical and statistical downscaling techniques are briefly introduced
hereafter. Specific limitations that need to be considered in the evaluation
of projections are also discussed in Section 3.2.3.2.
The most common approach to dynamical downscaling uses high-
resolution RCMs, currently at scales of 20 to 50 km, but in some cases
down to 10 to 15 km (e.g., Dankers et al., 2007), to represent regional
sub-domains, using either observed (reanalysis) or lower-resolution
GCM data to provide their boundary conditions. Using non-hydrostatic
mesoscale models, applications at 1- to 5-km resolution are also possible
for shorter periods (typically a few months, a few full years at most) – a
scale at which clouds and convection can be explicitly resolved and the
diurnal cycle tends to be better resolved (e.g., Grell et al., 2000; Hay et al.,
2006; Hohenegger et al., 2008; Kanada et al., 2010b). Less commonly
used approaches to dynamical downscaling involve the use of
stretched-grid (variable resolution) models and high-resolution ‘time-
slice’ models (e.g., Cubasch et al., 1995; Gibelin and Deque, 2003;
Coppola and Giorgi, 2005) with the latter including some simulations at
20 km globally (Kamiguchi et al., 2006; Kitoh et al., 2009; Kim et al.,
2010). The main advantage of dynamical downscaling is its potential for
capturing mesoscale nonlinear effects and providing information for
many climate variables at a relatively high spatial resolution, although
still not as high as some require. Dynamical downscaling cannot provide
information at the point (i.e., weather station) scale (a scale at which
the RCM and GCM parameterizations would not work). Like GCMs,
RCMs provide precipitation averaged over a grid cell, which means a
tendency to more days of light precipitation (Frei et al., 2003; Barring et
al., 2006) and reduced magnitude of extremes (Chen and Knutson,
2008; Haylock et al., 2008) compared with point values. These scaling
issues need to be considered when evaluating the ability of RCMs and
GCMs to simulate precipitation and other extremes.
Statistical downscaling methods use relationships between large-scale
fields (predictors) and local-scale surface variables (predictands) that
have been derived from observed data, and apply these to equivalent
large-scale fields simulated by climate models (Christensen et al., 2007).
They may also include weather generators that provide the basis for a
number of recently developed user tools that can be used to assess
changes in extreme events (Kilsby et al., 2007; Burton et al., 2008; Qian et
al., 2008; Semenov, 2008). Statistical downscaling has been demonstrated
to have potential in a number of different regions including Europe
(e.g., Schmidli et al., 2007), Africa (e.g., Hewitson and Crane, 2006),
Australia (e.g., Timbal et al., 2008, 2009), South America (e.g., D’Onofrio
et al., 2010) and North America (e.g., Vrac et al., 2007; Dibike et al.,
2008). Statistical downscaling methods are able to access finer spatial
scales than dynamical methods and can be applied to parameters that
cannot be directly obtained from RCMs. Seasonal indices of extremes
can, for example, be simulated directly without having to first produce
daily time series (Haylock et al., 2006a), or distribution functions of
extremes can be simulated (Benestad, 2007). However, statistical
downscaling methods require observational data at the desired scale
(e.g., the point or station scale) for a long enough period to allow the
model to be well trained and validated, and in some methods can lack
coherency among multiple climate variables and/or multiple sites. One
specific disadvantage of some, but not all, methods based on the
analog approach is that they cannot produce extreme events greater in
magnitude than have been observed before (Timbal et al., 2009).
Moreover, statistical downscaling does not allow for the possibility of
future process-based changes in relationships between predictors and
predictands (see Section 3.2.3.2). There have been few systematic
intercomparisons of dynamical and statistical downscaling approaches
focusing on extremes (Fowler et al., 2007b). Two examples focus on
extreme precipitation for the United Kingdom (Haylock et al., 2006a) and
the Alps (Schmidli et al., 2007), respectively. A few hybrid statistico-
dynamical downscaling methods also exist, including a two-step
approach used to downscale heavy precipitation events in southern
France (Beaulant et al., 2011). A conceptually similar cascading technique
has also been used to downscale tropical cyclones (Bender et al., 2010;
see Section 3.4.4).
In terms of temporal resolution, while GCMs and RCMs operate at
sub-daily time steps, model output at six-hourly or shorter temporal
resolutions, which is desirable for some applications such as urban
drainage, is less widely available than daily output. Where limited
studies have been undertaken, there is evidence that at the typical
spatial resolutions used (i.e., non-cloud/convection-resolving scales),
RCMs do not adequately represent sub-daily precipitation and the
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
130
diurnal cycle of convection (Gutowski et al., 2003; Brockhaus et al.,
2008; Lenderink and Van Meijgaard, 2008). Development of sub-daily
statistical downscaling methods is constrained by the availability of
long observed time series for calibration and validation and this
approach is not currently widely used for climate change applications,
although some weather generators, for example, do provide hourly
information (Maraun et al., 2010).
It is not possible in this chapter to provide assessments of projected
changes in extremes at spatial scales smaller than for large regions
(Table 3-3). These large-region projections provide a wider context for
national or more local projections, where they exist, and, where they do
not, a first indication of expected changes, their associated uncertainties,
and the evidence available. Several countries, for example in Europe,
North America, Australia, and some other regions, have developed
national or sub-national projections (generally based on dynamical
and/or statistical downscaling), including information about extremes,
and a range of other high-resolution information and tools are available
from national weather and hydrological services and academic institutions
to assist users and decisionmakers.
3.2.3.2. Uncertainty Sources in Climate Change Projections
Uncertainty in climate change projections arises at each of the steps
involved in their preparation: determination of greenhouse gas and
aerosol precursor emissions (driven by socioeconomic development
and represented through the use of multiple emissions scenarios),
concentrations of radiatively active species, radiative forcing, and climate
response including downscaling. Also, uncertainty in the estimation of
the true ‘signal’ of climate change is introduced by both errors in the
model representation of Earth system processes and by internal climate
variability.
As was noted in Section 3.2.3.1, most shortcomings in GCMs and
RCMs result from the fact that many important small-scale processes
(e.g., representations of clouds, convection, land surface processes) are
not represented explicitly (Randall et al., 2007). Some processes –
particularly those involving feedbacks (Section 3.1.4), and this is
especially the case for climate extremes and associated impacts – are
still poorly represented and/or understood (e.g., land-atmosphere
interactions, ocean-atmosphere interactions, stratospheric processes,
blocking dynamics) despite some improvements in the simulations of
others (see Box 3-2 and below). Therefore, limitations in computing
power and in the scientific understanding of some physical processes
currently restrict further global and regional climate model improvements.
In addition, uncertainty due to structural or parameter errors in GCMs
propagates directly from global model simulations as input to RCMs
and thus to downscaled information.
These problems limit quantitative assessments of the magnitude and
timing, as well as regional details, of some aspects of projected climate
change. For instance, even atmospheric models with approximately 20-km
horizontal resolution still do not resolve the atmospheric processes
sufficiently finely to simulate the high wind speeds and low pressure
centers of the most intense hurricanes (Knutson et al., 2010).
Realistically capturing details of such intense hurricanes, such as the
inner eyewall structure, would require models with 1-km horizontal
resolution, far beyond the capabilities of current GCMs and of most
current RCMs (and even global numerical weather prediction models).
Extremes may also be impacted by mesoscale circulations that GCMs
and even current RCMs cannot resolve, such as low-level jets and their
coupling with intense precipitation (Anderson et al., 2003; Menendez et
al., 2010). Another issue with small-scale processes is the lack of relevant
observations, such as is the case with soil moisture and vegetation
processes (Section 3.2.1) and relevant parameters (e.g., maps of soil types
and associated properties, see for instance Seneviratne et al., 2006b;
Anders and Rockel, 2009).
Since many extreme events, such as those associated with precipitation,
occur at rather small temporal and spatial scales, where climate
simulation skill is currently limited and local conditions are highly
variable, projections of future changes cannot always be made with a
high level of confidence (Easterling et al., 2008). The credibility of
projections of changes in extremes varies with extreme type, season, and
geographical region (Box 3-2). Confidence and credibility in projected
changes in extremes increase when the physical mechanisms producing
extremes in models are considered reliable, such as increases in specific
humidity in the case of the projected increase in the proportion of summer
precipitation falling as intense events in central Europe (Kendon et al.,
2010). The ability of a model to capture the full distribution of variables
– not just the mean – together with long-term trends in extremes,
implies that some of the processes relevant to a future warming world
may be captured (Alexander and Arblaster, 2009; van Oldenborgh et al.,
2009). It should nonetheless be stressed that physical consistency of
simulations with observed behavior provides necessary but not sufficient
evidence for credible projections (Gutowski et al., 2008a).
While downscaling provides more spatial detail (Section 3.2.3.1), the
added value of this step and the reliability of projections always needs
to be assessed (Benestad et al., 2007; Laprise et al., 2008). A potential
limitation and source of uncertainty in downscaling methods is that the
calibration of statistical models and the parameterization schemes used
in dynamical models are necessarily based on present (and past) climate
(as well as an understanding of physical processes). Thus they may not
be able to capture changes in extremes that are induced by future
mechanistic changes in regional (or global) climate, that is, if used
outside the range for which they were designed (Christensen et al.,
2007). Spatial inhomogeneity of both land use/land cover and aerosol
forcing adds to regional uncertainty. This means that the factors inducing
uncertainty in the projections of extremes in different regions may
differ considerably. Some specific issues inducing uncertainties in RCM
projections are the interactions with the driving GCM, especially in
terms of biases and climate change signal (e.g., de Elía et al., 2008;
Laprise et al., 2008; Kjellström and Lind, 2009; Déqué et al., 2011) and
the choice of regional domain (Wang et al., 2004; Laprise et al., 2008).
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
131
In the case of statistical downscaling, uncertainties are induced by,
inter alia, the definition and choice of predictors (Benestad, 2001;
Hewitson and Crane, 2006; Timbal et al., 2008) and the underlying
assumption of stationarity (Raje and Mujumdar, 2010). In general, both
approaches to downscaling are maturing and being more widely applied
but are still restricted in terms of geographical coverage (Maraun et al.,
2010). For many regions of the world, no downscaled information exists
at all and regional projections rely only on information from GCMs (see
Table 3-3).
For many user-driven applications, impact models need to be included
as an additional step for projections (e.g., hydrological or ecosystem
models). Because of the previously mentioned issues of scale discrepancies
and overall biases, it is necessary to bias-correct RCM data before input
to some impacts models (i.e., to bring the statistical properties of present-
day simulations in line with observations and to use this information to
correct projections). A number of bias correction methods, including
quantile mapping and gamma transform, have recently been developed
and exhibit promising skill for extremes of daily precipitation (Piani et
al., 2010; Themeßl et al., 2011).
3.2.3.3. Ways of Exploring and Quantifying Uncertainties
Uncertainties can be explored, and quantified to some extent, through
the combined use of observations and reanalyses, process understanding,
a hierarchy of climate models, and ensemble simulations. Ensembles of
model simulations represent a fundamental resource for studying the
range of plausible climate responses to a given forcing (Meehl et al.,
2007b; Randall et al., 2007). Such ensembles can be generated either by
(i) collecting results from a range of models from different modelling
centers (multi-model ensembles), to include the impact of structural
model differences; (ii) by generating simulations with different initial
conditions (intra-model ensembles) to characterize the uncertainties
due to internal climate variability; or (iii) varying multiple internal model
parameters within plausible ranges (perturbed and stochastic physics
ensembles), with both (ii) and (iii) aiming to produce a more systematic
estimate of single model uncertainty (Knutti et al., 2010b).
Many of the global models utilized for the AR4 were integrated as
ensembles, permitting more robust statistical analysis than is possible if a
model is only integrated to produce a single projection. Thus the available
CMIP3 Multi-Model Ensemble (MME) GCM simulations reflect both inter-
and intra-model variability. In advance of AR4, coordinated climate change
experiments were undertaken which provided information from 23 models
from around the world (Meehl et al., 2007a). The CMIP3 simulations
were made available at the Program for Climate Model Diagnosis and
Intercomparison (www-pcmdi.llnl.gov/ipcc/about_ipcc.php). However,
the higher temporal resolution (i.e., daily) data necessary to analyze
most extreme events were quite incomplete in the archive, with only
four models providing daily averaged output with ensemble sizes
greater than three realizations and many models not included at all.
GCMs are expensive to run, thus a compromise is needed between the
number of models, number of simulations, and the complexity of the
models (Knutti, 2010).
Besides the uncertainty due to randomness itself, which is the canonical
statistical definition, it is important to distinguish between the uncertainty
due to insufficient agreement in the model projections, the uncertainty
due to insufficient evidence (insufficient observational data to constrain
the model projections or insufficient number of simulations from different
models or insufficient understanding of the physical processes), and the
uncertainty induced by insufficient literature, which refers to the lack of
published analyses of projections. For instance, models may agree on a
projected change, but if this change is controlled by processes that are
not well understood and validated in the present climate, then there is
an inherent uncertainty in the projections, no matter how good the
model agreement may be. Similarly, available model projections may
agree in a given change, but the number of available simulations may
restrain the reliability of the inferred agreement (e.g., because the
analyses need to be based on daily data that may not be available from
all modelling groups). All these issues have been taken into account in
assessing the confidence and likelihood of projected changes in
extremes for this report (see Section 3.1.5).
Uncertainty analysis of the CMIP3 MME in AR4 focused essentially on the
seasonal mean and inter-model standard deviation values (Christensen et
al., 2007; Meehl et al., 2007b; Randall et al., 2007). In addition, confidence
was assessed in the AR4 through simple quantification of the number of
models that show agreement in the sign of a specific climate change
(e.g., sign of the change in frequency of extremes) – assuming that the
greater the number of models in agreement, the greater the robustness.
However, the shortcoming of this definition of model agreement is that
it does not take account of possible common biases among models.
Indeed, the ensemble was strictly an ‘ensemble of opportunity,’ without
sampling protocol, and the possible dependence of different models on
one another (e.g., due to shared parameterizations) was not assessed
(Knutti et al., 2010a). Furthermore, this particular metric, which assesses
sign agreement only, can provide misleading conclusions in cases, for
example, where the projected changes are near zero. For this reason, in
our assessments of projected changes in extreme indices we consider
the model agreement as a necessary but not a sufficient condition for
likelihood statements [e.g., agreement of 66% of the models, as indicated
with shading in several of the figures (Figures 3-3, 3-4, 3-6, 3-8, and
3-10), is a minimum but not a sufficient condition for a change being
considered ‘likely’].
Post-AR4 studies have concentrated more on the use of the MME in
order to better characterize uncertainty in climate change projections,
including those of extremes (Kharin et al., 2007; Gutowski et al., 2008a;
Perkins et al., 2009). New techniques have been developed for exploiting
the full ensemble information, in some cases using observational
constraints to construct probability distributions (Tebaldi and Knutti,
2007; Tebaldi and Sanso, 2009), although issues such as determining
appropriate metrics for weighting models are challenging (Knutti et al.,
2010a). Perturbed-physics ensembles have also become available (e.g.,
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
132
Collins et al., 2006; Murphy et al., 2007) and used to examine projected
changes in extremes and their uncertainties (Barnett et al., 2006; Clark
et al., 2006, 2010; Burke and Brown, 2008). Advances have also been
made in developing probabilistic information at regional scales from
the GCM simulations, but there has been rather less development
extending this to probabilistic downscaled regional information and to
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Box 3-2 | Variations in Confidence in Projections of Climate Change:
Mean versus Extremes, Variables, Scale
Comparisons of observed and simulated climate demonstrate good agreement for some climate variables such as mean temperature,
especially at large horizontal scales (e.g., Räisänen, 2007). For instance, Figure 9.12 of the AR4 (Hegerl et al., 2007) compares the ability
of 14 climate models to simulate the temporal variations of mean temperature through the 20th century. When the models included
both natural and anthropogenic forcings, they consistently reproduced the decadal variations in global mean temperature. Without the
anthropogenic influences the models consistently failed to reproduce the multi-decadal temperature variations. However, when the same
models’ abilities to simulate the temperature variations for smaller domains were assessed, although the mean temperature produced by
the ensemble generally tracked the observed temperature changes, the consistency among the models was poorer than was the case for
the global mean (Figure 9.12; Hegerl et al., 2007), partly because averaging over global scales smoothes internal variability or ‘noise’
more than averaging over smaller domains (see also Section 3.2.2.1). We can conclude that the smaller the spatial domain for which
simulations or projections are being prepared, the less confidence we should have in these projections (although in some limited cases
regional-scale projections can have higher reliability than larger-scale projections; see Section 3.1.6).
This increased uncertainty at smaller scales results from larger internal variability at smaller scales or ‘noise’ (i.e., natural variability
unrelated to external forcings) and increased model uncertainty, both of which lead to lower model consistency at these scales (Hawkins
and Sutton, 2009). The latter factor is largely due to the role of unresolved processes (representations of clouds, convection, land surface
processes; see also Section 3.2.3). Hawkins and Sutton (2009) also point out regional variations in these aspects: in the tropics the
temperature signal expected from anthropogenic factors is large relative to the model uncertainty and the natural variability, compared
with higher latitudes. Figure 9.12 from AR4 (Hegerl et al., 2007) also shows that the models are more consistent in reproducing decadal
temperature variations in the tropics than at higher latitudes, even though the magnitudes of the temperature trends are larger at higher
latitudes.
Uncertainty in projections also depends on the variables, phenomena, or impacts considered (Sections 3.3. to 3.5.). There is more model
uncertainty for variables other than temperature, for instance precipitation (Räisänen, 2007; Hawkins and Sutton, 2011; see also
Section 3.2.3). And the situation is more difficult again for extremes. For instance, climate models simulate observed changes in extreme
temperatures relatively well, but the frequency, distribution, and intensity of heavy precipitation is more poorly simulated (Randall et al.,
2007) as are observed changes in heavy precipitation (e.g., Alexander and Arblaster, 2009). Also, projections of changes in temperature
extremes tend to be more consistent across climate models (in terms of sign) than for (wet and dry) precipitation extremes (Tebaldi et
al., 2006; Orlowsky and Seneviratne, 2011; see also Figures 3-3 through 3-7 and 3-10) and significant inconsistencies are also found for
projections of agricultural (soil moisture) droughts (Wang, 2005; see also Box 3-3; Figure 3-10). For some other extremes, such as tropical
cyclones, differences in the regional-scale climate change projections between models can lead to marked differences in projected tropical
cyclone activity associated with anthropogenic climate change (Knutson et al., 2010), and thus decrease confidence in projections of
changes in that extreme.
The relative importance of various causes of uncertainties in projections is somewhat different for earlier compared with later future
periods. For some variables (mean temperature, temperature extremes), the choice of emission scenario becomes more critical than
model uncertainty for the second part of the 21st century (Tebaldi et al., 2006; Hawkins and Sutton, 2009, 2011) though this does not
apply for mean precipitation and some precipitation-related extremes (Tebaldi et al., 2006; Hawkins and Sutton, 2009, 2011), and has in
particular not been evaluated in detail for a wide range of extremes. Users need to be aware of such issues in deciding the range of
uncertainties that is appropriate to consider for their particular risk or impacts assessment
In summary, confidence in climate change projections depends on the (temporal and spatial) scale and variable being considered and
whether one considers extremes or mean quantities. Confidence is highest for temperature, especially at the global scale, and decreases
when other variables are considered, and when we focus on smaller spatial domains (Tables 3-1 and 3-3). Confidence in projections for
extremes is generally weaker than for projections of long-term averages.
133
extremes (Fowler et al., 2007a; Fowler and Ekstrom, 2009). Perhaps the
most comprehensive approach to date for quantifying the influence of
the cascade of uncertainties in regional projections is that used to
develop the recent United Kingdom Climate Projections (UKCP09;
Murphy et al., 2009). A complex Bayesian framework is used to combine
a perturbed physics ensemble exploring uncertainties in atmosphere
and ocean processes, and the carbon and sulfur cycles, with structural
uncertainty (represented by 12 CMIP3 models) and an 11-member RCM
perturbed physics ensemble. The published projections provide probability
distributions of changes in various parameters including the wettest and
hottest days of each season for 25-km grid squares across the United
Kingdom. These probabilities are conditional on the emissions scenario
(low, medium, high) and are described as representing the “relative degree
to which each climate outcome is supported by the evidence currently
available, taking into account our understanding of climate science and
observations, and using expert judgment” (Murphy et al., 2009).
Both statistical and dynamical downscaling methods are affected by
the uncertainties that affect the global models, and a further level of
uncertainty associated with the downscaling step also needs to be
taken into consideration (see also Sections 3.2.3.1 and 3.2.3.2). The
increasing availability of coordinated RCM simulations for different
regions permits more systematic exploration of dynamical downscaling
uncertainty. Such simulations are available for Europe (e.g., Christensen
and Christensen, 2007; van der Linden and Mitchell, 2009) and a few
other regions such as North America (Mearns et al., 2009) and West
Africa (van der Linden and Mitchell, 2009; Hourdin et al., 2010). RCM
intercomparisons have also been undertaken for a number of regions
including Asia (Fu et al., 2005), South America (Menendez et al., 2010) and
the Arctic (Inoue et al., 2006). A new series of coordinated simulations
covering the globe is planned (Giorgi et al., 2009). Increasingly, RCM
output from coordinated simulations is made available at the daily time
scale, facilitating the analysis of some extreme events. Nevertheless, it
is important to point out that ensemble runs with RCMs currently
involve a limited number of driving GCMs, and hence only subsample
uncertainty space. Ensuring adequate sampling of RCM simulations (both
in terms of the number of considered RCMs and number of considered
driving GCMs) may be more important for extremes than for changes in
mean values (Frei et al., 2006; Fowler et al., 2007a). Internal variability,
for example, has been shown to make a significant contribution to
the spectrum of variability on at least multi-annual time scales and
potentially up to multi-decadal time scales (Kendon et al., 2008;
Hawkins and Sutton, 2009, 2011; Box 3-2).
3.3. Observed and Projected Changes in
Weather and Climate Extremes
3.3.1. Temperature
Temperature is associated with several types of extremes, for example,
heat waves and cold spells, and related impacts, for example, on human
health, the physical environment, ecosystems, and energy consumption
(e.g., Chapter 4, Sections 3.5.6 and 3.5.7; see also Case Studies
9.2.1
and 9.2.10
). Temperature extremes often occur on weather time scales
that require daily or higher time scale resolution data to accurately
assess possible changes (Section 3.2.1). It is important to distinguish
between daily mean, maximum (i.e., daytime), and minimum (nighttime)
temperature, as well as between cold and warm extremes, due to their
differing impacts. Spell lengths (e.g., duration of heat waves) are
relevant for a number of impacts. Note that we do not consider
here changes in diurnal temperature range or frost days, which are not
typical ‘climate extremes’. There is an extensive body of literature
regarding the mechanisms of changes in temperature extremes (e.g.,
Christensen et al., 2007; Meehl et al., 2007b; Trenberth et al., 2007).
Heat waves are generally caused by quasi-stationary anticyclonic
circulation anomalies or atmospheric blocking (Xoplaki et al., 2003;
Meehl and Tebaldi, 2004; Cassou et al., 2005; Della-Marta et al., 2007b),
and/or land-atmosphere feedbacks (in transitional climate regions),
whereby the latter can act as an amplifying mechanism through reduction
in evaporative cooling (Section 3.1.4), but also induce enhanced
persistence due to soil moisture memory (Lorenz et al., 2010). Also snow
feedbacks (Section 3.1.4), and possibly changes in aerosols (Portmann et
al., 2009), are relevant for temperature extremes. Trends in temperature
extremes (either observed or projected) can sometimes be different for
the most extreme temperatures (e.g., annual maximum/minimum daily
maximum/minimum temperature) than for less extreme events [e.g.,
cold/warm days/nights; see, for instance, Brown et al. (2008) versus
Alexander et al. (2006)]. One reason for this is that ‘moderate extremes’
such as warm/cold days/nights are generally computed for each day
with respect to the long-term statistics for that day, thus, for example,
an increase in warm days for annual analyses does not necessarily imply
warming for the very warmest days of the year.
Observed Changes
Regional historical or paleoclimatic temperature reconstructions may
help place the recent instrumentally observed temperature extremes in
the context of a much longer period, but literature on this topic is very
sparse and most regional reconstructions are for Europe. For example
Dobrovolny et al. (2010) reconstructed monthly and seasonal temperature
over central Europe back to 1500 using a variety of temperature proxy
records. They concluded that the summer 2003 heat wave and the July
2006 heat wave exceeded the +2 standard deviation (associated with
the reconstruction method) of previous monthly temperature extremes
since 1500. Barriopedro et al. (2011) showed that the anomalously warm
summers of 2003 in western and central Europe and 2010 in eastern
Europe and Russia both broke the 500-year long seasonal temperature
record over 50% of Europe. The coldest periods within the last five
centuries occurred in the winter and spring of 1690. Another 500-year
temperature reconstruction was recently completed for the
Mediterranean basin by means of documentary data and instrumental
observations (Camuffo et al., 2010). It suggests strong natural variability
in the basin, possibly exceeding the recent warming, although
discontinuities in the records limit the interpretation of this finding.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
134
The AR4 (Trenberth et al., 2007, based on Alexander et al., 2006) reported
a statistically significant increase in the numbers of warm nights and a
statistically significant reduction in the numbers of cold nights for 70 to
75% of the land regions with data (for the spatial coverage of the
underlying data set and the definition of warm/cold days and nights, see
Section 3.2.1 and Box 3-1, respectively). Changes in the numbers of
warm days and cold days also showed warming, but less marked
than for nights, with about 40 to 50% of the area with data showing
statistically significant changes consistent with warming (Alexander et
al., 2006). Less than 1% of the area with data showed statistically
significant trends in cold/warm days and nights that were consistent with
cooling (Alexander et al., 2006). Trenberth et al. (2007)
also reported,
based on Vose et al. (2005), that from 1950 to 2004, the annual trends
in minimum and maximum land-surface air temperature averaged over
regions with data were 0.20°C per decade and 0.14°C per decade,
respectively, and that for 1979 to 2004, the corresponding linear trends
for the land areas with data were 0.29°C per decade for both maximum
and minimum temperature. Based on this evidence, the IPCC AR4 (SPM;
IPCC, 2007b) assessed that it was very likely that there had been trends
toward warmer and more frequent warm days and warm nights, and
warmer and less frequent cold days and cold nights in most land areas.
Regions that were found to depart from this overall behavior toward
more warm days and nights and fewer cold days and nights in
Alexander et al. (2006) were mostly central North America, the eastern
United States, southern Greenland (increase in cold days and decreases
in warm days), and the southern half of South America (decrease in
warm days; no data available for the northern half of the continent). In
central North America and the eastern United States this partial tendency
for a negative trend in extremes is also consistent with a reported mean
negative trend in temperatures, mostly in the spring to summer season
(also termed ‘warming hole’, e.g., Pan et al., 2004; Portmann et al.,
2009). Several explanations have been suggested for this behavior,
which seems partly associated with a change in the hydrological cycle,
possibly linked to soil moisture and/or aerosol feedbacks (Pan et al.,
2004; Portmann et al., 2009).
More recent analyses available since the AR4 include a global study (for
annual extremes) by Brown et al. (2008) based on the data set from
Caesar et al. (2006), and regional studies for North America (Peterson et
al., 2008a; Meehl et al., 2009c), Central-Western Europe (since 1880;
Della-Marta et al., 2007a), central and eastern Europe (Bartholy and
Pongracz, 2007; Kürbis et al., 2009), the eastern Mediterranean region
including Turkey (Kuglitsch et al., 2010), western Central Africa, Guinea
Conakry and Zimbabwe (Aguilar et al., 2009), the Tibetan Plateau (You
et al., 2008) and China (You et al., 2011), Uruguay (Rusticucci and
Renom, 2008), and Australia (Alexander and Arblaster, 2009). Further
references can also be found in Table 3-2. Overall, these studies are
consistent with the assessment of an increase in warm days and nights
and a reduction in cold days and nights on the global basis, although
they do not necessarily consider trends in all four variables, and a few
single studies report trends that are not statistically significant or even
trends opposite to the global tendencies in some extremes, subregions,
seasons, or decades. For instance, Rusticucci and Renom (2008) found
in Uruguay a reduction of cold nights, a positive but a statistically
insignificant trend in warm nights, statistically insignificant decreases in
cold days at most investigated stations, and inconsistent trends in
warm days. Together with the previous results from Alexander et al.
(2006) for southern South America (see above) and further regional
studies (Table 3-2), this suggests a less consistent warming tendency in
South America compared to other continents. Another notable feature is
that studies for central and southeastern Europe display a marked
change point in trends in temperature extremes at the end of the
1970s/beginning of 1980s (Table 3-2), which for some extremes can
lead to very small and/or statistically not significant overall trends since
the 1960s (e.g., Bartholy and Pongracz, 2007).
There are fewer studies available investigating changes in characteristics
of cold spells and warm spells, or cold waves and heat waves, compared
with studies of the intensity or frequency of warm and cold days or
nights. Alexander et al. (2006) provided an analysis of trends in warm
spells [based on the Warm Spell Duration Index (WSDI); see Table 3-2
and Box 3-1] mostly in the mid- and high-latitudes of the Northern
Hemisphere. The analysis displays a tendency toward a higher length or
number of warm spells (increase in number of days belonging to warm
spells) in much of the region, with the exception of the southeastern
United States and eastern Canada. Regional studies on trends in warm
spells or heat waves are also listed in Table 3-2. Kunkel et al. (2008)
found that the United States has experienced a general decline in cold
waves over the 20th century, with a spike of more cold waves in the
1980s. Further, they report a strong increase in heat waves since 1960,
although the heat waves of the 1930s associated with extreme drought
conditions still dominate the 1895-2005 time series. Kuglitsch et al.
(2009) reported an increase in heat wave intensity, number, and length
in summer over the 1960-2006 time period in the eastern Mediterranean
region. Ding et al. (2010) reported increasing numbers of heat waves
over most of China for the 1961-2007 period. The record-breaking heat
wave over western and central Europe in the summer of 2003 is an
example of an exceptional recent extreme (Beniston, 2004; Schär and
Jendritzky, 2004). That summer (June to August) was the hottest since
comparable instrumental records began around 1780 and perhaps the
hottest since at least 1500 (Luterbacher et al., 2004). Other examples of
recent extreme heat waves include the 2006 heat wave in Europe
(Rebetez et al., 2008), the 2007 heat wave in southeastern Europe
(Founda and Giannakopoulos, 2009), the 2009 heat wave in southeastern
Australia (National Climate Centre, 2009), and the 2010 heat wave in
Russia (Barriopedro et al., 2011). Both the 2003 European heat wave
(Andersen et al., 2005; Ciais et al., 2005) and the 2009 southeastern
Australian heat wave were also associated with drought conditions,
which can strongly enhance temperature extremes during heat waves in
some regions (see also Section 3.1.4).
Some recent analyses have led to revisions of previously reported
trends. For instance, Della-Marta et al. (2007a) found that mean summer
maximum temperature change over Europe was +1.6 ± 0.4°C during
1880 to 2005, a somewhat greater increase than reported in earlier
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
135
studies. Kuglitsch et al. (2009, 2010) homogenized and analyzed over
250 daily maximum and minimum temperature series in the
Mediterranean region since 1960, and found that after homogenization
the positive trends in the frequency of hot days and heat waves in the
Eastern Mediterranean region were higher than reported in earlier studies.
This was due to the correction of many warm-biased temperature data
in the region during the 1960s and 1970s.
In summary, regional and global analyses of temperature extremes on
land generally show recent changes consistent with a warming climate
at the global scale, in agreement with the previous assessment in AR4.
Only a few regions show changes in temperature extremes consistent
with cooling, most notably for some extremes in central North America,
the eastern United States, and also parts of South America. Based on the
available evidence we conclude that it is very likely that there has been
an overall decrease in the number of cold days and nights and very likely
that there has been an overall increase in the number of warm days and
nights in most regions, that is, for land areas with data (corresponding
to about 70 to 80% of all land areas; see Table 3-2). It is likely that this
statement applies at the continental scale in North America, Europe,
and Australia (Table 3-2). However, some subregions on these continents
have had warming trends in temperature extremes that were small or not
statistically significant (e.g., southeastern Europe), and a few subregions
have had cooling trends in some temperature extremes (e.g., central North
America and eastern United States). Asia also shows trends consistent
with warming in most of the continent, but which are assessed here to
be of medium confidence because of lack of literature for several regions
apart from the global study from Alexander et al. (2006). Most of Africa
is insufficiently well sampled to allow an overall likelihood statement to
be made at the continental scale, although most of the regions on this
continent for which data are available have exhibited warming in
temperature extremes (Table 3-2). In South America, both lack of data
and some inconsistencies in the reported trends imply low confidence in
the overall trends at the continental scale (Table 3-2). In many (but not
all) regions with sufficient data there is medium confidence that the
number of warm spells or heat waves has increased since the middle of
the 20th century (Table 3-2).
Causes of Observed Changes
The AR4 (Hegerl et al., 2007) concluded that surface temperature
extremes have likely been affected by anthropogenic forcing. This
assessment was based on multiple lines of evidence of temperature
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
5
10
20
40
100
ALL
ANT
Alaska
5
10
20
40
100
ALL ANT
Central North America
5
10
20
40
100
ALL ANT
Canada and Greenland
5
10
20
40
100
ALL ANT
Western North America
5
10
20
40
100
ALL ANT
Eastern North America
5
10
20
40
100
ALL
ANT
Central America and Mexico
5
10
20
40
100
ALL ANT
Southern South America
5
10
20
40
100
ALL
ANT
Southern Africa
5
10
20
40
100
ALL
ANT
North Europe
5
10
20
40
100
ALL ANT
North Asia
5
10
20
40
100
ALL ANT
South Europe
5
10
20
40
100
ALL
ANT
China
5
10
20
40
100
ALL
ANT
South Asia
5
10
20
40
100
ALL
ANT
Central Asia
5
10
20
40
100
ALL
ANT
Australia
5
10
20
40
100
ALL ANT
Global Land
Figure 3-2 | Estimated return periods (years) and their 5 and 95% uncertainty limits for 1960s 20-year return values of annual extreme daily temperatures in the 1990s climate
(see text for more details). ANT refers to model simulated responses with only anthropogenic forcing and ALL is both natural and anthropogenic forcing. Error bars are for annual
minimum daily minimum temperature (red: TNn), annual minimum daily maximum temperature (green: TXn), annual maximum daily minimum temperature (blue: TNx), and annual
maximum daily maximum temperature (pink: TXx), respectively. Grey areas have insufficient data. Source: Zwiers et al., (2011).
136
extremes at the global scale including the reported increase in the
number of warm extremes and decrease in the number of cold extremes
at that scale (Alexander et al., 2006). Hegerl et al. (2007) also state that
anthropogenic forcing may have substantially increased the risk of
extreme temperatures (Christidis et al., 2005) and of the 2003 European
heat wave (Stott et al., 2004).
Recent studies on attribution of changes in temperature extremes have
tended to reaffirm the conclusions reached in the AR4. Alexander and
Arblaster (2009) found that trends in warm nights over Australia could
only be reproduced by a coupled model that included anthropogenic
forcings. As part of the recent report of the US Climate Change Science
Program (CCSP, 2008), Gutowski et al. (2008a) concluded that most of
the observed changes in temperature extremes for the second half of
the 20th century over the United States can be attributed to human
activity. They compared observed changes in the number of frost days,
the length of growing season, the number of warm nights, and the heat
wave intensity with those simulated in a nine-member multi-model
ensemble simulation. The decrease in frost days, an increase in growing
season length, and an increase in heat wave intensity all show similar
changes over the United States in 20th-century experiments that
combine anthropogenic and natural forcings, though the relative
contributions of each are unclear.
Results from two global coupled climate models with separate
anthropogenic and natural forcing runs indicate that the observed
changes are simulated with anthropogenic forcings, but not with natural
forcings (even though there are some differences in the details of the
forcings). Zwiers et al. (2011) compared observed annual temperature
extremes including annual maximum daily maximum and minimum
temperatures, and annual minimum daily maximum and minimum
temperatures with those simulated responses to anthropogenic forcing or
anthropogenic and natural external forcings combined by multiple GCMs.
They fitted probability distributions (Box 3-1) to the observed extreme
temperatures with a time-evolving pattern of location parameters as
obtained from the model simulations, and found that both anthropogenic
influence and the combined influence of anthropogenic and natural
forcing can be detected in all four extreme temperature variables at the
global scale over the land, and also over many large land areas.
Globally, return periods for events that were expected to recur once
every 20 years in the 1960s are now estimated to exceed 30 years for
extreme annual minimum daily maximum temperature and 35 years for
extreme annual minimum daily minimum temperature, although these
estimates are subject to considerable uncertainty. Further, return peri-
ods were found to have decreased to less than 10 or 15 years for annual
maximum daily minimum and daily maximum temperatures respectively
(Figure 3-2).
However, the available detection and attribution studies for extreme
maximum and minimum temperatures (Christidis et al., 2011b; Zwiers
et al., 2011) suggest that the models overestimate changes in the
maximum temperatures and underestimate changes in the minimum
temperatures during the late 20th century.
Projected Changes and Uncertainties
Regarding projections of extreme temperatures, the AR4 (Meehl et al.,
2007b) noted that cold episodes were projected to decrease significantly
in a future warmer climate and considered it very likely that heat waves
would be more intense, more frequent, and last longer in a future warmer
climate. Post-AR4 studies of temperature extremes have utilized larger
model ensembles (Kharin et al., 2007; Sterl et al., 2008; Orlowsky and
Seneviratne, 2011) and generally confirm the conclusions of the AR4, while
also providing more specific assessments both in terms of the range of
considered extremes and the level of regional detail (see also Table 3-3).
There are few global analyses of multi-model projections of temperature
extremes available in the literature. The study by Tebaldi et al. (2006),
which provided the basis for extreme projections given in the AR4
(Figures 10.18 and 10.19 in Meehl et al., 2007b), provided global analyses
of projected changes (A1B scenario) in several extremes indices based
on nine GCMs (note that not all modelling groups that saved daily data
also calculated the indices). For temperature extremes, analyses were
provided for heat wave lengths (using only one index, see discussion in
Box 3-1) and warm nights. Stippling was used where five out of nine
models displayed statistically significant changes of the same sign.
Orlowsky and Seneviratne (2011) recently updated the analysis from
Tebaldi et al. (2006) for the full ensemble of GCMs that contributed A2
scenarios to the CMIP3, using a larger number of extreme indices
[including several additional analyses of daily extremes (see Figures 3-3
and 3-4), and three heat wave indices instead of one; see also discussion
of heat wave indices in Box 3-1], using other thresholds for display and
stippling of the figures (no results displayed if less than 66% of the
models agree on the sign of change; stippling used only for 90% model
agreement), and providing seasonal analyses. This analysis confirms
that strong agreement (in terms of sign of change) exists between the
various GCM projections for temperature-related extremes, with
projected increases in warm day occurrences (Figure 3-3) and heat wave
length, and decreases in cold extremes (Figure 3-4). Temperature
extremes on land are projected to warm faster than global annual mean
temperature in many regions and seasons, implying large changes in
extremes in some places, even for a global warming of 2 or 3°C (with
scaling factors for the SRES A2 scenario ranging between 0.5 and 2 for
moderate seasonal extremes; Orlowsky and Seneviratne, 2011). Based
on the analyses of Tebaldi et al. (2006) and Orlowsky and Seneviratne
(2011), as well as physical considerations, we assess that increases in
the number of warm days and nights and decreases in the number of
cold days and nights (defined with respect to present regional climate,
i.e., the 1961-1990 reference period, see Box 3-1) are virtually certain at
the global scale. Further, given the assessed changes in hot and cold
days and nights and available analyses of projected changes in heat
wave length in the two studies, we assess that it is very likely that the
length, frequency, and/or intensity of heat waves will increase over
most land areas.
Another global study of changes in extremes based on the CMIP3
ensemble is provided in Kharin et al. (2007), which focuses on changes
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
137
in annual extremes (20-year extreme values) based on 12 GCMs for
temperature extremes and 14 GCMs for precipitation extremes employing
the SRES A2, A1B, and B1 emissions scenarios. This analysis projects
increases in the temperature of the 1-in-20 year annual extreme hottest
day of about 2 to 6°C (depending on region and scenario; Figure 3-5
adapted from Kharin et al., 2007) and strong reductions in the return
periods of this extreme event by the end of the 21st century. However,
as noted above, the limited number of relevant detection and attribution
studies suggests that models may overestimate some changes in
temperature extremes, and our assessments take this into account by
reducing the level of certainty in the assessments from what would be
derived by uncritical acceptance of the projections in Figure 3-5. The
assessments are also weakened to reflect the possibility that some
important processes relevant to extremes may be missing or be poorly
represented in models, as well as the fact that the model projections
considered in this study did not correspond to the full CMIP3 ensemble.
Hence, we assess that in terms of absolute values, the 20-year extreme
annual daily maximum temperature (i.e., return value) will likely
increase by about 2 to 5°C by the late 21st century, and by about 1 to
3°C by mid-21st century, depending on the region and emissions scenario
(considering the B1, A1B, and A2 scenarios; Figure 3-5a). Furthermore,
we assess that globally under the A2 and A1B scenarios a 1-in-20 year
annual extreme hot day is likely to become a 1-in-2 year annual extreme
by the end of the 21st century in most regions, except in the high latitudes
of the Northern Hemisphere where it is likely to become a 1-in-5 year
annual extreme (Figure 3-5b, based on material from Kharin et al.,
2007). Further, we assess that under the more moderate B1 scenario a
current 1-in-20 year extreme would likely become a 1-in-5 year event
(and a 1-in-10 year event in Northern Hemisphere high latitudes).
Next, regional assessments of projected changes in temperature extremes
are provided. More details are found in Table 3-3. For North America, the
CCSP reached the following conclusions (using IPCC AR4 likelihood
terminology) regarding projected changes in temperature extremes by
the end of the 21st century (Gutowski et al., 2008a):
1) Abnormally hot days and warm nights and heat waves are very likely
to become more frequent.
2) Cold days and cold nights are very likely to become much less
frequent.
3) For a mid-range scenario (A1B) of future greenhouse gas emissions,
a day so hot that it is currently experienced only once every 20
years would occur every 3 years by the middle of the century over
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Fraction of Warm Days
ANN
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●
●●●●●● ●● ●●●●●●●●●●●●●●
● ●●●●●● ●● ●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●
●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●
●●●●●●●●●● ● ●● ●●● ●●●●●●●●●●●●●●●
●●●●●● ●●●
●●●●●●●●●●●
●●●●● ●●●● ●●●●
●●●● ●●●●●
●●●●●●●●
●●●●●●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●
DJF
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●● ● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●
●●●●●● ●● ●●●●●●●●●●●●●●
●● ●● ●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●
●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●
●●●●●●●●●● ● ●● ●●● ●●●●●●●●●●●●●●●
●●●●●● ●●●
●●●●●●●●●●●
●●●●● ●●●● ●●●●
●●●● ●●●●●
●●●●●●●●
●●●●●●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●● ● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●
JJA
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●● ● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●● ●● ●● ●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●●●●●●●● ● ●●●●●●●●●●●●
●●●●●● ●● ●●●●●●●●●●●●●●
● ●●●●●● ●● ●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●
●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●
●●●●●●●●●● ● ●● ●●● ●●●●●●●●●●●●●●●
●●●●●● ●●●
●●●●●●●●●●●
●●●●● ●●●● ●●●●
●●●● ●●●●●
●●●●●●●●
●●●●●●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●● ● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●● ●● ●● ●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●●●●●●●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●
Standard Deviation
6 3036
Fraction of Cold Days
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●●●●●●●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●
●●●●●● ●● ●●●●●●●●●●●●●●
● ●●●●●● ●● ●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●
●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●
●●●●●●●●●● ● ●● ●●● ●●●●●●●●●●●●●●●
●●●●●● ●●●
●●●●●●●●●●●
●●●●● ●●●● ●●●●
●●●● ●●●●●
●●●●●●●●
●●●●●●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●●●●●●●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●
●●●●●● ●● ●●●●●●●●●●●●●●
● ●●●●●● ●● ●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●
●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●
●●●●●●●●●● ● ●● ●●● ●●●●●●●●●●●●●●●
●●●●●● ●●●
●●●●●●●●●●●
●●●●● ●●●● ●●●●
●●●● ●●●●●
●●●●●●●●
●●●●●●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●● ●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ●●
●●●●●● ●●●●●●●●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●
●●●●●● ●● ●●●●●●●●●●●●●●
● ●●●●●● ●● ●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●
●●● ●●●●●●● ●● ●●●●●●●●●●●●●●●●●●●
●●●●●●●●●● ● ●● ●●● ●●●●●●●●●●●●●●●
●●●●●● ●●●
●●●●●●●●●●●
●●●●● ●●●● ●●●●
●●●● ●●●●●
●●●●●●●●
●●●●●●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●● ●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●
Standard Deviation
3.25 1.75 0 1 3.25
Percentage Days with Tmax>30
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●● ●●
●●●●● ●●● ●●●●● ●●●
●●●●● ●● ●●●●● ●● ●
●●●●● ●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●●●●●●●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ●●●●●●●●●●●●●●●●●
●●●● ●●●●●●●●●●●● ●●●●●
●●● ● ●●●●●●●●
●●● ●●●●
● ●●●●●●●●●●●
●●● ● ●●
●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●● ●●
●●●●● ●●● ●●●●● ●●●
●●●●● ●● ●●●●● ●● ●
●●●●● ●●
●●●●● ●●
●●●●●● ●●●●●●●●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ●●●●●●●●●●●●●●●●●
●●●● ●●●●●●●●●●●● ●●●●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●● ●●● ●●●●
●●● ●●● ●●●
●●● ●●●●●
●●●● ●●● ●●
●●●●● ●●●
●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●
●●● ●● ●●
●● ●●
●●● ●●●●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●
●●●●●●●●●●●●●
●● ●● ●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●● ●●● ●●●●
●●● ●●● ●●●
●●● ●●●●●
●●●● ●●● ●●
●●●●● ●●●
●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●
●●● ●● ●●
●● ●●
●●● ●●●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●
●●●●●●●●●●●●●
●● ●● ●● ●●
●●
●● ●●●
●● ●● ●●●●
●● ●●●●●
●●● ●●● ●●●●
●●● ●●● ●●
●●● ●●●●●
●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●● ●●
●●● ●● ●●
●●
●● ●●●●
●●● ●●●●●●
●● ●●●● ●●
●● ●●●●
●●●●● ●●● ● ●●● ●●●
●●●●● ●● ●●●●● ●● ●
●●●●● ●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ●●●●●●●●●●●●●●●●●
●●●● ●●●●●●●●●●●● ●●●●●
●●●● ●●●●●●●●● ●
●●● ●●●●●●●● ●
●●●●●●●●●●●
●●● ● ●●
●●
●●
●● ●●●
●● ●● ●●●●
●● ●●●●●
●●● ●●● ●●●●
●●● ●●● ●●
●●● ●●●●●
●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●● ●●
●●● ●● ●●
●●
●● ●●●●
●●● ●●●●●●
●● ●●●● ●●
●● ●●●●● ● ● ●●
●●●●● ●●● ● ●●● ●●●
●●●●● ●● ●●●●● ●● ●
●●●●● ●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ●●●●●●●●●●●●●●●●●
●●●● ●●●●●●●●●●●● ●●●●●
Percentage of Days
40 20 0 20 40
Figure 3-3 | Projected annual and seasonal changes in three indices for daily Tmax for 2081-2100 with respect to 1980-1999, based on 14 GCMs contributing to the CMIP3.
Left column: fraction of warm days (days in which Tmax exceeds the 90th percentile of that day of the year, calculated from the 1961-1990 reference period); middle column:
fraction of cold days (days in which Tmax is lower than the 10th percentile of that day of the year, calculated from the 1961-1990 reference period); right column: percentage of
days with Tmax >30°C. The changes are computed for the annual time scale (top row) and two seasons (December-January-February, DJF, middle row, and June-July-August, JJA,
bottom row) as the fractions/percentages in the 2081-2100 period (based on simulations for emission scenario SRES A2) minus the fractions/percentages of the 1980-1999 period
(from corresponding simulations for the 20th century). Warm day and cold day changes are expressed in units of standard deviations, derived from detrended per year annual or
seasonal estimates, respectively, from the three 20-year periods 1980-1999, 2046-2065, and 2081-2100 pooled together. Tmax >30°C changes are given directly as differences in
percentage points. Color shading is only applied for areas where at least 66% (i.e., 10 out of 14) of the GCMs agree on the sign of the change; stippling is applied for regions
where at least 90% (i.e.,13 out of 14) of the GCMs agree on the sign of the change. Adapted from Orlowsky and Seneviratne (2011); updating Tebaldi et al. (2006) for additional
number of indices and CMIP3 models, and including seasonal time frames. For more details, see Appendix 3.A.
138
much of the continental United States and every 5 years over most
of Canada; by the end of the century, it would occur every other
year or more.
Meehl et al. (2009c) examined changes in record daily high and low
temperatures in the United States and show that even with projected
strong warming resulting in many more record highs than lows, the
occasional record low is still set. For Australia, the CMIP3 ensemble
projected increases in warm nights (15-40% by the end of the 21st
century) and heat wave duration, together with a decrease in the number
of frost days (Alexander and Arblaster, 2009). Inland regions show
greater warming compared with coastal zones (Suppiah et al., 2007;
Alexander and Arblaster, 2009) and large increases in the number of
days above 35 or 40°C are indicated (Suppiah et al., 2007). For the
entire South American region, a study with a single RCM projected more
frequent warm nights and fewer cold nights (Marengo et al., 2009a).
Several studies of regional and global model projections of changes in
extremes are available for the European continent (see also Table 3-3).
Analyses of both global and regional model outputs show major
increases in warm temperature extremes across the Mediterranean
region including events such as hot days (Tmax >30°C) and tropical
nights (Tmin>20°C) (Giannakopoulos et al., 2009; Tolika et al., 2009).
Comparison of RCM projections using the A1B forcing scenario, with
data for 2007 (the hottest summer in Greece in the instrumental record
with a record daily Tmax observed value of 44.8°C) indicates that the
distribution for 2007 is closer to the distribution for 2071-2100 than for
the 2021-2050 period, thus 2007 might be considered a ‘normal’ summer
of the future (Founda and Giannakopoulos, 2009; Tolika et al., 2009).
Beniston et al. (2007) concluded from an analysis of RCM output that
regions such as France and Hungary may experience as many days per
year above 30°C as currently experienced in Spain and Sicily. In this
RCM ensemble, France was the area with the largest projected warming
in the uppermost percentiles of daily summer temperatures although
the mean warming was greatest in the Mediterranean region (Fischer
and Schär, 2009). New results from an RCM ensemble project increases
in the amplitude, frequency, and duration of health-impacting heat waves,
especially in southern Europe (Fischer and Schär, 2010). Overall these
regional assessments are consistent with the global assessments provided
above. It should be noted, however, that the assessed uncertainty is larger
at the regional level than at the continental or global level (see Box 3-2).
Global-scale trends in a specific extreme may be either more reliable or
less reliable than regional-scale trends, depending on the geographical
uniformity of the trends in the specific extreme (Section 3.1.6).
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Fraction of Warm Nights
ANN
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●●●●●●●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●
●●●●●● ●● ●●●●●●●●●●●●●●
●● ●● ●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●
●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●
●●●●●●●●●● ● ●● ●●● ●●●●●●●●●●●●●●●
●●●●●● ●●●
●●●●●●●●●●●
●●●●● ●●●● ●●●●
●●●● ●●●●●
●●●●●●●●
●●●●●●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●●●●●●●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●
DJF
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●●●●●●●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●
●●●●●● ●● ●●●●●●●●●●●●●●
●● ●● ●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●
●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●
●●●●●●●●●● ● ●● ●●● ●●●●●●●●●●●●●●●
●●●●●● ●●● ●●●●●●●●●●●
●●●●● ●●●● ●●●●
●●●● ●●●●●
●●●●●●●●
●●●●●●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●●●●●●●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
JJA
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●●●●●●●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●
●●●●●● ●● ●●●●●●●●●●●●●●
●● ●● ●●●●●●●●●● ●●●●
●●●●●●●●● ●●●●●●●●●●●●●● ●●●●●●
●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●
●●●●●●●●●● ● ●● ●●● ●●●●●●●●●●●●●●●
●●●●●● ●●●
●●●●●●●●●●●
●●●●● ●●●● ●●●●
●●●● ●●●●●
●●●●●●●●
●●●●●●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●●●●●●●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●
Standard Deviation
8 4048
Fraction of Cold Nights
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●
●●●●●● ●● ●●●●●●●●●●●●●●
●● ●● ●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●
●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●
●●●●●●●●●● ● ●● ●●● ●●●●●●●●●●●●●●●
●●●●●● ●●●
●●●●●●●●●●●
●●●●● ●●●● ●●●●
●●●● ●●●●●
●●●●●●●●
●●●●●●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●● ●●
●●●●● ● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●
●●●●●● ●● ●●●●●●●●●●●●●●
●● ●● ●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●
●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●
●●●●●●●●●● ● ●● ●●● ●●●●●●●●●●●●●●●
●●●●●● ●●● ●●●●●●●●●●●
●●●●● ●●●● ●●●●
●●●● ●●●●●
●●●●●●●●
●●●●●●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●●● ●●● ●●
●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●● ●●
●●●●● ● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●
●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●● ●●● ●●
●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●
●●●●●● ●● ●●●●●●●●●●●●●●
●●●● ●● ●●●●●●●●●●●●●●●
●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●
●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●
●●●●●●●●●● ● ●● ●●● ●●●●●●●●●●●●●●●
●●●●●● ●●●
●●●●●●●●●●●
●●●●● ●●●● ●●●●
●●●● ●●●●●
●●●●●●●●
●●●●●●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●●● ●●● ●●●●
●●● ●●● ●●●
●●●● ●●●●●
●●●●● ●●● ●●
●●●●● ●●●
●●●●●● ●●●●
●●●● ●●● ●●
●●● ●●●● ●●●
●●●● ●●●●
●●● ●●●● ●●
●● ●●●●●●●
●●● ●●●●●●●● ●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●●●●●●●
●●●●● ●●● ●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●● ●
●●●●● ●● ●●●●●●●●●
●●●●● ● ●● ●●●●●●●●
●●●●●● ●●
●●●●●●● ●●●●●● ●●●●●●●●●●
●●●●●●● ● ●●●●●●●●●●●●●●●●●
●●●●● ●● ●●●●●●●●●●●●●●●●●●
●●
Standard Deviation
3 1.5 0 1.5 3
Percentage Days with Tmin>20
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●● ●●● ●●●●
●●● ●●● ●●●
●●● ●●●●●
●●●●● ●●●
●●●●● ●●●
●●●●●● ●●●
●● ●●●
●●●
●●●●
●●● ●●●●
●● ●●●●●●●
●●●●●●●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●● ●●
●●●●● ●●● ●●●●● ●●●
●●●●● ●● ●●●● ●● ●
●●●●● ●● ●●
●●●●● ● ●● ● ●●●●●
●● ●●● ●●●●●
●●●●●●● ●●●●●● ●●●●● ● ●●
●●●●● ●●●●●●●●●● ●●●
●●●● ●●●●●●●●●●● ●●●
●● ●●
●●●●●●●
●●●●●● ● ●●
●●● ●
●●
●●
●●●●●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●●● ●● ●●●●●
●●● ●●● ●●●●
●●● ●●● ●●●
●●● ●●●●●
●●●●● ●●●
●●●●● ●●●
●●●●●● ●●●
●● ●●●
●●●
●●●●
●●● ●●●●
●● ●●●●●●●
●●●●●●●●
●●●●●●●
●●●●●●●● ● ●
●●●●●● ● ● ●
●●●●●●●● ●●●●
●●●●●● ●● ●●●●●●
●●●●●●●●●●●●●●●●
●● ●●●●●●●●●●● ●●
●●●●● ●●● ●●●●● ●●●
●●●●● ●● ●●●● ●● ●
●●●●● ●● ●●
●●●●● ●●●●
●● ●●● ●●●●●
●●●●●●● ●●●●●● ●●●●● ● ●●
●●●●● ●●●●●●●●●● ●●●
●●●● ●●●●●●●●●●● ●●●
●●
●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●● ●● ●●●●
●●● ●●● ●●
●● ●●●
●● ●●●
●● ●●●
●●●
●●●
●●●
●●●
●●●●
●● ●●●●
●● ●●●●●●●
●●●●●● ●
●●●●●●●
●●●●●●●● ● ●
●●● ● ● ● ●
●●
●● ●●
●●
●●
●● ●●
●●
●●● ●●
●●● ●●●●●
●● ●● ●●●●
●● ●● ●●●●
●●● ●●● ●●
●● ●●●
●● ●●●
●● ●●●
●●●
●●●
●●●
●●●
●●●●
●● ●●●●
●● ●●●●●●●
●●●●●● ●
●●●●●●●
●●●●●●●● ● ●
●●● ● ● ● ●
●●
●● ●●
●●
●●
●●
●●
●● ●●●●●
●●● ●●●●
●● ●●●
●●● ●● ●●
●●● ● ●●
●●●●● ●●●
●● ●●● ●●●
●● ●●
●● ●●●
●●
●● ●●●●
●● ●●●●●
●●●●
●●
●●
●● ●●
●● ●●●●
●● ●● ●●● ●● ●●
●●●●● ●●● ● ●●● ●●●
●●●●● ●● ●●●● ●● ●
●●●●● ●● ●●●
●●●●● ●●●●
●● ●●
●●●●●●● ●●●●●● ●●●●● ● ●●
●●●●● ●●●●●●●●●● ●●●
●●●● ●●●●●●●●●● ●●●
●● ●●
●●●●●●●
●●●●●● ● ●●
●● ●●
●●
●●
●●●●●●
●●
●●
●● ●●●●●
●●● ●●●●
●● ●●●
●●● ●● ●●
●●● ● ●●
●●●●● ●●●
●● ●●● ●●●
●● ●●
●● ●●●
●●
●● ●●●●
●● ●●●●●
●●●●
●●
●●
●● ●●
●● ●●●●
●● ●● ●●● ●● ●●
●●●●● ●●● ● ●●● ●●●
●●●●● ●● ●●●● ●● ●
●●●●● ●● ●●●
●●●●● ●●●●
● ●●●● ●●● ●●●●●
●●●●●●● ●●●●●● ●●●●● ● ●●
●●●●● ●●●●●●●●●● ●●●
●●●● ●●●●●●●●●● ●●●
●●
●●
Percentage of Days
28 14 0 14 28
Figure 3-4 | Projected annual and seasonal changes in three indices for daily Tmin for 2081-2100 with respect to 1980-1999, based on 14 GCMs contributing to the CMIP3.
Left column: fraction of warm nights (days at which Tmin exceeds the 90th percentile of that day of the year, calculated from the 1961-1990 reference period); middle column:
fraction of cold nights (days at which Tmin is lower than the 10th percentile of that day of the year, calculated from the 1961-1990 reference period); right column: percentage of
days with Tmin >20°C. The changes are computed for the annual time scale (top row) and two seasons (December-January-February, DJF, middle row, and June-July-August, JJA,
bottom row) as the fractions/percentages in the 2081-2100 period (based on simulations under emission scenario SRES A2) minus the fractions/percentages of the 1980-1999
period (from corresponding simulations for the 20th century). Warm night and cold night changes are expressed in units of standard deviations, derived from detrended per year
annual or seasonal estimates, respectively, from the three 20-year periods 1980-1999, 2046-2065, and 2081-2100 pooled together. Tmin >20°C changes are given directly as
differences of percentage points. Color shading is only applied for areas where at least 66% (i.e., 10 out of 14) of the GCMs agree in the sign of the change; stippling is applied
for regions where at least 90% (i.e.,13 out of 14) of the GCMs agree in the sign of the change. Adapted from Orlowsky and Seneviratne (2011); updating Tebaldi et al. (2006) for
additional number of indices and CMIP3 models, and including seasonal time frames. For more details, see Appendix 3.A.
139
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Full model range
Central 50%
intermodel range
Median
B1 A1B A2
Scenarios:
Temperature change
(°C)
204665 208100
0
2
4
6
8
10
Legend
204665 208100
0
2
4
6
8
10
Alaska/N.W. Canada - 1
E. Canada/Greenl./Icel. - 2
204665 208100
0
2
4
6
8
10
W. North America - 3
204665 208100
0
2
4
6
8
10
C. North America - 4
204665 208100
0
2
4
6
8
10
11
E. North America - 5
204665 208100
0
2
4
6
8
10
14
Central America/Mexico - 6
204665 208100
0
2
4
6
8
10
Amazon - 7
204665 208100
0
2
4
6
8
10
N.E. Brazil - 8
204665 208100
0
2
4
6
8
10
W. Coast South America - 9
204665 208100
0
2
4
6
8
10
S.E. South America - 10
204665 208100
0
2
4
6
8
10
N. Europe - 11
204665 208100
0
2
4
6
8
10
C. Europe - 12
204665 208100
0
2
4
6
8
10
11
S. Europe/Mediterranean - 13
204665 208100
0
2
4
6
8
10
Sahara - 14
204665 208100
0
2
4
6
8
10
W. Africa - 15
204665 208100
0
2
4
6
8
10
E. Africa - 16
204665 208100
0
2
4
6
8
10
S. Africa - 17
204665 208100
0
2
4
6
8
10
N. Asia - 18
204665 208100
0
2
4
6
8
10
E. Asia - 22
204665 208100
0
2
4
6
8
10
Tibetan Plateau - 21
204665 208100
0
2
4
6
8
10
C. Asia - 20
204665 208100
0
2
4
6
8
10
W. Asia - 19
204665 208100
0
2
4
6
8
10
S. Asia - 23
204665 208100
0
2
4
6
8
10
S.E. Asia - 24
204665 208100
0
2
4
6
8
10
N. Australia - 25
204665 208100
0
2
4
6
8
10
S. Australia/New Zealand - 26
204665 208100
0
2
4
6
8
10
Globe (Land only)
204665 208100
0
2
4
6
8
10
20-year return value (°C)
204665 208100
0
2
4
6
8
10
mean (°C)
Figure 3-5a | Projected changes (in °C) in 20-year return values of the annual maximum of the daily maximum temperature. The bar plots (see legend for more information) show results for regionally averaged projections for two
time horizons, 2046 to 2065 and 2081 to 2100, as compared to the late 20th century (1981-2000), and for three different SRES emission scenarios (B1, A1B, A2). Results are based on 12 GCMs contributing to the CMIP3. See
Figure 3-1 for defined extent of regions. Values are computed for land points only. The ‘Globe’ analysis (inset box) displays the change in 20-year return values of the annual maximum of the daily maximum temperature computed
using all land grid points (left), and the change in annual mean daily maximum temperature computed using all land grid points (right). Adapted from the analysis of Kharin et al. (2007). For more details, see Appendix 3.A.
140
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Full model range
Central 50%
intermodel range
Median
B1 A1B A2
Scenarios:
204665 208100
1
2
5
10
20
Return period (Years)
Legend
204665 208100
1
2
5
10
20
22
Alaska/N.W. Canada - 1
204665 208100
1
2
5
10
20
E. Canada/Greenl./Icel. - 2
204665 208100
1
2
5
10
20
W. North America - 3
204665 208100
1
2
5
10
20
C. North America - 4
204665 208100
1
2
5
10
20
E. North America - 5
204665 208100
1
2
5
10
20
Central America/Mexico - 6
204665 208100
1
2
5
10
20
Amazon - 7
204665 208100
1
2
5
10
20
N.E. Brazil - 8
204665 208100
1
2
5
10
20
W. Coast South America - 9
204665 208100
1
2
5
10
20
S.E. South America - 10
204665 208100
1
2
5
10
20
31
24
23
N. Europe - 11
204665 208100
1
2
5
10
20
C. Europe - 12
204665 208100
1
2
5
10
20
S. Europe/Mediterranean - 13
204665 208100
1
2
5
10
20
Sahara - 14
204665 208100
1
2
5
10
20
W. Africa - 15
204665 208100
1
2
5
10
20
E. Africa - 16
204665 208100
1
2
5
10
20
S. Africa - 17
204665 208100
1
2
5
10
20
N. Asia - 18
204665 208100
1
2
5
10
20
E. Asia - 22
204665 208100
1
2
5
10
20
Tibetan Plateau - 21
204665 208100
1
2
5
10
20
C. Asia - 20
204665 208100
1
2
5
10
20
W. Asia - 19
204665 208100
1
2
5
10
20
S. Asia - 23
204665 208100
1
2
5
10
20
S.E. Asia - 24
204665 208100
1
2
5
10
20
N. Australia - 25
204665 208100
1
2
5
10
20
S. Australia/New Zealand - 26
204665 208100
1
2
5
10
20
Globe (Land only)
Figure 3-5b | Projected return period (in years) of late 20th-century 20-year return values of the annual maximum of the daily maximum temperature. The bar plots (see legend for more information) show results for regionally
averaged projections for two time horizons, 2046 to 2065 and 2081 to 2100, as compared to the late 20th century (1981-2000), and for three different SRES emission scenarios (B1, A1B, A2). Results are based on 12 GCMs
contributing to the CMIP3. See Figure 3-1 for defined extent of regions. The ‘Globe’ analysis (inset box) displays the projected return period (in years) of late 20th-century 20-year return values of the annual maximum of the daily
maximum temperature computed using all land grid points. Adapted from the analysis of Kharin et al. (2007). For more details, see Appendix 3.A.
141
Temperature extremes were the type of extremes projected to change
with most confidence in the AR4 (IPCC, 2007a). This is confirmed
regarding the sign of change with more recent analyses (Figures 3-3
and 3-4), although there is a large spread with respect to the magnitude
of changes both due to emission scenario and climate model uncertainty
(Figures 3-5a,b). If changes in temperature extremes scale with changes
in mean temperature (i.e., simple shifts of the probability distribution),
we infer that it is virtually certain that hot extremes will increase and
cold extremes will decrease over the 21st century with respect to the
1960-1990 climate. Changes in the tails of the temperature distributions
may not scale with changes in the mean in some regions (Section 3.1.6),
though in most such reported cases hot extremes tend to increase and
cold extremes decrease more than mean temperature, and thus the
above statement for extremes (virtually certain increase in hot extremes
and decrease in cold extremes) still applies. Central and eastern Europe
is a region where the evidence suggests that projected changes in
temperature extremes result from both changes in the mean as well as
from changes in the shape of the probability distributions (Schär et al.,
2004). The main mechanism for the widening of the distribution is
linked to the drying of the soil in this region (Sections 3.1.4 and 3.1.6).
Furthermore, remote surface heating may induce circulation changes
that modify the temperature distribution (Haarsma et al., 2009). Other
local, mesoscale, and regional feedback mechanisms, in particular with
land surface conditions (beside soil moisture, also with vegetation and
snow; Section 3.1.4) and aerosol concentrations (Ruckstuhl and Norris,
2009) may enhance the uncertainties in temperature projections. Some
of these processes occur at a small scale unresolved by the models
(Section 3.2.3). In addition, lack of observational data (e.g., for soil
moisture and snow cover; see Section 3.2.1) reduces the possibilities to
evaluate climate models (e.g., Roesch, 2006; Boe and Terray, 2008; Hall
et al., 2008; Brown and Mote, 2009). Because of these various processes
and associated uncertainties, mean global warming does not necessarily
imply warming in all regions and seasons (see also Section 3.1.6).
Regarding mesoscale processes, lack of information also affects
confidence in projections. One example is changes in heat waves in the
Mediterranean region that are suggested to have the largest impact in
coastal areas, due to the role of enhanced relative humidity in health
impacts (Diffenbaugh et al., 2007; Fischer and Schär, 2010). But it is not
clear how this pattern may or may not be moderated by sea breezes
(Diffenbaugh et al., 2007).
In summary, since 1950 it is very likely that there has been an
overall decrease in the number of cold days and nights and an
overall increase in the number of warm days and nights at the
global scale, that is, for land areas with sufficient data. It is likely
that such changes have also occurred at the continental scale in
North America, Europe, and Australia. There is medium confidence
in a warming trend in daily temperature extremes in much of Asia.
Confidence in historical trends in daily temperature extremes in
Africa and South America generally varies from low to medium
depending on the region. Globally, in many (but not all) regions
with sufficient data there is medium confidence that the length or
number of warm spells or heat waves has increased since the
middle of the 20th century. It is likely that anthropogenic
influences have led to warming of extreme daily minimum and
maximum temperatures at the global scale. Models project
substantial warming in temperature extremes by the end of the
21st century. It is virtually certain that increases in the frequency
and magnitude of warm days and nights and decreases in the cold
days and nights will occur through the 21st century at the global
scale. This is mostly linked with mean changes in temperatures,
although changes in temperature variability can play an important
role in some regions. It is very likely that the length, frequency,
and/or intensity of warm spells or heat waves (defined with
respect to present regional climate) will increase over most land
areas. For the SRES A2 and A1B emission scenarios a 1-in-20 year
annual hottest day is likely to become a 1-in-2 year annual
extreme by the end of the 21st century in most regions, except in
the high latitudes of the Northern Hemisphere where it is likely
to become a 1-in-5 year annual extreme. In terms of absolute
values, 20-year extreme annual daily maximum temperature (i.e.,
return value) will likely increase by about 1 to 3°C by mid-21st
century and by about 2 to 5°C by the late 21st century, depending
on the region and emissions scenario (Figure 3-5). Moderate
temperature extremes on land are projected to warm faster than
global annual mean temperature in many regions and seasons.
Projected changes at subcontinental scales are less certain than
is the case for the global scale. Regional changes in temperature
extremes will differ from the mean global temperature change.
Mean global warming does not necessarily imply warming in all
regions and seasons.
3.3.2. Precipitation
This section addresses changes in daily extreme or heavy precipitation
events. Reductions in mean (or total) precipitation that can lead to
drought (i.e., associated with lack of precipitation) are considered in
Section 3.5.1.
Because climates are so diverse across different parts of
the world, it is difficult to provide a single definition of extreme or heavy
precipitation. In general, two different approaches have been used:
(1) relative thresholds such as percentiles (typically the 95th percentile)
and return values; and (2) absolute thresholds [e.g., 50.8 mm (2 inches)
day
-1
of rain in the United States, and 100 mm day
-1
of rain in China].
For more details on the respective drawbacks and advantages of these
two approaches, see Section 3.1 and Box 3-1. Note that we do not
distinguish between rain and snowfall (both considered as contributors
to overall extreme precipitation events) as they are not treated separately
in the literature, but do distinguish changes in hail from other precipitation
types. Increases in public awareness and changes in reporting practices
have led to inconsistencies in the record of severe thunderstorms and
hail that make it difficult to detect trends in the intensity or frequency
of these events (Kunkel et al., 2008). Furthermore, weather events such
as hail are not well captured by current monitoring systems and, in
some parts of the world, the monitoring network is very sparse (Section
3.2.1), resulting in considerable uncertainty in the estimates of extreme
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
142
precipitation. There are also known biases in precipitation measurements,
mostly leading to rain undercatch. Little evidence of paleoclimatic and
historical changes in heavy precipitation is available to place recent
variations into context.
Observed Changes
The AR4 (Trenberth et al., 2007) concluded that it was likely that there
had been increases in the number of heavy precipitation events (e.g.,
95th percentile) over the second half of the 20th century within many
land regions, even in those where there had been a reduction in total
precipitation amount, consistent with a warming climate and observed
significant increasing amounts of water vapor in the atmosphere.
Increases had also been reported for rarer precipitation events (1-in-50
year return period), but only a few regions had sufficient data to assess
such trends reliably. However, the AR4 (Trenberth et al., 2007) also stated
that “Many analyses indicate that the evolution of rainfall statistics
through the second half of the 20th century is dominated by variations
on the interannual to inter-decadal time scale and that trend estimates
are spatially incoherent (Manton et al., 2001; Peterson et al., 2002; Griffiths
et al., 2003; Herath and Ratnayake, 2004). Overall, as highlighted in
Alexander et al. (2006), the observed changes in precipitation extremes
were found at the time to be much less spatially coherent and statistically
significant compared to observed changes in temperature extremes:
although statistically significant trends toward stronger precipitation
extremes were generally found for a larger fraction of the land area
than trends toward weaker precipitation extremes, statistically significant
changes in precipitation indices for the overall land areas with data
were only found for the Simple Daily Intensity index, and not for other
considered indices such as Heavy Rainfall Days (Alexander et al., 2006).
Recent studies have updated the assessment of the AR4, with more
regional results now available (Table 3-2). Overall, this additional evidence
confirms that more locations and studies show an increase than a
decrease in extreme precipitation, but that there are also wide regional
and seasonal variations, and trends in many regions are not statistically
significant (Table 3-2).
Recent studies on past and current changes in precipitation extremes in
North America, some of which are included in the recent assessment of
the CCSP report (Kunkel et al., 2008), have reported an increasing trend
over the last half century. Based on station data from Canada, the
United States, and Mexico, Peterson et al. (2008a) reported that heavy
precipitation has been increasing over 1950-2004, as well as the average
amount of precipitation falling on days with precipitation. For the
contiguous United States, DeGaetano (2009) showed a 20% reduction
in the return period for extreme precipitation of different return levels
over 1950-2007; Gleason et al. (2008) reported an increasing trend in
the area experiencing a much above-normal proportion of heavy daily
precipitation from 1950 to 2006; and Pryor et al. (2009) provided evidence
of increases in the intensity of events above the 95th percentile during
the 20th century, with a larger magnitude of the increase at the end of the
century. The largest trends toward increased annual total precipitation,
number of rainy days, and intense precipitation (e.g., fraction derived
from events in excess of the 90th percentile value) were focused on the
Great Plains/northwestern Midwest (Pryor et al., 2009). In the core of
the North American monsoon region in northwest Mexico, statistically
significant positive trends were found in daily precipitation intensity
and seasonal contribution of daily precipitation greater than its 95th
percentile in the mountain sites for the period 1961-1998. However, no
statistically significant changes were found in coastal stations (Cavazos
et al., 2008). Overall, the evidence indicates a likely increase in observed
heavy precipitation in many regions in North America, despite statistically
non-significant trends and some decreases in some subregions (Table 3-2).
This general increase in heavy precipitation accompanies a general
increase in total precipitation in most areas of the country.
There is low to medium confidence in trends for Central and South
America, where spatially varying trends in extreme rainfall events have
been observed (Table 3-2). Positive trends in many areas but negative
trends in some regions are evident for Central America and northern
South America (Dufek and Ambrizzi, 2008; Marengo et al., 2009b; Re
and Ricardo Barros, 2009; Sugahara et al., 2009). For the western coast
of South America, a decrease in extreme rainfall in many areas and an
increase in a few areas are observed (Haylock et al., 2006b).
There is medium confidence in trends in heavy precipitation in Europe,
due to partly inconsistent signals across studies and regions, especially
in summer (Table 3-2). Winter extreme precipitation has increased in part
of the continent, in particular in central-western Europe and European
Russia (Zolina et al., 2009), but the trend in summer precipitation has
been weak or not spatially coherent (Moberg et al., 2006; Bartholy and
Pongracz, 2007; Maraun et al., 2008; Pavan et al., 2008; Zolina et al.,
2008; Costa and Soares, 2009; Kyselý, 2009; Durão et al., 2010; Rodda
et al., 2010). Increasing trends in 90th, 95th, and 98th percentiles of daily
winter precipitation over 1901-2000 were found (Moberg et al., 2006),
which has been confirmed by more detailed country-based studies for
the United Kingdom (Maraun et al., 2008), Germany (Zolina et al.,
2008), and central and eastern Europe (Bartholy and Pongracz, 2007;
Kyselý, 2009), while decreasing trends have been found in some regions
such as northern Italy (Pavan et al., 2008), Poland (Lupikasza, 2010),
and some Mediterranean coastal sites (Toreti et al., 2010). Uncertainties
are overall larger in southern Europe and the Mediterranean region,
where there is low confidence in the trends (Table 3-2). A recent study
(Zolina et al., 2010) has indicated that there has been an increase of
about 15 to 20% in the persistence of wet spells over most of Europe
over the last 60 years, which was not associated with an increase of the
total number of wet days.
There is low to medium confidence in trends in heavy precipitation in
Asia, both at the continental and regional scale for most regions (Table
3-2; see also Alexander et al., 2006). A weak increase in the frequency
of extreme precipitation events is observed in northern Mongolia
(Nandintsetseg et al., 2007). No systematic spatially coherent trends in
the frequency and duration of extreme precipitation events have been
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
143
found in Eastern and Southeast Asia (Choi et al., 2009), central and
south Asia (Klein Tank et al., 2006), and Western Asia (X. Zhang et al.,
2005; Rahimzadeh et al., 2009). However, statistically significant positive
and negative trends were observed at subregional scales within these
regions. Heavy precipitation increased in Japan during 1901-2004 (Fujibe
et al., 2006), and in India (Rajeevan et al., 2008; Krishnamurthy et al.,
2009) especially during the monsoon seasons (Sen Roy, 2009; Pattanaik
and Rajeevan, 2010). Both statistically significant increases and
decreases in extreme precipitation have been found in China over the
period 1951-2000 (Zhai et al., 2005) and 1978-2002 (Yao et al., 2008).
In Peninsular Malaysia during 1971-2005 the intensity of extreme
precipitation increased and the frequency decreased, while the trend in
the proportion of extreme rainfall over total precipitation was not
statistically significant (Zin et al., 2009). Heavy precipitation increased
over the southern and northern Tibetan Plateau but decreased in the
central Tibetan Plateau during 1961-2005 (You et al., 2008).
In southern Australia, there has been a likely decrease in heavy
precipitation in many areas, especially where mean precipitation has
decreased (Table 3-2). There were statistically significant increases in
the proportion of annual/seasonal rainfall stemming from heavy rain
days from 1911-2008 and 1957-2008 in northwest Australia (Gallant
and Karoly, 2010). Extreme summer rainfall over the northwest of the
Swan-Avon River basin in western Australia increased over 1950-2003
while extreme winter rainfall over the southwest of the basin decreased
(Aryal et al., 2009). In New Zealand, the trends are positive in the western
North and South Islands and negative in the east of the country (Mullan
et al., 2008).
There is low to medium confidence in regional trends in heavy
precipitation in Africa due to partial lack of literature and data, and due
to lack of consistency in reported patterns in some regions (Table 3-2).
The AR4 (Trenberth et al., 2007) reported an increase in heavy
precipitation over southern Africa, but this appears to depend on the
region and precipitation index examined (Kruger, 2006; New et al.,
2006; Seleshi and Camberlin, 2006; Aguilar et al., 2009). Central Africa
exhibited a decrease in heavy precipitation over the last half century
(Aguilar et al., 2009); however, data coverage for large parts of the
region was poor. Precipitation from heavy events has decreased in
western central Africa, but with low spatial coherence (Aguilar et al.,
2009). Rainfall intensity averaged over southern and west Africa has
increased (New et al., 2006). There is a lack of literature on changes in
heavy precipitation in East Africa (Table 3-2). Camberlin et al. (2009)
analyzed changes in components of rainy seasons’ variability over the
time period 1958-1987 in this region, but did not specifically address
trends in heavy precipitation. There were decreasing trends in heavy
precipitation over parts of Ethiopia during the period 1965-2002
(Seleshi and Camberlin, 2006).
Changes in hail occurrence are generally difficult to quantify because hail
occurrence is not well captured by monitoring systems and because of
historical data inhomogeneities. Sometimes, changes in environmental
conditions conducive to hail occurrence are used to infer changes in hail
occurrence. However, the atmospheric conditions are typically estimated
from reanalyses or from radiosonde data and the estimates are associated
with high uncertainty. As a result, assessment of changes in hail frequency
is difficult. For severe thunderstorms in the region east of the Rocky
Mountains in the United States, Brooks and Dotzek (2008) found strong
variability but no clear trend in the past 50 years. Cao (2008) identified
a robust upward trend in hail frequency over Ontario, Canada. Kunz et
al. (2009) found that both hail damage days and convective instability
increased during 1974-2003 in a state in southwest Germany. Xie et al.
(2008) identified no trend in the mean annual hail days in China from
1960 to the early 1980s but a statistically significant decreasing trend
afterwards.
Causes of Observed Changes
The observed changes in heavy precipitation appear to be consistent
with the expected response to anthropogenic forcing (increase due to
enhanced moisture content in the atmosphere; see, e.g., Section 3.2.2.1)
but a direct cause-and-effect relationship between changes in external
forcing and extreme precipitation had not been established at the time
of the AR4. As a result, the AR4 only concluded that it was more likely
than not that anthropogenic influence had contributed to a global trend
towards increases in the frequency of heavy precipitation events over
the second half of the 20th century (Hegerl et al., 2007).
New research since the AR4 provides more evidence of anthropogenic
influence on various aspects of the global hydrological cycle (Stott et al.,
2010; see also Section 3.2.2), which is directly relevant to extreme
precipitation changes. In particular, an anthropogenic influence on
atmospheric moisture content is detectable (Santer et al., 2007; Willett
et al., 2007; see also Section 3.2.2). Wang and Zhang (2008) show that
winter season maximum daily precipitation in North America appears to
be statistically significantly influenced by atmospheric moisture content,
with an increase in moisture corresponding to an increase in maximum
daily precipitation. This behavior has also been seen in model projections
of extreme winter precipitation under global warming (Gutowski et al.,
2008b). Climate model projections suggest that the thermodynamic
constraint based on the Clausius-Clapeyron relation is a good predictor
for extreme precipitation changes in a warmer world in regions where
the nature of the ambient flows change little (Pall et al., 2007). This
indicates that the observed increase in extreme precipitation in many
regions is consistent with the expected extreme precipitation response
to anthropogenic influences. However, the thermodynamic constraint
may not be a good predictor in regions with circulation changes, such as
mid- to higher latitudes (Meehl et al., 2005) and the tropics (Emori and
Brown, 2005), and in arid regions. Additionally, changes in precipitation
extremes with temperature also depend on changes in the moist-
adiabatic temperature lapse rate, in the upward velocity, and in the
temperature when precipitation extremes occur (O’Gorman and
Schneider, 2009a,b; Sugiyama et al., 2010). This may explain why there
have not been increases in precipitation extremes everywhere, although
a low signal-to-noise ratio may also play a role. However, even in
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
144
regions where the Clausius-Clapeyron constraint is not closely followed,
it still appears to be a better predictor for future changes in extreme
precipitation than the change in mean precipitation in climate model
projections (Pall et al., 2007). An observational study seems also to support
this thermodynamic theory. Analysis of daily precipitation from the
Special Sensor Microwave Imager over the tropical oceans shows a
direct link between rainfall extremes and temperature: heavy rainfall
events increase during warm periods (El Niño) and decrease during cold
periods (Allan and Soden, 2008). However, the observed amplification
of rainfall extremes is larger than that predicted by climate models
(Allan and Soden, 2008), due possibly to widely varying changes in
upward velocities associated with precipitation extremes (O’Gorman
and Schneider, 2008). Evidence from measurements in the Netherlands
suggests that hourly precipitation extremes may in some cases increase
14% per degree of warming, which is twice as fast as what would be
expected from the Clausius-Clapeyron relationship alone (Lenderink
and Van Meijgaard, 2008), though this is still under debate (Haerter and
Berg, 2009; Lenderink and van Meijgaard, 2009). A comparison between
observed and multi-model simulated extreme precipitation using an
optimal detection method suggests that the human-induced increase in
greenhouse gases has contributed to the observed intensification of
heavy precipitation events over large Northern Hemisphere land areas
during the latter half of the 20th century (Min et al., 2011). Pall et al.
(2011) linked human influence on global warming patterns with an
increased risk of England and Wales flooding in autumn (September-
November) 2000 that is associated with a displacement in the North
Atlantic jet stream. The present assessment based on evidence from new
studies and those used in the AR4 is that there is medium confidence
that anthropogenic influence has contributed to changes in extreme
precipitation at the global scale. However, this conclusion may be
dependent on the season and spatial scale. For example, there is now
about a 50% chance that an anthropogenic influence can be detected
in UK extreme precipitation in winter, but the likelihood of the detection
in other seasons is very small (Fowler and Wilby, 2010).
Projected Changes and Uncertainties
Regarding projected changes in extreme precipitation, the AR4 concluded
that it was very likely that heavy precipitation events, that is, the
frequency of heavy precipitation or proportion of total precipitation
from heavy precipitation, would increase over most areas of the globe
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Wet Day Intensity
ANN



 










 
 
   
    
    
     
 
   
   
 
   
    
   
  
 





 










 
 
   
    
    
     
 



DJF


 





  
   
   
  
  
 
 
  
   
  
  
  
  
 



 





  
   
   
  
    
 

JJA







 
  
 
  
  
      
    
 
  

 







 



Standard Deviation
1.2 0.4 0.4 1.20
Percentage Days with Pr>Q95


 
 







 
  
  
  
       
   
  
  
 
  
   
  
  
 




 
 







 
  
  
  
       
   



 
 



 
  
  
    
    
    
 
  
  
  
  
   
 
 

 


 
 



 
  
  
    
    
   
 







 
 
 
 
   
     
    
  
 









 


  
Percentage of Days
2 1012
Fraction of Days with Pr>10mm


 

  
  
  
   
  
  
  
 
  
   
 
   
 
 



 

  
  
    
   
  







 
   
  
  
  
  
  
 




 
   
  
  





  
   
       
 
 

 


 



Standard Deviation
1.2 0.6 0 0.6 1.2
Figure 3-6 | Projected annual and seasonal changes in three indices for daily precipitation (Pr) for 2081-2100 with respect to 1980-1999, based on 17 GCMs contributing to the
CMIP3. Left column: wet-day intensity; middle column: percentage of days with precipitation above the 95% quantile of daily wet day precipitation for that day of the year,
calculated from the 1961-1990 reference period; right column: fraction of days with precipitation higher than 10 mm. The changes are computed for the annual time scale (top row)
and two seasons (DJF, middle row, and JJA, bottom row) as the fractions/percentages in the 2081-2100 period (based on simulations under emission scenario SRES A2) minus the
fractions/percentages of the 1980-1999 period (from corresponding simulations for the 20th century). Changes in wet-day intensity and in the fraction of days with Pr >10 mm
are expressed in units of standard deviations, derived from detrended per year annual or seasonal estimates, respectively, from the three 20-year periods 1980-1999, 2046-2065,
and 2081-2100 pooled together. Changes in percentages of days with precipitation above the 95% quantile are given directly as differences in percentage points. Color shading is
only applied for areas where at least 66% (i.e., 12 out of 17) of the GCMs agree on the sign of the change; stippling is applied for regions where at least 90% (i.e., 16 out of 17)
of the GCMs agree on the sign of the change. Adapted from Orlowsky and Seneviratne (2011); updating Tebaldi et al. (2006) for additional number of indices and CMIP3 models,
and including seasonal time frames. For more details, see Appendix 3.A.
145
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Full model range
Central 50%
intermodel range
Median
B1 A1B A2
Scenarios:
Δ Precipitation (%)
204665 208100
20
0
20
40
60
80
Legend
204665 208100
20
0
20
40
60
80
Alaska/N.W. Canada - 1
204665 208100
20
0
20
40
60
80
E. Canada/Greenl./Icel. - 2
204665 208100
20
0
20
40
60
80
W. North America - 3
204665 208100
20
0
20
40
60
80
C. North America - 4
204665 208100
20
0
20
40
60
80
E. North America - 5
204665 208100
20
0
20
40
60
80
Central America/Mexico - 6
204665 208100
20
0
20
40
60
80
Amazon - 7
204665 208100
20
0
20
40
60
80
28
N.E. Brazil - 8
204665 208100
20
0
20
40
60
80
86
W. Coast South America - 9
204665 208100
20
0
20
40
60
80
S.E. South America - 10
204665 208100
20
0
20
40
60
80
N. Europe - 11
204665 208100
20
0
20
40
60
80
C. Europe - 12
204665 208100
20
0
20
40
60
80
S. Europe/Mediterranean - 13
204665 208100
20
0
20
40
60
80
37
Sahara - 14
204665 208100
20
0
20
40
60
80
W. Africa - 15
204665 208100
20
0
20
40
60
80
E. Africa - 16
204665 208100
20
0
20
40
60
80
S. Africa - 17
204665 208100
20
0
20
40
60
80
N. Asia - 18
204665 208100
20
0
20
40
60
80
E. Asia - 22
204665 208100
20
0
20
40
60
80
Tibetan Plateau - 21
204665 208100
20
0
20
40
60
80
C. Asia - 20
204665 208100
20
0
20
40
60
80
W. Asia - 19
204665 208100
20
0
20
40
60
80
89
94
S. Asia - 23
204665 208100
20
0
20
40
60
80
100
S.E. Asia - 24
204665 208100
20
0
20
40
60
80
N. Australia - 25
204665 208100
20
0
20
40
60
80
S. Australia/New Zealand - 26
204665 208100
20
0
20
40
60
80
mean (%)
204665 208100
20
0
20
40
60
80
20-year return value (%)
Globe (Land only)
Figure 3-7a | Projected changes (%) in 20-year return values of annual maximum 24-hour precipitation rates. The bar plots (see legend for more information) show results for regionally averaged projections for two time horizons,
2046 to 2065 and 2081 to 2100, as compared to the late 20th century (1981-2000), and for three different SRES emission scenarios (B1, A1B, A2). Results are based on 14 GCMs contributing to the CMIP3. See Figure 3-1 for defined
extent of regions. Values are computed for land points only. The ‘Globe’ analysis (inset box) displays the change in 20-year return values of the annual maximum 24-hour precipitation rates computed using all land grid points (left),
and the change in annual mean 24-hour precipitation rates computed using all land grid points (right). Adapted from the analysis of Kharin et al. (2007). For more details, see Appendix 3.A.
146
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Full model range
Central 50%
intermodel range
Median
B1 A1B A2
Scenarios:
Return period (Years)
204665 208100
3
5
10
20
50
Legend
204665 208100
3
5
10
20
50
Globe (Land only)
204665 208100
3
5
10
20
50
S. Australia/New Zealand - 26
204665 208100
3
5
10
20
50
N. Australia - 25
204665 208100
3
5
10
20
50
2.4
S.E. Asia - 24
204665 208100
3
5
10
20
50
S. Asia - 23
204665 208100
3
5
10
20
50
53
W. Asia - 19
204665 208100
3
5
10
20
50
C. Asia - 20
204665 208100
3
5
10
20
50
Tibetan Plateau - 21
204665 208100
3
5
10
20
50
E. Asia - 22
204665 208100
3
5
10
20
50
N. Asia - 18
204665 208100
3
5
10
20
50
S. Africa - 17
204665 208100
3
5
10
20
50
E. Africa - 16
204665 208100
3
5
10
20
50
W. Africa - 15
204665 208100
3
5
10
20
50
64
56
Sahara - 14
204665 208100
3
5
10
20
50
S. Europe/Mediterranean - 13
204665 208100
3
5
10
20
50
C. Europe - 12
204665 208100
3
5
10
20
50
N. Europe - 11
204665 208100
3
5
10
20
50
S.E. South America - 10
204665 208100
3
5
10
20
50
53
61
W. Coast South America - 9
204665 208100
3
5
10
20
50
57
N.E. Brazil - 8
204665 208100
3
5
10
20
50
Amazon - 7
204665 208100
3
5
10
20
50
Central America/Mexico - 6
204665 208100
3
5
10
20
50
E. North America - 5
204665 208100
3
5
10
20
50
C. North America - 4
204665 208100
3
5
10
20
50
W. North America - 3
204665 208100
3
5
10
20
50
E. Canada/Greenl./Icel. - 2
204665 208100
3
5
10
20
50
2.4
Alaska/N.W. Canada - 1
Figure 3-7b | Projected return period (in years) of late 20th century 20-year return values of annual maximum 24-hour precipitation rates. The bar plots (see legend for more information) show results for regionally averaged projections
for two time horizons, 2046 to 2065 and 2081 to 2100, as compared to the late 20th century (1981-2000), and for three different SRES emission scenarios (B1, A1B, A2). Results are based on 14 GCMs contributing to the CMIP3. See
Figure 3-1 for defined extent of regions. The ‘Globe’ analysis (inset box) displays the projected return period (in years) of late 20th-century 20-year return values of annual maximum 24-hour precipitation rates computed using all land
grid points. Adapted from the analysis of Kharin et al. (2007). For more details, see Appendix 3.A.
147
in the 21st century (IPCC, 2007a). The tendency for an increase in heavy
daily precipitation events was found in many regions, including some
regions in which the total precipitation was projected to decrease.
Post-AR4 analyses of climate model simulations partly confirm this
assessment but also highlight fairly large uncertainties and model biases
in projections of changes in heavy precipitation in some regions
(Section 3.2.3 and Table 3-3). On the other hand, more GCM and RCM
ensembles have now been analyzed for some regions (Table 3-3; see
also, e.g., Kharin et al., 2007; Kim et al., 2010). At the time of the AR4,
Tebaldi et al. (2006) was the main global study available on projected
changes in precipitation extremes (e.g., Figure 10.18 of Meehl et al.,
2007b). Orlowsky and Seneviratne (2011) extended this analysis to a
larger number of GCMs from the CMIP3 ensemble and for seasonal in
addition to annual time frames (see also Section 3.3.1). Figure 3-6 provides
corresponding analyses of projected annual and seasonal changes of
the wet-day intensity, the fraction of days with precipitation above the
95% quantile of daily wet-day precipitation, and the fraction of days
with precipitation above 10 mm day
-1
. It should be noted that the
10 mm day
-1
threshold cannot be considered extreme in several regions,
but highlights differences in projections for absolute and relative
thresholds (see also discussion in Box 3-1 and beginning of this section).
Figure 3-6 indicates that regions with model agreement (at least 66%)
with respect to changes in heavy precipitation are mostly found in the
high latitudes and in the tropics, and in some mid-latitude regions of the
Northern Hemisphere in the boreal winter. Regions with at least 90%
model agreement are even more limited and confined to the high
latitudes. Overall, model agreement in projected changes is found to be
stronger in boreal winter (DJF) than summer (JJA) for most regions.
Kharin et al. (2007) analyzed changes in annual maxima of 24-hour
precipitation in the outputs of 14 CMIP3 models. Figure 3-7a displays
the projected percentage change in the annual maximum of the 24-hour
precipitation rate from the late 20th-century 20-year return values,
while Figure 3-7b displays the corresponding projected return periods
for late 20th-century 20-year return values of the annual maximum
24-hour precipitation rates in the mid-21st century (left) and in late 21st
century (right) under three different emission scenarios (SRES B1, A1B,
and A2). Between the late 20th and the late 21st century, the projected
responses of extreme precipitation to future emissions show increased
precipitation rates in most regions, and decreases in return periods in
most regions in the high latitudes and the tropics and in some regions
in the mid-latitudes consistent with projected changes in several indices
related to heavy precipitation (see Figure 3-6 and Tebaldi et al., 2006),
although there are increases in return periods or only small changes
projected in several regions. Except for these regions, the return period
for an event of annual maximum 24-hour precipitation with a 20-year
return period in the late 20th century is projected to be about 5 to 15
years by the end of the 21st century. The greatest projected reductions
in return period are in high latitudes and some tropical regions. The
stronger CO
2
emissions scenarios (A1B and A2) lead to greater projected
decreases in return period. In some regions with projected decreases in
total precipitation (Christensen et al., 2007) such as southern Africa,
west Asia, and the west coast of South America, heavy precipitation is
nevertheless projected to increase (Figure 3-7, Table 3-3). In some other
areas with projected decreases in total precipitation (e.g., Central America
and northern South America), however, heavy precipitation is projected
to decrease or not change. It should be noted that Figure 3-7 addresses
very extreme heavy precipitation events (those expected to occur about
once in 20 years) whereas Figure 3-6 addresses less extreme, but still
heavy, precipitation events. Projections of changes for these differently
defined extreme events may differ.
Future precipitation projected by the CMIP3 models has also been
analyzed in a number of studies for various regions using different
combinations of the models (see next paragraphs and Table 3-3). In
general these studies confirm the findings of global-scale studies by
Tebaldi et al. (2006) and Kharin et al. (2007).
By analyzing simulations with a single GCM, Khon et al. (2007) reported
a projected general increase in extreme precipitation for the different
regions in northern Eurasia especially for winter. Su et al. (2009) found
that for the Yangtze River Basin region in 2001-2050, the 50-year heavy
precipitation events become more frequent, with return periods falling
to below 25 years (relative to 1951-2000 behavior). For the Indian
region, the Hadley Centre coupled model HadCM3 projects increases in
the magnitude of the heaviest rainfall with a doubling of atmospheric
CO
2
concentration (Turner and Slingo, 2009). Simulations by 12 GCMs
projected an increase in heavy precipitation intensity and mean
precipitation rates in east Africa, more severe precipitation deficits in
the southwest of southern Africa, and enhanced precipitation further
north in Zambia, Malawi, and northern Mozambique (Shongwe et al.,
2009, 2011). Rocha et al. (2008) evaluated differences in the precipitation
regime over southeastern Africa simulated by two GCMs under
present (1961-1990) and future (2071-2100) conditions as a result of
anthropogenic greenhouse gas forcing. They found that the intensity of
all episode categories of precipitation events is projected to increase
practically over the whole region, whereas the number of episodes is
projected to decrease in most of the region and for most episode
categories. Extreme precipitation is projected to increase over Australia in
2080-2099 relative to 1980-1999 in an analysis of the CMIP3 ensemble,
although there are inconsistencies between projections from different
models (Alexander and Arblaster, 2009).
High spatial resolution is important for studies of extreme precipitation
because the physical processes responsible for extreme precipitation
require high spatial resolution to resolve them (e.g., Kim et al., 2010).
Post-AR4 studies have employed three approaches to obtain high spatial
resolution to project precipitation extremes: high-resolution GCMs,
dynamical downscaling using RCMs, and statistical downscaling (see
also Section 3.2.3.1). Based on the Meteorological Research Institute
and Japan Meteorological Agency 20-km horizontal grid GCM, heavy
precipitation was projected to increase substantially in south Asia, the
Amazon, and west Africa, with increased dry spell persistence projected
in South Africa, southern Australia, and the Amazon at the end of the
21st century (Kamiguchi et al., 2006). In the Asian monsoon region,
heavy precipitation was projected to increase, notably in Bangladesh
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
148
and in the Yangtze River basin due to the intensified convergence of
water vapor flux in summer. Using statistical downscaling, Wang and
Zhang (2008) investigated possible changes in North American extreme
precipitation probability during winter from 1949-1999 to 2050-2099.
Downscaled results suggested a strong increase in extreme precipitation
over the south and central United States but decreases over the
Canadian prairies. Projected European precipitation extremes in high-
resolution studies tend to increase in northern Europe (Frei et al., 2006;
Beniston et al., 2007; Schmidli et al., 2007), especially during winter
(Haugen and Iversen, 2008; May, 2008), as also highlighted in Table 3-3.
Fowler and Ekström (2009) project increases in both short-duration
(1-day) and longer-duration (10-day) precipitation extremes across the
United Kingdom during winter, spring, and autumn. In summer, model
projections for the United Kingdom span the zero change line, although
there is low confidence due to poor model performance in this season.
Using daily statistics from various models, Boberg et al. (2009a,b)
projected a clear increase in the contribution to total precipitation from
more intense events together with a decrease in the number of days
with light precipitation. This pattern of change was found to be robust
for all European subregions. In double-nested model simulations with a
horizontal grid spacing of 10 km, Tomassini and Jacob (2009) projected
positive trends in extreme quantiles of heavy precipitation over
Germany, although they are relatively small except for the high-CO
2
A2
emission scenario. For the Upper Mississippi River Basin region during
October through March, the intensity of extreme precipitation is projected
to increase (Gutowski et al., 2008b). Simulations with a single RCM
project an increase in the intensity of extreme precipitation events over
most of southeastern South America and western Amazonia in 2071-2100,
whereas in northeast Brazil and eastern Amazonia smaller or no
changes are projected (Marengo et al., 2009a). Outputs from another
RCM indicate an increase in the magnitude of future extreme rainfall
events in the Westernport region of Australia, consistent with results
based on the CMIP3 ensemble (Alexander and Arblaster, 2009), and the
size of this increase is greater in 2070 than in 2030 (Abbs and Rafter,
2008). When both future land use changes and increasing greenhouse
gas concentrations are considered in the simulations, tropical and
northern Africa are projected to experience less extreme rainfall events
by 2025 during most seasons except for autumn (Paeth and Thamm,
2007). Simulations with high-resolution RCMs projected that the
frequency of extreme precipitation increases in the warm climate for
June through to September in Japan (Nakamura et al., 2008; Wakazuki
et al., 2008; Kitoh et al., 2009). An increase in 90th-percentile values of
daily precipitation on the Pacific side of the Japanese islands during July
in the future climate was projected with a 5-km mesh cloud-system-
resolving non-hydrostatic RCM (Kanada et al., 2010b).
Post-AR4 studies indicate that the projection of precipitation extremes
is associated with large uncertainties, contributed by the uncertainties
related to GCMs, RCMs, and statistical downscaling methods, and by
natural variability of the climate. Kyselý and Beranova (2009) examined
scenarios of change in extreme precipitation events in 24 future climate
runs of 10 RCMs driven by two GCMs, focusing on a specific area of
central Europe with complex orography. They demonstrated that the
inter- and intra-model variability and related uncertainties in the pattern
and magnitude of the change are large, although they also show that
the projected trends tend to agree with those recently observed in the
area, which may strengthen their credibility. May (2008) reported an
unrealistically large projected precipitation change over the Baltic Sea
in summer in an RCM, apparently related to an unrealistic projection of
Baltic Sea warming in the driving GCM. Frei et al. (2006) found large
model differences in summer when RCM formulation contributes
significantly to scenario uncertainty. In exploring the ability of two
statistical downscaling models to reproduce the direction of the projected
changes in indices of precipitation extremes, Hundecha and Bardossy
(2008) concluded that the statistical downscaling models seem to be
more reliable during seasons when local climate is determined by large-
scale circulation than by local convective processes. Themeßl et al.
(2011) merged linear and nonlinear empirical-statistical downscaling
techniques with bias correction methods, and demonstrated their
ability to drastically reduce RCM error characteristics. The extent to which
the natural variability of the climate affects our ability to project the
anthropogenically forced component of changes in daily precipitation
extremes was investigated by Kendon et al. (2008). They show that
annual to multidecadal natural variability across Europe may contribute to
substantial uncertainty. Also, Kiktev et al. (2009) performed an objective
comparison of climatologies and historical trends of temperature and
precipitation extremes using observations and 20th-century climate
simulations. They did not detect significant similarity between simulated
and actual patterns of the indices of precipitation extremes in most cases.
Moreover, Allan and Soden (2008) used satellite observations and model
simulations to examine the response of tropical precipitation events to
naturally driven changes in surface temperature and atmospheric
moisture content. The observed amplification of rainfall extremes was
larger than that predicted by models. The underestimation of rainfall
extremes by the models may be related to the coarse spatial resolution
used in the model simulations – the magnitude of changes in precipitation
extremes depends on spatial resolution (Kitoh et al., 2009) – suggesting
that projections of future changes in rainfall extremes in response to
anthropogenic global warming may be underestimated.
Confidence is still low for hail projections particularly due to a lack of
hail-specific modelling studies, and a lack of agreement among the few
available studies. There is little information in the AR4 regarding projected
changes in hail events, and there has been little new literature since the
AR4. Leslie et al. (2008) used coupled climate model simulations under
the SRES A1B scenario to estimate future changes in hailstorms in the
Sydney Basin, Australia. Their future climate simulations show an
increase in the frequency and intensity of hailstorms out to 2050, and
they suggest that the increase will emerge from the natural background
variability within just a few decades. This result offers a different
conclusion from the modelling study of Niall and Walsh (2005), which
simulated Convective Available Potential Energy (CAPE) for southeastern
Australia in an environment containing double the pre-industrial
concentrations of equivalent CO
2
. They found a statistically significant
projected decrease in CAPE values and concluded that “it is possible
that there will be a decrease in the frequency of hail in southeastern
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
149
Australia if current rates of CO
2
emission are sustained, assuming the
strong relationship between hail incidence and the CAPE for 1980-2001
remains unchanged under enhanced greenhouse conditions.
In summary, it is likely that there have been statistically significant
increases in the number of heavy precipitation events (e.g., 95th
percentile) in more regions than there have been statistically
significant decreases, but there are strong regional and subregional
variations in the trends (i.e., both between and within regions
considered in this report; Figure 3-1 and Tables 3-2 and 3-3). In
particular, many regions present statistically non-significant or
negative trends, and, where seasonal changes have been
assessed, there are also variations between seasons (e.g., more
consistent trends in winter than in summer in Europe). The overall
most consistent trends toward heavier precipitation events are
found in North America (likely increase over the continent). There
is low confidence in observed trends in phenomena such as hail
because of historical data inhomogeneities and inadequacies in
monitoring systems. Based on evidence from new studies and those
used in the AR4, there is medium confidence that anthropogenic
influence has contributed to intensification of extreme precipitation
at the global scale. There is almost no literature on the attribution
of changes in hail extremes, thus no assessment can be provided
for these at this point in time. Projected changes from both global
and regional studies indicate that it is likely that the frequency
of heavy precipitation or proportion of total rainfall from heavy
falls will increase in the 21st century over many areas on the
globe, especially in the high latitudes and tropical regions, and
northern mid-latitudes in winter. Heavy precipitation is projected
to increase in some (but not all) regions with projected decreases
of total precipitation (medium confidence). For a range of emission
scenarios (A2, A1B, and B1), projections indicate that it is likely that
a 1-in-20 year annual maximum 24-hour precipitation rate will
become a 1-in-5 to -15 year event by the end of 21st century in many
regions. Nevertheless, increases or statistically non-significant
changes in return periods are projected in some regions.
3.3.3. Wind
Extreme wind speeds pose a threat to human safety, maritime and
aviation activities, and the integrity of infrastructure. As well as extreme
wind speeds, other attributes of wind can cause extreme impacts. Trends
in average wind speed can influence potential evaporation and in turn
water availability and droughts (e.g., McVicar et al., 2008; see also
Section 3.5.1 and Box 3-3). Sustained mid-latitude winds can elevate
coastal sea levels (e.g., McInnes et al., 2009b), while longer-term
changes in prevailing wind direction can cause changes in wave climate
and coastline stability (Pirazzoli and Tomasin, 2003; see also Sections
3.5.4 and 3.5.5). Aeolian processes exert significant influence on the
formation and evolution of arid and semi-arid environments, being
strongly linked to soil and vegetation change (Okin et al., 2006). A rapid
shift in wind direction may reposition the leading edge of a forest fire
(see Section 4.2.2.2; Mills, 2005) while the fire itself may generate a
local circulation response such as tornado genesis (e.g., Cunningham
and Reeder, 2009). Unlike other weather and climate elements such as
temperature and rainfall, extreme winds are often considered in the
context of the extreme phenomena with which they are associated such
as tropical and extratropical cyclones (see also Sections 3.4.4 and 3.4.5),
thunderstorm downbursts, and tornadoes. Although wind is often not
used to define the extreme event itself (Peterson et al., 2008b), wind
speed thresholds may be used to characterize the severity of the
phenomenon (e.g., the Saffir-Simpson scale for tropical cyclones).
Changes in wind extremes may arise from changes in the intensity or
location of their associated phenomena (e.g., a change in local convective
activity) or from other changes in the climate system such as the
movement of large-scale circulation patterns. Wind extremes may be
defined by a range of quantities such as high percentiles, maxima over
a particular time scale (e.g., daily to yearly), or storm-related highest
values. Wind gusts, which are a measure of the highest winds in a short
time interval (typically 3 seconds), may be evaluated in models using
gust parameterizations that are applied to the maximum daily near-
surface wind speed (e.g., Rockel and Woth, 2007).
Over paleoclimatic time scales, proxy data have been used to infer
circulation changes across the globe from the mid-Holocene (~6000 years
ago) to the beginning of the industrial revolution (Wanner et al., 2008).
Over this period, there is evidence for changes in circulation patterns
across the globe. The Inter-Tropical Convergence Zone (ITCZ) moved
southward, leading to weaker monsoons across Asia (Haug et al., 2001).
The Walker circulation strengthened and Southern Ocean westerlies
moved northward and strengthened, affecting southern Australia, New
Zealand, and southern South America (Shulmeister et al., 2006; Wanner
et al., 2008), and an increase in ENSO variability and frequency occurred
(Rein et al., 2005; Wanner et al., 2008). There is also weaker evidence for
a change toward a lower Northern Atlantic Oscillation (NAO), implying
weaker westerly winds over the north Atlantic (Wanner et al., 2008).
While the changes in the Northern Hemisphere were attributed to
changes in orbital forcing, those in the Southern Hemisphere were more
complex, possibly reflecting the additional role on circulation of heat
transport in the ocean. Solar variability and volcanic eruptions may also
have contributed to decadal to multi-centennial fluctuations over this
time period (Wanner et al., 2008).
The AR4 did not specifically address changes in extreme wind although
it did report on wind changes in the context of other phenomena such as
tropical and extratropical cyclones and oceanic waves and concluded that
mid-latitude westerlies had increased in strength in both hemispheres
(Trenberth et al., 2007). Direct investigation of changes in wind
climatology has been hampered by the sparseness of long-term, high-
quality wind measurements from terrestrial anemometers arising from the
influence of changes in instrumentation, station location, and surrounding
land use (e.g., Cherry, 1988; Pryor et al., 2007; Jakob, 2010; see also
Section 3.2.1). Nevertheless, a number of recent studies report trends in
mean and extreme wind speeds in different parts of the world based on
wind observations and reanalyses.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
150
Over North America, a declining trend in 50th and 90th percentile wind
speeds has been reported for much of the United States over 1973 to
2005 (Pryor et al., 2007) and in 10-m hourly wind data over 1953-2006
over western and most of southern Canada (Wan et al., 2010). An
increasing trend has been reported in average winds over Alaska over
1955-2001 by Lynch et al. (2004) and over the central Canadian Arctic
in all seasons and in the Maritimes in spring and autumn by Wan et al.
(2010) as well as in annual maximum winds in a regional reanalysis
over the southern Maritimes from 1979-2003 (Hundecha et al., 2008).
Over China, negative trends have been reported in 10-m monthly mean
and 95th percentile winds over 1969-2005 (Guo et al., 2011), in daily
maximum wind speeds over 1956-2004 by Jiang et al. (2010a), and in
2-m average winds over the Tibetan plateau from 1966-2003 (Y. Zhang
et al., 2007), confirming earlier declining trends in mean and strong
10-m winds reported by Xu et al. (2006). Over Europe, Smits et al. (2005)
found declining trends in extreme winds (those occurring on average
10 and 2 times per year) in 10-m anemometer data over 1962-2002.
Pirazolli and Tomasin (2003) reported a generally declining trend in
both annual mean and annual maximum winds from 1951 to the mid-
1970s and an increasing trend since then, from observations in the
central Mediterranean region. Similar to the mostly declining trends
found in Northern Hemisphere studies of surface wind observations,
Vautard et al. (2010) also found mostly declining trends in surface wind
observations across the continental northern mid-latitudes and a
stronger decline in extreme winds compared to mean winds in surface
wind measurements. In the Southern Hemisphere, McVicar et al. (2008)
reported declines in 2-m mean wind speed over 88% of Australia
(significant over 57% of the country) over 1975-2006 and positive trends
over about 12% of the mainland interior and southern and eastern
coastal regions including Tasmania. In Antarctica, increasing trends in
mean wind speeds have been reported over the second half of the 20th
century (Turner et al., 2005). With the exception of the robust declines in
wind reported over China, studies in most areas are too few in number
to draw robust conclusions on wind speed change and even fewer
studies have addressed extreme wind change. Some studies report
opposite trends between anemometer winds and reanalysis data sets in
some areas (Smits et al., 2005; McVicar et al., 2008; Vautard et al., 2010);
however, comparisons of surface anemometer data at 10 m or lower
with reanalysis-derived 10-m data that do not resolve complex surface
features is problematic.
Trends in extreme winds have also been inferred from trends in particular
phenomena. With regards to tropical cyclones (Section 3.4.4.), no
statistically significant trends have been detected in the overall global
annual number although a trend has been reported in the intensity of
the strongest storms since 1980 [but there is low confidence that any
observed long-term (i.e., 40 years or more) increases in tropical cyclone
activity are robust, after accounting for past changes in observing
capabilities; see Section 3.4.4]. In the mid-latitudes, studies have used
proxies for wind such as pressure tendencies or geostrophic winds
calculated from triangles of pressure (geo-winds) over Europe (e.g.,
Barring and von Storch, 2004; Matulla et al., 2008; Allan et al., 2009;
Barring and Fortuniak, 2009; X.L. Wang et al., 2009b) and Australia (e.g.,
Alexander and Power, 2009; Alexander et al., 2011). For Europe, these
studies suggest that storm activity was higher around 1900 and in the
1990s and lower in the 1960s and 1970s, although X.L. Wang et al.
(2009b) note that seasonal trends behave differently than annual trends.
In general, long-term trends differ between the different available
studies as well as studies that focus on the period for which reanalysis
data exist (e.g., Raible, 2007; Leckebusch et al., 2008; Della-Marta et al.,
2009; Nissen et al., 2010), and strong inter-decadal variability is also
often reported (e.g., Allan et al., 2009; X.L. Wang et al., 2009b; Nissen et
al., 2010). Over southeast Australia, a decline in storm activity since
around 1885 has been reported (Alexander and Power, 2009; Alexander et
al., 2011). See Section 3.4.5 for more discussion of extratropical cyclones.
Regarding other phenomena associated with extreme winds, such as
thunderstorms, tornadoes, and mesoscale convective complexes, studies
are too few in number to assess the effect of their changes on extreme
winds. As well, historical data inhomogeneities mean that there is low
confidence in any observed trends in these small-scale phenomena.
The AR4 reported for the mid-latitudes that trends in the Northern and
Southern Annular Modes, which correspond to sea level pressure reductions
over the poles, are likely related in part to human activity, and this in
turn has affected storm tracks and wind patterns in both hemispheres
(Hegerl et al., 2007). The relationship between mean and severe winds
and natural modes of variability has been investigated in several post-
AR4 studies. On the Canadian west coast, Abeysirigunawardena et al.
(2009) found that higher extreme winds tend to occur during the negative
(i.e., cold) ENSO phase. The generally increasing trend in mean wind
speeds over recent decades in Antarctica is consistent with the change
in the nature of the Southern Annular Mode toward its high index state
(Turner et al., 2005). Donat et al. (2010b) concluded that 80% of storm
days in central Europe are connected with westerly flows that occur
primarily during the positive phase of the NAO. Declining trends in wind
over China have mainly been linked to circulation changes due to a
weaker land-sea thermal contrast (Xu et al., 2006; Jiang et al., 2010a;
Guo et al., 2011). Vautard et al. (2010) attribute the slowdown in mid to
high percentiles of surface winds over most of the continental northern
mid-latitudes to changes in atmospheric circulation (10-50%) and an
increase in surface roughness due to biomass increases (25-60%),
which are supported by RCM simulations. X.L. Wang et al. (2009a)
formally detected a link between external forcing and positive trends in
the high northern latitudes and negative trends in the northern mid-
latitudes using a proxy for wind (geostrophic wind energy) in the boreal
winter. Trends in mean and annual maximum winds in the central
Mediterranean region were found to be positively correlated with
temperature but not with the NAO index (Pirazzoli and Tomasin, 2003).
Nissen et al. (2010) used cyclone tracking to identify associated strong
winds in reanalysis data from 1957 to 2002 and found a positive trend
in the central Mediterranean region and southern Europe and a negative
trend over the western Mediterranean region.
Projections of wind speed changes and particularly wind extremes
were not specifically addressed in the AR4 although references to wind
speed were made in relation to other variables and phenomena such as
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
151
mid-latitude storm tracks, tropical cyclones, and ocean waves
(Christensen et al., 2007; Meehl et al., 2007b). Meehl et al. (2007b)
projected a likely increase in tropical cyclone extreme winds in the
future and provided more evidence for a projected poleward shift of the
storm tracks and associated changes in wind patterns. Since the AR4,
new studies have focused on future changes in winds. Gastineau and
Soden (2009) reported a decrease in 99th-percentile winds at 850 hPa
in the tropics and an increase in the extratropics in a 17-member multi-
model ensemble over 2081-2100 relative to 1981-2000. McInnes et al.
(2011) presented spatial maps of multi-model agreement in mean and
99th-percentile 10-m wind change between 1981-2000 and 2081-2100
in a 19-member ensemble (see Figure 3-8). These show an increase in
mean winds over Europe, parts of Central and North America, the tropical
South Pacific, and the Southern Ocean. Mean wind speed declines occur
along the equator reflecting a slowdown in the Walker circulation
(Collins et al., 2010) (and in the vicinity of the subtropical ridge in both
hemispheres which, together with the strengthening of winds further
poleward, reflect the contraction toward the poles of the mid-latitude
storm tracks; see Section 3.4.5). Seasonal differences are also apparent
with more extensive mean wind increases in the Arctic and parts of the
northern Pacific in DJF and decreases over most of the northern Pacific
in JJA. The 99th-percentile wind changes show declines over most ocean
areas except the northern Pacific and Arctic and Southern Ocean south
of 40°S in DJF, the south Pacific between about 10 and 25°S in JJA, and
the Southern Ocean south of 50°S in JJA. Increases in 99th-percentile
winds occur over the Arctic and large parts of the continental area in the
Northern Hemisphere in DJF and in Africa, northern Australia, and
Central and South America in JJA. Despite the projections displayed in
Figure 3-8, the relatively few studies of projected extreme winds,
combined with shortcomings in the simulation of extreme winds and
the different models, regions, and methods used to develop projections
of this quantity, means that we have low confidence in projections of
changes in strong winds.
Regional increases in winter wind storm risk over Europe due to
changes in storm tracks are also supported by a number of regional
studies (e.g., Pinto et al., 2007b; Debernard and Roed, 2008; Leckebusch
et al., 2008; Sterl et al., 2009; Donat et al., 2010a,b, 2011). However, GCMs
at their current resolution are unable to resolve small-scale phenomena
such as tropical cyclones, tornadoes, and mesoscale convective complexes
that are associated with particularly severe winds, although as noted by
McInnes et al. (2011) these winds would typically be more extreme than
99th percentile. There is evidence to suggest an increase in extreme
winds from tropical cyclones in the future (see Section 3.4.4). An
increase in atmospheric greenhouse gas concentrations may cause
some of the atmospheric conditions conducive to tornadoes such as
atmospheric instability to increasedue to increasing temperature and
humidity, while others such as vertical shear to decrease due to a
reduced pole-to-equator temperature gradient (Diffenbaugh et al.,
2008), but the literature on these phenomena is extremely limited at
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
% Change
-10 -5
DJFDJF
JJAJJA
10
05
Figure 3-8 | Averaged changes from a 19-member ensemble of CMIP3 GCMs in the mean of the daily averaged 10-m wind speeds (top) and 99th percentile of the daily averaged
10-m wind speeds (bottom) for the period 2081-2100 relative to 1981-2000 (% change) for December to February (left) and June to August (right) plotted only where more than
66% of the models agree on the sign of the change. Black stippling indicates areas where more than 90% of the models agree on the sign of the change. Red stippling indicates
areas where more than 66% of models agree on a small change between ±2%. Adapted from McInnes et al. (2011); for more details see Appendix 3.A.
152
this time. There is thus low confidence in projections of changes in such
small-scale systems because of limited studies, inability of climate models
to resolve these phenomena, and possible competing factors affecting
future changes. Confidence in the extreme wind changes is therefore
lower in the regions most influenced by these phenomena irrespective
of whether there is high agreement between GCMs on the sign of the
wind speed change.
In addition to studies using GCMs there have also been several recent
studies employing RCMs. Those focusing on Europe (e.g., Beniston et al.,
2007; Rockel and Woth, 2007; Haugen and Iversen, 2008; Rauthe et al.,
2010) also provide a general picture of an increasing trend in extreme
winds over northern Europe despite a range of different downscaling
models used, the different GCMs in which the downscaling is undertaken,
and different metrics used to quantify extreme winds. Small-scale polar
lows that typically form north of 60°N have been found to decline in
frequency in RCM simulations downscaled from a GCM under different
emission scenarios and this is related to greater stability over the region
due to mid-troposphere temperatures warming faster than sea surface
temperatures over the region (Zahn and von Storch, 2010). In other parts
of the world there have been very few studies. Over China, Jiang et al.
(2010b) projected decreases in annual and winter mean wind speed
based on two RCMs that downscale two different GCMs. Over North
America, statistical downscaling of winds from four GCMs over five
airports in the northwest United States indicated declines in summer
wind speeds and less certain changes in winter (Sailor et al., 2008).
A number of recent studies have addressed observed changes in
wind speed across different parts of the globe, but due to the
various shortcomings associated with anemometer data and the
inconsistency in anemometer and reanalysis trends in some regions,
we have low confidence in wind trends and their causes at this
stage. We also have low confidence in how the observed trends in
mean wind speed relate to trends in extreme winds. The few
studies of projected extreme winds, combined with shortcomings
in the simulation of extreme winds and the different models,
regions, and methods used to develop projections of this quantity,
mean that we have low confidence in projections of changes in
extreme winds (with the exception of changes associated with
tropical cyclones; Section 3.4.4). There is low confidence in
projections of small-scale phenomena such as tornadoes
because competing physical processes may affect future trends
and because climate models do not simulate such phenomena.
3.4. Observed and Projected Changes in
Phenomena Related to Weather and
Climate Extremes
3.4.1. Monsoons
Changes in monsoon-related extreme precipitation and winds due to
climate change are not well understood. Generally, precipitation is the
most important variable, but it is also a variable associated with larger
uncertainties in climate simulations and projections (Wang et al., 2005;
Kang and Shukla, 2006). Changes in monsoons should be better depicted
by large-scale dynamics, circulation, or moisture convergence more
broadly than via precipitation only. However, few studies have focused
on observed changes in the large-scale and regional monsoon circulations.
Hence, in this section, we focus mostly on monsoon-induced changes
in total and seasonal rainfall, with most discussions of intense rainfall
covered in Section 3.3.2.
Modeling experiments to assess paleo-monsoons suggest that in the
past, during the Holocene due to orbital forcing on a millennial time
scale, there was a progressive southward shift of the Northern
Hemisphere summer position of the ITCZ around 8,000 years ago. This
was accompanied by a pronounced weakening of the monsoon rainfall
systems in Africa and Asia and increasing dryness on both continents,
while in South America the monsoon was weaker and drier than in the
present, as suggested both by models and paleoclimatic indicators
(Wanner et al., 2008).
The delineation of the global monsoon has been mostly performed
using rainfall data or outgoing longwave radiation (OLR) fields (Kim et
al., 2008). Lau and Wu (2007) identified two opposite time evolutions in
the occurrence of rainfall events in the tropics: a negative trend in
moderate rain events and a positive trend in heavy and light rain
events. Positive trends in intense rain were located in deep convective
cores of the ITCZ, South Pacific Convergence Zone, Indian Ocean, and
monsoon regions.
In the Indo-Pacific region, covering the southeast Asian and north
Australian monsoon, Caesar et al. (2011) found low spatial coherence in
trends in precipitation extremes across the region between 1971 and
2003. In the few cases where statistically significant trends in precipitation
extremes were identified, there was generally a trend towards wetter
conditions, in common with the global results of Alexander et al. (2006).
Liu et al. (2011) reported a decline in recorded precipitation events in
China over 1960-2000, which was mainly accounted for by a decrease
in light precipitation events, with intensities of 0.1-0.3 mm day
-1
. Some
of the extreme precipitation appeared to be positively correlated with a
La Niña-like sea surface temperature (SST) pattern, but without
suggesting the presence of a trend. With regard to wind changes, Guo
et al. (2011) analyzed near-surface wind speed change in China and its
monsoon regions from 1969 to 2005 and showed a statistically significant
weakening in annual and seasonal mean wind.
For the Indian monsoon, Rajeevan et al. (2008) showed that extreme
rain events have an increasing trend between 1901 and 2005, but the
trend is much stronger after 1950. Sen Roy (2009) investigated changes
in extreme hourly rainfall in India, and found widespread increases in
heavy precipitation events across India, mostly in the high-elevation
regions of the northwestern Himalaya as well as along the foothills of the
Himalaya extending south into the Indo-Ganges basin, and particularly
during the summer monsoon season during 1980-2002.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
153
In the African monsoon region, Fontaine et al. (2011) investigated
recent observed trends using high-resolution gridded precipitation
(period 1979-2002), OLR, and reanalyses. Their results revealed a rainfall
increase in North Africa since the mid-1990s. Over the longer term,
however, Zhou et al. (2008a,b) and Wang and Ding (2006) reported an
overall decreasing long-term trend in global land monsoon rainfall
during the last 54 years, which was mainly caused by decreasing rainfall
in the North African and South Asian monsoons.
For the American monsoon regions, Cavazos et al. (2008) reported
increases in the intensity of precipitation in the mountain sites of the
northwestern Mexico section of the North American monsoon over the
1961-1998 period, apparently related to an increased contribution from
heavy precipitation derived from tropical cyclones. Arriaga-Ramírez and
Cavazos (2010) found that total and extreme rainfall in the monsoon
region of western Mexico and the US southwest presented a statistically
significant increase during 1961-1998, mainly in winter. Groisman and
Knight (2008) found that consecutive dry days (see Box 3-3 for definition)
have significantly increased in the US southwest. On the other hand,
increases in heavy precipitation during 1960-2000 in the South American
monsoon have been documented by Marengo et al. (2009a,b) and
Rusticucci et al. (2010). Studies using circulation fields such as 850 hPa
winds or moisture flux have been performed for the South American
monsoon system for assessments of the onset and end of the monsoon,
and indicate that the onset exhibits a marked interannual variability
linked to variations in SST anomalies in the eastern Pacific and tropical
Atlantic (Gan et al., 2006; da Silva and de Carvalho, 2007; Raia and
Cavalcanti, 2008; Nieto-Ferreira and Rickenbach, 2011).
Attributing the causes of changes in monsoons is difficult in part
because there are substantial inter-model differences in representing
Asian monsoon processes (Christensen et al., 2007). Most models
simulate the general migration of seasonal tropical rain, although the
observed maximum rainfall during the monsoon season along the west
coast of India, the North Bay of Bengal, and adjoining northeast India is
poorly simulated by many models due to limited resolution. Bollasina and
Nigam (2009) show the presence of large systematic biases in coupled
simulations of boreal summer precipitation, evaporation, and SST in the
Indian Ocean. Many of the biases are pervasive, being common to most
simulations.
The observed negative trend in global land monsoon rainfall is better
reproduced by atmospheric models forced by observed historical SST
than by coupled models without explicit forcing by observed ocean
temperatures (Kim et al., 2008). This trend in the east Asian monsoon is
strongly linked to the warming trend over the central eastern Pacific and
the western tropical Indian Ocean (Zhou et al., 2008b). For the west
African monsoon, Joly and Voldoire (2010) explore the role of Gulf of
Guinea SSTs in its interannual variability. In most of the studied CMIP3
simulations, the interannual variability of SST is very weak in the Gulf of
Guinea, especially along the Guinean Coast. As a consequence, the
influence on the monsoon rainfall over the African continent is poorly
reproduced. It is suggested that this may be due to the counteracting
effects of the Pacific and Atlantic basins over the last decades. The
decreasing long-term trend in north African summer monsoon rainfall may
be due to the atmosphere response to observed SST variations (Hoerling
et al., 2006; Zhou et al., 2008b; Scaife et al., 2009). A similar trend in
global monsoon precipitation in land regions is reproduced in CMIP3
models’ 20th-century simulations when they include anthropogenic
forcing, and for some simulations natural forcing (including volcanic
forcing) as well, though the trend is much weaker in general, with the
exception of one model (HadCM3) capable of producing a trend of
similar magnitude (Li et al., 2008). The decrease in east Asian monsoon
rainfall also seems to be related to tropical SST changes (Li et al., 2008),
and the less spatially coherent positive trends in precipitation extremes
in the southeast Asian and north Australian monsoons appear to be
positively correlated with a La Niña-like SST pattern (Caesar et al., 2011).
A variety of factors, natural and anthropogenic, have been suggested as
possible causes of variations in monsoons. Changes in regional monsoons
are strongly influenced by the changes in the states of dominant patterns
of climate variability such as ENSO, the Pacific Decadal Oscillation (PDO),
the Northern Annular Mode (NAM), the Atlantic Multi-decadal Oscillation
(AMO), and the Southern Annular Mode (SAM) (see also Sections 3.4.2
and 3.4.3). Additionally, model-based evidence has suggested that land
surface processes and land use changes could in some instances
significantly impact regional monsoons. Tropical land cover change in
Africa and southeast Asia appears to have weaker local climatic impacts
than in Amazonia (Voldoire and Royer, 2004; Mabuchi et al., 2005a,b).
Grimm et al. (2007) and Collini et al. (2008) explored possible feedbacks
between soil moisture and precipitation during the early stages of the
monsoon in South America, when the surface is not sufficiently wet, and
soil moisture anomalies may thus also modulate the development of
precipitation. However, the influence of historical land use on the
monsoon is difficult to quantify, due both to the poor documentation of
land use and difficulties in simulating the monsoon at fine scales. The
impact of aerosols (black carbon and sulfate) on changes in rainfall
variability and amounts in monsoon regions has been discussed by
Meehl et al. (2008), Lau et al. (2006), and Silva Dias et al. (2002). These
studies suggest that there are still large uncertainties and a strong
model dependency in the representation of the relevant land surface
processes and the role of aerosol direct forcing, and resulting interactions
(e.g., in the case of land use forcing; Pitman et al., 2009).
Regarding projections of change in the monsoons, the AR4 (Christensen et
al., 2007) concluded: “There is a tendency for monsoonal circulations to
result in increased precipitation due to enhanced moisture convergence,
despite a tendency towards weakening of the monsoonal flows
themselves. However, many aspects of tropical climatic responses remain
uncertain. Held and Soden (2006) demonstrate that an increase in the
hydrological cycle is accompanied by a global weakening of the large-
scale circulation. As global warming is projected to lead to faster
warming over land than over the oceans (e.g., Meehl et al., 2007b;
Sutton et al., 2007), the continental-scale land-sea thermal contrast, a
major factor affecting monsoon circulations, will become stronger in
summer. Based on this projection, a simple scenario is that the summer
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
154
monsoon will be stronger and the winter monsoon will be weaker in the
future than now. However, model results derived from the analyses of
15 CMIP3 global models are not as straightforward as implied by this
simple consideration (Tanaka et al., 2005), as they show a weakening of
these tropical circulations by the late 21st century compared to the late
20th century. In turn, such changes in circulation may lead to changes
in precipitation associated with monsoons. For instance, the monsoonal
precipitation in Mexico and Central America is projected to decrease in
association with increasing precipitation over the eastern equatorial
Pacific through changes in the Walker circulation and local Hadley
circulation (e.g., Lu et al., 2007). Furthermore, observations and models
suggest that changes in monsoons are related at least in part to
changes in observed SSTs, as noted above.
At regional scales, there is little consensus in GCM projections regarding
the sign of future change in monsoon characteristics, such as circulation
and rainfall. For instance, while some models project an intense
drying of the Sahel under a global warming scenario, others project an
intensification of the rains, and some project more frequent extreme
events (Cook and Vizy, 2006). Increases in precipitation are projected in
the Asian monsoon (along with an increase in interannual season-
averaged precipitation variability), and in the southern part of the west
African monsoon, but with some decreases in the Sahel in northern
summer. In the Australian monsoon in southern summer, an analysis by
Moise and Colman (2009) from the entire ensemble mean of CMIP3
simulations suggested no changes in Australian tropical rainfall during
the summer and only slightly enhanced interannual variability.
A study of 19 CMIP3 global models reported a projected increase in
mean south Asian summer monsoon precipitation of 8% and a possible
extension of the monsoon period (Kripalani et al., 2007). A study
(Ashfaq et al., 2009) from the downscaling of the National Center for
Atmospheric Research (NCAR) CCSM3 global model using the RegCM3
regional model suggests a weakening of the large-scale monsoon flow
and suppression of the dominant intra-seasonal oscillatory modes with
overall weakening of the south Asian summer monsoon by the end of
the 21st century, resulting in a decrease in summer precipitation in
southeast Asia.
Kitoh and Uchiyama (2006) used 15 models under the A1B scenario to
analyze the changes in intensity and duration of precipitation in the
Baiu-Changma-Meiyu rain band at the end of the 21st century. They
found a delay in early summer rain withdrawal over the region extending
from the Taiwan province of China, and across the Ryukyu Islands to the
south of Japan, contrasted with an earlier withdrawal over the Yangtze
Basin. They attributed this feature to El Niño-like mean state changes
over the monsoon trough and subtropical anticyclone over the western
Pacific region (Meehl et al., 2007b). A southwestward extension of the
subtropical anticyclone over the northwestern Pacific Ocean associated
with El Niño-like mean state changes and a dry air intrusion in the mid-
troposphere from the Asian continent to the northwest of Japan provides
favorable conditions for intense precipitation in the Baiu season in
Japan (Kanada et al., 2010a). Kitoh et al. (2009) projected changes in
precipitation characteristics during the east Asian summer rainy season,
using a 5-km mesh cloud-resolving model embedded in a 20-km mesh
global atmospheric model with CMIP3 mean SST changes. The frequency
of heavy precipitation was projected to increase at the end of the 21st
century for hourly as well as daily precipitation. Further, extreme hourly
precipitation was projected to increase even in the near future (2030s)
when the temperature increase is still modest, even though uncertainties
in the projection (and even the simulation) of hourly rainfall are still high.
Climate change scenarios for the 21st century show a weakening of the
North American monsoon through a weakening and poleward expansion
of the Hadley cell (Lu et al., 2007). The expansion of the Hadley cell is
caused by an increase in the subtropical static stability, which pushes
poleward the baroclinic instability zone and hence the outer boundary
of the Hadley cell. Simple physical arguments (Held and Soden, 2006)
predict a slowdown of the tropical overturning circulation under global
warming. A few studies (e.g., Marengo et al., 2009a) have projected over
the period 1960-2100 a weak tendency for an increase in dry spells. The
projections show an increase in the frequency of rainfall extremes in
southeastern South America by the end of the 21st century, possibly due
to an intensification of the moisture transport from Amazonia by a more
frequent/intense low-level jet east of the Andes in the A2 emissions
scenario (Marengo et al., 2009a; Soares and Marengo, 2009).
There are many deficiencies in model representation of the monsoons
and the processes affecting them, and this reduces confidence in their
ability to project future changes. Some of the uncertainty in global and
regional climate change projections in the monsoon regions results from
the limits in the model representation of resolved processes (e.g., moisture
advection), the parameterizations of sub-grid-scale processes (e.g.,
clouds, precipitation), and model simulations of feedback mechanisms
at the global and regional scale (e.g., changes in land use/cover; see
also Section 3.1.4). Kharin et al. (2007) made an intercomparison of
precipitation extremes in the tropical region in all AR4 models with
observed extremes expressed as 20-year return values. They found very
large disagreement in the tropics suggesting that some physical
processes associated with extreme precipitation are not well represented
by the models due to model resolution and physics. Shukla (2007) noted
that current climate models cannot even adequately predict the mean
intensity and the seasonal variations of the Asian summer monsoon. This
reduces confidence in the projected changes in extreme precipitation
over the monsoon regions. Many of the important climatic effects of the
Madden-Julian Oscillation (MJO, a natural mode of the climate system
operating on time scales of about a month), including its impacts on
rainfall variability in the monsoons, are still poorly simulated by
contemporary climate models (Christensen et al., 2007).
Current GCMs still have difficulties and display a wide range of skill in
simulating the subseasonal variability associated with the Asian summer
monsoon (Lin et al., 2008b). Most GCMs simulate westward propagation
of the coupled equatorial easterly waves, but relatively poor eastward
propagation of the MJO and overly weak variances for both the easterly
waves and the MJO. Most GCMs are able to reproduce the basic
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
155
characteristics of the precipitation seasonal cycle associated with
the South American Monsoon System (SAMS), but there are large
discrepancies in the South Atlantic Convergence Zone represented by
the models in both intensity and location, and in its seasonal evolution
(Vera et al., 2006). In addition, models exhibit large discrepancies in the
direction of the changes associated with the summer (SAMS) precipitation,
which makes the projections for that region highly uncertain. Lin et al.
(2008a) show that the coupled GCMs have significant problems and
display a wide range of skill in simulating the North American monsoon
and associated intra-seasonal variability.
Most of the models reproduce the monsoon rain belt, extending from
southeast to northwest, and its gradual northward shift in early summer,
but overestimate the precipitation over the core monsoon region
throughout the seasonal cycle and fail to reproduce the monsoon
retreat in the fall. The AR4 assessed that models fail in representing the
main features of the west African monsoon although most of them do
have a monsoonal climate albeit with some distortion (Christensen et
al., 2007). Other major sources of uncertainty in projections of monsoon
changes are the responses and feedbacks of the climate system to
emissions as represented in climate models. These uncertainties are
particularly related to the representation of the conversion of emissions
into concentrations of radiatively active species (i.e., via atmospheric
chemistry and carbon cycle models) and especially those derived from
aerosol products of biomass burning, which can affect the onset of the
rainy season (Silva Dias et al., 2002). The subsequent response of the
physical climate system complicates the nature of future projections of
monsoon precipitation. Moreover, the long-term variations of model
skill in simulating monsoons and their variations represent an additional
source of uncertainty for the monsoon regions, and indicate that the
regional reliability of long climate model runs may depend on the time
slice for which the output of the model is analyzed.
The AR4 (Hegerl et al., 2007) concluded that the current
understanding of climate change in the monsoon regions remains
one of considerable uncertainty with respect to circulation and
precipitation. With a few exceptions in some monsoon regions,
this has not changed. These conclusions have been based on very
few studies, there are many issues with model representation of
monsoons and the underlying processes, and there is little
consensus in climate models, so there is low confidence in
projections of changes in monsoons, even in the sign of the change.
However, one common pattern is a likely increase in extreme
precipitation in monsoon regions (see Section 3.3.2), though not
necessarily induced by changes in monsoon characteristics, and
not necessarily occurring in all monsoon regions.
3.4.2. El Niño-Southern Oscillation
The El Niño-Southern Oscillation (ENSO) is a natural fluctuation of the
global climate system caused by equatorial ocean-atmosphere interaction
in the tropical Pacific Ocean (Philander, 1990). The term ‘Southern
Oscillation’ refers to a tendency for above-average surface atmospheric
pressures in the Indian Ocean to be associated with below-average
pressures in the Pacific, and vice versa. This oscillation is associated
with variations in SSTs in the east equatorial Pacific. The oceanic and
atmospheric variations are collectively referred to as ENSO. An El Niño
episode is one phase of the ENSO phenomenon and is associated with
abnormally warm central and east equatorial Pacific Ocean surface
temperatures, while the opposite phase, a La Niña episode, is associated
with abnormally cool ocean temperatures in this region. Both phases
are associated with a characteristic spatial pattern of droughts and
floods. An El Niño episode is usually accompanied by drought in
southeastern Asia, India, Australia, southeastern Africa, Amazonia, and
northeast Brazil, with fewer than normal tropical cyclones around
Australia and in the North Atlantic. Wetter than normal conditions
during El Niño episodes are observed along the west coast of tropical
South America, subtropical latitudes of western North America, and
southeastern America. In a La Niña episode the climate anomalies are
usually the opposite of those in an El Niño. Pacific islands are strongly
affected by ENSO variations. Recent research (e.g., Kenyon and Hegerl,
2008; Ropelewski and Bell, 2008; Schubert et al., 2008a; Alexander et al.,
2009; Grimm and Tedeschi, 2009; Zhang et al., 2010) has demonstrated
that different phases of ENSO (El Niño or La Niña episodes) also are
associated with different frequencies of occurrence of short-term weather
extremes such as heavy rainfall events and extreme temperatures. The
relationship between ENSO and interannual variations in tropical cyclone
activity is well known (e.g., Kuleshov et al., 2008). The simultaneous
occurrence of a variety of climate extremes in an El Niño episode (or a
La Niña episode) may provide special challenges for organizations coping
with disasters induced by ENSO (see also Section 3.1.1). Monitoring and
predicting ENSO can lead to disaster risk reduction through early warning
(see Case Study 9.2.11).
The AR4 noted that orbital variations could affect the ENSO behavior
(Jansen et al., 2007). Cane (2005) found that a relatively simple coupled
model suggested that systematic changes in El Niño could be stimulat-
ed by seasonal changes in solar insolation. However, a more compre-
hensive model simulation (Wittenberg, 2009) has suggested that long-
term changes in the behavior of the phenomenon might occur even
without forcing from radiative changes. Vecchi and Wittenberg (2010)
concluded that the “tropical Pacific could generate variations in ENSO
frequency and intensity on its own (via chaotic behavior), respond to
external radiative forcings (e.g., changes in greenhouse gases, volcanic
eruptions, atmospheric aerosols, etc), or both.” Meehl et al. (2009a)
demonstrate that solar insolation variations related to the 11-year
sunspot cycle can affect ocean temperatures associated with ENSO.
ENSO has varied in strength over the last millennium with stronger
activity in the 17th century and late 14th century, and weaker activity
during the 12th and 15th centuries (Cobb et al., 2003; Conroy et al.,
2009). On longer time scales, there is evidence that ENSO may have
changed in response to changes in the orbit of the Earth (Vecchi and
Wittenberg, 2010), with the phenomenon apparently being weaker
around 6,000 years ago (according to proxy measurements from corals
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
156
and climate model simulations; Rein et al., 2005; Brown et al., 2006;
Otto-Bliesner et al., 2009), and model simulations suggest that it was
stronger at the last glacial maximum (An et al., 2004). Fossil coral
evidence indicates that the phenomenon continued to operate during
the last glacial interval (Tudhope et al., 2001). Thus the paleoclimatic
evidence indicates that ENSO can continue to operate, although altered
perhaps in intensity, in very different background climate states.
The AR4 noted that the nature of ENSO has varied substantially over the
period of instrumental data, with strong events from the late 19th
century through the first quarter of the 20th century and again after
1950. An apparent climate shift around 1976-1977 was associated with
a shift to generally above-normal SSTs in the central and eastern Pacific
and a tendency toward more prolonged and stronger El Niño episodes
(Trenberth et al., 2007). Ocean temperatures in the central equatorial
Pacific (the so-called NINO3 index) suggest a trend toward more frequent
or stronger El Niño episodes over the past 50 to 100 years (Vecchi and
Wittenberg, 2010). Vecchi et al. (2006) reported a weakening of the
equatorial Pacific pressure gradient since the 1960s, with a sharp drop
in the 1970s. Power and Smith (2007) proposed that the apparent
dominance of El Niño during the last few decades was due in part to a
change in the background state of the Southern Oscillation Index (SOI,
the standardized difference in surface atmospheric pressure between
Tahiti and Darwin), rather than a change in variability or a shift to more
frequent El Niño events alone. Nicholls (2008) examined the behavior of
the SOI and another index, the NINO3.4 index of central equatorial
Pacific SSTs, but found no evidence of trends in the variability or the
persistence of the indices [although Yu and Kao (2007) reported decadal
variations in the persistence barrier, the tendency for weaker persistence
across the Northern Hemisphere spring], nor in their seasonal patterns.
There was a trend toward what might be considered more ‘El Niño-like’
behavior in the SOI (and more weakly in NINO3.4), but only through the
period March to September and not in November to February, the season
when El Niño and La Niña events typically peak. The trend in the SOI
reflected only a trend in Darwin pressures, with no trend in Tahiti
pressures. Apart from this trend, the temporal/seasonal nature of ENSO has
been remarkably consistent through a period of strong global warming.
There is evidence, however, of a tendency for recent El Niño episodes to
be centered more in the central equatorial Pacific than in the east Pacific
(Yeh et al., 2009), and for these central Pacific episodes to be increasing
in intensity (Lee and McPhaden, 2010). In turn, these changes may
explain changes that have been noted in the remote influences of the
phenomenon on the climate over Australia and in the mid-latitudes
(Wang and Hendon, 2007; Weng et al., 2009). For instance, Taschetto et
al. (2009) demonstrated that episodes with the warming centered in the
central Pacific exhibit different patterns of Australian rainfall variations
relative to the east Pacific-centered El Niño events.
The possible role of increased greenhouse gases in affecting the behavior
of ENSO over the past 50 to 100 years is uncertain. Yeh et al. (2009)
suggested that changes in the background temperature associated with
increases in greenhouse gases should affect the behavior of El Niño,
such as the location of the strongest SST anomalies, because El Niño
behavior is strongly related to the average ocean temperature gradients
in the equatorial Pacific. Some studies (e.g., Q. Zhang et al., 2008) have
suggested that increased activity might be due to increased CO
2
;
however, no formal attribution study has yet been completed and some
other studies (e.g., Power and Smith, 2007) suggest that changes in the
phenomenon are within the range of natural variability (i.e., that no
change has yet been detected, let alone attributed to a specific cause).
Global warming is projected to lead to a mean reduction in the zonal
mean wind across the equatorial Pacific (Vecchi and Soden, 2007b).
However, this change should not be described as an ‘El Niño-like’ average
change even though during an El Niño episode these winds also weaken,
because there is only limited correspondence between these changes in
the mean state of the equatorial Pacific and an El Niño episode. The
AR4 determined that all models exhibited continued ENSO interannual
variability in projections through the 21st century, but the projected
behavior of the phenomenon differed between models, and it was
concluded that “there is no consistent indication at this time of
discernible changes in projected ENSO amplitude or frequency in the
21st century” (Meehl et al., 2007b). Models project a wide variety of
changes in ENSO variability and the frequency of El Niño episodes as a
consequence of increased greenhouse gas concentrations, with a range
between a 30% reduction to a 30% increase in variability (van Oldenborgh
et al., 2005). One model study even found that although ENSO activity
increased when atmospheric CO
2
concentrations were doubled or
quadrupled, a considerable decrease in activity occurred when CO
2
was
increased by a factor of 16 times, much greater than is possible through
the 21st century (Cherchi et al., 2008), suggesting a wide variety of
possible ENSO changes as a result of CO
2
changes. The remote impacts,
on rainfall for instance, of ENSO may change as CO
2
increases, even if
the equatorial Pacific aspect of ENSO does not change substantially. For
instance, regions in which rainfall increases in the future tend to show
increases in interannual rainfall variability (Boer, 2009), without any
strong change in the interannual variability of tropical SSTs. Also, since
some long-term projected changes in response to increased greenhouse
gases may resemble the climate response to an El Niño event, this may
enhance or mask the response to El Niño events in the future (Lau et al.,
2008b; Müller and Roeckner, 2008).
One change that models tend to project is an increasing tendency for El
Niño episodes to be centered in the central equatorial Pacific, rather
than the traditional location in the eastern equatorial Pacific. Yeh et al.
(2009) examined the relative frequency of El Niño episodes simulated in
coupled climate models with projected increases in greenhouse gas
concentrations. A majority of models, especially those best able to
simulate the current ratio of central Pacific locations to east Pacific
locations of El Niño events, projected a further increase in the relative
frequency of these central Pacific events. Such a change would also have
implications for the remote influence of the phenomenon on climate away
from the equatorial Pacific (e.g., Australia and India). However, even the
projection that the 21st century may see an increased frequency of central
Pacific El Niño episodes, relative to the frequency of events located
further east (Yeh et al., 2009), is subject to considerable uncertainty. Of
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
157
the 11 coupled climate model simulations examined by Yeh et al. (2009),
three projected a relative decrease in the frequency of these central
Pacific episodes, and only four of the models produced a statistically
significant change to more frequent central Pacific events.
A caveat regarding all projections of future behavior of ENSO arises
from systematic biases in the depiction of ENSO behavior through the
20th century by models (Randall et al., 2007; Guilyardi et al., 2009).
Leloup et al. (2008) for instance, demonstrate that coupled climate
models show wide differences in the ability to reproduce the spatial
characteristics of SST variations associated with ENSO during the 20th
century, and all models have failings. They concluded that it is difficult
to even classify models by the quality of their reproductions of the
behavior of ENSO, because models scored unevenly in their reproduction
of the different phases of the phenomenon. This makes it difficult to
determine which models to use to project future changes in ENSO.
Moreover, most of the models are not able to reproduce the typical
circulation anomalies associated with ENSO in the Southern
Hemisphere (Vera and Silvestri, 2009) and the Northern Hemisphere
(Joseph and Nigam, 2006).
There was no consistency in projections of changes in ENSO variability
or frequency at the time of the AR4 (Meehl et al., 2007b) and this
situation has not changed as a result of post-AR4 studies. The evidence
is that the nature of ENSO has varied in the past apparently sometimes
in response to changes in radiative forcing but also possibly due to
internal climatic variability. Since radiative forcing will continue to
change in the future, we can confidently expect changes in ENSO and
its impacts as well, although both El Niño and La Niña episodes will
continue to occur (e.g., Vecchi and Wittenberg, 2010). Our current limited
understanding, however, means that it is not possible at this time to
confidently predict whether ENSO activity will be enhanced or damped
due to anthropogenic climate change, or even if the frequency of El Niño
or La Niña episodes will change (Collins et al., 2010).
In summary, there is medium confidence in a recent trend toward
more frequent central equatorial Pacific El Niño episodes, but
insufficient evidence for more specific statements about
observed trends in ENSO. Model projections of changes in ENSO
variability and the frequency of El Niño episodes as a consequence
of increased greenhouse gas concentrations are not consistent,
and so there is low confidence in projections of changes in the
phenomenon. However, there is medium confidence regarding a
projected increase (projected by most GCMs) in the relative
frequency of central equatorial Pacific events, which typically
exhibit different patterns of climate variations than do the
classical East Pacific events.
3.4.3. Other Modes of Variability
Other natural modes of variability beside ENSO (Section 3.4.2) that are
relevant to extremes and disasters include the North Atlantic Oscillation
(NAO), the Southern Annular Mode (SAM), and the Indian Ocean Dipole
(IOD) (Trenberth et al., 2007). The NAO is a large-scale seesaw in
atmospheric pressure between the subtropical high and the polar low
in the North Atlantic region. The positive NAO phase has a strong
subtropical high-pressure center and a deeper than normal Icelandic
low. This results in a shift of winter storms crossing the Atlantic Ocean
to a more northerly track, and is associated with warm and wet winters
in northwestern Europe and cold and dry winters in northern Canada
and Greenland. Scaife et al. (2008) discuss the relationship between
the NAO and European extremes. Paleoclimatic data indicate that the
NAO was persistently in its positive phase during medieval times and
persistently in its negative phase during the cooler Little Ice Age (Trouet
et al., 2009). The NAO is closely related to the Northern Annular Mode
(NAM); for brevity we focus here on the NAO but much of what is said
about the NAO also applies to the NAM. The SAM is the largest mode of
Southern Hemisphere extratropical variability and refers to north-south
shifts in atmospheric mass between the middle and high latitudes. It
plays an important role in climate variability in these latitudes. The SAM
positive phase is linked to negative sea level pressure anomalies over
the polar regions and intensified westerlies. It has been associated with
cooler than normal temperatures over most of Antarctica and Australia,
with warm anomalies over the Antarctic Peninsula, southern South
America, and southern New Zealand, and with anomalously dry conditions
over southern South America, New Zealand, and Tasmania and wet
anomalies over much of Australia and South Africa (e.g., Hendon et al.,
2007). The IOD is a coupled ocean-atmosphere phenomenon in the Indian
Ocean. A positive IOD event is associated with anomalous cooling in the
southeastern equatorial Indian Ocean and anomalous warming in the
western equatorial Indian Ocean. Recent work (Ummenhofer et al., 2008,
2009a,b) has implicated the IOD as a cause of droughts in Australia, and
heavy rainfall in east Africa (Ummenhofer et al., 2009c). There is also
evidence of modes of variability operating on multi-decadal time scales,
notably the Pacific Decadal Oscillation (PDO) and the Atlantic Multi-
decadal Oscillation (AMO). Variations in the PDO have been related to
precipitation extremes over North America (Zhang et al., 2010).
Both the NAO and the SAM exhibited trends toward their positive phase
(strengthened mid-latitude westerlies) over the last three to four decades,
although the NAO has been in its negative phase in the last few years.
Goodkin et al. (2008) concluded that the variability in the NAO is linked
with changes in the mean temperature of the Northern Hemisphere.
Dong et al. (2011) demonstrated that some of the observed late 20th-
century decadal-scale changes in NAO behavior could be reproduced by
increasing the CO
2
concentrations in a coupled model, and concluded
that greenhouse gas concentrations may have played a role in forcing
these changes. The largest observed trends in the SAM occur in
December to February, and model simulations indicate that these are
due mainly to stratospheric ozone changes. However it has been argued
that anthropogenic circulation changes are poorly characterized by trends
in the annular modes (Woollings et al., 2008). Further complicating
these trends, Silvestri and Vera (2009) reported changes in the typical
hemispheric circulation pattern related to the SAM and its associated
impact on both temperature and precipitation anomalies, particularly
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
158
over South America and Australia, between the 1960s-1970s and
1980s-1990s. The time scales of variability in modes such as the AMO
and PDO are so long that it is difficult to diagnose any change in their
behavior in modern data, although some evidence suggests that the
PDO may be affected by anthropogenic forcing (Meehl et al., 2009b).
The AR4 (Hegerl et al., 2007) concluded that trends over recent decades
in the NAO and SAM are likely related in part to human activity. The
negative NAO phase of the last few years, however, with the lack of
formal attribution studies, means that attribution of changes in the NAO
to human activity in recent decades now can only be considered about
as likely as not (expert opinion). Attribution of the SAM trend to human
activity is still assessed to be likely (expert opinion) although mainly
attributable to trends in stratospheric ozone concentration (Hegerl et
al., 2007).
The AR4 noted that there was considerable spread among the model
projections of the NAO, leading to low confidence in NAO projected
changes, but the magnitude of the increase for the SAM is generally
more consistent across models (Meehl et al., 2007b). However, the ability
of coupled models to simulate the observed SAM impact on climate
variability in the Southern Hemisphere is limited (e.g., Miller et al., 2006;
Vera and Silvestri, 2009). Variations in the longer time-scale modes of
variability (AMO, PDO) might affect projections of changes in extremes
associated with the various natural modes of variability and global
temperatures (Keenlyside et al., 2008).
Sea level pressure is projected to increase over the subtropics and mid-
latitudes, and decrease over high latitudes (Meehl et al., 2007b). This
would equate to trends in the NAO and SAM, with a poleward shift of
the storm tracks of several degrees latitude and a consequent increase
in cyclonic circulation patterns over the Arctic and Antarctica. In the
Southern Hemisphere, two opposing effects, stratospheric ozone recovery
and increasing greenhouse gases, can be expected to affect the modes
such as the SAM (Arblaster et al., 2011). During the 21st century, although
stratospheric ozone concentrations are expected to recover, tending to
lead to a weakening of the SAM, models consistently project polar
vortex intensification to continue due to the increases in greenhouse
gases, except in summer where the competing effects of stratospheric
ozone recovery complicate this picture (Arblaster et al., 2011).
A recent study (Woollings et al., 2010) found a tendency toward a more
positive NAO under anthropogenic forcing through the 21st century
with one model, although they concluded that confidence in the model
projections was low because of deficiencies in its simulation of current-day
NAO regimes. Goodkin et al. (2008) predict continuing high variability,
on multi-decadal scales, in the NAO with continued global warming.
Keenlyside et al. (2008) proposed that variations associated with the
multi-decadal modes of variability may offset warming due to increased
greenhouse gas concentrations over the next decade or so. Conway et
al. (2007) reported that model projections of future IOD behavior showed
no consistency. Kay and Washington (2008) reported that under some
emissions scenarios, changes in a dipole mode in the Indian Ocean
could change rainfall extremes in southern Africa.
In summary, it is likely that there has been an anthropogenic
influence on recent trends in the SAM (linked with trends in
stratospheric ozone rather than changes in greenhouse gases),
but it is only about as likely as not that there have been
anthropogenic influences on observed trends in the NAO. Issues
with the ability of models to simulate current behavior of these
natural modes, the influence of competing factors (e.g.,
stratospheric ozone, greenhouse gases) on current and future
mode behavior, and inconsistency between the model projections
(and the seasonal dependence of these projections), means that
there is low confidence in the ability to project changes in the
modes including the NAO, SAM, and IOD. Models do, however,
consistently project a strengthening of the polar vortex in the
Southern Hemisphere from increasing greenhouse gases,
although in summer stratospheric ozone recovery is expected to
offset this intensification.
3.4.4. Tropical Cyclones
Tropical cyclones occur in most tropical oceans and pose a significant
threat to coastal populations and infrastructure, and marine interests
such as shipping and offshore activities. Each year, about 90 tropical
cyclones occur globally, and this number has remained roughly steady
over the modern period of geostationary satellites (since around the
mid-1970s). While the global frequency has remained steady, there can
be substantial inter-annual to multi-decadal frequency variability within
individual ocean basins (e.g., Webster et al., 2005). This regional variability,
particularly when combined with substantial inter-annual to multi-decadal
variability in tropical cyclone tracks (e.g., Kossin et al., 2010), presents a
significant challenge for disaster planning and mitigation aimed at
specific regions.
Tropical cyclones are perhaps most commonly associated with extreme
wind, but storm-surge and freshwater flooding from extreme rainfall
generally cause the great majority of damage and loss of life (e.g.,
Rappaport, 2000; Webster, 2008). Related indirect factors, such as the
failure of the levee system in New Orleans during the passage of
Hurricane Katrina (2005), or mudslides during the landfall of Hurricane
Mitch (1998) in Central America, represent important related impacts
(Case Study 9.2.5). Projected sea level rise will further compound tropical
cyclone surge impacts. Tropical cyclones that track poleward can undergo
a transition to become extratropical cyclones. While these storms have
different characteristics than their tropical progenitors, they can still be
accompanied by a storm surge that can impact regions well away from
the tropics (e.g., Danard et al., 2004).
Tropical cyclones are typically classified in terms of their intensity, which is
a measure of near-surface wind speed (sometimes categorized according
to the Saffir-Simpson scale). The strongest storms (Saffir-Simpson category
3, 4, and 5) are comparatively rare but are generally responsible for the
majority of damage (e.g., Landsea, 1993; Pielke Jr. et al., 2008).
Additionally, there are marked differences in the characteristics of both
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
159
observed and projected tropical cyclone variability when comparing
weaker and stronger tropical cyclones (e.g., Webster et al., 2005; Elsner
et al., 2008; Bender et al., 2010), while records of the strongest storms
are potentially less reliable than those of their weaker counterparts
(Landsea et al., 2006).
In addition to intensity, the structure and areal extent of the wind field
in tropical cyclones, which can be largely independent of intensity, also
play an important role on potential impacts, particularly from storm
surge (e.g., Irish and Resio, 2010), but measures of storm size are largely
absent in historical data. Other relevant tropical cyclone measures
include frequency, duration, and track. Forming robust physical links
between all of the metrics briefly mentioned here and natural or
human-induced changes in climate variability is a major challenge.
Significant progress is being made, but substantial uncertainties still
remain due largely to data quality issues (see Section 3.2.1 and below)
and imperfect theoretical and modeling frameworks (see below).
Observed Changes
Detection of trends in tropical cyclone metrics such as frequency,
intensity, and duration remains a significant challenge. Historical tropical
cyclone records are known to be heterogeneous due to changing
observing technology and reporting protocols (e.g., Landsea et al., 2004).
Further heterogeneity is introduced when records from multiple ocean
basins are combined to explore global trends because data quality and
reporting protocols vary substantially between regions (Knapp and
Kruk, 2010). Progress has been made toward a more homogeneous
global record of tropical cyclone intensity using satellite data (Knapp
and Kossin, 2007; Kossin et al., 2007), but these records are necessarily
constrained to the satellite era and so only represent the past 30 to 40
years.
Natural variability combined with uncertainties in the historical data
makes it difficult to detect trends in tropical cyclone activity. There have
been no significant trends observed in global tropical cyclone frequency
records, including over the present 40-year period of satellite observations
(e.g., Webster et al., 2005). Regional trends in tropical cyclone frequency
have been identified in the North Atlantic, but the fidelity of these trends
is debated (Holland and Webster, 2007; Landsea, 2007; Mann et al.,
2007a). Different methods for estimating undercounts in the earlier part
of the North Atlantic tropical cyclone record provide mixed conclusions
(Chang and Guo, 2007; Mann et al., 2007b; Kunkel et al., 2008; Vecchi
and Knutson, 2008). Regional trends have not been detected in other
oceans (Chan and Xu, 2009; Kubota and Chan, 2009; Callaghan and
Power, 2011). It thus remains uncertain whether any observed increases
in tropical cyclone frequency on time scales longer than about 40 years
are robust, after accounting for past changes in observing capabilities
(Knutson et al., 2010).
Frequency estimation requires only that a tropical cyclone be identified
and reported at some point in its lifetime, whereas intensity estimation
requires a series of specifically targeted measurements over the entire
duration of the tropical cyclone (e.g., Landsea et al., 2006). Consequently,
intensity values in the historical records are especially sensitive to
changing technology and improving methodology, which heightens the
challenge of detecting trends within the backdrop of natural variability.
Global reanalyses of tropical cyclone intensity using a homogenous
satellite record have suggested that changing technology has introduced
a non-stationary bias that inflates trends in measures of intensity
(Kossin et al., 2007), but a significant upward trend in the intensity of
the strongest tropical cyclones remains after this bias is accounted for
(Elsner et al., 2008). While these analyses are suggestive of a link
between observed global tropical cyclone intensity and climate change,
they are necessarily confined to a roughly 30-year period of satellite
observations, and cannot provide clear evidence for a longer-term trend.
Time series of power dissipation, an aggregate compound of tropical
cyclone frequency, duration, and intensity that measures total energy
consumption by tropical cyclones, show upward trends in the North
Atlantic and weaker upward trends in the western North Pacific over the
past 25 years (Emanuel, 2007), but interpretation of longer-term trends
in this quantity is again constrained by data quality concerns. The
variability and trend of power dissipation can be related to SST and
other local factors such as tropopause temperature and vertical wind
shear (Emanuel, 2007), but it is a current topic of debate whether local
SST or the difference between local SST and mean tropical SST is the
more physically relevant metric (Swanson, 2008). The distinction is an
important one when making projections of changes in power dissipation
based on projections of SST changes, particularly in the tropical Atlantic
where SST has been increasing more rapidly than in the tropics as a
whole (Vecchi et al., 2008). Accumulated cyclone energy, which is an
integrated metric analogous to power dissipation, has been declining
globally since reaching a high point in 2005, and is presently at a 40-
year low point (Maue, 2009). The present period of quiescence, as well
as the period of heightened activity leading up to the high point in 2005,
does not clearly represent substantial departures from past variability
(Maue, 2009).
Increases in tropical water vapor and rainfall (Trenberth et al., 2005; Lau
and Wu, 2007) have been identified and there is some evidence for
related changes in tropical cyclone-related rainfall (Lau et al., 2008a),
but a robust and consistent trend in tropical cyclone rainfall has not yet
been established due to a general lack of studies. Similarly, an increase
in the length of the North Atlantic hurricane season has been noted
(Kossin, 2008), but the uncertainty in the amplitude of the trends and
the lack of additional studies limits the utility of these results for a
meaningful assessment.
Estimates of tropical cyclone variability prior to the modern instrumental
historical record have been constructed using archival documents
(Chenoweth and Devine, 2008), coastal marsh sediment records, and
isotope markers in coral, speleothems, and tree rings, among other
methods (Frappier et al., 2007a). These estimates demonstrate centennial-
to millennial-scale relationships between climate and tropical cyclone
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
160
activity (Donnelly and Woodruff, 2007; Frappier et al., 2007b; Nott et al.,
2007; Nyberg et al., 2007; Scileppi and Donnelly, 2007; Neu, 2008;
Woodruff et al., 2008a,b; Mann et al., 2009; Yu et al., 2009), but generally
do not provide robust evidence that the observed post-industrial tropical
cyclone activity is unprecedented.
The AR4 Summary for Policymakers concluded that it is likely that an
increase had occurred in intense tropical cyclone activity since 1970 in
some regions (IPCC, 2007b). The subsequent CCSP assessment report
(Kunkel et al., 2008) concluded that it is likely that the frequency of
tropical storms, hurricanes, and major hurricanes in the North Atlantic
has increased over the past 100 years, a time in which Atlantic SSTs also
increased. Kunkel et al. (2008) also concluded that the increase in
Atlantic power dissipation is likely substantial since the 1950s. Based on
research subsequent to the AR4 and Kunkel et al. (2008), which further
elucidated the scope of uncertainties in the historical tropical cyclone data,
the most recent assessment by the World Meteorological Organization
(WMO) Expert Team on Climate Change Impacts on Tropical Cyclones
(Knutson et al., 2010) concluded that it remains uncertain whether past
changes in any tropical cyclone activity (frequency, intensity, rainfall)
exceed the variability expected through natural causes, after accounting
for changes over time in observing capabilities. The present assessment
regarding observed trends in tropical cyclone activity is essentially
identical to the WMO assessment (Knutson et al., 2010): there is low
confidence that any observed long-term (i.e., 40 years or more) increases
in tropical cyclone activity are robust, after accounting for past changes
in observing capabilities.
Causes of the Observed Changes
In addition to the natural variability of tropical SSTs, several studies
have concluded that there is a detectable tropical SST warming trend
due to increasing greenhouse gases (Karoly and Wu, 2005; Knutson et
al., 2006; Santer et al., 2006; Gillett et al., 2008a). The region where this
anthropogenic warming has occurred encompasses tropical cyclogenesis
regions, and Kunkel et al. (2008) stated that it is very likely that human-
caused increases in greenhouse gases have contributed to the increase
in SSTs in the North Atlantic and the Northwest Pacific hurricane formation
regions over the 20th century.
Changes in the mean thermodynamic state of the tropics can be directly
linked to tropical cyclone variability within the theoretical framework of
potential intensity theory (Bister and Emanuel, 1998). In this framework,
the expected response of tropical cyclone intensity to observed climate
change is relatively straightforward: if climate change causes an
increase in the ambient potential intensity that tropical cyclones move
through, the distribution of intensities in a representative sample of
storms is expected to shift toward greater intensities (Emanuel, 2000;
Wing et al., 2007). The fractional changes associated with such a shift
in the distribution would be largest in the upper quantiles of the
distribution as the strongest tropical cyclones become stronger (Elsner
et al., 2008).
Given the evidence that SST in the tropics has increased due to
increasing greenhouse gases, and the theoretical expectation that
increases in potential intensity will lead to stronger storms, it is essential
to fully understand the relationship between SST and potential intensity.
Observations demonstrate a strong positive correlation between SST
and the potential intensity. This relationship suggests that SST increases
will lead to increased potential intensity, which will then ultimately lead
to stronger storms (Emanuel, 2000; Wing et al., 2007). However, there is
a growing body of research suggesting that local potential intensity is
controlled by the difference between local SST and spatially averaged
SST in the tropics (Vecchi and Soden, 2007a; Xie et al., 2010; Ramsay
and Sobel, 2011). Since increases in SST due to global warming are not
expected to lead to continuously increasing SST gradients, this recent
research suggests that increasing SST due to global warming, by itself,
does not yet have a fully understood physical link to increasingly strong
tropical cyclones.
The present period of heightened tropical cyclone activity in the North
Atlantic, concurrent with comparative quiescence in other ocean basins
(e.g., Maue, 2009), is apparently related to differences in the rate of SST
increases, as global SST has been rising steadily but at a slower rate
than has the Atlantic (Holland and Webster, 2007). The present period of
relatively enhanced warming in the Atlantic has been proposed to be
due to internal variability (Zhang and Delworth, 2009), anthropogenic
tropospheric aerosols (Mann and Emanuel, 2006), and mineral (dust)
aerosols (Evan et al., 2009). None of these proposed mechanisms provide
a clear expectation that North Atlantic SST will continue to increase at
a greater rate than the tropical mean SST.
Changes in tropical cyclone intensity, frequency, genesis location,
duration, and track contribute to what is sometimes broadly defined as
‘tropical cyclone activity.’ Of these metrics, intensity has the most direct
physically reconcilable link to climate variability within the framework
of potential intensity theory, as described above (Kossin and Vimont,
2007). Statistical correlations between necessary ambient environmental
conditions (e.g., low vertical wind shear and adequate atmospheric
instability and moisture) and tropical cyclogenesis frequency have been
well documented (DeMaria et al., 2001) but changes in these conditions
due specifically to increasing greenhouse gas concentrations do not
necessarily preserve the same statistical relationships. For example, the
observed minimum SST threshold for tropical cyclogenesis is roughly
26°C. This relationship might lead to an expectation that anthropogenic
warming of tropical SST and the resulting increase in the areal extent of
the region of 26°C SST should lead to increases in tropical cyclone
frequency. However, there is a growing body of evidence that the
minimum SST threshold for tropical cyclogenesis increases at about the
same rate as the SST increase due solely to greenhouse gas forcing
(e.g., Ryan et al., 1992; Dutton et al., 2000; Yoshimura et al., 2006;
Bengtsson et al., 2007; Knutson et al., 2008; Johnson and Xie, 2010).
This is because the threshold conditions for tropical cyclogenesis are
controlled not just by surface temperature but also by atmospheric
stability (measured from the lower boundary to the tropopause), which
responds to greenhouse gas forcing in a more complex way than SST
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
161
alone. That is, when the SST changes due to greenhouse warming are
deconvolved from the background natural variability, that part of the SST
variability, by itself, has no manifest effect on tropical cyclogenesis. In
this case, the simple observed relationship between tropical cyclogenesis
and SST, while robust, does not adequately capture the relevant physical
mechanisms of tropical cyclogenesis in a warming world.
Another challenge to identifying causes behind observed changes in
tropical cyclone activity is introduced by uncertainties in the reanalysis
data used to identify environmental changes in regions where tropical
cyclones develop and evolve (Bister and Emanuel, 2002; Emanuel,
2010). In particular, heterogeneity in upper-tropospheric kinematic and
thermodynamic metrics complicates the interpretation of long-term
changes in vertical wind shear and potential intensity, both of which are
important environmental controls on tropical cyclones.
Based on a variety of model simulations, the expected long-term
changes in global tropical cyclone characteristics under greenhouse
warming is a decrease or little change in frequency concurrent with an
increase in mean intensity. One of the challenges for identifying these
changes in the existing data records is that the expected changes
predicted by the models are generally small when compared with
changes associated with observed short-term natural variability. Based
on changes in tropical cyclone intensity predicted by idealized numerical
simulations with CO
2
-induced tropical SST warming, Knutson and Tuleya
(2004) suggested that clearly detectable increases may not be manifest
for decades to come. Their argument was based on a comparison of the
amplitude of the modeled upward trend (i.e., the signal) in storm intensity
with the amplitude of the interannual variability (i.e., the noise). The
recent high-resolution dynamical downscaling study of Bender et al. (2010)
supports this argument and suggests that the predicted increases in the
frequency of the strongest Atlantic storms may not emerge as a clear
statistically significant signal until the latter half of the 21st century
under the SRES A1B warming scenario. Still, it should be noted that
while these model projections suggest that a statistically significant signal
may not emerge until some future time, the likelihood of more intense
tropical cyclones is projected to continually increase throughout the
21st century.
With the exception of the North Atlantic, much of the global tropical
cyclone data is confined to the period from the mid-20th century to
present. In addition to the limited period of record, the uncertainties in
the historical tropical cyclone data (Section 3.2.1 and this section) and
the extent of tropical cyclone variability due to random processes and
linkages with various climate modes such as El Niño, do not presently
allow for the detection of any clear trends in tropical cyclone activity
that can be attributed to greenhouse warming. As such, it remains
unclear to what degree the causal phenomena described here have
modulated post-industrial tropical cyclone activity.
The AR4 concluded that it is more likely than not that anthropogenic
influence has contributed to increases in the frequency of the most
intense tropical cyclones (Hegerl et al., 2007). Based on subsequent
research that further elucidated the scope of uncertainties in both the
historical tropical cyclone data as well as the physical mechanisms
underpinning the observed relationships, no such attribution conclusion
was drawn in the recent WMO assessment (Knutson et al., 2010). The
present assessment regarding detection and attribution of trends in
tropical cyclone activity is similar to the WMO assessment (Knutson et
al., 2010): the uncertainties in the historical tropical cyclone records, the
incomplete understanding of the physical mechanisms linking tropical
cyclone metrics to climate change, and the degree of tropical cyclone
variability – comprising random processes and linkages to various
natural climate modes such as El Niño – provide only low confidence for
the attribution of any detectable changes in tropical cyclone activity to
anthropogenic influences.
Projected Changes and Uncertainties
The AR4 concluded (Meehl et al., 2007b) that a broad range of modeling
studies project a likely increase in peak wind intensity and near-storm
precipitation in future tropical cyclones. A reduction of the overall
number of storms was also projected (but with lower confidence), with a
greater reduction in weaker storms in most basins and an increase in the
frequency of the most intense storms. Knutson et al. (2010) concluded
that it is likely that the mean maximum wind speed and near-storm
rainfall rates of tropical cyclones will increase with projected 21st-
century warming, and it is more likely than not that the frequency of the
most intense storms will increase substantially in some basins, but it is
likely that overall global tropical cyclone frequency will decrease or
remain essentially unchanged. The conclusions here are similar to those
of the AR4 and Knutson et al. (2010).
The spatial resolution of some models such as the CMIP3 coupled
ocean-atmosphere models used in the AR4 is generally not high enough
to accurately resolve tropical cyclones, and especially to simulate their
intensity (Randall et al., 2007). Higher-resolution global models have
had some success in reproducing tropical cyclone-like vortices (e.g.,
Chauvin et al., 2006; Oouchi et al., 2006; Zhao et al., 2009), but only
their coarse characteristics. Significant progress has been recently
made, however, using downscaling techniques whereby high-resolution
models capable of reproducing more realistic tropical cyclones are run
using boundary conditions provided by either reanalysis data sets or
output fields from lower-resolution climate models such as those used
in the AR4 (e.g., Knutson et al., 2007; Emanuel et al., 2008; Knutson et
al., 2008; Emanuel, 2010). A recent study by Bender et al. (2010) applies
a cascading technique that downscales first from global to regional
scale, and then uses the simulated storms from the regional model to
initialize a very high-resolution hurricane forecasting model. These
downscaling studies have been increasingly successful at reproducing
observed tropical cyclone characteristics, which provides increased
confidence in their projections, and it is expected that more progress
will be made as computing resources improve. Still, awareness that
limitations exist in the models used for tropical cyclone projections,
particularly the ability to accurately reproduce natural climate phenomena
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
162
that are known to modulate storm behavior (e.g., ENSO and MJO), is
important for context when interpreting model output (Sections 3.2.3.2
and 3.4.2).
While detection of long-term past increases in tropical cyclone activity
is complicated by data quality and signal-to-noise issues (as stated
above), theory (Emanuel, 1987) and idealized dynamical models
(Knutson and Tuleya, 2004) both predict increases in tropical cyclone
intensity under greenhouse warming. Recent simulations with high-
resolution dynamical models (Oouchi et al., 2006; Bengtsson et al., 2007;
Gualdi et al., 2008; Knutson et al., 2008; Sugi et al., 2009; Bender et al.,
2010) and statistical-dynamical models (Emanuel, 2007) consistently
find that greenhouse warming causes tropical cyclone intensity to shift
toward stronger storms by the end of the 21st century (2 to 11% increase
in mean maximum wind speed globally). These and other models also
consistently project little change or a reduction in overall tropical
cyclone frequency (e.g., Gualdi et al., 2008; Sugi et al., 2009; Murakami
et al., 2011), but with an accompanying substantial fractional increase
in the frequency of the strongest storms and increased precipitation
rates (in the models for which these metrics were examined). Current
models project changes in overall global frequency ranging from a
decrease of 6 to 34% by the late 21st century (Knutson et al., 2010). The
downscaling experiments of Bender et al. (2010) – which use an 18-
model ensemble-mean of CMIP3 simulations to nudge a high-resolution
dynamical model (Knutson et al., 2008) that is then used to initialize a
very high-resolution dynamical model – project a 28% reduction in the
overall frequency of Atlantic storms and an 80% increase in the frequency
of Saffir-Simpson category 4 and 5 Atlantic hurricanes over the next 80
years (A1B scenario).
The projected decreases in global tropical cyclone frequency may be due
to increases in vertical wind shear (Vecchi and Soden, 2007c; Zhao et
al., 2009; Bender et al., 2010), a weakening of the tropical circulation
(Sugi et al., 2002; Bengtsson et al., 2007) associated with a decrease in
the upward mass flux accompanying deep convection (Held and Soden,
2006), or an increase in the saturation deficit of the middle troposphere
(Emanuel et al., 2008). For individual basins, there is much more
uncertainty in projections of tropical cyclone frequency, with changes of
up to ±50% or more projected by various models (Knutson et al., 2010).
When projected SST changes are considered in the absence of projected
radiative forcing changes, Northern Hemisphere tropical cyclone frequency
has been found to increase (Wehner et al., 2010), which is congruent
with the hypothesis that SST changes alone do not capture the relevant
physical mechanisms controlling tropical cyclogenesis (e.g., Emanuel,
2010).
As noted above, observed changes in rainfall associated with tropical
cyclones have not been clearly established. However, as water vapor in
the tropics increases (Trenberth et al., 2005) there is an expectation for
increased heavy rainfall associated with tropical cyclones in response to
associated moisture convergence increases (Held and Soden, 2006). This
increase is expected to be compounded by increases in intensity as
dynamical convergence under the storm is enhanced. Models in which
tropical cyclone precipitation rates have been examined are highly
consistent in projecting increased rainfall within the area near the
tropical cyclone center under 21st century warming, with increases of
3 to 37% (Knutson et al., 2010). Typical projected increases are near
20% within 100 km of storm centers.
Another type of projection that is sometimes inferred from the literature
is based on extrapolation of an observed statistical relationship (see
also Section 3.2.3). These relationships are typically constructed on past
observed variability that represents a convolution of anthropogenically
forced variability and natural variability across a broad range of time
scales. In general, however, these relationships cannot be expected to
represent all of the relevant physics that control the phenomena of
interest, and their extrapolation beyond the range of the observed
variability they are built on is not reliable. As an example, there is a
strong observed correlation between local SST and tropical cyclone
power dissipation (Emanuel, 2007). If 21st-century SST projections are
applied to this relationship, power dissipation is projected to increase by
about 300% in the next century (Vecchi et al., 2008; Knutson et al.,
2010). Alternatively, there is a similarly strong relationship between
power dissipation and relative SST, which represents the difference
between local and tropical-mean SST and has been argued to serve as
a proxy for local potential intensity (Vecchi and Soden, 2007a). When
21st-century projections of relative SST are considered, this latter
relationship projects almost no change in power dissipation in the next
century (Vecchi et al., 2006). Both of these statistical relationships can
be reasonably defended based on physical arguments but it is not clear
which, if either, is correct (Ramsay and Sobel, 2011).
When simulating 21st-century warming under the A1B emission scenario
(or a close analog), the present models and downscaling techniques as a
whole are consistent in projecting (1) decreases or no change in tropical
cyclone frequency, (2) increases in intensity and fractional increases in
number of most intense storms, and (3) increases in tropical cyclone-
related rainfall rates. Differences in regional projections lead to lower
confidence in basin-specific projections of intensity and rainfall, and
confidence is particularly low for projections of frequency within
individual basins. More specifically, while projections under 21st-century
greenhouse warming indicate that it is likely that the global frequency
of tropical cyclones will either decrease or remain essentially unchanged,
an increase in mean tropical cyclone maximum wind speed is also likely,
although increases may not occur in all tropical regions. This assessment
is essentially identical with that of the recent WMO assessment (Knutson
et al., 2010). Furthermore, while it is likely that overall global frequency
will either decrease or remain essentially unchanged, it is more likely
than not that the frequency of the most intense storms (e.g., Saffir-
Simpson category 4 and 5) will increase substantially in some ocean
basins, again agreeing with the recent WMO assessment (Knutson et al.,
2010). Based on the level of consistency among models, and physical
reasoning, it is likely that tropical cyclone-related rainfall rates will
increase with greenhouse warming. Confidence in future projections for
particular ocean basins is undermined by the inability of global models
to reproduce accurate details at scales relevant to tropical cyclone
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
163
genesis, track, and intensity evolution. Of particular concern is the limited
ability of global models to accurately simulate upper-tropospheric wind
(Cordero and Forster, 2006; Bender et al., 2010), which modulates vertical
wind shear and tropical cyclone genesis and intensity evolution. Thus
there is low confidence in projections of changes in tropical cyclone
genesis, location, tracks, duration, or areas of impact, and existing
model projections do not show dramatic large-scale changes in these
features.
In summary, there is low confidence that any observed long-term
(i.e., 40 years or more) increases in tropical cyclone activity are
robust, after accounting for past changes in observing capabilities.
The uncertainties in the historical tropical cyclone records, the
incomplete understanding of the physical mechanisms linking
tropical cyclone metrics to climate change, and the degree of
tropical cyclone variability provide only low confidence for the
attribution of any detectable changes in tropical cyclone activity
to anthropogenic influences. There is low confidence in projections
of changes in tropical cyclone genesis, location, tracks, duration,
or areas of impact. Based on the level of consistency among
models, and physical reasoning, it is likely that tropical cyclone-
related rainfall rates will increase with greenhouse warming. It
is likely that the global frequency of tropical cyclones will either
decrease or remain essentially unchanged. An increase in mean
tropical cyclone maximum wind speed is likely, although increases
may not occur in all tropical regions. While it is likely that overall
global frequency will either decrease or remain essentially
unchanged, it is more likely than not that the frequency of the most
intense storms will increase substantially in some ocean basins.
3.4.5. Extratropical Cyclones
Extratropical cyclones (synoptic-scale low-pressure systems) exist
throughout the mid-latitudes in both hemispheres and mainly develop
over the oceanic basins in the proximity of the upper-tropospheric jet
streams, as a result of flow over mountains (lee cyclogenesis) or through
conversions from tropical to extratropical systems. It should be noted
that regionalized smaller-scale mid-latitude circulation phenomena such
as polar lows and mesoscale cyclones are not treated in this section (but
see Sections 3.3.3 and 3.4.3). Extratropical cyclones are the main poleward
transporter of heat and moisture and may be accompanied by adverse
weather conditions such as windstorms, the buildup of waves and storm
surges, or extreme precipitation events. Thus, changes in the intensity of
extratropical cyclones or a systematic shift in the geographical location
of extratropical cyclone activity may have a great impact on a wide
range of regional climate extremes as well as the long-term changes in
temperature and precipitation. Extratropical cyclones mainly form and
grow via atmospheric instabilities such as a disturbance along a zone of
strong temperature contrast (baroclinic instabilities), which is a reservoir
of available potential energy that can be converted into the kinetic energy
associated with extratropical cyclones. Intensification of the cyclones
may also take place due to processes such as release of energy due to
phase changes of water (latent heat release) (Gutowski et al., 1992;
Wernli et al., 2002). Why should we expect climate change to influence
extratropical cyclones? A simplified line of argument would be that both
the large-scale low and high level pole to equator temperature gradients
may change (possibly in opposite directions) in a climate change scenario
leading to a change in the atmospheric instabilities responsible for
cyclone formation and growth (baroclinicity). These changes may be
induced by a variety of mechanisms operating in different parts of the
atmospheric column ranging from changing surface conditions (Deser et
al., 2007; Bader et al., 2011) to stratospheric changes (Son et al., 2010).
In addition, changes in precipitation intensities within extratropical
cyclones may change the latent heat release. According to theories on
wave-mean flow interaction, changes in the extratropical storm tracks are
also associated with changes in the large-scale flow (Robinson, 2000;
Lorenz and Hartmann, 2003). A latitudinal shift of the upper tropospheric
jet would be accompanied by a latitudinal shift in the extratropical
storm track. It is, however, still unclear to what extent a latitudinal shift
in the jet changes the total storm track activity rather than shifting it
latitudinally (Wettstein and Wallace, 2010). Even within the very simplified
outline above the possible impacts of climate change on extratropical
cyclone development are many and clearly not trivial.
When validated using reanalyses with similar horizontal resolution,
climate models are found to represent the general structure of the
storm track pattern well (Bengtsson et al., 2006; Greeves et al., 2007;
Ulbrich et al., 2008; Catto et al., 2010). However, using data from five
different coupled models, the rate of transfer of zonal available potential
energy to eddy available potential energy in synoptic systems was found
to be too large, yielding too much energy and an overactive energy cycle
(Marques et al., 2011). Models tend to have excessively zonal storm
tracks and some show a poor extension of the storm tracks into Europe
(Pinto et al., 2006; Greeves et al., 2007; Orsolini and Sorteberg, 2009).
It has also been noted that r
epresentation of cyclone activity may
depend on the physics formulations and the horizontal resolution of the
model (Jung et al., 2006; Greeves et al., 2007)
.
Paleoclimatic proxies for extratropical cyclone variability are still few,
but progress is being made in using coastal dune field development and
sand grain content of peat bogs as proxies for storminess. Publications
covering parts of western Europe indicate enhanced sand movement in
European coastal areas during the Little Ice Age (Wilson et al., 2004; de
Jong et al., 2006, 2007; Clemmensen et al., 2007; Clarke and Rendell, 2009;
Sjogren, 2009). It should be noted that sand influx is also influenced by
sediment availability, which is controlled mainly by the degree of
vegetation cover and the moisture content of the sediment (Li et al.,
2004; Wiggs et al., 2004). Intense cultivation, overgrazing, and forest
disturbance make soils more prone to erosion, which can lead to
increased sand transport even under less windy conditions. Thus the
information gained from paleoclimatic proxies to put the last 100 years
of extratropical cyclone variability in context is limited.
Century-long time-series of estimates of extremes in geostrophic wind
deduced from triangles of pressure stations, pressure tendencies from
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
164
single stations (see Section 3.3.3 for details), or oceanic variables such
as extremes in non-tide residuals are (if these are located in the vicinity
of the main storm tracks) possible proxies for extratropical cyclone
activity. Trend detection in extratropical cyclone variables such as
number of cyclones, intensity, and activity (parameters integrating
cyclone intensity, number, and possibly duration) became possible with
the development of reanalyses, but remains challenging. Problems with
reanalyses have been especially pronounced in the Southern Hemisphere
(Hodges et al., 2003; Wang et al., 2006). Even though different reanalyses
correspond well in the Northern Hemisphere (Hodges et al., 2003;
Hanson et al., 2004), changes in the observing system giving artificial
trends in integrated water vapor and kinetic energy (Bengtsson et al.,
2004) may have influenced trends in both the number and intensity of
cyclones. In addition, studies indicate that the magnitude and even the
existence of the changes may depend on the choice of reanalysis (Trigo,
2006; Raible et al., 2008; Simmonds et al., 2008; Ulbrich et al., 2009)
and cyclone tracking algorithm (Raible et al., 2008).
The AR4 noted a likely net increase in the frequency/intensity of
Northern Hemisphere extreme extratropical cyclones and a poleward
shift in the tracks since the 1950s (Trenberth et al., 2007; Table 3.8), and
cited several papers showing increases in the number or strength of
intense extratropical cyclones both over the North Pacific and the North
Atlantic storm track (Trenberth et al., 2007, p. 312) during the last 50
years. Studies using reanalyses indicate a northward and eastward shift
in the Atlantic cyclone activity during the last 60 years with both more
frequent and more intense wintertime cyclones in the high-latitude
Atlantic (Weisse et al., 2005; Wang et al., 2006; Schneidereit et al., 2007;
Raible et al., 2008; Vilibic and Sepic, 2010) and fewer in the mid-latitude
Atlantic (Wang et al., 2006; Raible et al., 2008). The increase in high-
latitude cyclone activity was also reported in several studies of Arctic
cyclone activity (X.D. Zhang et al., 2004; Sorteberg and Walsh, 2008; Sepp
and Jaagus, 2011). Using ship-based trends in mean sea level pressure
(MSLP) variance (which is tied to cyclone intensity), Chang (2007) found
wintertime Atlantic trends to be consistent with National Centers for
Environmental Prediction (NCEP) reanalysis trends in the Atlantic, but
slightly weaker. There are inconsistencies among studies of extreme
cyclones in reanalyses, since some studies show an increase in intensity
and number of extreme Atlantic cyclones (Geng and Sugi, 2001; Paciorek
et al., 2002; Lehmann et al., 2011) while others show a reduction (Gulev
et al., 2001). These differences may in part be due to sensitivities of the
identification schemes and different definitions of an extreme cyclone
(Leckebusch et al., 2006; Pinto et al., 2006). New studies have confirmed
that a positive NAM/NAO (see Section 3.4.3) corresponds to stronger
Atlantic/European cyclone activity (e.g., Chang, 2009; Pinto et al., 2009;
X.L. Wang et al., 2009b). However, studies using long historical records
seem to suggest that some of these links may be statistically intermittent
(Hanna et al., 2008; Matulla et al., 2008; Allan et al., 2009) due to
interdecadal shifts in the location of the positions of the NAO pressure
centers (Vicente-Serrano and Lopez-Moreno, 2008; X.D. Zhang et al.,
2008). It is unclear to what extent the statistical intermittency implies
that the underlying physical processes creating the connection act only
intermittently. A possible influence of the Pacific North America (PNA)
pattern on the entrance of the North Atlantic storm track (over
Newfoundland) has been reported by Pinto et al. (2011). It should be
noted that there is some suggestion that the reanalyses cover a time
period that starts with relatively low cyclonic activity in northern coastal
Europe in the 1960s and reaches a maximum in the 1990s. Long-term
European storminess proxies show no clear trends over the last century
(Hanna et al., 2008; Allan et al., 2009; see Section 3.3.3 for details).
Studies using reanalyses and in situ data for the last 50 years have noted
an increase in the number and intensity of north Pacific wintertime
intense extratropical cyclone systems since the 1950s (Graham and Diaz,
2001; Simmonds and Keay, 2002; Raible et al., 2008) and cyclone activity
(X.D. Zhang et al., 2004), but signs of some of the trends disagreed
when different tracking algorithms or reanalysis products were used
(Raible et al., 2008). A slight positive trend has been found in north
Pacific extreme cyclones (Geng and Sugi, 2001; Gulev et al., 2001;
Paciorek et al., 2002). Using ship measurements, Chang (2007) found
intensity-related wintertime trends in the Pacific to be about 20 to 60%
of that found in the reanalysis. Long-term in situ observations of north
Pacific cyclones based on observed pressure data are considerably
fewer than for coastal Europe. However, using hourly tide gauge records
from the western coast of the United States as a proxy for storminess,
an increasing trend in the extreme winter Non-Tide Residuals (NTR) has
been observed in the last decades (Bromirski et al., 2003; Menendez et al.,
2008). Years having high NTR were linked to a large-scale atmospheric
circulation pattern, with intense storminess associated with a broad,
south-easterly displaced, deep Aleutian low that directed storm tracks
toward the US West Coast. North Pacific cyclonic activity has been
linked to tropical SST anomalies (NINO3.4; see Section 3.4.2) and the
PNA (Eichler and Higgins, 2006; Favre and Gershunov, 2006; Seierstad
et al., 2007), showing that the PNA and NINO3.4 influence storminess,
in particular over the eastern North Pacific with an equatorward shift in
storm tracks in the North Pacific basin, as well as an increase in storm
track activity along the US East Coast during El Niño events.
Based on reanalyses, North American cyclone numbers have increased
over the last 50 years, with no statistically significant change in cyclone
intensity (X.D. Zhang et al., 2004). Hourly MSLP data from Canadian
stations showed that winter cyclones have become significantly more
frequent, longer lasting, and stronger in the lower Canadian Arctic over
the last 50 years (1953-2002), but less frequent and weaker in the south,
especially along the southeast and southwest Canadian coasts (Wang
et al., 2006). Further south, a tendency toward weaker low-pressure
systems over the past few decades was found for US East Coast winter
cyclones using reanalyses, but no statistically significant trends in the
frequency of occurrence of systems (Hirsch et al., 2001).
Studies on extratropical cyclone activity in northern Asia are few. Using
reanalyses, a decrease in extratropical cyclone activity (X.D. Zhang et al.,
2004) and intensity (X.D. Zhang et al., 2004; X. Wang et al., 2009) over
the last 50 years has been reported for northern Eurasia (60-40°N) with
a possible northward shift with increased cyclone frequency in the higher
latitudes (50-45°N) and decrease in the lower latitudes (south of 45°N),
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
165
based on a study with reanalyses. The low-latitude (south of 45°N)
decrease was also noted by Zou et al. (2006), who reported a decrease
in the number of severe storms for mainland China based on an analysis
of extremes of observed 6-hourly pressure tendencies over the last 50
years.
Alexander and Power (2009) showed that the number of observed
severe storms at Cape Otway (south-east Australia) has decreased
since the mid-19th century, strengthening the evidence of a southward
shift in Southern Hemisphere storm tracks previously noted using
reanalyses (Fyfe, 2003; Hope et al., 2006; Wang et al., 2006). Frederiksen
and Frederiksen (2007) linked the reduction in cyclogenesis at 30°S
and southward shift to a decrease in the vertical mean meridional
temperature gradient. Using reanalyses, both Pezza et al. (2007) and
Lim and Simmonds (2009) have confirmed previous studies showing a
trend toward more intense low-pressure systems. However, the trend of
a decreasing number of cyclones seems to depend on the choice of
reanalysis and pressure level (Lim and Simmonds, 2009), emphasizing
the weaker consistency among reanalysis products for the Southern
Hemisphere extratropical cyclones. Recent studies support the notion
of more cyclones around Antarctica when the SAM (see Section 3.4.3)
is in its positive phase and a shift of cyclones toward mid-latitudes
when the SAM is in its negative phase (Pezza and Simmonds, 2008).
Additionally, more intense (and fewer) cyclones seem to occur when
the PDO (see Section 3.4.3) is strongly positive and vice versa (Pezza et
al., 2007).
In conclusion, it is likely that there has been a poleward shift in the
main northern and southern storm tracks during the last 50 years. There
is strong agreement with respect to this change between several
reanalysis products for a wide selection of cyclone parameters and
cyclone identification methods and European and Australian pressure-
based storminess proxies are consistent with a poleward shift over the
last 50 years, which indicates that the evidence is robust. Advances have
been made in documenting the observed decadal and multi-decadal
variability of extratropical cyclones using proxies for storminess. So the
recent poleward shift should be seen in light of new studies with longer
time spans that indicate that the last 50 years coincide with relatively
low cyclonic activity in northern coastal Europe in the beginning of the
period. Several studies using reanalyses suggest an intensification of
high-latitude cyclones, but there is still insufficient knowledge of how
changes in the observational systems are influencing the cyclone
intensification in reanalyses so even in cases of high agreement among
the studies the evidence cannot be considered to be robust, thus we
have only low confidence in these changes. Other regional changes in
intensity and the number of cyclones have been reported. However, the
level of agreement between different studies using different tracking
algorithms, different reanalyses, or different cyclone parameters is still
low. Thus, we have low confidence in the amplitude, and in some
regions in the sign, of the regional changes.
Regarding possible causes of the observed poleward shift, the AR4
concluded that trends over recent decades in the Northern and
Southern Annular Modes, which correspond to sea level pressure
reductions over the poles, are likely related in part to human activity,
but an anthropogenic influence on extratropical cyclones had not been
formally detected, owing to large internal variability and problems due
to changes in observing systems (Hegerl et al., 2007). Anthropogenic
influences on these modes of variability are also discussed in Section
3.4.3.
Seasonal global sea level pressure changes have been shown to be
inconsistent with simulated internal variability (Giannini et al., 2003;
Gillett et al., 2005; Gillett and Stott, 2009; X.L. Wang et al., 2009a), but
changes in sea level pressure in regions of extratropical cyclones (mid-
and high latitudes) have not formally been attributed to anthropogenic
forcings (Gillett and Stott, 2009). However, the trend pattern in
atmospheric storminess as inferred from geostrophic wind energy and
ocean wave heights has been found to contain a detectable response to
anthropogenic and natural forcings with the effect of external forcings
being strongest in the winter hemisphere (X.L. Wang et al., 2009a).
Nevertheless, the models generally simulate smaller changes than
observed and also appear to underestimate the internal variability,
reducing the robustness of their detection results. New idealized studies
have advanced the physical understanding of how storm tracks may
respond to changes in the underlying surface conditions, indicating that
a uniform SST increase weakens (reduced cyclone intensity or number
of cyclones) and shifts the storm track poleward and strengthened SST
gradients near the subtropical jet may lead to a meridional shift in the
storm track either toward the poles or the equator depending on the
location of the SST gradient change (Deser et al., 2007; Brayshaw et al.,
2008; Semmler et al., 2008; Kodama and Iwasaki, 2009), but the average
global cyclone activity is not expected to change much under moderate
greenhouse gas forcing (O’Gorman and Schneider, 2008; Bengtsson et al.,
2009). Studies have also emphasized the important role of stratospheric
changes (induced by ozone or greenhouse gas changes) in explaining
latitudinal shifts in storm tracks and several mechanisms have been
proposed (Son et al., 2010). This has particularly strengthened the
understanding of the Southern Hemisphere changes. According to Fogt
et al. (2009) both coupled climate models and observed trends in the
SAM were found to be outside the range of internal climate variability
during the austral summer. This was mainly attributed to stratospheric
ozone depletion (see Section 3.4.3).
In summary, there is medium confidence in an anthropogenic influence
on the observed poleward shift in extratropical cyclone activity. It has
not formally been attributed. However indirect evidence such as global
anthropogenic influence on the sea level pressure distribution and trend
patterns in atmospheric storminess inferred from geostrophic wind and
ocean wave heights has been found. While physical understanding of
how anthropogenic forcings may influence extratropical cyclone storm
tracks has strengthened, the importance of the different mechanisms in
the observed shifts is still unclear.
The AR4 reported that in a future warmer climate, a consistent projection
from the majority of the coupled atmosphere-ocean GCMs is fewer
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
166
mid-latitude storms averaged over each hemisphere (Meehl et al.,
2007b) and a poleward shift of storm tracks in both hemispheres
(particularly evident in the Southern Hemisphere), with greater storm
activity at higher latitudes (Meehl et al., 2007b).
A poleward shift in the upper level tropospheric storm track due to
increased greenhouse gas forcing is supported by post-AR4 studies
(Lorenz and DeWeaver, 2007; O’Gorman, 2010; Wu et al., 2011). It
should be noted that other studies indicate that the poleward shift is
less clear when models including a full stratosphere or ozone recovery
are used (Huebener et al., 2007; Son et al., 2008; Morgenstern et al.,
2010; Scaife et al., 2011) and the strength of the poleward shift is often
seen more clearly in upper-level quantities than in low-level transient
parameters (Ulbrich et al., 2008). Post-AR4 single model studies support
the projection of a reduction in extratropical cyclones averaged over the
Northern Hemisphere during future warming (Finnis et al., 2007;
Bengtsson et al., 2009; Orsolini and Sorteberg, 2009). However, neither
the global changes in storm frequency or intensity were found to be
statistically significant by Bengtsson et al. (2009), although they were
accompanied by significant increases in total and extreme precipitation.
Models tend to project a reduction of winter cyclone activity throughout
the mid-latitude North Pacific and for some models a north-eastern
movement of the North Pacific storm track (Loeptien et al., 2008; Ulbrich
et al., 2008; Favre and Gershunov, 2009; McDonald, 2011). However, the
exact geographical pattern of cyclone frequency anomalies exhibits
large variations across models (Teng et al., 2008; Favre and Gershunov,
2009; Laine et al., 2009).
Using band-passed sea level pressure data from 16 CMIP3 coupled
GCMs, Ulbrich et al. (2008) showed regional increases in the storm track
activity over the Eastern North Atlantic/Western European area. This
eastward or southeastward extension of the storm track is also found in
other studies (Ulbrich et al., 2008; Laine et al., 2009; McDonald, 2011) and
may be attributed to a local minimum in ocean warming in the central
North Atlantic and subsequent local changes in baroclinicity (McDonald,
2011). In line with the eastward shift, Donat et al. (2010a) projected an
increase in wind storm days for central Europe by the end of the 21st
century. The increase varies according to the definition of storminess
and one model projects a decrease. A common deficiency among many
AR4 models is a coarsely resolved stratosphere and there are still
concerns that this may lead to systematic biases in the Atlantic storm
track response to increased anthropogenic forcing (Scaife et al., 2011).
A reduction in cyclone frequency along the Canadian east coast has
been reported (Bengtsson et al., 2006; Watterson, 2006; Pinto et al.,
2007a; Teng et al., 2008; Long et al., 2009). New results for Southern
Hemisphere cyclones confirm the previously projected poleward shift in
storm tracks under increased greenhouse gases (Lim and Simmonds,
2009). That study projected a reduction of Southern Hemisphere
extratropical cyclone frequency and intensity in mid-latitudes but a
slight increase at high latitudes. The poleward shift due to increased
greenhouse gases may be partly opposed by ozone recovery (Son et al.,
2010).
Detailed analyses of changes in physical mechanisms related to cyclone
changes in coupled climate models are still few. O’Gorman (2010) showed
that changes in mean available potential energy of the atmosphere can
account for much of the varied response in storm-track intensity to
global warming, implying that changes in storm-track intensity are
sensitive to competing effects of changes in temperature gradients and
static stability in different atmospheric levels. Using two coupled climate
models, Laine et al. (2009) indicate that the primary cause for synoptic
activity changes at the western end of the Northern Hemisphere storm
tracks is related to the baroclinic conversion processes linked to mean
temperature gradient changes in localized regions of the western
oceanic basins. They also found downstream changes in latent heat
release during the developing and mature stages of the cyclone to be of
importance and indicated that changes in diabatic process may be
amplified by the upstream baroclinic changes [stronger (weaker)
baroclinic activity in the west gives stronger (weaker) latent heat
release downstream]. Pinto et al. (2009) found that regional increases
in track density and intensity of extreme cyclones close to the British
Isles using a single model was associated with an eastward shift of the
jet stream into Europe, more frequent extreme values of baroclinicity,
and stronger upper level divergence.
The modeled reduction in Southern Hemisphere extratropical cyclone
frequency and intensity in the mid-latitudes has been attributed to the
tropical upper tropospheric warming enhancing static stability and
decreasing baroclinicity while an increased meridional temperature
gradient in the high latitudes is suggested to be responsible for the
increase in cyclone activity in this region (Lim and Simmonds, 2009). In
addition to details in the modeled changes in local baroclinicity and
diabatic changes, the geographical pattern of modeled response in
cyclone activity has been reported to be influenced by the individual
model’s structure of intrinsic modes of variability (Branstator and Selten,
2009) and biases in the climatology (Kidston and Gerber, 2010).
In summary it is likely that there has been a poleward shift in the
main Northern and Southern Hemisphere extratropical storm
tracks during the last 50 years. There is medium confidence in an
anthropogenic influence on this observed poleward shift. It has
not formally been attributed. There is low confidence in past
changes in regional intensity. There is medium confidence that
an increased anthropogenic forcing will lead to a reduction in the
number of mid-latitude cyclones averaged over each hemisphere,
and there is also medium confidence in a poleward shift of the
tropospheric storm tracks due to future anthropogenic forcings.
Regional changes may be substantial and CMIP3 simulations show
some regions with medium agreement. However, there are still
uncertainties related to how the poorly resolved stratosphere in
many CMIP3 models may influence the regional results. In addition,
studies using different analysis techniques, different physical
quantities, different thresholds, and different atmospheric vertical
levels to represent cyclone activity and storm tracks result in
different projections of regional changes. This leads to low
confidence in region-specific projections.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
167
3.5. Observed and Projected Impacts on the
Natural Physical Environment
3.5.1. Droughts
Drought is generally “a period of abnormally dry weather long enough to
cause a serious hydrological imbalance” (see the Glossary and Box 3-3).
While lack of precipitation (i.e., meteorological drought; Box 3-3) is often
the primary cause of drought, increased potential evapotranspiration
induced by enhanced radiation, wind speed, or vapor pressure deficit (itself
linked to temperature and relative humidity), as well as pre-conditioning
(pre-event soil moisture; lake, snow, and/or groundwater storage)
can contribute to the emergence of soil moisture and hydrological
drought (Box 3-3). Actual evapotranspiration is additionally controlled
by soil moisture, which constitutes a limiting factor for further drying
under drought conditions, and other processes that impact vegetation
development and phenology (e.g., temperature) are also relevant. As
noted in the AR4 (Trenberth et al., 2007), there are few direct observations
of drought-related variables, in particular of soil moisture, available for
a global analysis (see also Section 3.2.1). Hence, proxies for drought
(‘drought indices’) are often used to infer changes in drought conditions.
Box 3-3 provides a discussion of the issue of drought definition and a
description of commonly used drought indices. In order to understand the
impact of droughts (e.g., on crop yields, general ecosystem functioning,
water resources, and electricity production), their timing, duration,
intensity, and spatial extent need to be characterized. Several weather
elements may interact to increase the impact of droughts: enhanced air
temperature can indirectly lead to enhanced evaporative demand
(through enhanced vapor pressure deficit), although enhanced wind
speed or increased incoming radiation are generally more important
factors. Moreover, climate phenomena such as monsoons (Section 3.4.1)
and ENSO (Section 3.4.2) affect changes in drought occurrence in some
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Box 3-3 | The Definition of Drought
Though a commonly used term, drought is defined in various ways, and these definitional issues make the analysis of changes in
drought characteristics difficult. This explains why assessments of (past or projected) changes in drought can substantially differ between
published studies or chosen indices (see Section 3.5.1). Some of these difficulties and their causes are highlighted in this box.
What is Drought or Dryness?
The Glossary defines drought as follows: A period of abnormally dry weather long enough to cause a serious hydrological imbalance.
Drought is a relative term, therefore any discussion in terms of precipitation deficit must refer to the particular precipitation-related
activity that is under discussion. For example, shortage of precipitation during the growing season impinges on crop production or
ecosystem function in general (due to soil moisture drought, also termed agricultural drought), and during the runoff and percolation
season primarily affects water supplies (hydrological drought). Storage changes in soil moisture and groundwater are also affected by
increases in actual evapotranspiration in addition to reductions in precipitation. A period with an abnormal precipitation deficit is
defined as a meteorological drought. A megadrought is a very lengthy and pervasive drought, lasting much longer than normal, usually a
decade or more.
As highlighted in the above definition, drought can be defined from different perspectives, depending on the stakeholders involved. The
scientific literature commonly distinguishes meteorological drought, which refers to a deficit of precipitation, soil moisture drought
(often called agricultural drought), which refers to a deficit of (mostly root zone) soil moisture, and hydrological drought, which refers to
negative anomalies in streamflow, lake, and/or groundwater levels (e.g., Heim Jr., 2002). We use here the term ‘soil moisture drought’
instead of ‘agricultural drought,’ despite the widespread use of the latter term (e.g., Heim Jr., 2002; Wang, 2005), because soil moisture
deficits have several additional effects beside those on agroecosystems, most importantly on other natural or managed ecosystems
(including both forests and pastures), on building infrastructure through soil mechanical processes (e.g., Corti et al., 2009), and health
through impacts on heat waves (Section 3.1.4). Water scarcity (linked to socioeconomic drought), which may be caused fully or in part
by use from human activities, does not lie within the scope of this chapter (see Section 4.2.2); however, it should be noted that changing
pressure on water resources by human uses may itself influence climate and possibly the drought conditions, for example, via declining
groundwater levels, or enhanced local evapotranspiration and associated land-atmosphere feedbacks. Drought should not be confused
with aridity, which describes the general characteristic of an arid climate (e.g., desert). Indeed, drought is considered a recurring feature
of climate occurring in any region and is defined with respect to the average climate of the given region (e.g., Heim Jr., 2002; Dai, 2011).
Nonetheless, the effects of droughts are not linear, given the existence of, for example, discrete soil moisture thresholds affecting
vegetation and surface fluxes (e.g., Koster et al., 2004b; Seneviratne et al., 2010), which means that the same precipitation deficit or
radiation excess relative to normal will not affect different regions equally (e.g., short-term lack of precipitation in a very humid region
may not be critical for agriculture because of the ample soil moisture supply). In this chapter we often use the term ‘dryness’ instead of
‘drought’ as a more general term.
Continued next page
168
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Drought Drivers
For soil moisture or hydrological droughts, the main drivers are reduced precipitation and/or increased evapotranspiration (Figure 3-9).
Although the role of deficits in precipitation is generally considered more prominently in the literature, several drought indicators also
explicitly or indirectly consider effects of evapotranspiration. In the context of climate projections, analyses suggest that changes in
simulated soil moisture drought are mostly driven by changes in precipitation, with increased evapotranspiration from higher vapor pressure
deficit (often linked to increased temperature) and available radiation modulating some of the changes (e.g., Burke and Brown, 2008;
Sheffield and Wood, 2008a; Orlowsky and Seneviratne, 2011). It should nonetheless be noted that under strong drought conditions, soil
moisture becomes limiting for evapotranspiration, thus limits further soil moisture depletion. Other important aspects for soil moisture
and hydrological droughts are persistence and pre-conditioning. Because soil moisture, groundwater, and surface waters are associated
with water storage, they have a characteristic memory (e.g., Vinnikov et al., 1996; Eltahir and Yeh, 1999; Koster and Suarez, 2001;
Seneviratne et al., 2006b) and thus specific response times to drought forcing (e.g., Begueria et al., 2010; Fleig et al., 2011). The memory
is also a function of the atmospheric forcing and system’s feedbacks
(Koster and Suarez, 2001; A.H. Wang et al., 2009), and the relevant
storage is dependent on soil characteristics and rooting depth of
the considered ecosystems. This means that drought has a different
persistence depending on the affected system, and that it is also
sensitive to pre-conditioning (Figure 3-9). Effects of pre-conditioning
also explain the possible occurrence of multi-year droughts, whereby
soil moisture anomalies can be carried over from one year to the
next (e.g., Wang, 2005). However, other features can induce
drought persistence, such as persistent circulation anomalies,
possibly strengthened by land-atmosphere feedbacks (Schubert et
al., 2004; Rowell and Jones, 2006). The choice of variable (e.g.,
precipitation, soil moisture, or streamflow) and time scale can
strongly affect the ranking of drought events (Vidal et al., 2010).
Drought Indices
Because of the complex definition of droughts, and the lack of soil moisture observations (Section 3.2.1), several indices have been
developed to characterize (meteorological, soil moisture, and hydrological) drought (see, e.g., Heim Jr., 2002; Dai, 2011). These indicators
include land surface, hydrological, or climate model simulations (providing estimates of, e.g., soil moisture or runoff) and indices based
on measured meteorological or hydrological variables. We provide here a brief overview of the wide range of drought indices used in the
literature for the analysis of recent and projected changes. Note that information on paleoclimate proxies such as tree rings,
speleothems, lake sediments, or historical evidence (e.g., harvest dates) is not detailed here.
Some indices are based solely on precipitation data. A widely used index is the Standard Precipitation Index (SPI) (McKee et al., 1993;
Lloyd-Hughes and Saunders, 2002), which consists of fitting and transforming a long-term precipitation record into a normal distribution
that has zero mean and unit standard deviation. SPI values of -0.5 to -1 correspond to mild droughts, -1 to -1.5 to moderate droughts,
-1.5 to -2 to severe droughts, and below -2 to extreme droughts. Similarly, values from 0 to 2 correspond to mildly wet to severely wet
conditions, and values above 2 to extremely wet conditions (Lloyd-Hughes and Saunders, 2002). SPI can be computed over several time
scales (e.g., 3, 6, 12, or more months) and thus indirectly considers effects of accumulating precipitation deficits, which are critical for
soil moisture and hydrological droughts. Another index commonly used in the analysis of climate model simulations is the Consecutive
Dry Days (CDD) index, which considers the maximum consecutive number of days without rain (i.e., below a given threshold, typically
1 mm day
-1
) within a considered period (i.e., year in general; Frich et al., 2002; Alexander et al., 2006; Tebaldi et al., 2006). For seasonal
time frames, the CDD periods can either be considered to be bound to the respective seasons (e.g., Figure 3-10) or considered in their
entirety (across seasons) but assigned to a specific season. Though SPI and CDD are both only based on precipitation, they do not
necessarily only consider the effects of meteorological drought, since periods without rain (thus less cloud cover) are bound to have
higher daytime radiation forcing and generally higher temperatures, thus possibly positive evapotranspiration anomalies (unless soil
moisture conditions are too dry and limit evapotranspiration).
Some indices reflect both precipitation and estimates of actual or potential evapotranspiration, in some cases also accounting for some
temporal accumulation of the forcings or persistence of the drought anomalies. These include the Palmer Drought Severity Index (PDSI)
Precipitation deficit
(meteorological drought)
Evapotranspiration
Pre-event soil moisture,
surface water, and/or
groundwater storage
Critical soil moisture deficit
(soil moisture drought)
Critical streamflow and
groundwater deficit
(hydrological drought)
Figure 3-9 | Simplified sketch of processes and drivers relevant for meteorological,
soil moisture (agricultural), and hydrological droughts.
Continued next page
169
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
(Palmer, 1965), which measures the departure of moisture balance from normal conditions using a simple water balance model (e.g., Dai,
2011), as well as other indices such as the Precipitation Potential Evaporation Anomaly (PPEA, based on the cumulative difference
between precipitation and potential evapotranspiration) used in Burke and Brown (2008) and the Standardized Precipitation-
Evapotranspiration Index (SPEI, which considers cumulated anomalies of precipitation and potential evapotranspiration) described in
Vicente-Serrano et al. (2010). PDSI has been widely used for decades (in particular in the United States), and also in climate change
analyses (e.g., Dai et al., 2004; Burke and Brown, 2008; Dai, 2011); however, it has some shortcomings for climate change monitoring
and projection. PDSI was originally calibrated for the central United States, which can impair the comparability of the index across
regions (and also across time periods if drought mechanisms change over time). Thus it is often of advantage to renormalize the local
PDSI (Dai, 2011), which can also be done using the self-calibrated PDSI (Wells et al., 2004), but several studies do not apply these steps.
Moreover, the land surface model underlying the computation of the PDSI is essentially a simple bucket-type model, which is less
sophisticated than more recent land surface and hydrological models and thus implies several limitations (e.g., Dai et al., 2004; Burke et
al., 2006). Another important issue is that the parameterization of potential evapotranspiration as empirically (and solely) dependent on
air temperature, which is often applied for these various indices (e.g., in the study of Dai et al., 2004) can lead to biased results (e.g.,
Donohue et al., 2010; Milly and Dunne, 2011; Shaw and Riha, 2011). Temperature is only an indirect driver of evapotranspiration, via its
effect on vapor pressure deficit and via effects on vegetation phenology. Furthermore, approaches using potential evapotranspiration as
a proxy for actual evapotranspiration do not consider soil moisture and vegetation control on evapotranspiration, which are important
mechanisms limiting drought development.
For the assessment of soil moisture drought, simulated soil moisture anomalies also can be considered (Wang et al., 2005; Burke and
Brown, 2008; Sheffield and Wood, 2008a; A.H. Wang et al., 2009; Dai, 2011; Orlowsky and Seneviratne, 2011). Simulated soil moisture
anomalies integrate the effects of precipitation forcing, simulated actual evapotranspiration (resulting from atmospheric forcing and
simulated soil moisture limitation on evapotranspiration), and simulated soil moisture persistence. Although the soil moisture simulated
by (land-surface, hydrological, and climate) models often exhibits strong discrepancies in absolute terms, soil moisture anomalies can be
compared with simple scaling and generally match reasonably well (e.g., Koster et al., 2009; A.H. Wang et al., 2009). Soil moisture
persistence is found to be an important component in projected changes in soil moisture drought, with some regions displaying year-
round dryness compared to reference (late 20th or pre-industrial) conditions due to the carry-over effect of soil moisture storage from
season to season, leading to year-round soil moisture deficits compared to late 20th century climate (e.g., Wang et al., 2005, Figure 3-10).
However, it should be noted that some land surface and hydrological models (used offline or coupled to climate models) suffer from
similar shortcomings as noted above for PDSI – that is, they use simple bucket models or simplified representations of potential
evapotranspiration. The latter issue has been suggested as being particularly critical for models used in offline mode (Milly and Dunne,
2011). Nonetheless, for the assessment of soil moisture drought, using simulated soil moisture anomalies seems less problematic than
many other indices for the reasons highlighted in the above paragraphs.
The indices listed above have been used in various studies analyzing drought in the context of climate change, but with a few exceptions
most available studies are based only on one index, which makes their comparison difficult. Nonetheless, these studies suggest that
projections can be highly dependent on the choice of drought index. For instance, one study projected changes in drought area possibly
varying between a negligible impact and a 5 to 45% increase depending on the drought index considered (Burke and Brown, 2008).
Other drought indices are used to quantify hydrological drought (e.g., Heim Jr., 2002; Vidal et al., 2010; Dai, 2011), but are less
commonly used in climate change studies. Further analyses or indices also consider the area affected by droughts (e.g., Burke et al.,
2006; Sheffield and Wood, 2008a; Dai, 2011) or additional variables (such as snow or vegetation indices from satellite measurements,
e.g., Heim Jr., 2002). As for the definition of other indices (Box 3-1), the determination of the reference period is critical for the assessment
of changes in drought patterns independently of the chosen index. In general, late 20th-century conditions are used as reference (e.g.,
Figure 3-10).
In summary, drought indices often integrate precipitation, temperature, and other variables, but may emphasize different
aspects of drought and should be carefully selected with respect to the drought characteristic in mind. In particular, some
indices have specific shortcomings, especially in the context of climate change. For this reason, assessments of changes in
drought characteristics with climate change should consider several indices including a specific evaluation of their relevance
to the addressed question to support robust conclusions. In this assessment we focus on the following indices: consecutive
dry days (CDD) and simulated soil moisture anomalies (SMA), although evidence based on other indices (e.g., PDSI for
present climate) is also considered (Section 3.5.1; Tables 3-2 and 3-3).
170
regions. Hence, drought is a complex phenomenon that is strongly
affected by other extremes considered in this chapter, but that is also
affected by changes in mean climate features (Section 3.1.6). In addition,
via land-atmosphere interactions, drought also has the potential to
impact other weather and climate elements such as temperature and
precipitation and associated extremes (Koster et al., 2004b; Seneviratne
et al., 2006a; Hirschi et al., 2011; see also Section 3.1.4). Case Study 9.2.3
addresses aspects related to the management of adverse consequences
of droughts; while Case Study 9.2.2 considers the possible impacts of
high temperatures and drought on wildfire.
Observed Changes
There are still large uncertainties regarding observed global-scale trends in
droughts. The AR4 reported based on analyses using PDSI (see Box 3-3)
that very dry areas had more than doubled in extent since 1970 at the
global scale (Trenberth et al., 2007). This assessment was, however,
largely based on the study by Dai et al. (2004) only. These trends in the
PDSI proxy were found to be largely affected by changes in temperature,
not precipitation (Dai et al., 2004). On the other hand, based on soil
moisture simulations with an observation-driven land surface model for
the time period 1950-2000, Sheffield and Wood (2008a) have inferred
trends in drought duration, intensity, and severity predominantly
decreasing, but with strong regional variation and including increases in
some regions. They concluded that there was an overall moistening trend
over the considered time period, but also a switch since the 1970s to a
drying trend, globally and in many regions, especially in high northern
latitudes. Some regional studies are consistent with the results from
Sheffield and Wood (2008a), regarding, for example, less widespread
increase (or statistically insignificant changes or decreases) in some
regions compared to the study of Dai et al. (2004) (e.g., in Europe, see
below). More recently, Dai (2011) by extending the record did, however,
find widespread increases in drought both based on various versions of
PDSI (for 1950-2008) and soil moisture output from a land surface model
(for 1948-2004). Hence there are still large uncertainties with respect to
global assessments of past changes in droughts. Nonetheless, there is
some agreement between studies over the different time frames (i.e.,
since 1950 versus 1970) and using different drought indicators regarding
increasing drought occurrence in some regions (e.g., southern Europe,
West Africa; see below and Table 3-2), although other regions also indicate
opposite trends (e.g., central North America, northwestern Australia; see
below and Table 3-2). As mentioned in Section 3.1.6, spatially coherent
shifts in drought regimes are expected with changing global circulation
patterns. Table 3-2 provides regional and continental-scale assessments
of observed trends in dryness based on different indices (Box 3-3). The
following paragraphs provide more details by continent.
From a paleoclimate perspective recent droughts are not unprecedented,
with severe ‘megadroughts’ reported in the paleoclimatic record for
Europe, North America, and Australia (Jansen et al., 2007). Recent studies
extend this observation to African and Indian droughts (Sinha et al.,
2007; Shanahan et al., 2009): much more severe and longer droughts
occurred in the past centuries with widespread ecological, political, and
socioeconomic consequences. Overall, these studies confirm that in the
last millennium several extreme droughts have occurred (Breda and
Badeau, 2008; Kallis, 2008; Büntgen et al., 2010).
In North America, there is medium confidence that there has been an
overall slight tendency toward less dryness (wetting trend with more soil
moisture and runoff; Table 3-2), although analyses for some subregions
also indicate tendencies toward increasing dryness. This assessment is
based on several lines of evidence, including simulations with different
hydrological models as well as PDSI and CDD estimates (Alexander et
al., 2006; Andreadis and Lettenmaier, 2006; van der Schrier et al., 2006a;
Kunkel et al., 2008; Sheffield and Wood, 2008a; Dai, 2011). The most
severe droughts in the 20th century have occurred in the 1930s and
1950s, where the 1930s Dust Bowl was most intense and the 1950s
drought most persistent (Andreadis et al., 2005) in the United States,
while in Mexico the 1950s and late 1990s were the driest periods.
Recent regional trends toward more severe drought conditions were
identified over southern and western Canada, Alaska, and Mexico, with
subregional exceptions (Dai, 2011).
In Europe, there is medium confidence regarding increases in dryness
based on some indices in the southern part of the continent, but large
inconsistencies between indices in this region, and inconsistent or
statistically insignificant trends in the rest of the continent (Table 3-2).
Although Dai et al. (2004) found an increase in dryness for most of the
European continent based on PDSI, Lloyd-Hughes and Saunders (2002)
and van der Schrier et al. (2006b) concluded, based on the analysis of
SPI and self-calibrating PDSI for the 20th century (for 1901-1999 and
1901-2002, respectively), that no statistically significant changes were
observed in extreme and moderate drought conditions in Europe [with
the exception of the Mediterranean region in van der Schrier et al.
(2006b)]. Sheffield and Wood (2008a) also found contrasting dryness
trends in Europe, with increases in the southern and eastern part of the
continent, but decreases elsewhere. Beniston (2009b) reported a strong
increase in warm-dry conditions over all central-southern (including
maritime) Europe via a quartile analysis from the middle to the end of
the 20th century. Alexander et al. (2006) found trends toward increasing
CDD mostly in the southern and central part of the continent. Trends of
decreasing precipitation and discharge are consistent with increasing
salinity in the Mediterranean Sea, indicating a trend toward freshwater
deficits (Mariotti et al., 2008), but this could also be partly caused by
increased human water use. In France, an analysis based on a variation
of the PDSI model also reported a significant increasing trend in drought
conditions, in particular from the 1990s onward (Corti et al., 2009).
Stahl et al. (2010) investigated streamflow data across Europe and
found negative trends (lower streamflow) in southern and eastern
regions, and generally positive trends (higher streamflow) elsewhere
(especially in northern latitudes). Low flows have decreased in most
regions where the lowest mean monthly flow occurs in summer, but
vary for catchments that have flow minima in winter and secondary low
flows in summer. The exceptional 2003 summer heat wave on the
European continent (see Section 3.3.1) was also associated with a
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
171
major soil moisture drought, as could be inferred from satellite
measurements (Andersen et al., 2005), model simulations (Fischer et al.,
2007a,b), and impacts on ecosystems (Ciais et al., 2005; Reichstein et
al., 2007).
There is low confidence in dryness trends in South America (Table 3-2),
partly due to lack of data and partly due to inconsistencies. For the
Amazon, repeated intense droughts have been occurring in the last
decades but no particular trend has been reported. The 2005 and 2010
droughts in Amazonia are, however, considered the strongest in the last
century as inferred from integrating precipitation records and water
storage estimates via satellite (measurements from the Gravity
Recovery and Climate Experiment; Chen et al., 2009; Lewis et al., 2011).
For other parts of South America, analyses of the return intervals
between droughts in the instrumental and reconstructed precipitation
series indicate that the probability of drought has increased during the
late 19th and 20th centuries, consistent with selected long instrumental
precipitation records and with a recession of glaciers in the Chilean and
Argentinean Andean Cordillera (Le Quesne et al., 2006, 2009).
Changes in drought patterns have been reported for the monsoon regions
of Asia and Africa with variations at the decadal time scale (e.g., Janicot,
2009). In Asia there is overall low confidence in trends in dryness both
at the continental and regional scale, mostly due to spatially varying
trends, except in East Asia where a range of studies, based on different
indices, show increasing dryness in the second half of the 20th century,
leading to medium confidence (Table 3-2).
In the Sahel, recent years have been characterized by greater interannual
variability than the previous 40 years (Ali and Lebel, 2009; Greene et al.,
2009), and by a contrast between the western Sahel remaining dry and
the eastern Sahel returning to wetter conditions (Ali and Lebel, 2009).
Giannini et al. (2008) report a drying of the African monsoon regions,
related to warming of the tropical oceans, and variability related to ENSO.
In the different subregions of Africa there is overall low to medium
confidence regarding regional dryness trends (Table 3-2).
For Australia, Sheffield and Wood (2008a) found very limited increases
in dryness from 1950 to 2000 based on soil moisture simulated using
existing climate forcing (mostly in southeastern Australia) and some
marked decreases in dryness in central Australia and the northwestern
part of the continent. Dai (2011), for an extended period until 2008 and
using different PDSI variants as well as soil moisture output from a land
surface model, found a more extended drying trend in the eastern half
of the continent, but also a decrease in dryness in most of the western
half. Jung et al. (2010) inferred from a combination of remote sensing
and quasi-globally distributed eddy covariance flux observations that in
particular the decade after 1998 became drier in Australia (and parts of
Africa and South America), leading to decreased evapotranspiration, but
it is not clear if this is a trend or just decadal variation.
Following the assessment of observed changes in the AR4 (Chapter 3),
which was largely based on one study (Dai et al., 2004), subsequent
work has drawn a more differentiated picture both regionally and
temporally. There is not enough evidence at present to suggest high
confidence in observed trends in dryness due to lack of direct observations,
some geographical inconsistencies in the trends, and some dependencies
of inferred trends on the index choice. There is medium confidence that
since the 1950s some regions of the world have experienced more
intense and longer droughts (e.g., southern Europe, west Africa) but
also opposite trends exist in other regions (e.g., central North America,
northwestern Australia).
Causes of the Observed Changes
The AR4 (Hegerl et al., 2007) concluded that it is more likely than not
that anthropogenic influence has contributed to the increase in the
droughts observed in the second half of the 20th century. This assessment
was based on several lines of evidence, including a detection study that
identified an anthropogenic fingerprint in a global PDSI data set with
high significance (Burke et al., 2006), although the model trend was
weaker than observed and the relative contributions of natural external
forcings and anthropogenic forcings were not assessed.
There is now a better understanding of the potential role of land-
atmosphere feedbacks versus SST forcing for meteorological droughts
(e.g., Schubert et al., 2008a,b), and some modeling studies have also
addressed potential impacts of land use changes (e.g., Deo et al., 2009),
but large uncertainties remain in the field of land surface modeling and
land-atmosphere interactions, in part due to lack of observations
(Seneviratne et al., 2010), inter-model discrepancies (Koster et al., 2004b;
Dirmeyer et al., 2006; Pitman et al., 2009), and model resolution of
orographic and other effects. Nonetheless, a new set of climate modeling
studies show that US drought response to SST variability is consistent
with observations (Schubert et al., 2009). Inferred trends in drought are
also consistent with trends in global precipitation and temperature, and
the latter two are consistent with expected responses to anthropogenic
forcing (Hegerl et al., 2007; X. Zhang et al., 2007). The change in the pattern
of global precipitation in the observations and in model simulations is
also consistent with the theoretical understanding of hydrological
response to global warming that wet regions become overall wetter
and dry regions drier in a warming world (Held and Soden, 2006; see
also Section 3.1.6), though some regions also display shifts in climate
regimes (Section 3.1.6). Nonetheless, some single events have been
reported as differing from projections (Seager et al., 2009), though this is
not necessarily incompatible given the superimposition of anthropogenic
climate change and natural climate variability (Section 3.1). For soil
moisture and hydrological drought it has been suggested that the
stomatal ‘antitranspirant’ responses of plants to rising atmospheric CO
2
may lead to a decrease in evapotranspiration (Gedney et al., 2006). This
could mean that increasing CO
2
levels alleviate soil moisture and
streamflow drought, but this result is still debated (e.g., Piao et al.,
2007; Gerten et al., 2008), in particular due to the uncertainty in
observed runoff trends used to infer these effects (e.g., Peel and
McMahon, 2006; see also Section 3.2.1).
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
172
Overall, though new studies have furthered the understanding of the
mechanisms leading to drought, there is still relatively limited evidence
to provide an attribution of observed changes, in particular given the
issues associated with the availability of observational data (Section 3.2.1)
and the definition and computation of drought indicators (Box 3-3). This
latter point was mostly identified in post-AR4 studies (Box 3-3). Moreover,
regions where consistent increases in drought are identified (see ‘Observed
Changes’) are only partly consistent with those where projections indicate
an enhancement of drought conditions in coming decades (see next
paragraphs). We thus assess that there is medium confidence (see also
Section 3.1.5) that anthropogenic influence has contributed to some
changes in the drought patterns observed in the second half of the 20th
century, based on its attributed impact on precipitation and temperature
changes (though temperature can only be indirectly related to drought
trends; see Box 3-3). However there is low confidence in the attribution
of changes in droughts at the level of individual regions.
Projected Changes and Uncertainties
The AR4 assessed that projections at the time indicated an increase in
droughts, in particular in subtropical and mid-latitude areas
(Christensen et al., 2007). An increase in dry spell length and frequency
was considered very likely over the Mediterranean region, southern
areas of Australia, and New Zealand and likely over most subtropical
regions, with little change over northern Europe. Continental drying and
the associated risk of drought were considered likely to increase in
summer over many mid-latitude continental interiors (e.g., central and
southern Europe, the Mediterranean region), in boreal spring, and dry
periods of the annual cycle over Central America.
More recent global and regional climate simulations and hydrological
models mostly support the projections from the AR4, as summarized in the
following paragraphs (see also Table 3-3), although we assess the overall
confidence in drought projections as medium given the definitional
issues associated with dryness and the partial lack of agreement in
model projections when based on different dryness indices (Box 3-3).
Indeed, particular care is needed in inter-comparing ‘drought’ projections
since very many different definitions are employed (corresponding to
different types of droughts), from simple climatic indices such as CDD
to more complex indices of soil moisture and hydrological drought (Box
3-3). A distinction also needs to be made between short-term and
longer-term events. Blenkinsop and Fowler (2007a) and Burke et al.
(2010), for example, show different trend strength, and sometimes sign
(Blenkinsop and Fowler, 2007a), for changes in short- and long-term
droughts with RCM ensembles applied to the United Kingdom
(although uncertainties in the latter projections are large; see below).
These various distinctions are generally not considered and most
currently available studies only assess changes in very few (most
commonly one or two) dryness indices.
On the global scale, Burke and Brown (2008) provided an analysis of
projected changes in drought based on four indices (SPI, PDSI, PPEA,
and SMA; for definitions, see Box 3-3) using two model ensembles: one
based on a GCM expressing uncertainty in parameter space, and a multi-
model ensemble of 11 GCM simulations from CMIP3. Their analysis
revealed that SPI, based solely on precipitation, showed little change in
the proportion of the land surface in drought, and that all other indices,
which include a measure of the atmospheric demand for moisture,
showed a statistically significant increase with an additional 5 to 45% of
the land surface in drought. This study also highlighted large uncertainties
in regional changes in drought. For reasons highlighted in Box 3-3, using
simulated soil moisture anomalies from the climate models avoids some
shortcomings of other commonly used indices (although the quality of
simulated soil moisture cannot be well evaluated due to lack of
observations; Section 3.2 and Box 3-3). In the study of Burke and Brown
(2008), this index showed weaker drying compared to PDSI and PPEA
indices (but more pronounced drying than the SPI index). In this report,
we display projected changes in soil moisture anomalies and CDD
(Figure 3-10), this latter index being chosen for continuity with the AR4
(see Figure 10.18 of that report). It can be seen that the two indices
partly agree on increased drought in some large regions (e.g., on the
annual time scale, in Southern Europe and the Mediterranean region,
central Europe, central North America, Central America and Mexico,
northeast Brazil, and southern Africa), but some regions where the models
show consistent increases in CDD (e.g., southeast Asia) do not show
consistent decreases in soil moisture. Conversely, regions displaying a
consistent decrease in CDD (e.g., in northeastern Asia) do not show a
consistent increase in soil moisture. The substantial uncertainty of drought
projections is particularly clear from the soil moisture projections, with,
for example, no agreement among the models regarding the sign of
changes in December to February over most of the globe. These results
regarding changes in CDD and soil moisture are consistent with other
published studies (Wang, 2005; Tebaldi et al., 2006; Burke and Brown,
2008; Sheffield and Wood, 2008b; Sillmann and Roeckner, 2008) and the
areas that display consistent increasing drought tendencies for both
indices have also been reported to display such tendencies for additional
indices (e.g., Burke and Brown, 2008; Dai, 2011; Table 3-3). Sheffield and
Wood (2008b) examined projections in drought frequency (for droughts
of duration of 4 to 6 months and longer than 12 months, estimated from
soil moisture anomalies) based on CMIP3 simulations with eight GCMs
and the SRES scenarios A2, A1B, and B1. They concluded that drought was
projected to increase in several regions under these three scenarios
(mostly consistent with those displayed in Figure 3-10 for SMA),
although the projections of drought intensification were stronger for the
high CO
2
emissions scenarios (A2 and A1B) than for the more moderate
scenario (B1). Regions showing statistically significant increases in drought
frequency were found to be broadly similar for all three scenarios,
despite the more moderate signal in the B1 scenario (their Figures 8 and
9). This study also highlighted the large uncertainty of scenarios for
drought projections, as scenarios were found to span a large range of
changes in drought frequency in most regions, from close to no change
to two- to three-fold increases (their Figure 10).
Regional climate simulations and high-resolution global atmospheric
model simulations over Europe also highlight the Mediterranean region
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
173
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
20462065 (ANN)
Consecutive dry days (CDD)





 
 

  
  
  


 


  
     
   
    
 





20462065 (ANN)
Soil moisture anomalies (SMA)





20812100 (ANN)



 

 


 
 

 
 

 
 
  
   
 
  
   
    
     
  

 
  
   
   
    
   
   
  
 






20812100 (ANN)






20812100 (DJF)






 



 
   
   
 
  
  

 
 
 
  
 
   
        
       
    
    
   
  
  






20812100 (DJF)








20812100 (JJA)


  
  
 






  
  
  
  
    





20812100 (JJA)






 


 


Standard Deviation
0.6 0.2 0.2 0.60
Standard Deviation
0.75 0.25 0.25 0.750
Figure 3-10 | Projected annual and seasonal changes in dryness assessed from two indices for 2081-2100 (bottom three rows, showing the annual time scale and two
seasons, DJF and JJA) and 2046-2065 (top, annual time scale) with respect to 1980-1999. Left column: changes in the maximum number of CDD (days with precipitation
<1 mm), based on 17 GCMs contributing to the CMIP3. Right column: changes in soil moisture (soil moisture anomalies, SMA), based on 15 GCMs contributing to
the CMIP3. Increased dryness is indicated with warm colors (positive changes in CDD and negative SMA values). The maps show differences between the annual and
seasonal averages over the respective 20-year periods, that is, the average of 2081-2100 or 2046-2065, respectively (based on simulations under emission scenario
SRES A2), minus the average of 1980-1999 (from corresponding simulations for the 20th century). Differences are expressed in units of standard deviations, derived
from detrended per year annual or seasonal estimates, respectively, from the three 20-year periods 1980-1999, 2046-2065, and 2081-2100 pooled together. Color
shading is only applied for areas where at least 66% of the GCMs (12 out of 17 for CDD, 10 out of 15 for soil moisture) agree on the sign of the change; stippling is
applied for regions where at least 90% of the GCMs (16 out of 17 for CDD, 14 out of 15 for soil moisture) agree on the sign of the change. Adapted from Orlowsky and
Seneviratne (2011); updating Tebaldi et al. (2006) for SMA and for additional CMIP3 models, and including seasonal time frames. For more details, see Appendix 3.A.
174
as being affected by more severe droughts, consistent with available
global projections (Table 3-3; see also Giorgi, 2006; Rowell and Jones,
2006; Beniston et al., 2007; Mariotti et al., 2008; Planton et al., 2008).
Mediterranean (summer) droughts are projected to start earlier in the
year and last longer. Also, increased variability during the dry and warm
season is projected (Giorgi, 2006). One GCM-based study projected one
to three weeks of additional dry days for the Mediterranean region by
the end of the century (Giannakopoulos et al., 2009). For North America,
intense and heavy episodic rainfall events with high runoff amounts
are interspersed with longer relatively dry periods with increased
evapotranspiration, particularly in the subtropics. There is a consensus of
most climate model projections for a reduction in cool season precipitation
across the US southwest and northwest Mexico (Christensen et al., 2007),
with more frequent multi-year drought in the US southwest (Seager et al.,
2007; Cayan et al., 2010). Reduced cool season precipitation promotes
drier summer conditions by reducing the amount of soil water available
for evapotranspiration in summer. For Australia, Alexander and Arblaster
(2009) project increases in consecutive dry days, although consensus
between models is only found in the interior of the continent. African
studies indicate the possibility of relatively small-scale (500-km)
heterogeneity of changes in precipitation and drought, based on climate
model simulations (Funk et al., 2008; Shongwe et al., 2009). Regional
climate simulations of South America project spatially coherent increases
in CDD, particularly large over the Brazilian Plateau, and northern Chile
and the Altiplano (Kitoh et al., 2011).
Available global and regional studies of hydrological drought (Hirabayashi
et al., 2008b; Feyen and Dankers, 2009) project a higher likelihood of
hydrological drought by the end of this century, with a substantial
increase in the number of drought days (defined as streamflow below a
specific threshold) during the last 30 years of the 21st century over
North and South America, central and southern Africa, the Middle East,
southern Asia from Indochina to southern China, and central and western
Australia. Some regions, including eastern Europe to central Eurasia,
inland China, and northern North America, project increases in drought.
In contrast, wide areas over eastern Russia project a decrease in drought
days. At least in Europe, hydrological drought is primarily projected to
occur in the frost-free season.
Increased confidence in modeling drought stems from consistency
between models and satisfactory simulation of drought indices during
the past century (Sheffield and Wood, 2008a; Sillmann and Roeckner,
2008). Inter-model agreement is stronger for long-term droughts and
larger spatial scales (in some regions, see above discussion), while local to
regional and short-term precipitation deficits are highly spatially variable
and much less consistent between models (Blenkinsop and Fowler,
2007b). Insufficient knowledge of the physical causes of meteorological
droughts, and of the links to the large-scale atmospheric and ocean
circulation, is still a source of uncertainty in drought simulations and
projections. For example, plausible explanations have been proposed for
projections of both a worsening drought and a substantial increase in
rainfall in the Sahara (Biasutti et al., 2009; Burke et al., 2010). Another
example is illustrated with the relationship of rainfall in southern
Australia with SSTs around northern Australia. On annual time scales,
low rainfall is associated with cooler than normal SSTs. Yet the warming
observed in SST over the past few decades has not been associated with
increased rainfall, but with a trend toward more drought-like conditions
(N. Nicholls, 2010).
There are still further sources of uncertainties affecting the projections
of trends in meteorological drought for the coming century. The two
most important may be uncertainties in the development of the ocean
circulation and feedbacks between land surface and atmospheric
processes. These latter processes are related to the effects of drought on
vegetation physiology and dynamics (e.g., affecting canopy conductance,
albedo, and roughness), with resulting (positive or negative) feedbacks
to precipitation formation (Findell and Eltahir, 2003a,b; Koster et al.,
2004b; Cook et al., 2006; Hohenegger et al., 2009; Seneviratne et al.,
2010; van den Hurk and van Meijgaard, 2010), and possibly – as only
recently highlighted – also feedbacks between droughts, fires, and
aerosols (Bevan et al., 2009). Furthermore, the development of soil
moisture that results from complex interactions among precipitation,
water storage as soil moisture (and snow), and evapotranspiration by
vegetation is still associated with large uncertainties, in particular
because of lack of observations of soil moisture and evapotranspiration
(Section 3.2.1), and issues in the representation of soil moisture-
evapotranspiration coupling in current climate models (Dirmeyer et al.,
2006; Seneviratne et al., 2010). Uncertainties regarding soil moisture-
climate interactions are also due to uncertainties regarding the behavior
of plant transpiration, growth, and water use efficiency under enhanced
atmospheric CO
2
concentrations, which could potentially have impacts
on the hydrological cycle (Betts et al., 2007), but are not well understood
yet (Hungate et al., 2003; Piao et al., 2007; Bonan, 2008; Teuling et al.,
2009; see also above discussion on the causes of observed changes).
The space-time development of hydrological drought as a response to a
meteorological drought and the associated soil moisture drought
(drought propagation, e.g., Peters et al., 2003) also needs more attention.
There is some understanding of these issues at the catchment scale
(e.g., Tallaksen et al., 2009), but these need to be extended to the
regional and continental scales. This would lead to better understanding
of the projections of hydrological droughts, which would contribute to
a better identification and attribution of droughts and help to improve
global hydrological models and land surface models.
In summary, there is medium confidence that since the 1950s
some regions of the world have experienced trends toward more
intense and longer droughts, in particular in southern Europe
and West Africa, but in some regions droughts have become less
frequent, less intense, or shorter, for example, central North
America and northwestern Australia. There is medium confidence
that anthropogenic influence has contributed to some changes
in the drought patterns observed in the second half of the 20th
century, based on its attributed impact on precipitation and
temperature changes (though temperature can only be indirectly
related to drought trends; see Box 3-3). However there is low
confidence in the attribution of changes in droughts at the level
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
175
of single regions due to inconsistent or insufficient evidence.
Post-AR4 studies indicate that there is medium confidence in a
projected increase in duration and intensity of droughts in some
regions of the world, including southern Europe and the
Mediterranean region, central Europe, central North America,
Central America and Mexico, northeast Brazil, and southern
Africa. Elsewhere there is overall low confidence because of
insufficient agreement of projections of drought changes
(dependent both on model and dryness index). Definitional
issues and lack of data preclude higher confidence than medium
in observations of drought changes, while these issues plus the
inability of models to include all the factors likely to influence
droughts preclude stronger confidence than medium in the
projections.
3.5.2. Floods
A flood is “the overflowing of the normal confines of a stream or other
body of water, or the accumulation of water over areas that are not
normally submerged (some specific examples are discussed in Case Study
9.2.6). Floods include river (fluvial) floods, flash floods, urban floods,
pluvial floods, sewer floods, coastal floods, and glacial lake outburst
floods” (see Glossary). The main causes of floods are intense and/or
long-lasting precipitation, snow/ice melt, a combination of these causes,
dam break (e.g., glacial lakes), reduced conveyance due to ice jams or
landslides, or by a local intense storm (Smith and Ward, 1998). Floods
are affected by various characteristics of precipitation, such as intensity,
duration, amount, timing, and phase (rain or snow). They are also affected
by drainage basin conditions such as water levels in the rivers, the
presence of snow and ice, soil character and status (frozen or not, soil
moisture content and vertical distribution), rate and timing of snow/ice
melt, urbanization, and the existence of dikes, dams, and reservoirs (Bates
et al., 2008). Along coastal areas, flooding may be associated with storm
surge events (Section 3.5.5). A change in the climate physically changes
many of the factors affecting floods (e.g., precipitation, snow cover, soil
moisture content, sea level, glacial lake conditions, vegetation) and thus
may consequently change the characteristics of floods. Engineering
developments such as dikes and reservoirs regulate flow, and land use
may also affect floods. Therefore the assessment of causes of changes
in floods is complex and difficult. The focus in this section is on changes
in floods that might be related to changes in climate (i.e., referred to as
‘climate-driven’), rather than changes in engineering developments or
land use. However, because of partial lack of documentation, these can
be difficult to distinguish in the instrumental record.
Literature on the impact of climate change on pluvial floods (e.g., flash
floods and urban floods) is scarce, although the changes in heavy
precipitation discussed in Section 3.3.2 may imply changes in pluvial
floods in some regions. This chapter focuses on the spatial, temporal,
and seasonal changes in high flows and peak discharge in rivers related
to climate change, which cause changes in fluvial (river) floods. River
discharge simulation under a changing climate scenario requires a set
of GCM or RCM outputs (e.g., precipitation and surface air temperature)
and a hydrological model. A hydrological model may consist of a land
surface model of a GCM or RCM and a river routing model. Different
hydrological models may yield quantitatively different river discharge,
but they may not yield different signs of the trend if the same GCM/
RCM outputs are used. So the ability of models to simulate floods, in
particular regarding the signs of the past and future trends, depends on
the ability of the GCM or RCM to simulate precipitation changes. The
ability of a GCM or RCM to simulate temperature is important for river
discharge simulation in snowmelt- and glacier-fed rivers. Downscaling
and/or bias-correction are frequently applied to GCM/RCM outputs
before hydrological simulations are conducted, which becomes a source
of uncertainty. More details on the feasibility and uncertainties in
hydrological projections are described later in this section. Coastal
floods are discussed in Sections 3.5.3 and 3.5.5. Glacial lake outburst
floods are discussed in Section 3.5.6. The impact of floods on human
society and ecosystems and related changes are discussed in Chapter 4.
Case Study 9.2.6 discusses the management of floods.
Worldwide instrumental records of floods at gauge stations are limited
in spatial coverage and in time, and only a limited number of gauge
stations have data that span more than 50 years, and even fewer more
than 100 years (Rodier and Roche, 1984; see also Section 3.2.1). However,
this can be overcome partly or substantially by using pre-instrumental
flood data from documentary records (archival reports, in Europe
continuous over the last 500 years) (Brázdil et al., 2005), and from
geological indicators of paleofloods (sedimentary and biological records
over centennial to millennial scales) (Kochel and Baker, 1982). Analysis
of these pre-instrumental flood records suggest that (1) flood magnitude
and frequency can be sensitive to modest alterations in atmospheric
circulation, with greater sensitivity for ‘rare’ floods (e.g., 50-year flood and
higher) than for smaller and more frequent floods (e.g., 2-year floods)
(Knox, 2000; Redmond et al., 2002); (2) high interannual and interdecadal
variability can be found in flood occurrences both in terms of frequency
and magnitude although in most cases, cyclic or clusters of flood
occurrence are observed in instrumental (Robson et al., 1998), historical
(Vallve and Martin-Vide, 1998; Benito et al., 2003; Llasat et al., 2005),
and paleoflood records (Ely et al., 1993; Benito et al., 2008); (3) past
flood records may contain analogs of unusual large floods, similar to
some recorded recently, sometimes considered to be the largest on
record. For example, pre-instrumental flood data show that the 2002
summer flood in the Elbe did not reach the highest flood levels recorded
in 1118 and 1845 although it was higher than other disastrous floods
of 1432, 1805, etc. (Brázdil et al., 2006). However, the currently available
pre-instrumental flood data is also limited, particularly in spatial coverage.
The AR4 and the IPCC Technical Paper VI based on the AR4 concluded
that no gauge-based evidence had been found for a climate-driven
globally widespread change in the magnitude/frequency of floods during
the last decades (Rosenzweig et al., 2007; Bates et al., 2008). However,
the AR4 also pointed to possible changes that may imply trends in flood
occurrence with climate change. For instance, Trenberth et al. (2007)
highlighted a catastrophic flood that occurred along several central
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
176
European rivers in 2002, although neither flood nor mean precipitation
trends could be identified in this region; however, there was a trend
toward increasing precipitation variability during the last century which
itself could imply an enhanced probability of flood occurrence.
Kundzewicz et al. (2007) argued that climate change (i.e., observed
increase in precipitation intensity and other observed climate changes)
might already have had an impact on floods. Regarding the spring peak
flows, the AR4 concluded with high confidence that abundant evidence
was found for an earlier occurrence in snowmelt- and glacier-fed rivers
(Rosenzweig et al., 2007; Bates et al., 2008), though we expressly note
here that a change in the timing of peak flows does not necessarily
imply nor preclude changes in flood magnitude or frequency in the
affected regions.
Although changes in flood magnitude/frequency might be expected in
regions where temperature change affects precipitation type (i.e., rain/
snow separation), snowmelt, or ice cover (in particular northern high-
latitude and polar regions), widespread evidence of such climate-driven
changes in floods is not available. For example, there is no evidence of
widespread common trends in the magnitude of floods based on the
daily river discharge of 139 Russian gauge stations for the last few to
several decades, though a significant shift in spring discharge to earlier
dates has been found (Shiklomanov et al., 2007). Lindström and
Bergström (2004) noted that it is difficult to conclude that flood levels
are increasing from an analysis of runoff trends in Sweden for 1807 to
2002.
In the United States and Canada during the 20th century and in the
early 21st century, there is no compelling evidence for climate-driven
changes in the magnitude or frequency of floods (Lins and Slack, 1999;
Douglas et al., 2000; McCabe and Wolock, 2002; Cunderlik and Ouarda,
2009; Villarini et al., 2009). There are relatively abundant studies on the
changes and trends for rivers in Europe such as rivers in Germany and
its neighboring regions (Mudelsee et al., 2003; Tu et al., 2005; Yiou et
al., 2006; Petrow and Merz, 2009), in the Swiss Alps (Allamano et al.,
2009), in France (Renard et al., 2008), in Spain (Benito et al., 2005), and
in the United Kingdom (Robson et al., 1998; Hannaford and Marsh, 2008),
but a continental-scale assessment of climate-driven changes in the
flood magnitude and frequency for Europe is difficult to provide
because geographically organized patterns are not seen in the reported
changes.
Available (limited) analyses for Asia suggest the following changes: the
annual flood maxima of the lower Yangtze region show an upward
trend over the last 40 years (Jiang et al., 2008), the likelihood for
extreme floods in the Mekong River has increased during the second
half of the 20th century although the probability of an average flood
has decreased (Delgado et al., 2009), and both upward and downward
trends are identified over the last four decades in four selected river
basins of the northwestern Himalaya (Bhutiyani et al., 2008). In the
Amazon region in South America, the 2009 flood set record highs in the
106 years of data for the Rio Negro at the Manaus gauge site in July
2009 (Marengo et al., 2011). Recent increases have also been reported
in flood frequency in some other river basins in South America
(Camilloni and Barros, 2003; Barros et al., 2004). Conway et al. (2009)
concluded that robust identification of hydrological change was severely
limited by data limitations and other issues for sub-Saharan Africa.
Di Baldassarre et al. (2010) found no evidence that the magnitude of
African floods has increased during the 20th century. However, such
analyses cover only limited parts of the world. Evidence in the scientific
literature from the other parts of the world, and for other river basins,
appears to be very limited.
Many river systems are not in their natural state anymore, making it
difficult to separate changes in the streamflow data that are caused by
the changes in climate from those caused by human regulation of the
river systems. River engineering and land use may have altered flood
probability. Many dams are designed to reduce flooding. Large dams
have resulted in large-scale land use change and may have changed the
effective rainfall in some regions (Hossain et al., 2009).
The above analysis indicates that research subsequent to the AR4 still
does not show clear and widespread evidence of climate-driven
observed changes in the magnitude or frequency of floods at the global
level based on instrumental records, and there is thus low confidence
regarding the magnitude and frequency and even the sign of these
changes. The main reason for this lack of confidence is due to limited
evidence in many regions, since available instrumental records of floods
at gauge stations are limited in space and time, which limits the number
of analyses. Moreover, the confounding effects of changes in land use
and engineering mentioned above also make the identification of
climate-driven trends difficult. There are limited regions with medium
evidence, where no ubiquitous change is apparent (low agreement).
Pre-instrumental flood data can provide information for longer periods,
but current availability of these data is even scarcer particularly in spatial
coverage. There is abundant evidence for an earlier occurrence of spring
peak flows in snowmelt- and glacier-fed rivers (high confidence), though
this feature may not necessarily be linked with changes in the magnitude
of spring peak flows in the concerned regions.
The possible causes for changes in floods were discussed in the AR4 and
Bates et al. (2008), but cause-and-effect between external forcing and
changes in floods was not explicitly assessed. A rare example considered
in Rosenzweig et al. (2007) and Bates et al. (2008) was a study by Milly
et al. (2002) which, based on monthly river discharge, reported an
impact of anthropogenic climate change on changes (mostly increases)
in ‘large’ floods during the 20th century in selected extratropical river
basins larger than 20,000 km
2
, but they did not endorse the study
because of the lack of widespread observed evidence of such trends in
other studies. More recent literature has detected the influence of
anthropogenically induced climate change in variables that affect
floods, such as aspects of the hydrological cycle (see Section 3.2.2.2)
including mean precipitation (X. Zhang et al., 2007), heavy precipitation
(see Section 3.3.2), and snowpack (Barnett et al., 2008), though a direct
statistical link between anthropogenic climate change and trends in the
magnitude and frequency of floods is still not established.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
177
In climates where seasonal snow storage and melting play a significant
role in annual runoff, the hydrologic regime is affected by changes in
temperature. In a warmer world, a smaller portion of precipitation falls
as snow (Hirabayashi et al., 2008a) and the melting of winter snow
occurs earlier in spring, resulting in a shift in peak river runoff to winter
and early spring. This has been observed in the western United States
(Regonda et al., 2005; Clow, 2010), in Canada (Zhang et al., 2001), and
in other cold regions (Rosenzweig et al., 2007; Shiklomanov et al.,
2007), along with an earlier breakup of river ice in Arctic rivers (Smith,
2000; Beltaos and Prowse, 2009). The observed trends toward earlier
timing of snowmelt-driven streamflows in the western United States
since 1950 are detectably different from natural variability (Barnett et
al., 2008; Hidalgo et al., 2009). Thus, observed warming over several
decades that is attributable to anthropogenic forcing has likely been
linked to earlier spring peak flows in snowmelt- and glacier-fed rivers. It
is unclear if observed warming over several decades has affected the
magnitude of the snowmelt peak flows, but warming may result either
in an increase in spring peak flows where winter snow depth increases
(Meehl et al., 2007b) or a decrease in spring peak flows because of
decreased snow cover and amounts (Hirabayashi et al., 2008b; Dankers
and Feyen, 2009).
There is still a lack of studies identifying an influence of anthropogenic
climate change over the past several decades on rain-generated peak
streamflow trends because of availability and uncertainty in the
observed streamflow data and low signal-to-noise ratio. Evidence has
recently emerged that anthropogenic climate change could have
increased the risk of rainfall-dominated flood occurrence in some river
basins in the United Kingdom in autumn 2000 (Pall et al., 2011). Overall,
there is low confidence (due to limited evidence) that anthropogenic
climate change has affected the magnitude and frequency of floods,
though it has detectably influenced several components of the
hydrological cycle, such as precipitation and snowmelt, that may impact
flood trends. The assessment of causes behind the changes in floods is
inherently complex and difficult.
The number of studies that investigated projected flood changes in
rivers especially at a regional or a continental scale was limited when
the AR4 was published. Projections of flood changes at the catchment/
river-basin scale were also not abundantly cited in the AR4. Nevertheless,
Kundzewicz et al. (2007) and Bates et al. (2008) argued that more
frequent heavy precipitation events projected over most regions would
affect the risk of rain-generated floods (e.g., flash flooding and urban
flooding).
The number of regional- or continental-scale studies of projected
changes in floods is still limited. Recently, a few studies for Europe
(Lehner et al., 2006; Dankers and Feyen, 2008, 2009) and a study for the
globe (Hirabayashi et al., 2008b) have indicated changes in the frequency
and/or magnitude of floods in the 21st century at large scale using daily
river discharge calculated from RCM or GCM outputs and hydrological
models. A notable change is projected to occur in northeastern Europe
in the late 21st century because of a reduction in snow accumulation
(Dankers and Feyen, 2008, 2009; Hirabayashi et al., 2008b), that is, a
decrease in the probability of floods, that generally corresponds to
lower flood peaks. For other parts of the world, Hirabayashi et al.
(2008b) show an increase in the risk of floods in most humid Asian
monsoon regions, tropical Africa, and tropical South America with a
decrease in the risk of floods in non-negligible areas of the world such
as most parts of northern North America.
Projections of flood changes at the catchment/river-basin scale are also
not abundant in the scientific literature. Several studies have been
undertaken for UK catchments (Cameron, 2006; Kay et al., 2009;
Prudhomme and Davies, 2009) and catchments in continental Europe
and North America (Graham et al., 2007; Thodsen, 2007; Leander et al.,
2008; Raff et al., 2009; van Pelt et al., 2009). However, projections for
catchments in other regions such as Asia (Asokan and Dutta, 2008;
Dairaku et al., 2008), the Middle East (Fujihara et al., 2008), South
America (Nakaegawa and Vergara, 2010; Kitoh et al., 2011), and Africa
(Taye et al., 2011) are rare.
Uncertainty is still large in the projected changes in the magnitude and
frequency of floods. It has been recently recognized that the choice of
GCMs is the largest source of uncertainties in hydrological projections
at the catchment/river-basin scale, and that uncertainties from emission
scenarios and downscaling methods are also relevant but less important
(Graham et al., 2007; Leander et al., 2008; Kay et al., 2009; Prudhomme
and Davies, 2009), although, in general, hydrological projections require
downscaling and/or bias-correction of GCM outputs (e.g., precipitation
and temperature). Also the choice of hydrological models was found to
be relevant but less important (Kay et al., 2009; Taye et al., 2011).
However, the relative importance of downscaling, bias-correction, and
the choice of hydrological models may depend on the selected region/
catchment, the selected downscaling and bias-correction methods, and
the selected hydrological models (Wilby et al., 2008). For example, the
sign of the above-mentioned flood changes in northeastern Europe is
affected by differences in temporal downscaling and bias-correction
methods applied in the different studies (Dankers and Feyen, 2009).
Chen et al. (2011) demonstrated considerable uncertainty caused by
several downscaling methods in a hydrological projection for a
snowmelt-dominated Canadian catchment. Downscaling (see Section
3.2.3) and bias-correction are also a major source of uncertainty in rain-
dominated catchments (van Pelt et al., 2009). We also note that bias-
correction and statistical downscaling tend to ignore the energy closure
of the climate system, which could be a non-negligible source of
uncertainty in hydrological projections (Milly and Dunne, 2011).
The number of projections of flood magnitude and frequency changes is
still limited at regional and continental scales. Projections at the
catchment/river-basin scale are also not abundant in the peer-reviewed
scientific literature, especially for regions outside Europe and North
America. In addition, considerable uncertainty remains in the projections
of flood changes, especially regarding their magnitude and frequency.
Therefore, our assessment is that there is low confidence (due to limited
evidence) in future changes in flood magnitude and frequency derived
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
178
from river discharge simulations. Nevertheless, as was argued by
Kundzewicz et al. (2007) and Bates et al. (2008), physical reasoning
suggests that projected increases in heavy rainfall in some catchments
or regions would contribute to increases in rain-generated local floods
(medium confidence). We note that heavy precipitation may be projected
to increase despite a projected decrease of total precipitation depending
on the regions considered (Section 3.3.2), and that changes in several
variables (e.g., precipitation totals, frequency, and intensity, snow cover
and snowmelt, soil moisture) are relevant for changes in floods.
Confidence in change in one of these components alone may thus not
be sufficient to confidently project changes in flood occurrence. Hence,
medium confidence is attached to the above statement based on
physical reasoning, although the link between increases in heavy
rainfall and increases in local flooding seems apparent. The earlier shifts
of spring peak flows in snowmelt- and glacier-fed rivers are robustly
projected (Kundzewicz et al., 2007; Bates et al., 2008); so these are
assessed as very likely, though this may not necessarily be relevant for
flood occurrence. There is low confidence (due to limited evidence) in
the projected magnitude of the earlier peak flows in snowmelt- and
glacier-fed rivers.
In summary, there is limited to medium evidence available to
assess climate-driven observed changes in the magnitude and
frequency of floods at a regional scale because the available
instrumental records of floods at gauge stations are limited in
space and time, and because of confounding effects of changes
in land use and engineering. Furthermore, there is low agreement
in this evidence, and thus overall low confidence at the global
scale regarding even the sign of these changes. There is low
confidence (due to limited evidence) that anthropogenic climate
change has affected the magnitude or frequency of floods,
though it has detectably influenced several components of the
hydrological cycle such as precipitation and snowmelt (medium
confidence to high confidence), which may impact flood trends.
Projected precipitation and temperature changes imply possible
changes in floods, although overall there is low confidence in
projections of changes in fluvial floods. Confidence is low due to
limited evidence and because the causes of regional changes are
complex, although there are exceptions to this statement. There is
medium confidence (based on physical reasoning) that projected
increases in heavy rainfall (Section 3.3.2) would contribute to
increases in rain-generated local flooding, in some catchments or
regions. Earlier spring peak flows in snowmelt- and glacier-fed
rivers are very likely, but there is low confidence in their projected
magnitude.
3.5.3. Extreme Sea Levels
Transient sea level extremes and extreme coastal high water are caused
by severe weather events or tectonic disturbances that cause tsunamis.
Since tsunamis are not climate-related, they are not addressed here. The
drop in atmospheric pressure and strong winds that accompany severe
weather events such as tropical or extratropical cyclones (Sections 3.4.4
and 3.4.5) can produce storm surges at the coast, which may be further
elevated by wave setup caused by an onshore flux of momentum due to
wave breaking in the surf zone. Various metrics are used to characterize
extreme sea levels including storm-related highest values, annual
maxima, or percentiles. Extreme sea levels may change in the future as
a result of both changes in atmospheric storminess and mean sea level
rise. However, neither contribution will be spatially uniform across the
globe. For severe storm events such as tropical and extratropical
cyclones, changes may occur in the frequency, intensity, or genesis
regions of severe storms and such changes may vary between ocean
basins (see Sections 3.4.4 and 3.4.5). Along some coastlines, land
subsidence due to glacial isostatic adjustment (e.g., Lambeck et al.,
2010) is causing a relative fall in sea levels. Variations in the rate of sea
level rise can be large relative to mean sea level (Yin et al., 2010) and
will occur as a result of variations in wind change (e.g., Timmermann et
al., 2010), changes in atmospheric pressure and oceanic circulation
(e.g., Tsimplis et al., 2008), and associated differences in water density
and rates of thermal expansion (e.g., Bindoff et al., 2007; Church et al.,
2010; Yin et al., 2010). In addition, if rapid melting of ice sheets occurs
it would lead to non-uniform rates of sea level rise across the globe due
to adjustments in the Earth’s gravitational field (e.g., Mitrovica et al.,
2010). On some coastlines, higher mean sea levels may alter the
astronomical tidal range and the evolution of storm surges, and
increase the wave height in the surf zones. As well as gradual increases
in mean sea level that contribute to extreme impacts from transient
extreme sea levels, rapid changes in sea level arising from, for example,
collapse of ice shelves could be considered to be an extreme event with
the potential to contribute to extreme impacts in the future. However,
knowledge about the likelihood of such changes occurring is limited
and so does not allow an assessment at this time.
Mean sea level has varied considerably over glacial time scales as the
extent of ice caps and glaciers have fluctuated with global temperatures.
Sea levels have risen around 120 to 130 m since the last glacial maximum
19 to 23 ka before present to around 7,000 years ago, and reached a
level close to present at least 6,000 years ago (Lambeck et al., 2010). As
well as the influence on sea level extremes caused by rapidly changing
coastal bathymetries (Clarke and Rendell, 2009) and large-scale circulation
patterns (Wanner et al., 2008), there is some evidence that changes in
the behavior of severe tropical cyclones has changed on centennial time
scales, which points to non-stationarity in extreme sea level events
(Nott et al., 2009). Woodworth et al. (2011) use tide gauge records dating
back to the 18th century, and salt marsh data, to show that sea level
rise has accelerated over this time frame.
The AR4 reported that there was high confidence that the rate of observed
sea level rise increased from the 19th to the 20th century (Bindoff et al.,
2007). It also reported that the global mean sea level rose at an average
rate of 1.7 (1.2 to 2.2) mm yr
-1
over the 20th century, 1.8 (1.3 to 2.3)
mm yr
-1
over 1961 to 2003, and at a rate of 3.1 (2.4 to 3.8) mm yr
-1
over 1993 to 2003. With updated satellite data to 2010, Church and
White (2011) show that satellite-measured sea levels continue to rise at
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
179
a rate close to that of the upper range of the AR4 projections. Whether
the faster rate of increase during the latter period reflects decadal
variability or an increase in the longer-term trend is not clear. However,
there is evidence that the contribution to sea level due to mass loss
from Greenland and Antarctica is accelerating (Velicogna, 2009; Rignot
et al., 2011; Sørensen et al., 2011). The AR4 also reported that the rise
in mean sea level and variations in regional climate led to a likely
increase in the trend of extreme high water worldwide in the late 20th
century (Bindoff et al., 2007), it was very likely that humans contributed
to sea level rise during the latter half of the 20th century (Hegerl et al.,
2007), and therefore that it was more likely than not that humans
contributed to the trend in extreme high sea levels (IPCC, 2007a). Since
the AR4, Menendez and Woodworth (2010), using data from 258 tide
gauges across the globe, have confirmed the earlier conclusions of
Woodworth and Blackman (2004) that there was an increasing trend in
extreme sea levels globally, more pronounced since the 1970s, and that
this trend was consistent with trends in mean sea level (see also Lowe
et al., 2010). Additional studies at particular locations support this finding
(e.g., Marcos et al., 2009; Haigh et al., 2010).
Various studies also highlight the additional influence of climate
variability on extreme sea level trends. Menendez and Woodworth
(2010) report that ENSO (see Section 3.4.2) has a large influence on
interannual variations in extreme sea levels in the Pacific Ocean and
the monsoon regions based on sea level records since the 1970s. In
southern Europe, Marcos et al. (2009) report that changes in extremes
are also significantly negatively correlated with the NAO (see Section
3.4.3). Ullmann et al. (2007) concluded that maximum annual sea levels
in the Camargue had risen twice as fast as mean sea level during the
20th century due to an increase in southerly winds associated with a
general rise in sea level pressure over central Europe (Ullmann et al.,
2008). Sea level trends from two tide gauges on the north coast of
British Columbia from 1939 to 2003 were twice that of mean sea level
rise, the additional contribution being due to the strong positive phase
of the PDO (see Section 3.4.3), which has lasted since the mid-1970s
(Abeysirigunawardena and Walker, 2008). Cayan et al. (2008) reported
an increase of 20-fold at San Francisco since 1915 and 30-fold at La Jolla
since 1933 in the frequency of exceedance of the 99.99th percentile sea
level. They also noted that positive sea level anomalies of 10 to 20 cm
that often persisted for several months during El Niño events produced
an increase in storm surge peaks over this time. The spatial extent of
these oscillations and their influence on extreme sea levels across the
Pacific has been discussed by Merrifield et al. (2007). Church et al.
(2006a) examined changes in extreme sea levels before and after 1950
in two tide gauge records of approximately 100 years on the east and
west coasts of Australia, respectively. At both locations a stronger
positive trend was found in the sea level exceeded by 0.01% of the
observations than the median sea level, suggesting that in addition to
mean sea level rise, other modes of variability or climate change are
contributing to the extremes. At Mar del Plata, Argentina, Fiore et al.
(2009) noted an increase in the number and duration of positive storm
surges in the decade 1996 to 2005 compared to previous decades,
which may be due to a combination of mean sea level rise and changes
in wind climatology resulting from a southward shift in the South
Atlantic high.
Thus, studies since the AR4 conclude that trends in extreme sea level are
generally consistent with changes in mean sea level (e.g., Marcos et al.,
2009; Haigh et al., 2010; Menendez and Woodworth, 2010) although
some studies note that the trends in extremes are larger than the
observed trend in mean sea levels (e.g., Church et al., 2006a; Ullmann et
al., 2007; Abeysirigunawardena and Walker, 2008) and may be influenced
by modes of climate variability, such as the PDO on the Canadian west
coast (e.g., Abeysirigunawardena and Walker, 2008). These studies are
consistent with the conclusions from the AR4 that increases in extremes
are related to trends in mean sea level and modes of variability in the
regional climate.
The AR4 (Meehl et al., 2007b) projected sea level rise for 2090-2099
relative to 1980-1999 due to ocean thermal expansion, glaciers and
ice caps, and modeled ice sheet contributions of 18 to 59 cm, which
incorporates a 90% uncertainty range across all scenarios. An additional
contribution to the sea level rise projections was taken into account for
a possible rapid dynamic response of the Greenland and West Antarctic
ice sheets, which could result in an accelerating contribution to sea level
rise. This was estimated to be 10 to 20 cm of sea level rise by 2090-2099
using a simple linear relationship with projected temperature. Because
of insufficient understanding of the dynamic response of ice sheets,
Meehl et al. (2007b) also noted that a larger contribution could not be
ruled out.
Several studies since the AR4 have developed statistical models that
relate 20th-century (e.g., Rahmstorf, 2007; Horton et al., 2008) or longer
(e.g.,Vermeer and Rahmstorf, 2009; Grinsted et al., 2010) temperature
and sea level rise to extrapolate future global mean sea level. These
alternative approaches yield projections of sea level rise under a range
of SRES scenarios by 2100 of 0.47 to 1.00 m (B1 to A2 scenarios; Horton
et al., 2008), 0.50 to 1.40 m (B1 to A1FI scenarios; Rahmstorf, 2007),
0.75 to 1.90 m (B1 to A1FI scenarios; Vermeer and Rahmstorf, 2009),
and 0.90 to 1.30 m (A1B scenario only; Grinsted et al., 2010). However,
future rates of sea level rise may be less closely associated with global
mean temperature if ice sheet dynamics play a larger role in the future
(Cazenave and Llovel, 2010). Furthermore, Church et al. (2011) note that
these models may overestimate future sea levels because non-climate
related contributions to trends over the observational period such as
groundwater depletion may not have been removed, and non-linear
effects such as the reduction in glacier area as glaciers contract and the
reduction in the efficiency of ocean heat uptake with global warming in
the future are not accounted for. Pfeffer et al. (2008), using a dynamical
model of glaciers, found that sea level rise of more than 2 m by 2100 is
physically implausible. An estimate of 0.8 m by 2100 that included
increased ice dynamics was considered most plausible.
New studies, whose focus is on quantifying the effect of storminess
changes on storm surge, have been carried out over northern Europe
since the AR4. Debernard and Roed (2008) used hydrodynamic models
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
180
to investigate storm surge changes over Europe in four regionally
downscaled GCMs including two runs with B2, one with A2, and one
with an A1B emission scenario. Despite large inter-model differences,
statistically significant changes between 1961-1990 and 2071-2100
consisted of decreases in the 99th percentile surge heights south of
Iceland, and an 8 to 10% increase along the coastlines of the eastern
North Sea and the northwest British Isles, which occurred mainly in the
winter season. Wang et al. (2008) projected a significant increase in
wintertime storm surges around Ireland except the south Irish coast
over 2031-2060 relative to 1961-1990 using a downscaled GCM under
an A1B scenario. Sterl et al. (2009) joined the output from an ensemble
of 17 GCM (CMIP3) simulations using the A1B emissions scenario over
the model periods 1950-2000 and 2050-2100 into a single longer time
series to estimate 10,000-year return values of surge heights along the
Dutch coastline. No statistically significant change in this value was
projected for the 21st century because projected wind speed changes
were not associated with the surge-generating northerlies but rather
non-surge generating south-westerlies.
Other studies have undertaken a sensitivity approach to compare the
relative impact on extreme sea levels of severe weather changes and
mean sea level rise. Over southeastern Australia, McInnes et al. (2009b)
found that a 10% increase in wind speeds, consistent with the upper
end of the range under an A1FI scenario from a multi-model ensemble
for 2070 together with an A1FI sea level rise scenario, would produce
extreme sea levels that were 12 to 15% higher than those including
just the A1FI sea level rise projection alone. Brown et al. (2010) also
investigated the relative impact of sea level rise and wind speed change
on an extreme storm surge in the eastern Irish Sea. Both studies
concluded that sea level rise rather than meteorological changes has
the greater potential to increase extreme sea levels in these locations in
the future.
The degree to which climate models (GCM or RCM) have sufficient
resolution and/or internal physics to realistically capture the meteorological
forcing responsible for storm surges is regionally dependent. For example
current GCMs are unable to realistically represent tropical cyclones (see
Section 3.4.4). This has led to the use of alternative approaches for
investigating the impact of climate change on storm surges in tropical
locations whereby large numbers of cyclones are generated using
statistical models that govern the cyclones’ characteristics over the
observed period (e.g., McInnes et al., 2003). These models are then
perturbed to represent projected future cyclone characteristics and used
to force a hydrodynamic model. Recent studies on the tropical east
coast of Australia reported in Harper et al. (2009) that employ these
approaches show a relatively small impact of a 10% increase in tropical
cyclone intensity on the 1-in-100 year storm tide (the combined sea level
due to the storm surge and tide), and mean sea level rise being found
to produce the larger contribution to changes in future 1-in-100 year
sea level extremes. However, one study that has incorporated scenarios
of sea level rise in the hydrodynamic modeling of hurricane-induced sea
level extremes on the Louisiana coast found that increased coastal
water depths had a large impact on surge propagation over land,
increasing storm surge heights by two to three times the sea level rise
scenario, particularly in wetland-fronted areas (J.M. Smith et al., 2010).
To summarize, post-AR4 studies provide additional evidence that
trends in extreme coastal high water across the globe reflect the
increases in mean sea level, suggesting that mean sea level rise
rather than changes in storminess are largely contributing to this
increase (although data are sparse in many regions and this lowers
the confidence in this assessment). It is therefore considered likely
that sea level rise has led to a change in extreme coastal high
water levels. It is likely that there has been an anthropogenic
influence on increasing extreme coastal high water levels via
mean sea level contributions. While changes in storminess
may contribute to changes in sea level extremes, the limited
geographical coverage of studies to date and the uncertainties
associated with storminess changes overall (Sections 3.4.4 and
3.4.5) mean that a general assessment of the effects of storminess
changes on storm surge is not possible at this time. On the basis
of studies of observed trends in extreme coastal high water
levels it is very likely that mean sea level rise will contribute to
upward trends in the future.
3.5.4. Waves
Severe waves threaten the safety of coastal inhabitants and those
involved in maritime activities and can damage and destroy coastal
and marine infrastructure. Waves play a significant role in shaping a
coastline by transporting energy from remote areas of the ocean to the
coast. Energy dissipation via wave breaking contributes to beach erosion,
longshore currents, and elevated coastal sea levels through wave set-up
and wave run-up. Wave properties that influence these processes
include wave height, the wave energy directional spectrum, and period.
Studies of past and future changes in wave climate to date have tended
to focus on wave height parameters such as ‘Significant Wave Height’
(SWH, the average height from trough to crest of the highest one-third
of waves) and metrics of extreme waves, such as high percentiles or
wave heights above particular thresholds, although one study (Dodet et
al., 2010) also examines trends in mean wave direction and peak wave
period. It should also be noted that waves may become an increasingly
important factor along coastlines experiencing a decline in coastal
protection afforded by sea ice (see Sections 3.5.5 and 3.5.7).
Wave climates have changed over paleoclimatic time scales. Wave
modeling using paleobathymetries over the past 12,000 years indicates
an increase in peak annual SWH of around 40% due to the increase in
mean sea level, which redefines the location of the coastline, and hence
progressively extends the fetch length in most of the shelf sea regions
(Neill et al., 2009). Major circulation changes that result in changes in
storminess and wind climate (see Section 3.3.3) have also affected
wave climates. Evidence of enhanced storminess determined from sand
drift and dune building along the western European coast indicates that
enhanced storminess occurred over the period of the Little Ice Age
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
181
(1570-1900) and the mid Holocene (~8,200 years before present; Clarke
and Rendell, 2009).
The AR4 reported statistically significant positive trends in SWH over the
period 1950 to 2002 over most of the mid-latitudinal North Atlantic and
North Pacific, as well as in the western subtropical South Atlantic, the
eastern equatorial Indian Ocean and the East China and South China
Sea, and declining trends around Australia, and parts of the Philippine,
Coral, and Tasman Seas (Trenberth et al., 2007), based on voluntary
observing ship data (e.g., Gulev and Grigorieva, 2004). Several studies
that address trends in extreme wave conditions have been completed
since the AR4 and the new studies generally provide more evidence for
the previously reported positive trends in SWH and extreme waves in
the north Atlantic and north Pacific. Global trends in 99th-percentile
satellite-measured wave heights show a mostly significant positive
trend of between 0.5 and 1.0% per year in the mid-latitude oceans but
less clear trends over the tropical oceans from 1985 to 2008 (Young et
al., 2011). X.L. Wang et al. (2009b) found that SWH increased in the
boreal winter over the past half century in the high latitudes of the
Northern Hemisphere (especially the northeast Atlantic), and decreased
in more southerly northern latitudes based on the European Centre for
Medium Range Weather Forecasts 40-year reanalysis (ERA-40). They
also found that storminess around the 1880s was of similar magnitude
to that in the 1990s. This is also found using the same data set by
Le Cozannet et al. (2011), who relate the change in waves to the NAO
pattern that is moderated by an east Atlantic pattern of climate variability
during winter. A wave hindcast over the north-eastern Atlantic Ocean
over the period 1953 to 2009 revealed a significant positive trend in
SWH, as well as a counterclockwise shift in mean direction in the north
and a slight but not significant increase in peak wave period in the
northeast. In the south, no trend was found for SWH or wave period
while a clockwise trend in mean direction was found (Dodet et al.,
2010). In a regional North Sea hindcast, Weisse and Günther (2007)
found a positive trend in 99th-percentile wave height from 1958 to the
early 1990s followed by a declining trend to 2002 over the southern
North Sea, except on the UK North Sea coast where negative trends
occurred over much of the hindcast period.
On the North American Atlantic coast, Komar and Allan (2008) found a
statistically significant trend of 0.059 m yr
-1
in waves exceeding 3 m
during the summer months over 30 years since the mid-1970s at
Charleston, South Carolina, with weaker but statistically significant
trends at wave buoys further north. These trends were associated with
an increase in intensity and frequency of hurricanes over this period (see
Section 3.4.4). In contrast, winter waves, generated by extratropical
storms, were not found to have experienced a statistically significant
change. In the eastern North Pacific, SWH is strongly correlated with
El Niño (Section 3.4.2). However positive trends were also found in SWH
and extreme wave height from the mid-1970s to 2006 in wave buoy
data (Allan and Komar, 2006), for excesses of the 98th percentile SWH
over 1985 to 2007 (Menendez et al., 2008) along the US west coast, and
in hindcast SWH over 1948 to 1998 in the Southern Californian Bight
(Adams et al., 2008). Positive though not statistically significant trends
in annual mean SWH were found over south-eastern South America for
in situ wave data over the 1996-2006 period and in satellite wave data
over 1993 to 2001, while simulated wave fields using reanalysis wind
forcing over the period 1971 to 2005 produced statistically significant
trends in SWH (Dragani et al., 2010). Trends at particular locations may
be also influenced by local factors. For example, Suursaar and Kullas
(2009) reported a slight decreasing trend in mean SWHs from 1966 to
2006 in the Gulf of Riga within the Baltic Sea, while the frequency and
intensity of high wave events (i.e., the difference between the maximum
and 99th-percentile wave height) showed rising trends. These changes
were associated with a decrease in local average wind speed, but an
intensification of westerly winds and storm events occurring further to
the west.
In the Southern Ocean, SWH derived from satellite observations was
found to be strongly positively correlated with the SAM, particularly from
March to August (Hemer et al., 2010). However, the analysis of reliable
long-term trends in the Southern Hemisphere remains challenging due
to limited in situ data and problems of temporal homogeneity in
reanalysis products (Wang et al., 2006). For example, Hemer et al. (2010)
also found that trends in SWH derived from satellite data over 1998-2000
relative to 1993-1996 were positive only over the Southern Ocean south
of 45°S whereas trends were positive across most of the Southern
Hemisphere in the Corrected ERA-40 reanalysis (C-ERA-40; Hemer, 2010).
Hemer (2010) found that the frequency of wave events exceeding the
98th percentile over the period 1985 to 2002 using data from a wave buoy
situated on the west coast of Tasmania showed no statistically significant
trend whereas a strong positive trend was found in equivalent fields of
C-ERA-40 data.
New studies have demonstrated strong links between wave climate and
natural modes of climate variability (Section 3.4.3). For example, along
the US west coast and the western North Pacific, SWH was found to be
strongly correlated with El Niño (Allan and Komar, 2006; Sasaki and
Toshiyuki, 2007) and, in the Southern Ocean, SWH was positivity
correlated with the SAM (Hemer et al., 2010). On the US east coast,
positive trends in summer SWH were linked to increasing numbers of
hurricanes (Komar and Allan, 2008). In the northeast Atlantic, trends in
SWH exhibited significant positive (negative) correlations with the NAO
in the north (south) and more generally, trends in SWH, mean wave
direction, and peak wave period over the period 1953 to 2009 were
related to the increase in the NAO index over this time (Dodet et al.,
2010). One study (X.L. Wang et al., 2009a) reported a link between
external forcing (i.e., anthropogenic forcing due to greenhouse gases
and aerosols, and natural forcing due to solar and volcanic forcing) and
an increase in SWH in the boreal winter in the high latitudes of the
Northern Hemisphere (especially the northeast North Atlantic), and a
decrease in more southerly northern latitudes over the past half century.
The AR4 projected an increase in extreme wave height in many regions
of the mid-latitude oceans as a result of projected increases in wind speeds
associated with more intense mid-latitude storms in these regions in a
future warmer climate (Meehl et al., 2007b). At the regional scale,
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
182
increases in wave height were projected for most mid-latitude areas
analyzed, including the North Atlantic, North Pacific, and Southern Ocean
(Christensen et al., 2007) but with low confidence due to low confidence
in projected changes in mid-latitude storm tracks and intensities (see
Section 3.4.5). Several studies since then have developed wave climate
projections that provide stronger evidence for future wave climate
change. Global-scale projections of SWH were developed by Mori et al.
(2010), using a 1.25° resolution wave model forced with projected winds
from a 20-km global GCM, in which ensemble-averaged SST changes
from the CMIP3 models provided the climate forcing. The spatial pattern
of projected SWH change between 1979-2004 and 2075-2100 reflects
the changes in the forcing winds, which are generally similar to the mean
wind speed changes shown in Figure 3-8. Extreme waves (measured by
a spatial and temporal average of the top 10 values over the 25-year
period) were projected to exhibit large increases in the northern Pacific,
particularly close to Japan due to an increase in strong tropical cyclones
and also the Indian Ocean despite decreases in SWH.
A number of regional studies have also been completed since the AR4
in which forcing conditions were obtained for a few selected emission
scenarios (typically B2 and A2, representing low-high ranges) from GCMs
or RCMs. These studies provide additional evidence for positive projected
trends in SWH and extreme waves along the western European coast
(e.g., Debernard and Roed, 2008; Grabemann and Weisse, 2008) and the
UK coast (Leake et al., 2007), declines in extreme wave height in the
Mediterranean sea (Lionello et al., 2008) and the southeast coast of
Australia (Hemer et al., 2010), and little change along the Portuguese
coast (Andrade et al., 2007). However, considerable variation in projections
can arise from the different climate models and scenarios used to force
wave models, which lowers the confidence in the projections. For example,
along the European North Sea coast, 99th-percentile wave height over
the late 21st century relative to the late 20th century is projected to
increase by 6 to 8% by Debernard and Roed (2008) based on wave
model simulations with forcing from several GCMs under A2, B2, and
A1B greenhouse gas scenarios, whereas they are projected to increase
by up to 18% by Grabemann and Weisse (2008), who downscaled two
GCMs under A2 and B2 emission scenarios. In one region, opposite
trends in extreme waves were projected. Grabemann and Weisse (2008)
project negative trends in 99th-percentile wave height along the UK
North Sea coast, whereas Leake et al. (2007) downscaled the same
GCM for the same emission scenarios, using a different RCM, and found
positive changes in high percentile wave heights offshore of the East
Anglia coastline. A wave projection study by Hemer et al. (2010)
concluded that uncertainties arising from the method by which climate
model winds were applied to wave model simulations (e.g., by applying
bias-correction to winds or perturbing current climate winds with wind
changes derived from climate models) made a larger contribution to the
spread of RCM projections than the forcing from different GCMs or
emission scenarios.
In summary, although post-AR4 studies are few and their regional
coverage is limited, their findings generally support the evidence
from earlier studies of wave climate trends. Most studies find a
link between variations in waves (both SWH and extremes) and
internal climate variability. There is low confidence that there
has been an anthropogenic influence on extreme wave heights
(because of insufficient literature). Despite the existence of
downscaling studies for some regions such as the eastern North
Sea, there is overall low confidence in wave height projections
because of the small number of studies, the lack of consistency
of the wind projections between models, and limitations in their
ability to simulate extreme winds. However, the strong linkages
between wave height and winds and storminess means that it is
likely that future negative or positive changes in SWH will reflect
future changes in these parameters.
3.5.5. Coastal Impacts
Severe coastal hazards such as erosion and inundation are important in
the context of disaster risk management and may be affected by climate
change through rising sea levels and changes in extreme events.
Increasing sea levels will also increase the potential for saltwater intrusion
into coastal aquifers. Coastal inundation occurs during periods of extreme
sea levels due to storm surges and high waves, particularly when
combined with high tides. Although tropical and extratropical cyclones
(Sections 3.4.4 and 3.4.5) are the most common causes of sea level
extremes, other weather events that cause persistent winds such as
anticyclones and fronts can also influence coastal sea levels (Green et
al., 2009; McInnes et al., 2009b). In many parts of the world, sea levels
are influenced by modes of large scale variability such as ENSO (Section
3.4.2). In the western equatorial Pacific, sea levels can fluctuate up to half
a meter between ENSO phases (Church et al., 2006b) and in combination
with extremes of the tidal cycle, can cause extensive inundation in low-
lying atoll nations even in the absence of extreme weather events
(Lowe et al., 2010).
Shoreline position can change from the combined effects of various
factors such as:
1) Rising mean sea levels, which cause landward recession of coastlines
made up of erodible materials (e.g., Ranasinghe and Stive, 2009)
2) Changes in coastal height due to isostatic rebound (Blewitt et al.,
2010; Mitrovica et al., 2010), or sediment compaction from the
removal of oil, gas, and water (Syvitski et al., 2009)
3) Changes in the frequency or severity of transient storm erosion
events (K.Q. Zhang et al., 2004)
4) Changes in sediment supply to the coast (Stive et al., 2003;
Nicholls et al., 2007; Tamura et al., 2010)
5) Changes in wave speed due to sea level rise, which alters wave
refraction, or in wave direction, which can cause realignment of
shorelines (Ranasinghe et al., 2004; Bryan et al., 2008; Tamura et
al., 2010)
6) The loss of natural protective structures such as coral reefs (e.g.,
Sheppard et al., 2005; Gravelle and Mimura, 2008) due to
increased ocean temperatures (Hoegh-Guldberg, 1999) and ocean
acidification (Bongaerts et al., 2010) or the reduction in permafrost
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
183
or sea ice in mid- and high latitudes, which exposes soft shores to
the effects of waves and severe storms (see Section 3.5.7; Manson
and Solomon, 2007).
For example, permafrost degradation and sea ice retreat may contribute
to coastal erosion in Arctic regions (see Section 3.5.7).
The susceptibility of coastal regions to erosion and inundation is related
to various physical (e.g., shoreline slope), and geomorphological and
ecosystem attributes, and therefore may be inferred to some extent
from broad coastal characterizations. These include the presence of
beaches, rocky shorelines, or coasts with cliffs; deltas; back-barrier
environments such as estuaries and lagoons; the presence of mangroves,
salt marshes, or sea grasses; and shorelines flanked by coral reefs (e.g.,
Nicholls et al., 2007) or by permafrost or seasonal sea ice, each of which
are characterized by different vulnerability to climate change-driven
hazards. For example, deltas are low-lying and hence generally prone to
inundation, while beaches are comprised of loose particles and therefore
erodible. However, the degree to which these systems are impacted by
erosion and inundation will also be influenced by other factors affecting
disaster responses. For example, reduced protection from high waves
during severe storms could occur as a result of depleted mangrove
forests or the degradation of coral reefs (e.g., Gravelle and Mimura,
2008), or loss of sea ice or permafrost (e.g., Manson and Solomon, 2007);
there may be a loss of ecosystem services brought about by saltwater
contamination of already limited freshwater reserves due to rising sea
levels and these will amplify the risks brought about by climate change
(McGranahan et al., 2007), and also reduce the resilience of coastal
settlements to disasters. Dynamical processes such as vertical land
movement also contribute to inundation potential (Haigh et al., 2009).
Coastal regions may be rising or falling due to post-glacial rebound or
slumping due to aquifer drawdown (Syvitski et al., 2009). Multiple
contributions to coastal flooding such as heavy rainfall and flooding in
coastal catchments that coincide with elevated sea levels may also be
important. Ecosystems such as coral reefs also play an important role in
providing material on which atolls are formed. Large-scale oceanic
changes that are particularly relevant to both coral reefs and small
island countries are discussed in Box 3-4.
As discussed in Section 3.5.3, mean sea level has risen by 120 to 130 m
since the end of the last glacial maximum (Jansen et al., 2007), and this
has had a profound effect on coastline position around the world.
Coastlines have also evolved over this time frame due to changes in the
action of the ocean on the coast through changes in wave climate
(Neill et al., 2009) and tides (Gehrels et al., 1995), which arise from the
changing geometries of coastlines over glacial time scales and changes
in storminess (e.g., Clarke and Rendell, 2009).
The AR4 (Nicholls et al., 2007) reported that coasts are experiencing the
adverse consequences of impacts such as increased coastal inundation,
erosion, and ecosystem losses. However, attributing these changes to
sea level rise is difficult due to the multiple drivers of change over the
20th century (R.J. Nicholls, 2010) and the scarcity and fragmentary
nature of data sets that contribute to the problem of identifying and
attributing changes (e.g., Defeo et al., 2009). Since the AR4 there have
been several new studies that examine coastline changes. In the
Caribbean, the beach profiles at 200 sites across 113 beaches and eight
islands were monitored on a three-monthly basis from 1985 to 2000,
with most beaches found to be eroding and faster rates of erosion
generally found on islands that had been impacted by a higher number
of hurricanes (Cambers, 2009). However, the relative importance of
anthropogenic factors, climate variability, and climate change on the
eroding trends could not be separated quantitatively. In Australia,
Church et al. (2008) report that despite the positive trend in sea levels
during the 20th century, beaches have generally been free of chronic
coastal erosion, and where it has been observed it has not been possible
to unambiguously attribute it to sea level rise in the presence of other
anthropogenic activities. Webb and Kench (2010) argue that the
commonly held view of atoll nations being vulnerable to erosion must
be reconsidered in the context of physical adjustments to the entire
island shoreline, because erosion of some sectors may be balanced by
progradation on other sectors. In their survey of 27 atoll islands across
three central Pacific Nations (Tuvalu, Kiribati, and Federated States of
Micronesia) over a 19- to 61-year period using photography and
satellite imagery, they found that 43% of islands remained stable and
43% increased in area, with largest rates of increase in island area
ranging from 0.1 to 5.6 ha per decade. Only 14% of islands studied
exhibited a net reduction in area. On islands exhibiting either no net
change or an increase in area, a larger redistribution of land area was
evident in 65% of cases, consisting of mainly a shoreline recession on
the ocean side and an elongation of the island or progradation of the
shoreline on the lagoon side. Human settlements were present on 7 of
the 27 atolls surveyed and the majority of those exhibited net accretion
due in part to coastal protection works. For a coral reef island at the
northern end of the Great Barrier Reef, Australia, Dawson and Smithers
(2010) report a 6% increase in area and 4% increase in volume between
1967 and 2007 but with a net retreat on the east-southeast shoreline
and advance on the western side. Chust et al. (2009) evaluated the
relative contribution of local anthropogenic (non-climate change related)
and sea level rise impacts on the coastal morphology and habitats in
the Basque coast, northern Spain, for the period 1954 to 2004. They
found that the impact from local anthropogenic influences was about
an order of magnitude greater than that due to sea level rise over this
period. Increased rates of coastal erosion have also been observed since
1935 in Canada’s Gulf of St. Lawrence (Forbes et al., 2004).
The AR4 stated with very high confidence that the impact of climate
change on coasts is exacerbated by increased pressures on the physical
environment arising from human settlements in the coastal zone (Nicholls
et al., 2007). The small number of studies that have been completed
since the AR4 have been either unable to attribute coastline changes to
specific causes in a quantitative way or else find strong evidence for
non-climatic causes that are natural and/or anthropogenic.
The AR4 reported with very high confidence that coasts will be exposed
to increasing impacts, including coastal erosion, over coming decades
due to climate change and sea level rise, both of which will be
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
184
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Box 3-4 | Small Island States
Small island states represent a distinct category of locations owing to their small size and highly maritime climates, which means that
their concerns and information needs in relation to future climate change differ in many ways from those of the larger continental
regions that are addressed in this chapter. Their small land area and often low elevation makes them particularly vulnerable to rising sea
levels and impacts such as inundation, shoreline change, and saltwater intrusion into underground aquifers (Mimura, 1999). Their
maritime environments lead to an additional emphasis on oceanic information to understand the impacts of climate change (see Case
Study 9.2.9). Particular challenges exist for the assessment of past changes in climate given the sparse regional and temporal coverage
of terrestrial-based observation networks and the limited in situ ocean observing network, although observations have improved
somewhat in recent decades with the advent of satellite-based observations of meteorological and oceanic variables. However, the short
length of these records hampers the investigation of long-term trends in the region. The resolution of GCMs is insufficient to represent
small islands and few studies have been undertaken to provide projections for small islands using RCMs (Campbell et al., 2011). In
regions such as the Pacific Ocean, large-scale climate features such as the South Pacific Convergence Zone ENSO (Section 3.4.2) have a
substantial influence on the pattern and timing of precipitation, yet these features and processes are often poorly represented in GCMs
(Collins et al., 2010). The purpose of this box is to present available information on observed trends and climate change projections that
are not covered in the other sections of this chapter as well as discuss key aspects of the climate system that are particularly relevant for
small islands. The very likely contribution of mean sea level rise to increased extreme sea levels (see Section 3.5.3), coupled with the
likely increase in tropical cyclone maximum wind speed (see Section 3.4.4), is a specific issue for tropical small island states.
Although the underlying data sources are limited, some data for the Indian Ocean, South Pacific (Fiji), and Caribbean were available in
the studies of Alexander et al. (2006) and Caesar et al. (2011). Problems of data availability and homogeneity for the Caribbean are
discussed by T.S. Stephenson et al. (2008). Based on standard extremes indices, positive trends in warm days and warm nights and
negative trends in cold days and cold nights
2
have occurred across the Indian Ocean and South Pacific region for the period 1971 to
2005 (Caesar et al., 2011) and the Caribbean for the period 1951 to 2003 (based on data from Alexander et al., 2006). Based on the
same data sources, trends in average total wet-day precipitation were positive and statistically significant over the Indian Ocean region,
negative over the South Pacific region, and weakly negative over the Caribbean. Trends in heavy and very heavy precipitation were
positive over the Indian Ocean, negative over the South Pacific region, and close to zero over the Caribbean. We have low confidence in
temperature trends over the Indian Ocean and South Pacific region due to the shorter record over which trends were assessed,
whereas for the Caribbean, we have medium confidence in the temperature trends due to the longer records available for assessment.
Because of the spatial heterogeneity exhibited in precipitation trends in general, there is insufficient evidence to assess observed
rainfall trends. For the Caribbean, temperatures are projected to increase across the region by 1 to 4°C over 2071-2100 relative to
1961-1990 under the A2 and B2 scenarios and rainfall is mainly projected to decrease by 25 to 50% except in the north (Campbell et al.,
2011). Based on this study and the evidence for projected temperature increases reported for other regions (see Table 3-3) we have
medium confidence in the projected temperature increases for the Caribbean. However, due to the range of processes that contribute to
rainfall change, some of which are poorly resolved by GCMs, there is insufficient evidence to assess projected rainfall changes on these
small islands.
Given the low elevation of many small islands, sea level extremes are of particular relevance. Sea level extremes are strongly influenced
by tidal extremes (Chowdhury et al., 2007; Merrifield et al., 2007). When the tide behavior is mostly semi-diurnal (two high and low tides
per day), there will be a clustering of high spring tides around the time of the equinoxes whereas when the tide behavior is diurnal (one
high and low tide per day), the clustering of high spring tides will occur around the time of the solstices. In addition, ENSO has a strong
influence such that sea levels and their extremes are positively (negatively) correlated with the SOI in the tropical Pacific west (east) of
180° (Church et al., 2006b; Menendez et al., 2010). Tides and ENSO have contributed to the more frequent occurrence of sea level
extremes and associated flooding experienced at some Pacific Islands such as Tuvalu in recent years, and make the task of determining
the relative roles of these natural effects and mean sea level rise difficult (Lowe et al., 2010). Furthermore, the steep shelf margins that
surround many islands and atolls in the Pacific support larger wave-induced contributions to sea level anomalies. Unfortunately, wave
observations (including wave direction) that would facilitate more comprehensive studies of tide, surge, and wave extremes in the region
are sparse, including those that are co-located with tide gauges (Lowe et al., 2010).
____________
2
Termed “cool days” and “cool nights” in that study.
Continued next page
185
exacerbated by increasing human-induced pressures (Nicholls et al.,
2007). However it was also noted that since coasts are dynamic systems,
adaptation to climate change required understanding of processes
operating on decadal to century time scales, yet this understanding was
least developed.
Because of the diverse and complex nature of coastal impacts, assessments
of the future impacts of climate change have focused on a wide range
of questions and employed a diverse range of methods, making direct
comparison of studies difficult (R.J. Nicholls, 2010). Two types of studies
are examined here: the first are assessments, typically undertaken at the
country or regional scale and which combine information on physical
changes with the socioeconomic implications (e.g., Nicholls and de la
Vega-Leinert, 2008); the second type are studies oriented around improved
scientific understanding of the impacts of climate change. In terms of
coastal assessments, Aunan and Romstad (2008) reported that Norway’s
generally steep and resistant coastlines contribute to a low physical
susceptibility to accelerated sea level rise. Nicholls and de la Vega-Leinert
(2008) reported that large parts of the coasts in Great Britain (including
England, Wales, and Scotland) are already experiencing widespread
sediment starvation and erosion, loss/degradation of coastal ecosystems,
and significant exposure to coastal flooding. Lagoons, river deltas, and
estuaries are assessed as being particularly vulnerable in Poland (Pruszak
and Zawadzka, 2008). In Estonia, Kont et al. (2008) reported increased
beach erosion, which is believed to be the result of increased storminess
in the eastern Baltic Sea since 1954, combined with a decline in sea ice
cover during the winter. Sterr (2008) reported that for Germany there is
a high level of reliance on hard coastal protection against extreme sea
level hazards, which will increase ecological vulnerability over time. In
France, the Atlantic coast Aquitaine region was considered more
resilient to rising sea levels over the coming century because of the
sediment storage in the extensive dune systems whereas the sandy
coast regions of the Languedoc Roussillon region on the Mediterranean
coast were considered more vulnerable because of narrow dune
systems that are also highly urbanized (Vinchon et al., 2009). A coastal
vulnerability assessment for Australia (Department of Climate Change,
2009) characterized future vulnerability in terms of coastal geomorphology,
sediment type, and tide and wave characteristics, from which it concluded
that the tropical northern coastline would be most sensitive to changes
in tropical cyclone behavior while health of the coral reefs may also
influence the tropical eastern coastline. The mid-latitude southern and
eastern coastlines were expected to be most sensitive to changes in
mean sea level, wave climate, and changes in storminess. A comparative
study of the impact of sea level rise on coastal inundation across 84
developing countries showed that the greatest vulnerability to a 1 m
sea level rise in terms of inundation of land area was located in East
Asia and the Pacific, followed by South Asia, Latin America, and the
Caribbean, the Middle East and North Africa, and finally sub-Saharan
Africa (Dasgupta et al., 2009).
New models have been developed for the assessment of coastal
vulnerability at the global to national level (Hinkel and Klein, 2009). At
the local to regional scale, new techniques and approaches have also
been developed to better quantify impacts from inundation due to
future sea level rise. Bernier et al. (2007) evaluated spatial maps of
extreme sea level for different return periods on a seasonal basis that
were used to estimate seasonal risk of inundation under future sea level
scenarios. McInnes et al. (2009a) developed spatial maps of storm tide
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Coral reefs are a feature of many small islands and healthy reef systems mitigate against erosion and inundation by not only providing a
buffer zone for the shoreline during extreme surge and wave events but also providing a source of carbonate sand and gravel, which are
delivered to the shores by storms and swell to maintain the atoll (Woodroffe, 2008; Webb and Kench, 2010). Anthropogenic oceanic
changes may indirectly contribute to extreme impacts for coral atolls by affecting the health of the surrounding reef system. Such
changes include: (1) warming of the surface ocean, which slows or prevents growth in temperature-sensitive species and causes more
frequent coral bleaching events (e.g., Hoegh-Guldberg, 1999; see also Chapter 4); (2) ocean acidification, caused by increases in
atmospheric CO
2
being absorbed into the oceans, which lowers coral growth rates (Bongaerts et al., 2010); and (3) reduction in oxygen
concentration in the ocean due to a combination of changes in temperature-driven gas solubility (Whitney et al., 2007), ocean ventilation
due to circulation changes, and biological cycling of organic material (Keeling et al., 2010). Quantifying these changes and understanding
their impact on coral reef health will be important to understanding the impact of anthropogenic climate change on atolls.
In summary, the small land area and often low elevation of small island states make them particularly vulnerable to rising
sea levels and impacts such as inundation, shoreline change, and saltwater intrusion into underground aquifers. Short
record lengths and the inadequate resolution of current climate models to represent small island states limit the assessment
of changes in extremes. There is insufficient evidence to assess observed trends and future projections in rainfall across the
small island regions considered here. The reported increases in warm days and nights and decreases in cold days and
nights are of medium confidence over the Caribbean and of low confidence over the Pacific and Indian Oceans. There is
medium confidence in the projected temperature increases across the Caribbean. The unique situation of small islands
states and their maritime environments leads to an additional emphasis on oceanic information to understand the impacts
of climate change. The very likely contribution of mean sea level rise to increased coastal high water levels, coupled with
the likely increase in tropical cyclone maximum wind speed, is a specific issue for tropical small island states.
186
and using a simple inundation model with high-resolution Light Detection
and Ranging (LIDAR) data and a land subdivision database, identified
the impact of inundation on several coastal towns along the southeastern
Australian coastline under future sea level and wind speed scenarios.
Probabilistic approaches have also been used to evaluate extreme sea
level exceedance under uncertain future sea level rise scenarios. Purvis
et al. (2008) constructed a probability distribution around the range of
future sea level rise estimates and used Monte Carlo sampling to apply
the sea level change to a two-dimensional coastal inundation model.
They showed that by evaluating the possible flood-related losses in
this framework they were able to represent spatially the higher losses
associated with the low-frequency but high-impact inundation events
instead of considering only a single midrange scenario. Hunter (2010)
combined sea level extremes evaluated from observations with projections
of sea level rise to 2100 and showed, for example, that planning levels
in Sydney, Australia, would need to be increased substantially to cope
with increased risk of flooding. Along the Portuguese coast, Andrade et
al. (2007) found that projected future climate in the HadCM3 model
would not affect wave height along this coastline but the projected
rotation in wave direction would increase the net littoral drift and the
erosional response. Along a section of the southeast coast of the United
Kingdom, the effect of sea level rise, surge, and wave climate change on
the inshore wave climate was evaluated and the frequency and height
of extreme waves was projected to increase in the north of the domain
(Chini et al., 2010). On the basis of modeling the 25-year beach response
along a stretch of the Portuguese coast to various climate change
scenarios, Coelho et al. (2009) concluded that the projected stormier
wave climate led to higher rates of beach erosion than mean sea level
rise. Modeling of the evolution of soft rock shores with rising sea levels
has revealed a relatively simple relationship between sea level rise and
the equilibrium cliff profile (Walkden and Dickson, 2008).
To summarize, recent observational studies that identify trends
and impacts at the coast are limited in regional coverage, which
means there is low confidence, due to insufficient evidence,
that anthropogenic climate change has been a major cause of
any observed changes. However, recent coastal assessments at
the national and regional scale and process-based studies have
provided further evidence of the vulnerability of low-lying
coastlines to rising sea levels and erosion, so that in the absence
of adaptation there is high confidence that locations currently
experiencing adverse impacts such as coastal erosion and
inundation will continue to do so in the future due to increasing
sea levels in the absence of changes in other contributing factors.
3.5.6. Glacier, Geomorphological, and Geological Impacts
Mountains are prone to mass movements including landslides, avalanches,
debris flows, and flooding that can lead to disasters. Changes in the
cryosphere affect such extremes, but also water supply and hydropower
generation. Many of the world’s high mountain ranges are situated at
the margins of tectonic plates, increasing the possibility of potentially
hazardous interactions between climatic and geological processes. The
principal drivers are glacier ice mass loss, mountain permafrost
degradation, and possible increases in the intensity of precipitation
(Liggins et al., 2010; McGuire, 2010). The possible consequences are
changes in mass movement on short contemporary time scales, and
modulations of seismicity and volcanic activity on longer, century to
millennium time scales.
The AR4 assessed that “the late 20th century glacier wastage likely has
been a response to post-1970 warming” (Lemke et al., 2007). However,
the impacts of glacier retreat on the natural physical system in the
context of changes in extreme events were not assessed in detail.
Additionally, the AR4 did not assess geomorphological and geological
impacts that might result from anthropogenic climate change. The most
studied change in the high-mountain environment has been the retreat
of glaciers (Paul et al., 2004; Kaser et al., 2006; Larsen et al., 2007;
Rosenzweig et al., 2007). Alpine glaciers around the world were at
maximum extent by the end of the Little Ice Age (~1850), and have
retreated since then (Leclercq et al., 2011), with an accelerated decay
during the past several decades (Zemp et al., 2007). Most glaciers have
retreated since the mid-19th century (Francou et al., 2000; Cullen et al.,
2006; Thompson et al., 2006; Larsen et al., 2007; Schiefer et al., 2007; Paul
and Haeberli, 2008). Rates of retreat that exceed historical experience and
internal (natural) variability have become apparent since the beginning of
the 21st century (Reichert et al., 2002; Haeberli and Hohmann, 2008).
Outburst floods from lakes dammed by glaciers or unstable moraines [or
‘glacial lake outburst floods’ (GLOFs)] are commonly a result of glacier
retreat and formation of lakes behind unstable natural dams (Clarke,
1982; Clague and Evans, 2000; Huggel et al., 2004; Dussaillant et al.,
2010). In the past century, GLOFs have caused disasters in many high-
mountain regions of the world (Rosenzweig et al., 2007), including the
Andes (Reynolds et al., 1998; Carey, 2005; Hegglin and Huggel, 2008),
the Caucasus and Central Asia (Narama et al., 2006; Aizen et al., 2007),
the Himalayas (Vuichard and Zimmermann, 1987; Richardson and
Reynolds, 2000; Xin et al., 2008; Bajracharya and Mool, 2009; Osti and
Egashira, 2009), North America (Clague and Evans, 2000; Kershaw et al.,
2005), and the European Alps (Haeberli, 1983; Haeberli et al., 2001;
Vincent et al., 2010). However, because GLOFs are relatively rare, it is
unclear whether their frequency of occurrence is changing at either the
regional or global scale. Clague and Evans (2000) argue that outburst
floods from moraine-dammed lakes in North America may have peaked
due to a reduction in the number of the lakes since the end of the Little
Ice Age. In contrast, a small but not statistically significant increase of
GLOF events was observed in the Himalayas over the period 1940 to
2000 (Richardson and Reynolds, 2000), but the event documentation
may not be complete. Over the past several decades, human mitigation
measures at unstable glacier lakes in the Himalaya and European Alps
may have prevented some potential GLOF events (Reynolds, 1998;
Haeberli et al., 2001).
Evidence of degradation of mountain permafrost and attendant slope
instability has emerged from recent studies in the European Alps
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
187
(Gruber and Haeberli, 2007; Huggel, 2009) and other mountain regions
(Niu et al., 2005; Geertsema et al., 2006; Allen et al., 2011). This evidence
includes several recent rock falls, rock slides, and rock avalanches in
areas where permafrost thaw in steep bedrock is occurring. Landslides
with volumes ranging up to a few million cubic meters have occurred in
the Mont Blanc region (Barla et al., 2000), in Italy (Sosio et al., 2008;
Huggel, 2009; Fischer et al., 2011), in Switzerland, and in British
Columbia (Evans and Clague, 1998; Geertsema et al., 2006). Very large
rock and ice avalanches with volumes of 30 to over 100 million m
3
include the 2002 Kolka avalanche in the Caucasus (Haeberli et al., 2004;
Kotlyakov et al., 2004; Huggel et al., 2005), the 2005 Mt. Steller rock
avalanche in the Alaska Range (Huggel et al., 2008), the 2007 Mt. Steele
ice and rock avalanche in the St. Elias Mountains, Yukon (Lipovsky et al.,
2008), and the 2010 Mt. Meager rock avalanche and debris flow in the
Coast Mountains of British Columbia.
Quantification of possible trends in the frequency of landslides and ice
avalanches in mountains is difficult due to incomplete documentation
of past events, especially those that happened before regular satellite
observations became available. Nevertheless, there has been an apparent
increase in large rock slides during the past two decades, and especially
during the first years of the 21st century in the European Alps (Ravanel
and Deline, 2011), in the Southern Alps of New Zealand (Allen et al., 2011),
and in northern British Columbia (Geertsema et al., 2006) in combination
with temperature increases, glacier shrinkage, and permafrost degradation.
Research, however, has not yet provided any clear indication of a
change in the frequency of debris flows due to recent deglaciation.
Debris flow activity at a local site in the Swiss Alps was higher during
the 19th century than today (Stoffel et al., 2005). In the French Alps no
significant change in debris flow frequency has been observed since the
1950s in terrain above elevations of 2,200 m (Jomelli et al., 2004).
Processes not, or not directly, driven by climate, such as sediment yield,
can also be important for changes in the magnitude or frequency of
alpine debris flows (Lugon and Stoffel, 2010).
Debris flows from both glaciated and unglaciated volcanoes, termed
lahars, can be particularly large and hazardous. Lahars produced by
volcanic eruptions on the glacier-clad Nevado del Huila volcano in
Colombia in 2007 and 2008 were the largest rapid mass flows on Earth
in recent years. Similarly, large mass flows occur on ice-covered active
volcanoes in Iceland (Björnsson, 2003), including Eyjafjallajökull in 2010.
Large rock and ice avalanches, with volumes up to 30 million m
3
, have
happened frequently (on average about one every four years) on the
glaciated Alaskan volcano, Iliamna, and are thought to be related to
elevated volcanic heat flow and possibly meteorological conditions
(Huggel et al., 2007). Glacier decay on active volcanoes can lead to a
reduction of lahar hazards due to less potential meltwater available for
lahar generation, but it is difficult to make a general conclusion as local
conditions also play important roles. In 1998, intense rainfall mobilized
pyroclastic material on the flanks of Vesuvius and Campi Flegrei
volcanoes, feeding approximately 150 debris flows that damaged nearby
communities and resulted in 160 fatalities (Bondi and Salvatori, 2003).
In the same year, intense precipitation associated with Hurricane Mitch
triggered a small flank collapse at Casita volcano in Nicaragua. This slope
failure transformed into debris flows that destroyed two towns and
claimed 2,500 lives (Scott et al., 2005). Following the 1991 Pinatubo
eruption in the Philippines, heavy rains associated with tropical storms
moved large volumes of volcanic sediment. The sediment dammed rivers,
causing massive flooding across the region that continued for several
years after the eruption ended (Newhall and Punongbayan, 1996).
A variety of climate and weather events can have geomorphological
and geological impacts. Warming and degradation of mountain
permafrost affect slope stability through a reduction in the shear
strength of ice-filled rock discontinuities. For example, the 2003
European summer heat wave (Section 3.3.1) caused rapid thaw and
thickening of the active layer, triggering a large number of mainly small
rock falls (Gruber et al., 2004; Gruber and Haeberli, 2007). Permafrost thaw
in sediment such as in talus slopes may increase both the frequency and
magnitude of debris flows (Zimmermann et al., 1997; Rist and Phillips,
2005). The frost table at the base of the active layer is a barrier to
groundwater infiltration and can cause the overlying non-frozen sediment
to become saturated. Snow cover can also affect debris flow activity by
supplying additional water to the soil, increasing pore water pressure
and initiating slope failure (Kim et al., 2004). Many of the largest debris
flows in the Alps in the past 20 years were triggered by intense rainfall
in summer or fall when the snowline was elevated (Rickenmann and
Zimmermann, 1993; Chiarle et al., 2007). Warming may increase the
flow speed of frozen bodies of sediment (Kääb et al., 2007; Delaloye et
al., 2008; Roer et al., 2008). Rock slopes can fail after they have been
steepened by glacial erosion or unloaded (debuttressed) following glacier
retreat (Augustinus, 1995). Although it may take centuries or even
longer for a slope to fail following glacier retreat, recent landslides
demonstrate that some slopes can respond to glacier down-wasting
within a few decades or less (Oppikofer et al., 2008). Twentieth-century
warming may have penetrated some decameters into thawing steep rock
slopes in high mountains (Haeberli et al., 1997). Case studies indicate that
both small and large slope failures can be triggered by exceptionally
warm periods of weeks to months prior to the events (Gruber et al.,
2004; Huggel, 2009; Fischer et al., 2011).
The spatial and temporal patterns of precipitation, the intensity and
duration of rainfall, and antecedent rainfall are important factors in
triggering shallow landslides (Iverson, 2000; Wieczorek et al., 2005;
Sidle and Ochiai, 2006). In some regions antecedent rainfall is probably
a more important factor than rainfall intensity (Kim et al., 1991; Glade,
1998), whereas in other regions rainfall duration and intensity are the
critical factors (Jakob and Weatherly, 2003). Landslides in temperate
and tropical mountains that have no seasonal snow cover are not
temperature-sensitive and may be more strongly influenced by human
activities such as poor land use practices, deforestation, and overgrazing
(Sidle and Ochiai, 2006).
Rock and ice avalanches on glaciated volcanoes can be triggered by
heat generated by volcanic activity. Their incidence may increase with
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
188
rising air and rock temperatures (Gruber and Haeberli, 2007) or during
or following brief, anomalously warm events (Huggel et al., 2010) due
to meltwater infiltration and shear strength reduction. Debuttressing
effects due to glacier retreat can also destabilize or over-steepen slopes
(Tuffen, 2010). Furthermore, on volcanoes, geothermal heat flow can
enhance ice melting and thus create weak zones at the ice-bedrock
interface; and hydrothermal alteration of rocks can decrease the slope
stability (Huggel, 2009). On unglaciated high volcanoes in the Caribbean,
Central America, Europe, Indonesia, the Philippines, and Japan, an
increase in total rainfall or an increase in the frequency or magnitude of
severe rainstorms (see Section 3.3.2) could cause more frequent debris
flows by mobilizing unconsolidated, volcanic regolith and by raising pore-
water pressures, which could lead to deep-seated slope failure. Heavy
rainfall events could also influence the behavior of active volcanoes. For
example, Mastin (1994) attributes the violent venting of volcanic gases
at Mount St. Helens between 1989 and 1991 to slope instability or
accelerated growth of cooling fractures within the lava dome following
rainstorms, and Matthews et al. (2002) link episodes of intense tropical
rainfall with collapses of the Soufriere Hills lava dome on Montserrat in
the Caribbean. It is well established that ice mass wastage following the
end of the last glaciations led to increased levels of seismicity associated
with post-glacial rebound of the lithosphere (e.g., Muir-Wood, 2000;
Stewart et al., 2000). There has been a large reduction in glacier cover
in southern Alaska. Sauber and Molnia (2004) reported several hundred
meters vertical reduction. This ice reduction may be responsible for an
increase in seismicity in the region where earthquake faults are at the
threshold of failure (Sauber and Molnia, 2004; Doser et al., 2007). An
increase in the frequency of small earthquakes in the Icy Bay area, also
in southeast Alaska, is interpreted to be a crustal response to glacier
wastage between 2002 and 2006 (Sauber and Ruppert, 2008). Large-
scale ice mass loss in glaciated volcanic terrain reduces the load on the
crust and uppermost mantle, facilitating magma formation and its
ascent into the crust (Jull and McKenzie, 1996) and allowing magma to
reach the surface more easily (Sigmundsson et al., 2010). At the end of
the last glaciation, this mechanism resulted in a more than 10-fold
increase in the frequency of volcanic eruptions in Iceland (Sinton et al.,
2005).
The AR4 projected that glaciers in mountains will lose additional mass
over this century because more ice will be lost due to summer melting
than is replenished by winter precipitation (Meehl et al., 2007b). The
total area of glaciers in the European Alps may decrease by 20 to more
than 50% by 2050 (Zemp et al., 2006; Huss et al., 2008). The projected
glacier retreat in the 21st century may form new potentially unstable
lakes. Probable sites of new lakes have been identified for some alpine
glaciers (Frey et al., 2010). Rock slope and moraine failures may trigger
damaging surge waves and outburst floods from these lakes. The
temperature rise also will result in gradual degradation of mountain
permafrost (Haeberli and Burn, 2002; Harris et al., 2009). The zone of warm
permafrost (mean annual rock temperature approximately -2 to 0°C),
which is more susceptible to slope failures than cold permafrost, may
rise in elevation a few hundred meters during the next 100 years
(Noetzli and Gruber, 2009). This in turn may shift the zone of enhanced
instability and landslide initiation toward higher-elevation slopes that in
many regions are steeper, and therefore predisposed to failure. The
response of bedrock temperatures to surface warming through thermal
conduction will be slow, but warming will eventually penetrate to
considerable depths in steep rock slopes (Noetzli et al., 2007). Other heat
transport processes such as advection, however, may induce warming of
bedrock at much faster rates (Gruber and Haeberli, 2007). The response
of firn and ice temperatures to an increase in air temperature is faster
and nonlinear (Haeberli and Funk, 1991; Suter et al., 2001; Vincent et al.,
2007). Latent heat effects from refreezing meltwater can amplify the
increase in air temperature in firn and ice (Huggel, 2009; Hoelzle et al.,
2010). At higher temperatures, more ice melts and the strength of the
remaining ice is lower; as a result, the frequency and perhaps size of ice
avalanches may increase (Huggel et al., 2004; Caplan-Auerbach and
Huggel, 2007). Warm extremes can trigger large rock and ice avalanches
(Huggel et al., 2010).
Current low levels of seismicity in Antarctica and Greenland may be a
consequence of ice-sheet loading, and isostatic rebound associated
with accelerated deglaciation of these regions may result in an increase
in earthquake activity, perhaps on time scales as short as 10 to 100
years (Turpeinen et al., 2008; Hampel et al., 2010). Future ice mass loss
on glaciated volcanoes, notably in Iceland, Alaska, Kamchatka, the
Cascade Range in the northwest United States, and the Andes, could
lead to eruptions, either as a consequence of reduced load pressures on
magma chambers or through increased magma-water interaction.
Reduced ice load arising from future thinning of Iceland’s Vatnajökull
Ice Cap is projected to result in an additional 1.4 km
3
of magma
produced in the underlying mantle every century (Pagli and Sigmundsson,
2008). Ice unloading may also promote failure of shallow magma
reservoirs with a potential consequence of a small perturbation of the
natural eruptive cycle (Sigmundsson et al., 2010). Initially, ice thinning
of 100 m or more on volcanoes with glaciers more than 150-m thick,
such as Sollipulli in Chile, may cause more explosive eruptions, with
increased tephra hazards (Tuffen, 2010). Additionally, the potential for
edifice lateral collapse could be enhanced by loss of support previously
provided by ice (Tuffen, 2010) or to elevated pore-water pressures
arising from meltwater (Capra, 2006; Deeming et al., 2010). Ultimately
the loss of ice cover on glaciated volcanoes may reduce opportunities
for explosions arising from magma-ice interaction. The incidence of ice-
sourced lahars may also eventually fall, although exposure of new
surfaces of volcanic debris due to ice wastage may provide the raw
material for precipitation-related lahars. The likelihood of both volcanic
and non-volcanic landslides may also increase due to greater availability
of water, which could destabilize slopes. Many volcanoes provide a
ready source of unconsolidated debris that can be rapidly transformed
into potentially hazardous lahars by extreme precipitation events.
Volcanoes in coastal, near-coastal, or island locations in the tropics are
particularly susceptible to torrential rainfall associated with tropical
cyclones, and the rainfall rate associated with tropical cyclones is
projected to increase though the number of tropical cyclones is projected
to decrease or stay essentially unchanged (see Section 3.4.4). The impact
of future large explosive volcanic eruptions may also be exacerbated by an
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
189
increase in extreme precipitation events (see Section 3.3.2) that provide
an effective means of transferring large volumes of unconsolidated
ash and pyroclastic flow debris from the flanks of volcanoes into
downstream areas.
Quantification of possible trends in the frequency of landslides
and ice avalanches in mountains is difficult due to incomplete
documentation of past events. There is high confidence that
changes in heat waves, glacial retreat, and/or permafrost
degradation will affect high mountain phenomena such as slope
instabilities, mass movements, and glacial lake outburst floods,
and medium confidence that temperature-related changes will
influence bedrock stability. There is also high confidence that
changes in heavy precipitation will affect landslides in some
regions. There is medium confidence that high-mountain debris
flows will begin earlier in the year because of earlier snowmelt,
and that continued mountain permafrost degradation and glacier
retreat will further decrease the stability of rock slopes. There is
low confidence regarding future locations and timing of large rock
avalanches, as these depend on local geological conditions and
other non-climatic factors. There is low confidence in projections
of an anthropogenic effect on phenomena such as shallow
landslides in temperate and tropical regions, because these are
strongly influenced by human activities such as poor land use
practices, deforestation, and overgrazing. It is well established
that ice mass wastage following the end of the last glaciations
led to increased levels of seismicity, but there is low confidence
in the nature of recent and projected future seismic responses to
anthropogenic climate change.
3.5.7. High-latitude Changes Including Permafrost
Permafrost is widespread in Arctic, in subarctic, in ice-free areas of
Antarctica, and in high-mountain regions, and permafrost regions occupy
approximately 23 million km
2
of land area in the Northern Hemisphere
(Zhang et al., 1999). Melting of massive ground ice and thawing of
ice-rich permafrost can lead to subsidence of the ground surface and to
the formation of uneven topography known as thermokarst, having
implications for ecosystems, landscape stability, and infrastructure
performance (Walsh, 2005). See also Case Study 9.2.10 for discussion of
the impacts of cold events in high latitudes. The active layer (near-
surface layer that thaws and freezes seasonally over permafrost) plays
an important role in cold regions because most ecological, hydrological,
biogeochemical, and pedogenic (soil-forming) activity takes place within
it (Hinzman et al., 2005).
Observations show that permafrost temperatures have increased since the
1980s (IPCC, 2007b). Temperatures in the colder permafrost of northern
Alaska, the Canadian Arctic, and Russia have increased up to 3°C near the
permafrost table and up to 1 to 2°C at depths of 10 to 20 m (Osterkamp,
2007; Romanovsky et al., 2010; S.L. Smith et al., 2010) since the late
1970s/early 1980s. Temperature increases have generally been less than
1°C in the warmer permafrost of the discontinuous permafrost zone of
the polar regions (Osterkamp, 2007; Romanovsky et al., 2010; S.L. Smith
et al., 2010), and also in the high-altitude permafrost of Mongolia and
the Tibetan Plateau (Zhao et al., 2010). When the other conditions
remain constant, active layer thickness is expected to increase in
response to warming. Active layer thickness has increased by about
20 cm in the Russian Arctic between the early 1960s and 2000 (T. Zhang
et al., 2005) and by up to 1.0 m over the Qinghai-Tibetan Plateau since
the early 1980s (Wu and Zhang, 2010), with no significant trend in the
North American Arctic since the early 1990s (Shiklomanov et al., 2010).
However, over extreme warm summers, active layer thickness may
increase substantially (Smith et al., 2009), potentially triggering active-
layer detachment failures on slopes (Lewkowicz and Harris, 2005).
Extensive thermokarst development has been found in Alaska (Jorgenson
et al., 2006; Osterkamp et al., 2009), Canada (Vallée and Payette, 2007),
and central Yakutia (Gavriliev and Efremov, 2003). Increased rates of
retrogressive thaw slump activities have been reported on slopes over
the Qinghai-Tibetan Plateau (Niu et al., 2005) and adjacent to tundra
lakes over the Mackenzie Delta region of Canada (Lantz and Kokelj,
2008). Substantial expansion and deepening of thermokarst lakes was
observed near Yakutsk with subsidence rates of 17 to 24 cm yr
-1
from
1992 to 2001 (Fedorov and Konstantinov, 2003). Satellite remote sensing
data show that thaw lake surface area has increased in continuous
permafrost regions and decreased in discontinuous permafrost regions
(Smith et al., 2005). Coasts with ice-bearing permafrost that are exposed
to the Arctic Ocean are very sensitive to permafrost degradation. Some
Arctic coasts are retreating at a rapid rate of 2 to 3 m yr
–1
and the rate
of erosion along Alaska’s northeastern coastline has doubled over the
past 50 years, related to declining sea ice extent, increasing sea surface
temperature, rising sea level, thawing coastal permafrost, and possibly
increases in storminess and waves (Jones et al., 2009; Karl et al., 2009)
Increases in air temperature are in part responsible for the observed
increase in permafrost temperature over the Arctic and subarctic, but
changes in snow cover also play a critical role (Osterkamp, 2005; Zhang,
2005; T. Zhang et al., 2005; S.L. Smith et al., 2010). Trends toward earlier
snowfall in autumn and thicker snow cover during winter have resulted
in a stronger snow insulation effect, and as a result a much warmer
permafrost temperature than air temperature in the Arctic. On the other
hand, permafrost temperature may decrease even if air temperature
increases, if there is also a decrease in the duration and thickness of
snow cover (Taylor et al., 2006). The lengthening of the thaw season and
increases in summer air temperature have resulted in changes in active
layer thickness. Model simulations have projected thickening of the
active layer, a northward shift of the permafrost boundary, reductions
in permafrost area, and an increase in permafrost temperature in the
21st century and beyond (Saito et al., 2007; Schaefer et al., 2011). The
projected permafrost degradation may result in ancient carbon currently
frozen in permafrost being released into the atmosphere, providing a
positive feedback to the climate system (Schaefer et al., 2011). Expansion
of lakes in the continuous permafrost zone may be due to thawing of
ice-rich permafrost and melting of massive ground ice, while decreases
in lake area in the discontinuous permafrost zone may be due to lake
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
190
bottom drainage (Smith et al., 2005). Overall, increased air temperature
over high latitudes is primarily responsible for the development of
thermokarst terrains and thaw lakes.
In summary, it is likely that there has been warming of permafrost
in recent decades. There is high confidence that permafrost
temperatures will continue to increase, and that there will be
increases in active layer thickness and reductions in the area of
permafrost in the Arctic and subarctic.
3.5.8. Sand and Dust Storms
Sand and dust storms are widespread natural phenomena in many parts
of the world. Heavy dust storms disrupt human activities. Dust aerosols in
the atmosphere can cause a suite of health impacts including respiratory
problems (Small et al., 2001). The long-range transport of dust can
affect conditions at long distances from the dust sources, linking the
biogeochemical cycles of land, atmosphere, and ocean (Martin and
Gordon, 1988; Bergametti and Dulac, 1998; Kellogg and Griffin, 2006).
For example, dust from the Saharan region and from Asia may reach
North America and South America (McKendry et al., 2007). Some climate
models have representations of dust aerosols (Textor et al., 2006).
Climate variables that are most important to dust emission and transport
such as soil moisture (see also Section 3.5.1), precipitation, wind, and
vegetation cover are still subject to large uncertainties in climate model
simulations. As a result, the sand and dust storm simulations have large
uncertainties as well.
The Sahara (especially the Bodélé Depression in Chad) and east Asia
have been recognized as the largest dust sources globally (Goudie,
2009). Over the few decades before the 1990s, the frequency of dust
events increased in some regions such as the Sahel zone of Africa
(Goudie and Middleton, 1992), and decreased in some other regions
such as China (Zhang et al., 2003). There seems to be an increase in more
recent years in China (Shao and Dong, 2006). Despite the importance of
African dust, studies on long-term change in Sahel dust are limited.
However, dust transported far away from the source region may provide
some evidence of long-term changes in the Sahel region. The African
dust transported to Barbados began to increase in the late 1960s and
through the 1970s; transported dust reached a peak in the early 1980s
but remains high into the present (Prospero and Lamb, 2003; Prospero
et al., 2009).
Surface soil dust concentration during a sand and dust storm is
controlled by a number of factors. The driving force for the production
of dust storms is the surface wind associated with cold frontal systems
sweeping across arid and semi-arid regions and lifting soil particles in
the atmosphere. Dust emissions are also controlled by the surface
conditions in source regions such as the desert coverage distributions,
snow cover, and soil moisture. For example, in the Sahel region, the
elevated high level of dust emission prior to the 1990s was related to
the persistent drought during that time, and to long-term changes in the
NAO (Ginoux et al., 2004; Chiapello et al., 2005; Engelstaedter et al.,
2006), and perhaps to North Atlantic SST as well (Wong et al., 2008).
Further evidence of the importance of climate on dust emission is that
despite an increase of approximately 2 to 7% in desert areas in China
over the four decades since 1960, dust storm frequency decreased in
that period (Zhong, 1999). Studies on Asian soil dust production from
1960 to 2003 suggest that climatic variations have played a major role
in the declining trends in dust emission and storm frequencies in China
(Zhang et al., 2003; Zhou and Zhang, 2003; Zhao et al., 2004; Gong et
al., 2006). Overall, changes in dust activity are affected by changes in
the climate, such as wind and moisture conditions in the dust source
regions. Changes in large-scale circulation play an additional role in the
long-distance transport of dust. However, understanding of the physical
mechanisms of the long-term trends in dust activity is not complete; for
example, the relative importance of the various factors affecting dust
frequency as outlined above is uncertain.
Future dust activity depends on two main factors: land use in the dust
source regions, and climate both in the dust source region and large-
scale circulation that affects long distance dust transport. Studies on
projected future dust activity are very limited. It is difficult to project
future land use. Precipitation, soil moisture, and runoff have been
projected to decrease in major dust source regions (Figure 10.12 in Meehl
et al., 2007b). Thomas et al. (2005) suggest that dune fields in southern
Africa can become active again, and sand will become significantly
exposed and move, as a consequence of 21st-century warming. A study
based on simulations from two climate models also suggests increased
desertification in arid and semi-arid China, especially in the second half
of the 21st century (X.M. Wang et al., 2009). However, confident projected
changes in wind are lacking (see Section 3.3.3).
In summary, there is low confidence in projecting future dust
storm changes, although an increase could be expected where
aridity increases. There is a lack of data and studies on past
changes. There is also a lack of understanding of processes such
as the relative importance of different climate variables affecting
dust storms, as well as a high uncertainty in simulating important
climate variables such as soil moisture, precipitation, and wind
that affect dust storms.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
191
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
High confidence: Likely overall increase in WD,
decrease in CD (Aguilar et al., 2005; Alexander et
al., 2006).
High confidence: Likely overall decrease in
CN, increase in WN (Aguilar et al., 2005;
Alexander et al., 2006).
Medium confidence: Overall increase
since 1960 (Kunkel et al., 2008). Some
areas with significant WSDI increase,
others with insignificant WSDI increase
or decrease (Alexander et al., 2006).
High confidence: Likely increase in many
areas since 1950 (Aguilar et al., 2005;
Alexander et al., 2006; Trenberth et al.,
2007; Kunkel et al., 2008).
Medium confidence: Overall slight decrease
in dryness (SMA, PDSI, CDD) since 1950;
regional variability and 1930s drought
dominate the signal (Aguilar et al., 2005;
Alexander et al., 2006; Kunkel et al., 2008;
Sheffield and Wood, 2008a; Dai, 2011).
High confidence: Very likely large increases in
WD, large decreases in CD (Robeson, 2004;
Vincent and Mekis, 2006; Kunkel et al., 2008;
Peterson et al., 2008a).
High confidence: Very likely large decreases
in CN, large increases in WN (Robeson,
2004; Vincent and Mekis, 2006; Kunkel et
al., 2008; Peterson et al., 2008a).
Medium confidence: Increase in WSDI
(Alexander et al., 2006).
Medium confidence: Spatially varying
trends. General increase, decrease in some
areas (Alexander et al., 2006).
Medium confidence: No overall or slight
decrease in dryness (SMA, PDSI, CDD) since
1950; large variability, large drought of
1930s dominates (Alexander et al., 2006;
Kunkel et al., 2008; Sheffield and Wood,
2008a; Dai, 2011).
Medium confidence: Spatially varying trends.
Small increases in WD, decreases in CD in north
CNA. Small decreases in WD, increases in CD in
south CNA (Robeson, 2004; Vincent and Mekis,
2006; Kunkel et al., 2008; Peterson et al., 2008a).
Medium confidence: Spatially varying
trends. Small decreases in CN, increases in
WN in north CNA. Small increases in CN,
decreases in WN in south CNA (Robeson,
2004; Vincent and Mekis, 2006; Kunkel et
al., 2008; Peterson et al., 2008a).
Medium confidence: Spatially varying
trends. Some areas with WSDI increase,
others with WSDI decrease (Alexander et
al., 2006).
High confidence: Very likely increase since
1950 (Alexander et al., 2006).
Medium confidence: Decrease in dryness
(SMA, PDSI, CDD) and increase in mean
precipitation since 1950; large variability,
large drought of 1930s dominates
(Alexander et al., 2006; Kunkel et al., 2008;
Sheffield and Wood, 2008a; Dai, 2011).
Medium confidence: Spatially varying trends.
Overall increases in WD, decreases in CD;
opposite or insignificant signal in a few areas
(Robeson, 2004; Vincent and Mekis, 2006; Kunkel
et al., 2008; Peterson et al., 2008a).
Medium confidence: Weak and spatially
varying trends. (Robeson, 2004; Vincent
and Mekis, 2006; Kunkel et al., 2008;
Peterson et al., 2008a).
Medium confidence: Spatially varying
trends. Many areas with WSDI increase,
some areas with WSDI decrease
(Alexander et al., 2006).
High confidence: Very likely increase since
1950 (Alexander et al., 2006).
Medium confidence: Slight decrease in
dryness (SMA, PDSI, CDD) since 1950, large
variability, large drought of 1930s
dominates (Alexander et al., 2006; Kunkel
et al., 2008; Sheffield and Wood, 2008a;
Dai, 2011).
High confidence: Very likely large increases in
WD, large decreases in CD (Robeson, 2004;
Vincent and Mekis, 2006; Kunkel et al., 2008;
Peterson et al., 2008a).
High confidence: Very likely large decreases
in CN, large increases in WN (Robeson,
2004; Vincent and Mekis, 2006; Kunkel et
al., 2008; Peterson et al., 2008a).
Low confidence: Insufficient evidence. Medium confidence: Slight tendency for
increase, in southern Alaska; no significant
trend (Kunkel et al., 2008).
Medium confidence: Inconsistent trends;
increase in dryness (SMA, PDSI, CDD) since
1950 in part of the region. (Alexander et
al., 2006; Kunkel et al., 2008; Sheffield and
Wood, 2008a; Dai, 2011).
High confidence: Likely increases in WD in some
areas, decrease in others. Decreases in CD in
some areas, increase in others (Robeson, 2004;
Alexander et al., 2006; Vincent and Mekis, 2006;
Trenberth et al., 2007; Kunkel et al., 2008;
Peterson et al., 2008a).
Medium confidence: Small increases in
unusually cold nights, decreases in WN in
northeastern Canada. Small decreases in
CN, increases in WN in southeastern and
south central Canada. (Robeson, 2004;
Vincent and Mekis, 2006; Kunkel et al.,
2008; Peterson et al., 2008a).
Medium confidence: Some areas with
WSDI increase, most others with WSDI
decrease (Alexander et al., 2006).
Medium confidence: Increase in a few areas
(Alexander et al., 2006).
Low confidence: Insufficient evidence.
Medium confidence: Increases in WD, decreases
in CD (Aguilar et al., 2005; Alexander et al.,
2006).
Medium confidence: Decreases in CN,
increases in WN (Aguilar et al., 2005;
Alexander et al., 2006).
Low confidence: Spatially varying trends.
A few areas increase, a few others
decrease (Aguilar et al., 2005; Alexander
et al., 2006).
Medium confidence: Spatially varying
trends. Increase in many areas, decrease in
a few areas, (Aguilar et al., 2005;
Alexander et al., 2006).
Low confidence: Spatially varying trends,
inconsistencies in trends in dryness (SMA,
PDSI, CDD). (Aguilar et al., 2005; Sheffield
and Wood, 2008a; Dai, 2011).
Continued next page
Tmin
[WN = Warm Nights
CN = Cold Nights; see Box 3-1] (using
late 20th-century extreme values as
reference, e.g., 90th/10th percentile)
Heavy Precipitation (HP)
(using late 20th-century extreme values
as reference, e.g., 90th percentile)
Dryness
[CDD = Consecutive Dry Days
SMA = (Simulated) Soil Moisture
Anomalies
PDSI = Palmer Drought Severity Index;
see Box 3-3 for definitions]
Tmax
[WD = Warm Days
CD = Cold Days; see Box 3-1] (using late
20th-century extreme values as reference,
e.g., 90th/10th percentile)
Regions
Heat Waves (HW)/
Warm Spells (WS)
[WSDI = Warm Spell Duration Index,
i.e., number or fraction of days
belonging to spells of at least 6
days with Tmax >90th percentile]
(using late 20th-century extreme
values as reference)
All North
America and
Central
America
W. North
America
(WNA, 3)
Central
North
America
(CNA, 4)
E. North
America
(ENA, 5)
Alaska/
N.W. Canada
(ALA, 1)
E. Canada,
Greenland,
Iceland
(CGI, 2)
Central
America and
Mexico
(CAM, 6)
A. North America and Central America
Table 3-2 | Regional observed changes in temperature and precipitation extremes, including dryness, since 1950 unless indicated otherwise, and using late 20th-century values as reference (see Box 3-1), generally 1961-1990.
See Figure 3-1 for definitions of regions. For assessments for small island states refer to Box 3-4.
192
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
High confidence: Overall likely increase in WD
and likely decrease of CD over most of the
continent since 1950. Strong increasing tendency
in WD in most regions since 1976 onward; small
or insignificant decrease in CD over same period
(Alexander et al., 2006; see also entries for
individual subregions).
High confidence: Overall likely increase in
WN and likely decrease in CN over most of
the continent since 1950. Strong increasing
tendency in WN in most regions since 1976
onward; small or insignificant decrease in
CN over same period (Klein Tank and
Können, 2003; Alexander et al., 2006; see
also entries for individual subregions).
Medium confidence: Increase of HW
since 1950. Overall consistent positive
trend of WSDI across Europe, but no
coherent region with significant trends
(Alexander et al., 2006). Availability of a
few single studies for specific regions
(see below).
Medium confidence: Increase in part of the
region, mostly in winter, insignificant or
inconsistent changes elsewhere, in
particular in summer. Some inconsistencies
in overall patterns between studies
depending on considered indices. Most
consistent signal over central W. Europe
and European Russia (Klein Tank and
Können, 2003; Haylock and Goodess, 2004;
Alexander et al., 2006; Zolina et al., 2009).
Medium confidence: Inconsistent trends.
Increase in dryness (SMA, PDSI, CDD) in
part of the region; insignificant,
inconsistent, or no changes elsewhere.
Most consistent signal for increase in
dryness in central and S. Europe since the
1950s. No signal in N. Europe (Kiktev et al.,
2003; Haylock and Goodess, 2004;
Alexander et al., 2006; Sheffield and Wood,
2008a; Dai, 2011).
Medium confidence: Increase in WD and
decrease in CD. Consistent signals for whole
region, but generally not significant at the local
scale (Alexander et al., 2006).
Medium confidence: Increase in WN and
decrease in CN. Consistent signals over
whole region but generally not significant at
the local scale (Klein Tank and Können,
2003; Alexander et al., 2006).
Medium confidence: Increase in HW.
Consistent tendency for increase in
WSDI, but no significant trends
(Alexander et al., 2006).
Medium confidence: Increase in winter in
some areas, but often insignificant or
inconsistent trends at subregional scale, in
particular in summer (Fowler and Kilsby,
2003; Kiktev et al., 2003; Klein Tank and
Können, 2003; Alexander et al., 2006;
Maraun et al., 2008; Zolina et al., 2009).
Medium confidence: Spatially varying
trends. Overall only slight or no increase in
dryness (SMA, PDSI, CDD), slight decrease
in dryness in part of the region (Kiktev et
al., 2003; Alexander et al., 2006; Sheffield
and Wood, 2008a; Dai, 2011).
High confidence: Likely overall increase in WD
and likely decrease in CD since 1950 in most
regions. Some regional and temporal variations
in significance of trends.
High confidence: Very likely increase in WD since
1950, 1901 and 1880 and likely decrease in CD
since 1950 and 1901 in west Central Europe
(Alexander et al., 2006; Della-Marta et al.,
2007a; Laurent and Parey, 2007).
Medium confidence: Lower confidence in trends
in east Central Europe due to lack of literature,
partial lack of access to observations, overall
weaker signals, and change point in trends at
the end of the 1970s / beginning of 1980s.
Strongest increase in WD since 1976 (Alexander
et al., 2006; Bartholy and Pongracz, 2007;
Hirschi et al., 2011).
High confidence: Likely overall increase in
WN and likely overall decrease in CN at the
yearly time scale. Some regional and
seasonal variations in significance and in a
few cases also the sign of the trends.
High confidence: Very likely increase in WN
and very likely decrease in CN since 1950
and 1901 in west Central Europe (Kiktev et
al., 2003; Alexander et al., 2006).
Medium confidence: Lower confidence in
trends in east Central Europe due to lack of
literature, partial lack of access to
observations, overall weaker signals, and
change point in trends at the end of the
1970s / beginning of 1980s. (Klein Tank and
Können, 2003; Alexander et al., 2006;
Bartholy and Pongracz, 2007).
Medium confidence: Increase in heat
waves. Consistent tendency for WSDI
increase but no significant trends
(Alexander et al., 2006). Significant
increase in max HW duration since 1880
in west Central Europe in summer (JJA)
(Della-Marta et al., 2007a). Less
significant signal in heat wave indices in
east Central Europe due to presence of
change point (Bartholy and Pongracz,
2007; Hirschi et al., 2011).
Medium confidence: Increase in part of the
domain, in particular in central W. Europe
and European Russia, especially in winter.
Insignificant or inconsistent trends
elsewhere, in particular in summer (Kiktev
et al., 2003; Klein Tank and Können, 2003;
Schmidli and Frei, 2005; Alexander et al.,
2006; Bartholy and Pongracz, 2007; Kyselý,
2009; Tomassini and Jacob, 2009; Zolina et
al., 2009).
Medium confidence: Spatially varying
trends. Increase in dryness (SMA, PDSI,
CDD) in part of the region but some
regional variation in dryness trends and
dependence of trends on considered
studies (index, time period) (Kiktev et al.,
2003; Alexander et al., 2006; Bartholy and
Pongracz, 2007; Sheffield and Wood,
2008a; Brázdil et al., 2009; Dai, 2011).
High confidence: Likely increase in WD and likely
decrease in CD in most of the region. Some
regional and temporal variations in significance
of trends. Likely strongest and most significant
trends in the Iberian Peninsula and southern
France (Alexander et al., 2006; Brunet et al.,
2007; Della-Marta et al., 2007a; Bartolini et al.,
2008; Kuglitsch et al., 2010; Rodríguez-Puebla et
al., 2010; Hirschi et al., 2011).
Medium confidence: Smaller or less significant
trends in S.E. Europe and Italy due to change
point in trends at the end of the 1970s /
beginning of 1980s; sometimes linked with
changes in sign of trends; strongest WD increase
since 1976 (Bartholy and Pongracz, 2007;
Bartolini et al., 2008; Toreti and Desiato, 2008;
Kuglitsch et al., 2010; Hirschi et al., 2011).
High confidence: Likely increase in WN and
likely decrease in CN in most of the region.
Some regional variations in significance of
trends.
Very likely overall increase in WN and very
likely overall decrease in CN in S.W. Europe
and W. Mediterranean; likely strongest
signals in Spain and southern France (Kiktev
et al., 2003; Klein Tank and Können, 2003;
Alexander et al., 2006; Brunet et al., 2007;
Rodríguez-Puebla et al., 2010). Likely overall
tendency for increase in WN and likely
overall tendency for decrease in CN in S.E.
Europe and E. Mediterranean (Kiktev et al.,
2003; Klein Tank and Können, 2003;
Alexander et al., 2006).
High confidence: Likely overall increase
in HW in summer (JJA). Significant
increase in max HW duration since 1880
in Iberian Peninsula and west Central
Europe in JJA (Della-Marta et al.,
2007a). Significant increase in max HW
duration in Tuscany (Italy) (Bartolini et
al., 2008). Significant increase in HW
indices in Turkey and to a smaller extent
in S.E. Europe and Turkey in JJA
(Kuglitsch et al., 2010). Less significant
signal in HW indices in S.E. Europe due
to presence of change point in trends
(Bartholy and Pongracz, 2007; Hirschi et
al., 2011).
Low confidence: Inconsistent trends within
domain and across studies (Kiktev et al.,
2003; Klein Tank and Können, 2003;
Alexander et al., 2006; García et al., 2007;
Pavan et al., 2008; Zolina et al., 2009;
Rodrigo, 2010).
Medium confidence: Overall increase in
dryness (SMA, PDSI, CDD), but partial
dependence on index and time period
(Kiktev et al., 2003; Alexander et al., 2006;
Sheffield and Wood, 2008a; Dai, 2011).
All Europe and
Mediterranean
Region
N. Europe
(NEU, 11)
Central
Europe
(CEU, 12)
S. Europe and
Mediterranean
(MED, 13)
Continued next page
Tmin Heavy Precipitation Dryness
Tmax
Regions
Heat Waves / Warm Spells
B. Europe and Mediterranean Region
Table 3-2 (continued)
193
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Low confidence to medium confidence: Low
confidence due to insufficient evidence (lack of
literature) in many regions. Medium confidence
in increase in frequency of WD and decrease in
frequency of CD in southern part of continent
(Alexander et al., 2006). See also regional
assessments.
Low confidence to medium confidence
depending on region: Low confidence due to
insufficient evidence (lack of literature) in
many regions. Medium confidence in
increase in frequency of WN in northern and
southern part of continent (Alexander et al.,
2006). Medium confidence in decrease in
frequency of CN in southern part of
continent (Alexander et al., 2006). See also
regional assessments.
Low confidence: Insufficient evidence
(lack of literature). Some analyses for
localized regions (see regional
assessments).
Low confidence: Partial lack of data and
literature and inconsistent patterns in
existing studies (New et al., 2006; Aguilar
et al., 2009). See also regional assessments.
Medium confidence: Overall increase in
dryness (SMA, PDSI); regional variability,
1970s prolonged Sahel drought dominates
(Sheffield and Wood, 2008a; Dai, 2011). No
apparent continent-wide trends in change
in rainfall over the 20th century, although
there was a continent-wide drought in
1983 and 1984 (Hulme et al., 2001). Wet
season arrives 9–21 days later, large inter-
annual variability of wet season start, local-
scale geographical variability (Kniveton et
al., 2009).
Medium confidence: Significant increase in
temperature of warmest day and coldest day,
significant increase in frequency of WD, and
significant decrease in frequency of CD in
western central Africa, Guinea Conakry, Nigeria,
and Gambia (New et al., 2006; Aguilar et al.,
2009).
Low confidence: Lack of literature in other parts
of the region.
Medium confidence: Decreases in frequency
of CN in western central Africa, Nigeria, and
Gambia; insignificant decreases in frequency
of CN in Guinea Conakry (New et al., 2006;
Aguilar et al., 2009).
Low confidence: Lack of literature on
changes in CN in other parts of the region.
Medium confidence: Increases in frequency
of WN (Alexander et al., 2006; New et al.,
2006; Aguilar et al., 2009).
Low confidence: Insufficient evidence
(lack of literature) for most of the
region; increases in WSDI in Nigeria and
Gambia (New et al., 2006).
Medium confidence: Precipitation from
heavy events has decreased (western
central Africa, Guinea Conakry) but low
spatial coherence (Aguilar et al., 2009),
rainfall intensity increased (New et al.,
2006).
Medium confidence: 1970s prolonged
Sahel drought dominates, conditions are
still drier (SMA, PDSI, precipitation
anomalies) than during the humid 1950s
(L'Hôte et al., 2002; Dai et al., 2004;
Sheffield and Wood, 2008a; Dai, 2011). Dry
spell duration (CDD) overall increased from
1961 to 2000 (New et al., 2006). Recent
years characterized by a greater
interannual variability than previous 40
years, western Sahel remaining dry and the
eastern Sahel returning to wetter
conditions (Ali and Lebel, 2009).
Low confidence: Lack of evidence due to lack of
literature and spatially non-uniform trends. Over
time period 1939–1992 spatially non-uniform
trends in daytime temperature, some areas with
cooling (King’uyu et al., 2000). In southern tip of
domain increases in WD, decreases in CD
(Alexander et al., 2006).
Medium confidence: Over time period 1939–
1992, spatially non-uniform trends, rise of
nighttime temperature at several locations,
but with many coastal areas and stations
near large water bodies showing a
significant decrease (King’uyu et al., 2000).
In southern tip of domain, decreases in CN,
increases in WN (Alexander et al., 2006).
Low confidence: Insufficient evidence
(lack of literature) for most of region;
increase in WSDI in southern tip of
domain (New et al., 2006).
Low confidence: Insufficient evidence (lack
of literature) to assess trends.
Low confidence: Spatially varying trends in
dryness (SMA, PDSI) (Sheffield and Wood,
2008a; Dai, 2011).
Medium confidence: Increases in WD, decreases
in CD (Alexander et al., 2006).
Medium confidence: Decreases in CN,
increases in WN (King’uyu et al., 2000;
Alexander et al., 2006).
Medium confidence: Increase in WSDI
(New et al., 2006).
Low confidence: No spatially coherent
patterns of trends in precipitation extremes
(Kruger, 2006; New et al., 2006; Trenberth
et al., 2007).
Medium confidence: Slight dry spell
duration increase (Alexander et al., 2006;
Kruger, 2006; New et al., 2006). General
increase in dryness (SMA, PDSI) (Sheffield
and Wood, 2008a; Dai, 2011).
Low confidence: Lack of literature. Medium confidence: Increases in WN
(Alexander et al., 2006).
Low confidence: Lack of literature on trends
in CN.
Low confidence: Insufficient evidence
(lack of literature).
Low confidence: Insufficient evidence. Low confidence: Limited data, spatial
variation in the trends (Dai, 2011).
Tmin Heavy Precipitation Dryness
Tmax
Regions
Heat Waves / Warm Spells
All Africa
W. Africa
(WAF, 15)
S. Africa
(SAF, 17)
Sahara
(SAH, 14)
E. Africa
(EAF, 16)
Continued next page
C. Africa
Table 3-2 (continued)
194
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Low confidence to high confidence depending on
region: On continental scale, medium confidence
in overall increase in WD and decrease in CD
(Alexander et al., 2006). See also individual
regional entries.
Low confidence to high confidence
depending on region: On continental scale,
medium confidence in overall increase in
WN and decrease in CN (Alexander et al.,
2006). See also individual regional entries.
Low confidence to medium confidence
depending on region: Low confidence
due to insufficient evidence in several
regions; medium confidence in trends in
other regions (Alexander et al., 2006).
See also individual regional entries.
Low confidence to medium confidence
depending on region: Low confidence due
to insufficient evidence or inconsistent
trends in several regions; medium
confidence in trends in HP in a few regions
(Alexander et al., 2006). See also individual
regional entries.
Low confidence to medium confidence
depending on region: Low confidence in
most regions due to spatially varying
trends. Some areas have consistent
increases, but others display decreases in
dryness indicated by different measures
(SMA, PDSI, CDD) (Alexander et al., 2006;
Sheffield and Wood, 2008a; Dai, 2011).
High confidence: Likely increases in WD, likely
decreases in CD (Alexander et al., 2006).
High confidence: Likely decreases in CN,
likely increases in WN (Alexander et al.,
2006).
Medium confidence: Spatially varying
trends. Overall WSDI increase, WSDI
decrease in a few areas (Alexander et
al., 2006).
Low confidence: Increase in some regions,
but spatial variations (Alexander et al.,
2006). Some increase in western Russia,
especially in winter (DJF), 1950–2000
(Zolina et al., 2009).
Low confidence: Spatially varying trends.
Tendency for increased dryness (SMA, PDSI,
CDD) in central and northeastern N. Asia,
other areas decreased dryness (Alexander
et al., 2006; Sheffield and Wood, 2008a;
Dai, 2011).
Low confidence to medium confidence
depending on region: Low confidence in trends
in WD and CD due to insufficient evidence in
many regions, in particular in northern half of
continent. Medium confidence in trends in
southern half of continent but often spatially
varying trends. See regional assessments for
details and basis for continental assessment.
Low confidence to medium confidence
depending on region: Low confidence in
trends in WN and CN due to insufficient
evidence in many regions, in particular in
northern half of continent. Medium
confidence in decrease in CN and increase in
WN in southern half of continent. See
regional assessments for details and basis
for continental assessment.
Low confidence: Insufficient evidence
(most of continent) or lack of coherent
signal (Southeast South America). See
regional assessments for details and
basis for continental assessment.
Low confidence: Insufficient evidence or
spatially varying trends. See regional
assessments for details and basis for
continental assessment.
Low confidence: Spatially varying trends,
inconsistencies between studies (Sheffield
and Wood, 2008a; Dai, 2011).
Low confidence: Insufficient, scattered, evidence
(Alexander et al., 2006; Dufek et al., 2008).
Low confidence: Insufficient, scattered,
evidence (Alexander et al., 2006; Dufek et
al., 2008).
Low confidence: Insufficient evidence. Medium confidence: Spatially varying
trends. Increase in many areas, decrease in
a few areas (Alexander et al., 2006;
Haylock et al., 2006b).
Low confidence: Spatially varying trends,
mixed results. Slight decrease in CDD
(Dufek et al., 2008). Tendency to decreased
dryness in much of region, but some
opposite trends and inconsistencies
between studies (Sheffield and Wood,
2008a; Dai, 2011).
Medium confidence: Increases in WD (Silva and
Azevedo, 2008).
Medium confidence: Increases in WN (Silva
and Azevedo, 2008).
Low confidence: Insufficient evidence. Medium confidence: Increase in many
areas, decrease in a few areas, (Alexander
et al., 2006; Haylock et al., 2006b; Santos
and Brito, 2007; Silva and Azevedo, 2008;
Santos et al., 2009).
Low confidence: Spatially varying trends.
Inconsistent trends in CDD (Santos and
Brito, 2007; Dufek et al., 2008; Silva and
Azevedo, 2008; Santos et al., 2009).
Inconsistent trends in dryness (SMA, PDSI)
between studies (Sheffield and Wood,
2008a; Dai, 2011).
Medium confidence: Spatially varying trends.
Increases in WD in some areas, decrease in
others. Decreases in CD in some areas, increase
in others (Rusticucci and Barrucand, 2004;
Vincent et al., 2005; Alexander et al., 2006;
Rusticucci and Renom, 2008; Marengo et al.,
2009b).
Medium confidence: Decreases in CN,
increases in WN (Rusticucci and Barrucand,
2004; Vincent et al., 2005; Alexander et al.,
2006; Rusticucci and Renom, 2008;
Marengo et al., 2009b).
Low confidence: Spatially varying
trends. Some areas increase, others
decrease (Alexander et al., 2006).
Low confidence to medium confidence,
depending on subregion: Medium
confidence: Increase in northern portion of
domain (Alexander et al., 2006; Dufek et
al., 2008; Sugahara et al., 2009; Penalba
and Robledo, 2010). Low confidence in
southern portion of domain: Insufficient
evidence.
Low confidence: Slight increase in dryness,
large variability (Haylock et al., 2006b;
Dufek and Ambrizzi, 2008; Dufek et al.,
2008; Llano and Penalba, 2011). Decrease
in dryness (SMA, PDSI) in much of region
(Sheffield and Wood, 2008a; Dai, 2011).
Medium confidence: Increases in WD in some
areas, decrease in others. Decreases in CD in
some areas, increase in others (Rosenbluth et al.,
1997; Vincent et al., 2005; Alexander et al., 2006).
Medium confidence: Decreases in CN,
increases in WN (Rosenbluth et al., 1997;
Vincent et al., 2005; Alexander et al., 2006).
Low confidence: Insufficient evidence. Medium confidence: Spatially varying
trends. Decrease in many areas, increase in
a few areas (Alexander et al., 2006;
Haylock et al., 2006b).
Low confidence: Overall inconsistent and
spatially varying signal (SMA, PDSI, CDD)
(Dufek et al., 2008; Sheffield and Wood,
2008a; Dai, 2011).
Tmin Heavy Precipitation Dryness
Tmax
Regions
Heat Waves / Warm Spells
Tmin Heavy Precipitation Dryness
Tmax
Regions
Heat Waves / Warm Spells
All South
America
Amazon
(AMZ, 7)
S.E. South
America
(SSA, 10)
W. Coast
South
America
(WSA, 9)
All Asia
N. Asia
(NAS, 18)
N.E. Brazil
(NEB, 8)
Continued next page
D. South America
E. Asia
Table 3-2 (continued)
195
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
All Australia
and New
Zealand
F. Australia/New Zealand
Low confidence to high confidence
depending on region: Insufficient studies
for assessment in N. Australia, likely
decrease in HP in many areas in S.
Australia. See individual regional entries for
assessment basis and details.
High confidence: Overall likely increases in WD,
likely decreases in CD. See individual regional
entries for assessment basis and details.
High confidence: Overall likely decreases in
CN, likely increases in WN. See individual
regional entries for assessment basis and
details.
Low confidence to medium confidence
depending on region: See individual
regional entries for assessment basis
and details.
Medium confidence: Some regions with
dryness decreases, others with dryness
increases. See individual regional entries for
assessment basis and details.
High confidence: Likely increases in WD, likely
decreases in CD. Weaker trends in northwest
(Alexander et al., 2006).
High confidence: Likely decreases in CN,
likely increases in WN (Alexander et al.,
2006; Alexander and Arblaster, 2009).
Low confidence: Insufficient literature
for assessment.
Low confidence: Insufficient studies for
assessment.
Medium confidence: Decrease in dryness
(SMA, PDSI) in northwest since mid-20th
century (Sheffield and Wood, 2008a; Dai,
2011).
High confidence: Very likely increases in WD,
very likely decreases in CD (Alexander et al.,
2006). NZ positive trends vary across country,
related to circulation changes (Chambers and
Griffiths, 2008; Mullan et al., 2008).
High confidence: Very likely decreases in
CN, very likely increases in WN (Alexander
et al., 2006; Alexander and Arblaster, 2009).
General decrease in frosts in NZ but trends
vary across country, related to circulation
changes (Chambers and Griffiths, 2008;
Mullan et al., 2008).
Medium confidence: Increase in warm
spells across southern Australia
(Alexander and Arblaster, 2009).
High confidence: Likely decrease in heavy
precipitation in many areas, especially
where mean precipitation has decreased
(CSIRO, 2007; Gallant et al., 2007;
Alexander and Arblaster, 2009). NZ trends
are positive in western N. and S. Islands
and negative in east of country, and are
strongly correlated with changes in mean
rainfall (Mullan et al., 2008).
Medium confidence:
Increase in dryness
(SMA, PDSI, CDD) in southeastern part and
southwestern tip of Australia since mid-20th
century. Decrease in dryness in central part
of Australia (Alexander et al., 2006;
Sheffield and Wood, 2008a; Dai, 2011).
High confidence: Likely increases in WD, likely
decreases in CD (Alexander et al., 2006).
High confidence: Likely decreases in CN,
likely increases in WN (Alexander et al.,
2006).
Medium confidence: WSDI increase in a
few areas, insufficient evidence
elsewhere (Alexander et al., 2006).
Low confidence: Spatially varying trends.
Increase in a few areas, decrease in a few
areas (Alexander et al., 2006).
Low confidence: Spatially varying trends in
dryness (SMA, PDSI, CDD); partial lack of
coverage in some studies (Alexander et al.,
2006; Sheffield and Wood, 2008a; Dai,
2011).
High confidence: Likely increases in WD, likely
decreases in CD (Alexander et al., 2006; Ding et
al., 2010).
Medium confidence: Decreases in CN,
increases in WN (Alexander et al., 2006).
Medium confidence: Increase in warm
season heat waves in China (Ding et al.,
2010); increase in WSDI in northern
China, but decline in southern China
(Alexander et al., 2006).
Low confidence: Spatially varying trends.
Increase in a few areas, decrease in a few
areas (Alexander et al., 2006).
Medium confidence: Overall tendency for
increased dryness (SMA, PDSI, CDD); few
areas with opposite trends (Alexander et al.,
2006; Sheffield and Wood, 2008a; Dai,
2011).
Medium confidence: Increases in WD, decreases
in CD in northern part of domain (Alexander et
al., 2006).
Low confidence: Insufficient evidence for Malay
Archipelago.
Medium confidence: Decreases in CN,
increases in WN in northern part of domain.
(Alexander et al., 2006).
Low confidence: Insufficient evidence for
Malay Archipelago.
Low confidence: Insufficient evidence. Low confidence: Spatially varying trends
and partial lack of evidence. Some areas
increase, some areas decrease (Alexander
et al., 2006).
Low confidence: Spatially varying trends,
inconsistent trends in dryness (SMA, PDSI)
between studies (Sheffield and Wood,
2008a; Dai, 2011).
Medium confidence: Increase in WD and
decrease in CD (Alexander et al., 2006).
Medium confidence: Decreases in CN,
increases in WN (Alexander et al., 2006).
Low confidence: Insufficient evidence. Low confidence: Mixed signal in India
(Alexander et al., 2006).
Low confidence: Inconsistent signal for
different studies and indices. Decrease in
CDD over India (Alexander et al., 2006).
Increased dryness (SMA, PDSI) in central
India (Sheffield and Wood, 2008a; Dai,
2011).
High confidence: More likely than not decrease
in CD and very likely increase in WD
(Rahimzadeh et al., 2009; Rehman, 2010).
High confidence: Likely decrease in CN and
likely increase in WN (Rehman, 2010).
Medium confidence: WSDI increase
(Alexander et al., 2006).
Medium confidence: Decrease in heavy
precipitation events (Kwarteng et al., 2009;
Rahimzadeh et al., 2009).
Low confidence: Lack of studies for part of
the region; mixed results (Sheffield and
Wood, 2008a; Rahimzadeh et al., 2009).
High confidence: Likely increase in WD and likely
decrease in CD (Alexander et al., 2006).
High confidence: Likely decreases in CN,
likely increases in WN (Alexander et al.,
2006).
Low confidence: Spatially varying trends
(Alexander et al., 2006).
Low confidence: Insufficient evidence. Low confidence: Lack of studies. Tendency
to decreased dryness (PDSI, SMA) in Dai
(2011).
Tmin Heavy Precipitation
Tmax
Regions
Heat Waves / Warm Spells
Dryness
Tmin Heavy Precipitation Dryness
Tmax
Regions
Heat Waves / Warm Spells
Central Asia
(CAS, 20)
E. Asia
(EAS, 22)
S. Asia
(SAS, 23)
W. Asia
(WAS, 19)
Tibetan
Plateau
(TIB, 21)
S.E. Asia
(SEA, 24)
S. Australia/
New Zealand
(SAU, 26)
N. Australia
(NAU, 25)
E. Asia (Continued)
Table 3-2 (continued)
196
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
High confidence to medium confidence
depending on subregion. High confidence:
Likely more frequent, longer, and/or more
intense heat waves and warm spells over all
North American subregions. Medium
confidence in increase in warm spells in part
of Central America, but lack of agreement in
signal of change in heat waves (T06;
Christensen et al., 2007; Karl et al., 2008;
Clark et al., 2010; OS11).
High confidence: Likely more frequent,
longer, and/or more intense heat waves and
warm spells (T06; Christensen et al., 2007;
Meehl et al., 2007b; Karl et al., 2008; Clark
et al., 2010; OS11).
Medium confidence: RCM simulations for
2030–2039 and 2090–2099 consistent with
projected long-term increase in frequency
and/or intensity of HW (Diffenbaugh and
Ashfaq, 2010; Kunkel et al., 2010).
High confidence: Likely more frequent,
longer, and/or more intense heat waves and
warm spells (T06; Christensen et al., 2007;
Karl et al., 2008; Clark et al., 2010; OS11).
Medium confidence: RCM simulations for
2030–2039 and 2090–2099 consistent with
projected long-term increase in frequency
and/or intensity of HW (Diffenbaugh and
Ashfaq, 2010; Kunkel et al., 2010).
Low confidence to high confidence depending on
region and index: Likely increase in HP, including
HPD, HPC and RV20HP, over Canada and Alaska;
Low confidence to medium confidence in the south
(particularly CAM) due to smaller and less
consistent changes, and inconsistencies between
%DP10 (decreases in winter and spring) and other
indices (T06; Christensen et al., 2007; Kharin et al.,
2007; Meehl et al., 2007b; Karl et al., 2008; OS11;
Fig. 3-6).
Low confidence to medium
confidence depending on region.
Medium confidence regarding
increase in CDD, and SMA drought
in Texas and N. Mexico (T06;
SW08b; Fig. 3-10). Low confidence:
Inconsistent change in other regions
(SMA, CDD) (T06; SW08b; Fig. 3-10).
Low confidence to medium confidence depending
on subregion and index: Medium confidence in
increase in HPD/HPC over northern part of domain
(Canada); low confidence due to no signal or
inconsistent signal in HPD/HPC changes over
southern part of domain (T06; Fig. 3-6). Medium
confidence in increase in RV20HP (Fig. 3-7).
Low confidence: Inconsistent signal
in CDD and SMA changes (T06;
SW08b; Fig. 3-10).
Low confidence to medium confidence depending
on index:
Low confidence in changes in HPD/HPC due to
inconsistent or no signal (T06; Fig. 3-6).
Medium confidence in increase in RV20HP (Fig. 3-
7).
Medium confidence: Increase in CDD
and decrease in SMA in southern
part of the domain (SW08b; Fig. 3-
10). Low confidence: inconsistent
signal elsewhere (Fig. 3-10).
High confidence: WN very likely to increase
and CN very likely to decrease (T06;
Christensen et al., 2007; Kharin et al., 2007;
Meehl et al., 2007b; Karl et al., 2008; Fig. 3-
4).
High confidence: WD very likely or likely to
increase and CD very likely or likely to decrease
in all regions (Christensen et al., 2007; Meehl
et al., 2007b; Karl et al., 2008; Fig. 3-3). Very
likely increase in RV20AHD in all regions except
CAM (Fig. 3-5).
Medium confidence: Largest increases in WD in
summer and fall particularly over the United
States; largest decrease in CD in Canada in fall
and winter (OS11).
High confidence: WN very likely or likely to
increase and CN very likely or likely to
decrease depending on subregion (T06;
Christensen et al., 2007; Kharin et al., 2007;
Meehl et al., 2007b; Karl et al., 2008; Fig. 3-
4).
Medium confidence: Largest increase in WN
and decrease in CN in summer, particularly
in the United States (OS11).
High confidence: WD very likely to increase and
CD very likely to decrease in all seasons
(Christensen et al., 2007; Karl et al., 2008;
Clark et al., 2010; Fig. 3-3). Very likely increase
in RV20AHD (Fig. 3-5).
Medium confidence: Overall weaker signal in
spring and winter for both CD and WD (OS11).
RCM simulations for 2030–2039 consistent
with projected long-term increase in WD
(Diffenbaugh and Ashfaq, 2010).
High confidence: WN very likely to increase
and CN very likely to decrease (T06;
Christensen et al., 2007; Kharin et al., 2007;
Meehl et al., 2007b; Karl et al., 2008; Fig. 3-
4).
Medium confidence: Largest WN increases
and CN decreases in summer (OS11).
High confidence: WD very likely to increase and
CD very likely to decrease in all seasons
(Christensen et al., 2007; Karl et al., 2008;
Clark et al., 2010; Fig. 3-3). Very likely increase
in RV20AHD (Fig. 3-5).
Medium confidence: Weaker signal for CD in
spring and winter (OS11). RCM simulations for
2030–2039 consistent with projected long-term
increase in WD (Diffenbaugh and Ashfaq,
2010).
GGG GG
G R GG R
GG
G R GG R
GG
Continued next page
All North
America and
Central
America
W. North
America
(WNA, 3)
Central
North
America
(CNA, 4)
Tmin
[WN = Warm Nights
CN = Cold Nights]
(using late 20th-century extreme values
as reference; see Box 3-1)
Heavy Precipitation (HP)
[HPD = Heavy Precipitation Days, e.g.,
precipitation >95th percentile (p)
%DP10 = Percentage of Days with
Precipitation >10 mm
HPC = Heavy Precipitation Contribution,
generally fraction from precipitation >95th
percentile
RV20HP = 20-year return value of annual
maximum daily precipitation rates] (using late
20th-century extreme values as reference; see
Box 3-1)
Dryness
[CDD = Consecutive Dry Days
SMA = (Simulated) Soil Moisture
Anomalies; see Box 3-3 for
definitions]
Tmax
[WD = Warm Days
CD = Cold Days
RV20AHD: 20-year return value of annual
maximum hottest day]
(using late 20th-century extreme values as
reference; see Box 3-1)
Regions
Heat Waves (HW)/Warm Spells (WS)
(using late 20th-century extreme values
as reference; see Box 3-1)
A. North America and Central America
Table 3-3 | Projected regional changes in temperature and precipitation (including dryness) extremes. See Figure 3-1 for definitions of regions (numbers indicated next to regions’ names). For assessments for small island states, refer
to Box 3-4. Projections are for the end of the 21st century versus the end of the 20th century (e.g., 1961-1990 or 1980-2000 versus 2071-2100 or 2080-2100) and for the A2/A1B emissions scenarios (except if noted otherwise).
Late 20th-century extreme values (generally either 1961-1990 or ~1980-2000) are used as reference (see Box 3-1 for discussion). Codes for the source of modelling evidence: G: multiple GCMs; R: single RCM forced by single GCM;
R: multiple RCMs forced by single GCM; R: multiple RCMs forced by multiple GCMs. T06 stands for Tebaldi et al. (2006), SW08b stands for Sheffield and Wood (2008b), and OS11 stands for Orlowsky and Seneviratne (2011).
197
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Medium confidence: Increase in HPD/HPC in
northern part of domain but no signal or
inconsistent signal in southern part (T06; Fig. 3-6).
Medium confidence in increase in RV20HP (Fig. 3-
7).
Low confidence: Inconsistent signal
in CDD, some consistent decrease in
SMA (SW08b; Fig. 3-10).
High confidence: Likely increase in HPD and HPC
(T06; Fig. 3-6). Likely increase of RV20HP (Fig. 3-7).
Low confidence: Inconsistent signal
in change of CDD and SMA (T06;
SW08b; Fig. 3-10).
High confidence: Likely increase in HPD and HPC
(T06; Fig. 3-6). Likely increase of RV20HP (Fig. 3-7).
Low confidence: Inconsistent signal
in CDD and/or SMA changes (T06;
SW08b; Fig. 3-10).
Low confidence: Lack of agreement between models
and indices regarding changes in %DP10, HPC,
RV20HP, and other HP indicators (Kamiguchi et al.,
2006; T06; Campbell et al., 2011; Figs. 3-6 and 3-7).
Low confidence to medium
confidence depending on region.
Medium confidence: Increased
dryness (CDD, SMA) in Central
America and Mexico; low
confidence in change in dryness
(CDD, SMA) in the extreme south of
region due to inconsistent signal
(Kamiguchi et al., 2006; T06;
Campbell et al., 2011; Fig. 3-10).
High confidence: WD very likely to increase and
CD very likely to decrease in all seasons
(Christensen et al., 2007; Karl et al., 2008;
Clark et al., 2010; Fig. 3-3). Very likely increase
in RV20AHD (Fig. 3-5).
Medium confidence: Largest WD increase in
summer and fall; weaker CD decrease in spring
(OS11). RCM simulations for 2030–2039
consistent with projected long-term increase in
WD (Diffenbaugh and Ashfaq, 2010).
High confidence: WN very likely to increase
and CN very likely to decrease (T06;
Christensen et al., 2007; Kharin et al., 2007;
Meehl et al., 2007b; Karl et al., 2008; Fig. 3-
4).
Medium confidence: Largest WN increases
and CN decreases in summer (OS11).
High confidence: Likely more frequent,
longer, and/or more intense heat waves and
warm spells (T06; Christensen et al., 2007;
Meehl et al., 2007b; Karl et al., 2008; Clark
et al., 2010; OS11)
Medium confidence: RCM simulations for
2030–2039 and 2090–2099 consistent with
projected long-term increase in frequency of
HW (Diffenbaugh and Ashfaq, 2010; Kunkel
et al., 2010).
High confidence: WD very likely to increase and
CD very likely to decrease (Christensen et al.,
2007; Karl et al., 2008; Fig. 3-3). Very likely
increase in RV20AHD (Fig. 3-5).
Medium confidence: Strongest increase in WD
in the fall (OS11).
High confidence: WN very likely to increase
and CN very likely to decrease (T06;
Christensen et al., 2007; Kharin et al., 2007;
Meehl et al., 2007b; Karl et al., 2008; Fig. 3-
4).
High confidence: Likely more frequent and/or
longer heat waves and warm spells (T06;
Christensen et al., 2007; Meehl et al.,
2007b; Karl et al., 2008; OS11).
High confidence: WD very likely to increase and
CD very likely to decrease (Christensen et al.,
2007; Karl et al., 2008; Fig. 3-3). Very likely
increase in RV20AHD (Fig. 3-5).
Medium confidence: Strongest increase of WD
in fall (in summer in Greenland), weakest in
spring, weaker increase of CD in summer
(OS11).
High confidence: WN very likely to increase
and CN very likely to decrease (T06;
Christensen et al., 2007; Kharin et al., 2007;
Meehl et al., 2007b; Fig. 3-4).
High confidence: Likely more frequent and/or
longer heat waves and warm spells (T06;
OS11).
High confidence: WD likely to increase and CD
likely to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: WN likely to increase (T06;
Fig. 3-4) and CN likely to decrease (Fig. 3-4).
Medium confidence to high confidence
depending on region: Likely more frequent,
longer, and/or more intense heat waves and
warm spells in most of the region; medium
confidence in increase in warm spells in part
of Central America, but lack of agreement in
signal of change in heat waves (T06; OS11;
Clark et al., 2010).
G R GG R GG
GGG GG
GGG GG
GGG
G R G R
Continued next page
Tmin
Heavy Precipitation
Dryness
TmaxRegions
Heat Waves / Warm Spells
E. North
America
(ENA, 5)
Alaska/
N.W. Canada
(ALA, 1)
E. Canada,
Greenland,
Iceland
(CGI, 2)
Central
America and
Mexico
(CAM, 6)
A. North America and Central America (Continued)
Table 3-3 (continued)
198
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
High confidence: WD very likely to increase –
largest increases in summer and Central/S.
Europe and smallest in N. Europe (Scandinavia)
(Goubanova and Li, 2007; Kjellström et al.,
2007; Koffi and Koffi, 2008; Fischer and Schär,
2010; Fig. 3-3) and CD very likely to decrease
(Fig. 3-3). Likely increase in RV20AHD (Fig. 3-
5).
Medium confidence: Changes in higher
quantiles of Tmax generally greater than
changes in lower quantiles of Tmax in summer
in Central Europe and Mediterranean
(Diffenbaugh et al., 2007; Kjellström et al.,
2007; Fischer and Schär, 2009, 2010; OS11).
High confidence: CN very likely to decrease
– largest decreases in winter in E. Europe
and Scandinavia (Goubanova and Li, 2007;
Kjellström et al., 2007; Sillmann and
Roeckner, 2008). WN very likely to increase
(T06; Fig. 3-4).
High confidence: Likely more frequent,
longer, and/or more intense heat waves and
warm spells but little change over
Scandinavia (Beniston et al., 2007; Koffi and
Koffi, 2008; Clark et al., 2010; OS11).
Low confidence to high confidence, depending on
region: Likely overall increases in HPD, %DP10, and
RV20HP and decreases in return periods of long (5-
day) and short (1-day) events; strong signals in N.
Europe particularly in winter, but lower confidence
in changes in Central Europe and in particular the
Mediterranean (T06; Beniston et al., 2007; Fowler et
al., 2007a; Sillmann and Roeckner, 2008; Kendon et
al., 2010; Figs. 3-6 and 3-7).
Likely increase in HPC in some regions (Boberg et
al., 2009b; Kendon et al., 2010).
Likely greater changes in extremes than mean in
many regions. Increase in HP intensity (and increase
in HPC) despite decrease in summer mean in some
regions – e.g., Central Europe (Beniston et al., 2007;
Fowler et al., 2007a; Haugen and Iversen, 2008;
May, 2008; Kyselý and Beranová, 2009).
Medium confidence: European area
affected by stronger dryness
(reduced SMA and CDD) with largest
and most consistent changes in
Mediterranean Europe (T06; Burke
and Brown, 2008; May, 2008;
SW08b; Sillmann and Roeckner,
2008; Fig. 3-10).
High confidence: Very likely increase in
frequency of WD, but smaller than in Central
and S. Europe (Fischer and Schär, 2010; Fig. 3-
3). Very likely decrease in CD (Fig. 3-3). Likely
increase in RV20AHD (Fig. 3-5).
Medium confidence: Changes in lower
quantiles of Tmax generally greater than for
changes in higher quantiles of Tmax in fall,
winter, and spring in Scandinavia and
northeastern Europe (OS11).
High confidence: CN very likely to decrease
(Kjellström et al., 2007; Sillmann and
Roeckner, 2008; Fig. 3-4); WN very likely to
increase (T06; Fig. 3-4).
Medium confidence: Changes in lower
quantiles of Tmin generally greater than
changes in higher quantiles of Tmin in
Scandinavia and northeastern Europe
(Kjellström et al., 2007; OS11).
High confidence: Likely more frequent,
longer and/or more intense heat waves and
warm spells, but summer increases smaller
than in S. Europe and little change over
Scandinavia (Beniston et al., 2007; Koffi and
Koffi, 2008; Fischer and Schär, 2010; OS11).
Medium confidence: Some dependency of
projections of changes in HW intensity on
parameterization choice (Clark et al., 2010).
High confidence: Very likely increases in HP
(intensity and frequency) and %DP10 north of 45°N
in winter (Frei et al., 2006; T06; Beniston et al.,
2007; Kendon et al., 2008; Fig. 3-6). Likely increase
in RV20HP (Fig. 3-7).
Medium confidence: No major
changes in dryness (CDD, SMA) in N.
Europe (T06; SW08b; Sillmann and
Roeckner, 2008; Fig. 3-10).
High confidence: Very likely increase in
frequency and intensity of WD (Fischer and
Schär, 2010; Fig. 3-3) and decrease in
frequency of CD (Fig. 3-3). Very likely increase
in RV20AHD (Fig. 3-5).
Medium confidence: Changes in higher
quantiles of Tmax much larger than changes in
lower quantiles of Tmax in summer; results in
very large increase in Tmax variability
(Diffenbaugh et al., 2007; Kjellström et al.,
2007; Fischer and Schär, 2009, 2010; OS11).
High confidence: CN very likely to decrease
(Goubanova and Li, 2007; Kjellström et al.,
2007; Sillmann and Roeckner, 2008); WN
very likely to increase (T06; Fig. 3-4).
High confidence: Likely more frequent,
longer, and/or more intense heat waves and
warm spells (Beniston et al., 2007; Koffi and
Koffi, 2008; Clark et al., 2010; Fischer and
Schär, 2010; OS11).
Medium confidence: Some dependency of
projections of changes in HW intensity on
parameterization choice (Clark et al., 2010).
High confidence: Likely increases in HP (intensity
and frequency) in large part of the region in winter
(Frei et al., 2006; Beniston et al., 2007; Kendon et
al., 2008; Kyselý and Beranová, 2009; Fig. 3-6).
Likely increase in RV20HP (Fig. 3-7).
Medium confidence: Inconsistent evidence for
summer: increase in HP in summer evident in RCMs
(Christensen and Christensen, 2003; Frei et al.,
2006) versus no signal in GCMs (Fig. 3-6).
Medium confidence: Increase in
dryness (CDD, SMA) in Central
Europe (Seneviratne et al., 2006a;
T06; Fig. 3-10).
Medium confidence: Increase in
short-term droughts (SW08b).
G R G R G R G R G R
G R
G R G R G R G
G R
G R G R G R G R
Continued next page
Tmin
Heavy Precipitation
Dryness
Tmax
Regions
Heat Waves / Warm Spells
All Europe and
Mediterranean
Region
N. Europe
(NEU, 11)
Central
Europe
(CEU, 12)
B. Europe and Mediterranean Region
Table 3-3 (continued)
199
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Low confidence: Lack of agreement/signal in %DP10
and other HP indicators for the region as a whole
(T06; Fig. 3-6). Some model agreement in increase in
RV20HP (Fig. 3-7). Some evidence of increased HP
intensity in the S.E. (Hewitson and Crane, 2006;
Rocha et al., 2008; Shongwe et al., 2009).
Low confidence to high confidence depending on
region: Inconsistent change or no signal in HP
indicators across much of continent (T06; Fig. 3-6,
Fig. 3-7). Strongest and most consistent signal is
likely increase in HP in E. Africa (T06; Figs. 3-6 and
3-7; Shongwe et al., 2011).
Low confidence to medium confidence depending
on subregion: Medium confidence in slight or no
change in HP indicators in most of region; Low
confidence due to low model agreement in northern
part of region (T06; Figs. 3-6 and 3-7).
Low confidence: Inconsistent signal
of change in CDD and SMA (T06;
Fig. 3-10).
High confidence: Likely increase in HP indicators
(T06; Fig. 3-6; Fig. 3-7; Shongwe et al., 2011).
Medium confidence: Decreasing
dryness in large part of region,
especially based on change in SMA,
and partly also in CDD (T06; Fig. 3-
10; Shongwe et al., 2011).
Medium confidence: Increase in
dryness (CDD, SMA), except eastern
part (T06; Shongwe et al., 2009; Fig.
3-10). Consistent increase in area of
drought (Burke and Brown, 2008).
Low confidence: Low agreement/no signal in
%DP10, RV20HP, and other HP indicators (T06;
Figs. 3-6 and 3-7).
Low confidence: Inconsistent signal
of change in CDD and SMA (T06;
Fig. 3-10).
High confidence: WD likely to increase and CD
likely to decrease in all regions (Fig. 3-3). Likely
increase in RV20AHD in all regions (Fig. 3-5).
Medium confidence: Increase in WD largest in
summer and fall (OS11).
High confidence: WN likely to increase (T06;
Kharin et al., 2007; Fig. 3-4) and CN likely to
decrease (Fig. 3-4).
High confidence: Likely more frequent and/or
longer heat waves and warm spells (T06;
OS11).
High confidence: WD likely to increase and CD
likely to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: WN likely to increase (T06;
Fig. 3-4) and CN likely to decrease (Fig. 3-4).
High confidence: Likely
more frequent and/or
longer heat waves and warm spells (T06;
OS11).
High confidence: WD likely to increase and CD
likely to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: WN likely to increase (T06;
Fig. 3-4) and CN likely to decrease (Fig. 3-4).
High confidence: Likely more frequent and/or
longer heat waves and warm spells (T06;
OS11).
High confidence: WD likely to increase and CD
likely to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: WN likely to increase (T06;
Fig. 3-4) and CN likely to decrease (Fig. 3-4).
High confidence: Likely more frequent and/or
longer heat waves and warm spells (T06;
OS11).
High confidence: WD likely to increase and CD
likely to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: WN likely to increase (T06;
Fig. 3-4) and CN likely to decrease (Fig. 3-4).
High confidence: Likely more frequent and/or
longer heat waves and warm spells (T06;
OS11).
G R G R G R G R G R
GGG GG
GGG GG
GGG GG
GGG GG
GGG GG
E. Africa
(EAF, 16)
Continued next page
Tmin Heavy Precipitation DrynessTmaxRegions Heat Waves / Warm Spells
Tmin Heavy Precipitation DrynessTmaxRegions Heat Waves / Warm Spells
S. Europe and
Mediterranean
(MED, 13)
All Africa
W. Africa
(WAF, 15)
S. Africa
(SAF, 17)
Sahara
(SAH, 14)
B. Europe and Mediterranean Region (Continued)
C. Africa
High confidence: CN very likely to decrease
(Goubanova and Li, 2007; Kjellström et al.,
2007; Sillmann and Roeckner, 2008).
High confidence: WN very likely to increase
(T06; Sillmann and Roeckner, 2008;
Giannakopoulos et al., 2009; Fig. 3-4).
High confidence: Tropical nights very likely to
increase (Sillmann and Roeckner, 2008).
Number of days with combined hot summer
days (>35°C) and tropical nights (>20°C)
very likely to increase (Fischer and Schär,
2010).
Medium confidence: Changes in higher
quantiles of Tmin generally greater than
changes in lower quantiles of Tmin in
summer in the Mediterranean (Diffenbaugh
et al., 2007; OS11).
High confidence: Likely more frequent
and/or longer heat waves and warm spells
(also increases in intensity); likely largest
increases in S.W., S., and E. (Beniston et al.,
2007; Diffenbaugh et al., 2007; Koffi and
Koffi, 2008; Giannakopoulos et al., 2009;
Clark et al., 2010; Fischer and Schär, 2010;
OS11).
Low confidence: Inconsistent change in HP intensity
and %D10, depends on region and season; increase
in HP intensity in all seasons except summer over
parts of the region, but decrease in other parts, e.g.,
Iberian Peninsula (T06; Goubanova and Li, 2007;
Giorgi and Lionello, 2008; Giannakopoulos et al.,
2009; Fig. 3-6). Low confidence in changes in
RV20HP (Fig. 3-7).
Medium confidence: Increase in
dryness (CDD, SMA) in
Mediterranean (T06; Beniston et al.,
2007; SW08b; Sillmann and
Roeckner, 2008; Giannakopoulos et
al., 2009; Fig. 3-10). Consistent
increase in area of drought (Burke
and Brown, 2008).
High confidence: Very likely increase in
frequency and intensity of WD (Fischer and
Schär, 2009, 2010; Giannakopoulos et al.,
2009; Fig. 3-3) and decrease in CD (Fig. 3-4).
Very likely increase in RV20AHD (Fig. 3-5).
High confidence: Number of days with
combined hot summer days (>35°C) and
tropical nights (>20°C) very likely to increase
(Fischer and Schär, 2010).
Medium confidence: Changes in higher
quantiles of Tmax greater than changes in
lower quantiles of Tmax in summer
(Diffenbaugh et al., 2007; Kjellström et al.,
2007; Fischer and Schär, 2009, 2010; OS11).
Low confidence to medium
confidence depending on region: Low
confidence in most regions, medium
confidence of increase in dryness
(CDD, SMA) in southern Africa except
eastern part (T06; SW08b; Fig. 3-10).
Table 3-3 (continued)
200
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Medium confidence: Increase of %DP10,
precipitation intensity, precip >95th p, and RV20HP
in the northern portion. Low confidence in the
southern portion: inconsistent change or no signal
in HP indicators (Kamiguchi et al., 2006; T06;
Marengo et al., 2009a; Nunez et al., 2009; Figs. 3-6
and 3-7).
High confidence: WD likely to increase and CD
likely to decrease in all regions (Fig. 3-3). Likely
increase in RV20AHD in all regions (Fig. 3-5).
High confidence: WN likely to increase (T06;
Kharin et al., 2007; Marengo et al., 2009a;
Fig. 3-4) and CN likely to decrease (Marengo
et al., 2009a; Fig. 3-4).
Medium confidence to high confidence
depending on region: Likely more frequent
and/or longer heat waves and warm spells
on annual time scale; only medium
confidence in increase in HW duration in
southeastern South America (T06; OS11).
Low confidence to medium confidence depending
on region: Inconsistent sign of change in HP
indicators in some regions; some regions with
model agreement (T06; Figs. 3-6 and 3-7).
Low confidence to medium
confidence depending on region:
Inconsistent signal except for
dryness increase (CDD and SMA) in
northeastern Brazil (SW08b; Fig. 3-
10).
High confidence: WD likely to increase and CD
likely to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: Very likely increase in WN
(T06; Marengo et al., 2009a; Fig. 3-4) and
CN likely to decrease (Fig. 3-4).
High confidence: Likely more frequent and/or
longer heat waves and warm spells based
on multi-model assessments (T06, OS11);
non-significant signal in single-model and
single-index study by Uchiyama et al.
(2006).
Medium confidence: Tendency to increase in precip
>95th p and in RV20HP but less consistent increase
for some other HP indicators and single studies
(Kamiguchi et al., 2006; T06; Marengo et al., 2009a;
Sorensson et al., 2010; Figs. 3-6 and 3-7).
Low confidence: Inconsistent signal
in SMA (SW08b; Fig. 3-10) and
inconsistent signal in CDD
(Kamiguchi et al., 2006; T06;
Marengo et al., 2009a; Sorensson et
al., 2010; Fig. 3-10).
High confidence: WD likely to increase and CD
likely
to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: Likely increase of WN (T06;
Marengo et al., 2009a; Fig. 3-4) and CN
likely to decrease (Fig. 3-4).
High confidence: Likely more frequent and/or
longer heat waves and warm spells (T06;
OS11).
Low confidence: Slight or no change in %DP10 and
other HP indicators (Kamiguchi et al., 2006; T06;
Marengo et al., 2009a; Sorensson et al., 2010; Figs.
3-6 and 3-7).
Medium confidence: Dryness
increase (CDD, SMA) (Kamiguchi et
al., 2006; T06; SW08b; Marengo et
al., 2009a; Sorensson et al., 2010;
Fig. 3-10).
High confidence: WD likely to increase and CD
likely to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: Very likely increase in WN
(T06; Marengo et al., 2009a; Fig. 3-4) and
CN likely to decrease (Fig. 3-4).
Medium confidence: Tendency for more
frequent and/or longer heat waves and
warm spells on annual time scale; weak
signal compared to other regions but robust
sign (increase) across majority of models
(T06; OS11).
Low confidence: Inconsistent signal
in SMA (SW08b).
Low confidence: Inconsistent signal
in CDD (Kamiguchi et al., 2006; T06;
Marengo et al., 2009a; Sorensson et
al., 2010; Fig. 3-10).
High confidence: WD likely to increase and CD
likely to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: Likely increase in WN (T06;
Marengo et al., 2009a; Fig. 3-4) and CN
likely to decrease (Fig. 3-4).
High confidence: Likely more frequent and/or
longer heat waves and warm spells on
annual time scale (T06; OS11).
Medium confidence: Increase in %DP10, precip
>95th p and other HP indicators in the tropics. Low
confidence for the extratropics (Kamiguchi et al.,
2006; T06; Marengo et al., 2009a; Sorensson et al.,
2010; Fig. 3-6). Low confidence in changes in
RV20HP (Fig. 3-7).
Low confidence: Inconsistent
changes in SMA across domain (Fig.
3-10), CDD decrease in the tropics
and increase in the extratropics
(Kamiguchi et al., 2006; T06;
Marengo et al., 2009a; Sorensson et
al., 2010; Fig. 3-10).
Medium confidence: CDD increase
and SMA decrease in southwest SA
(SW08b; Fig. 3-10).
GG R GG R G R
GG R GGG
GG R GG R G R
GG R GG R G R
GG R GG R G R
N.E. Brazil
(NEB, 8)
Amazon
(AMZ, 7)
All South
America
S.E. South
America
(SSA, 10)
W. Coast
South
America
(WSA, 9)
Continued next page
Tmin
Heavy Precipitation
Dryness
Tmax
Regions
Heat Waves / Warm Spells
D. South America
Table 3-3 (continued)
201
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
High confidence: WD likely to increase and CD
likely to decrease in all regions (Fig. 3-3). Likely
increase in RV20AHD in all regions (Fig. 3-5).
High confidence: WN likely to increase (T06;
Kharin et al., 2007; Fig. 3-4) and CN likely to
decrease (Fig. 3-4).
Low confidence to high confidence
depending on region:
Likely more frequent and/or longer heat
waves and warm spells in most regions
(continental) on the annual time scale; low
confidence due to inconsistent signal in
Indonesia, Philippines, Malaysia, Papua New
Guinea, and neighboring islands for some
HW indices (T06; OS11).
Low confidence to high confidence depending on
region and index: High confidence regarding likely
increase in HP in N. Asia; medium confidence
regarding increase in HP in S.E. Asia, E. Asia, and
Tibetan Plateau; low confidence regarding increase
in HP in S. and W. Asia (T06; Figs. 3-6 and 3-7).
Low confidence: Inconsistent change
in CDD and SMA between models in
large part of domain (T06; Fig. 3-
10).
High confidence: WD likely to increase and CD
likely to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: WN likely to increase (T06;
Fig. 3-4) and CN likely to decrease (Fig. 3-4).
High confidence: Likely more frequent and/or
longer heat waves and warm spells (T06;
OS11).
High confidence: Likely increase in HP (T06; Figs. 3-
6 and 3-7) including more frequent and intense HPD
over most regions (Emori and Brown, 2005;
Kamiguchi et al., 2006).
Low confidence: Inconsistent change
in dryness (CDD, SMA) between
models in large part of domain (T06;
SW08b; Fig. 3-10).
High confidence: WD likely to increase and CD
likely to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: WN likely to increase (T06;
Fig. 3-4) and CN likely to decrease (Fig. 3-4).
High confidence: Likely more frequent,
longer and/or more intense heat waves and
warm spells (T06; Clark et al., 2010; OS11).
Low confidence: Inconsistent signal in models
regarding changes in HP (T06; Fig. 3-6). Some
model consistency in projections of RV20HP (Fig. 3-
7).
Low confidence: Inconsistent signals
across indices (CDD, SMA) (T06;
SW08b; Fig. 3-10).
High confidence: WD likely to increase and CD
likely to decrease across the region (Clark et
al., 2010; Fig. 3-3), including in Korea (Boo et
al., 2006; Im and Kwon, 2007; Im et al., 2008,
2011; Koo et al., 2009). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: WN likely to increase (T06;
Fig. 3-4) and CN likely to decrease (Fig. 3-4),
including in Korea (Boo et al., 2006; Koo et
al., 2009; Im et al., 2011).
High confidence: Likely more frequent,
longer, and/or more intense heat waves and
warm spells (T06; Clark et al., 2010; OS11).
Medium confidence: Increases in HP (less consistent
in %DP10 than other indicators) across the region
(T06; Figs. 3-6 and 3-7), including increase in Japan
and Korea (Emori and Brown, 2005; Kimoto et al.,
2005; Boo et al., 2006; Kamiguchi et al., 2006;
Kusunoki and Mizuta, 2008; Kitoh et al., 2009; Su et
al., 2009; Kim et al., 2010; Im et al., 2011).
Low confidence: Inconsistent signal
across indices (CDD, SMA) (T06;
SW08b; Fig. 3-10).
High confidence: WD likely to increase and CD
likely to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: WN likely to increase (T06;
Fig. 3-4) and CN likely to decrease (Fig. 3-4).
Low confidence to high confidence
depending on subregion and index: Likely
more frequent and/or longer heat waves and
warm spells on the annual time scale over
continental areas; Low confidence in
changes in some HW indices but model
agreement of increases in a WS index in
Indonesia, Philippines, Malaysia, Papua New
Guinea, and neighboring islands (T06;
OS11).
Medium confidence: Inconsistent signal in change
of %DP10 across models (T06; Fig. 3-6) but more
frequent and intense HPD, including increase in
RV20HP, suggested by other indicators over most
regions especially non-continental parts (Emori and
Brown, 2005; Kamiguchi et al., 2006; Figs. 3-6 and
3-7).
Low confidence: Inconsistent signal
of change in CDD and/or SMA (T06;
SW08b; Fig. 3-10).
High confidence
: WD likely to increase and CD
likely to decrease (Kumar et al., 2006;
Rajendran and Kitoh, 2008; Fig. 3-3). Likely
increase in RV20AHD (Fig. 3-5).
High confidence: WN likely to increase (T06;
Fig. 3-4) and CN likely to decrease (Fig. 3-4).
Medium confidence: Extreme nighttime
temperature warms faster than daytime
(Kumar et al., 2006).
High confidence: Likely more frequent and/or
longer heat waves and warm spells on
annual time scale (T06; OS11).
Medium confidence: Some dependency of
magnitude of signal on index choice (OS11).
Low confidence: Slight or no increase in %DP10
(T06; Fig. 3-6). Some model consistency regarding
increase in RV20HP (Fig. 3-7).
Low confidence: More frequent and intense HPD
over parts of S. Asia (Emori and Brown, 2005;
Kamiguchi et al., 2006; Kharin et al., 2007;
Rajendran and Kitoh, 2008).
Low confidence: Inconsistent signal
of change in CDD and SMA (T06;
SW08b; Fig. 3-10).
GGG GG
GG G GG
GGG GG
G R
G R GG R G
GGG GG
G R G R GG R G
Central Asia
(CAS, 20)
N. Asia
(NAS, 18)
All Asia
E. Asia
(EAS, 22)
S.E. Asia
(SEA, 24)
S. Asia
(SAS, 23)
Continued next page
Tmin Heavy Precipitation DrynessTmaxRegions Heat Waves / Warm Spells
E. Asia
Table 3-3 (continued)
202
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
Low confidence: Lack of agreement regarding sign
of change for different models and different indices,
and spatial variations in signal (T06; Figs. 3-6 and 3-
7).
Low confidence: HPD tend to increase in E. and
decrease in W. half of country – but considerable
inter-model inconsistencies; HPC tends to increase
everywhere – but considerable inter-model
inconsistencies (Alexander and Arblaster, 2009).
High confidence: WD very likely to increase and
CD very likely to decrease in all regions (CSIRO,
2007; Mullan et al., 2008; Fig. 3-3). Very likely
increase in RV20AHD (Fig. 3-5).
High confidence: WN very likely to increase
everywhere (T06; Kharin et al., 2007;
Alexander and Arblaster, 2009; Fig. 3-4) and
CN very likely to decrease (Fig. 3-4).
Medium confidence: WN increase
everywhere. Largest increases in WN in N.
compared with S. and most consistent
changes in inland regions (Alexander and
Arblaster, 2009).
High confidence: Likely more frequent and/or
longer heat waves and warm spells (T06;
OS11; Alexander and Arblaster, 2009).
Medium confidence: Strongest increases in
HW duration in N.W. and most consistent
increases inland (Alexander and Arblaster,
2009).
Low confidence to medium
confidence depending on region:
Models agree on increase in CDD in
S. Australia, but inconsistent signal
over most of S. Australia in SMA;
inconsistent signal in CDD and SMA
in N. Australia (T06; SW08b; Fig. 3-
10). Strongest CDD increases in W.
half of Australia (Alexander and
Arblaster, 2009). Inconsistent
change in area of drought
depending on index used (Burke and
Brown, 2008).
High confidence: WD very likely to increase and
CD very likely to decrease (CSIRO, 2007; Fig. 3-
3). Very likely increase in RV20AHD (Fig. 3-5).
High confidence: WN very likely to increase
(T06; Alexander and Arblaster, 2009; Fig. 3-
4) and CN very likely to decrease (Fig. 3-4).
Medium confidence: Changes larger than in
S. Australia (Alexander and Arblaster, 2009).
High confidence: Likely more frequent and/or
longer heat waves and warm spells (T06;
Alexander and Arblaster, 2009; OS11).
Medium confidence: Strongest increases in
N.W. and most consistent increases inland
(Alexander and Arblaster, 2009).
Low confidence: Lack of agreement regarding sign
of change for different models and different indices
(T06; Fig. 3-6, Fig. 3-7).
Low confidence: Inconsistent signal
in CDD and SMA (T06; SW08b; Fig.
3-10).
High confidence: WD very likely to increase and
CD very likely to decrease (CSIRO, 2007; Fig. 3-
3). Very likely increase in RV20AHD (Fig. 3-5).
Low confidence to medium confidence:
Strongest New Zealand increases in WD in
North Island and largest decreases in frost days
in South Island (Mullan et al., 2008).
High confidence: WN very likely to increase
(T06; Alexander and Arblaster, 2009; Fig. 3-
4) and CN very likely to decrease (Fig. 3-4).
Medium confidence: Changes smaller than
in N. Australia (Alexander and Arblaster,
2009).
High confidence: Likely more frequent and/or
longer heat waves and warm spells (T06;
Alexander and Arblaster, 2009; OS11).
Medium confidence: Most consistent
increases inland (Alexander and Arblaster,
2009).
Low confidence: Lack of agreement regarding sign
of change for different models and different indices,
and spatial variations in signal (T06; Fig. 3-6, Fig. 3-
7).
Low confidence to medium confidence: In New
Zealand, increase in HP events at most locations
(Mullan et al., 2008; Carey-Smith et al., 2010).
Medium confidence: Models agree
on increase in CDD in southern
Australia including S.W. (T06;
Alexander and Arblaster, 2009; Fig.
3-10), but inconsistent signal in
SMA over most of the region, slight
decrease in S.W. (SW08b; Fig. 3-10).
High confidence: WD likely to increase and CD
likely to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: WN likely to increase (T06;
Fig. 3-4) and CN likely to decrease (Fig. 3-4).
High confidence: Likely more frequent,
longer and/or more intense heat waves and
warm spells (T06; Clark et al., 2010; OS11).
Low confidence: Inconsistent signal of change in HP
(T06; Fig. 3-6; Fig. 3-7).
Low confidence: Inconsistent signal
of change in CDD and SMA (T06;
SW08b; Fig. 3-10).
High confidence: WD likely to increase and CD
likely to decrease (Fig. 3-3). Likely increase in
RV20AHD (Fig. 3-5).
High confidence: WN likely to increase (T06;
Fig. 3-4) and CN likely to decrease (Fig. 3-4).
High confidence: Likely more frequent,
longer, and/or more intense heat waves and
warm spells (T06; Clark et al., 2010; OS11).
Medium confidence: Increase in HP (T06; Figs. 3-6
and 3-7).
Low confidence: Inconsistent signal
of change in CDD (T06; SW08b; Fig.
3-10).
GG G GG
GG G GG
GGG GG
GGG GG
G R GG G R G
Tmin Heavy Precipitation DrynessTmaxRegions Heat Waves / Warm Spells
Tmin Heavy Precipitation DrynessTmaxRegions Heat Waves / Warm Spells
All Australia
and New
Zealand
Tibetan
Plateau
(TIB, 21)
W. Asia
(WAS, 19)
N. Australia
(NAU, 25)
S. Australia/
New Zealand
(SAU, 26)
Medium confidence: Some dependency of
magnitude of signal on index choice (OS11).
F. Australia/New Zealand
E. Asia (Continued)
Table 3-3 (continued)
203
References
A digital library of non-journal-based literature cited in this chapter that
may not be readily available to the public has been compiled as part of
the IPCC review and drafting process, and can be accessed via either the
IPCC Secretariat or IPCC Working Group II web sites.
Abbs, D., and T. Rafter, 2008: The Effect of Climate Change on Extreme Rainfall
Events in the Western Port Region. CSIRO Marine and Atmospheric Research,
Aspendale, Australia, 23 pp.
Abeysirigunawardena, D.S. and I.J. Walker, 2008: Sea level responses to climatic
variability and change in northern British Columbia. Atmosphere-Ocean, 46(3),
277-296.
Abeysirigunawardena, D.S., E. Gilleland, D. Bronaugh, and W. Wong, 2009:
Extreme wind regime responses to climate variability and change in the inner
south coast of British Columbia, Canada. Atmosphere-Ocean, 47(1), 41-62.
Adams, P.N., D.L. Imnan, and N.E. Graham, 2008: Southern California deep-water
wave climate: Characterization and application to coastal processes. Journal of
Coastal Research, 24(4), 1022-1035.
Aguilar, E., T.C. Peterson, P. Ramírez Obando, R. Frutos, J.A. Retana, M. Solera, J. Soley,
I. González García, R.M. Araujo, A. Rosa Santos, V.E. Valle, M. Brunet, L. Aguilar,
L. Álvarez, M. Bautista, C. Castañón, L. Herrara, E. Ruano, J.J. Sinay, E. Sánchez,
G.I. Hernández Oviedo, F. Obed, J.E. Salgado, J.L. Vázquez, M. Baca, M. Gutiérrez,
C. Centella, J. Espinosa, D. Martinez, B. Olmedo, C.E. Ojeda Espinoza, R. Núñez,
M. Haylock, H. Benavides, and R. Mayorga, 2005: Changes in precipitation and
temperature extremes in Central America and northern South America, 1961-
2003. Journal of Geophysical Research – Atmospheres, 110, D23107.
Aguilar, E., A. Aziz Barry, M. Brunet, L. Ekang, A. Fernandes, M. Massoukina, J. Mbah,
A. Mhanda, D.J. do Nascimento, T.C. Peterson, M. Thamba Umba, M. Tomou, and
X. Zhang, 2009: Changes in temperature and precipitation extremes in western
central Africa, Guinea Conakry, and Zimbabwe, 1955-2006. Journal of
Geophysical Research – Atmospheres, 114, D02115.
Aizen, V.B., V.A. Kuzmichenok, A.B. Surazakov, and E.M. Aizen, 2007: Glacier changes
in the Tien Shan as determined from topographic and remotely sensed data.
Global and Planetary Change, 56(3-4), 328-340.
Alexander, L.V. and J.M. Arblaster, 2009: Assessing trends in observed and modelled
climate extremes over Australia in relation to future projections. International
Journal of Climatology, 29(3), 417-435.
Alexander, L.V. and S. Power, 2009: Severe storms inferred from 150 years of sub-
daily pressure observations along Victoria’s “Shipwreck Coast.” Australian
Meteorological and Oceanographic Journal, 58(2), 129-133.
Alexander, L.V., X. Zhang, T.C. Peterson, J. Caesar, B. Gleason, A.M.G. Klein Tank, M.
Haylock, D. Collins, B. Trewin, F. Rahimzadeh, A. Tagipour, K.R. Kumar, J.
Revadekar, G. Griffiths, L. Vincent, D.B. Stephenson, J. Burn, E. Aguilar, M. Brunet,
M. Taylor, M. New, P. Zhai, M. Rusticucci, and J.L. Vazquez-Aguirre, 2006: Global
observed changes in daily climate extremes of temperature and precipitation.
Journal of Geophysical Research – Atmospheres, 111, D05109.
Alexander, L.V., P. Uotila, and N. Nicholls, 2009: Influence of sea surface temperature
variability on global temperature and precipitation extremes. Journal of
Geophysical Research – Atmospheres, 114, D18116.
Alexander, L.V., X.L. Wang, H. Wan, and B. Trewin, 2011: Significant decline in
storminess over southeast Australia since the late 19th century. Australian
Meteorological and Oceanographic Journal, 61(1), 23-30.
Ali, A. and T. Lebel, 2009: The Sahelian standardized rainfall index revisited.
International Journal of Climatology, 29(12), 1705-1714.
Allamano, P., P. Claps, and F. Laio, 2009: Global warming increases flood risk in
mountainous areas. Geophysical Research Letters, 36
, L24404.
Allan, J.C. and P.D. Komar, 2006: Climate controls on US West Coast erosion processes.
Journal of Coastal Research, 22(3), 511-529.
Allan, R., S. Tett, and L. Alexander, 2009: Fluctuations in autumn-winter severe storms
over the British Isles: 1920 to present. International Journal of Climatology,
29(3), 357-371.
Allan, R.P. and B.J. Soden, 2008: Atmospheric warming and the amplification of
precipitation extremes. Science, 321(5895), 1481-1484.
Allen, M., 2003: Liability for climate change: will it ever be possible to sue anyone
for damaging the climate? Nature, 421(6926), 891-892.
Allen, M.R., D.J. Frame, C. Huntingford, C.D. Jones, J.A. Lowe, M. Meinshausen, and
N. Meinshausen, 2009: Warming caused by cumulative carbon emissions
towards the trillionth tonne. Nature, 458(7242), 1163-1166.
Allen, S.K., S.C. Cox, and I.F. Owens, 2011: Rock-avalanches and other landslides in
the central Southern Alps of New Zealand: A regional assessment of possible
climate change impacts. Landslides, 8(1), 33-48.
An, S.I., A. Timmermann, L. Bejarano, F.F. Jin, F. Justino, Z. Liu, and A.W. Tudhope, 2004:
Modeling evidence for enhanced El Niño-Southern Oscillation amplitude during
the Last Glacial Maximum. Paleoceanography, 19, PA4009.
Anders, I., and B. Rockel, 2009: The influence of prescribed soil type distribution on
the representation of present climate in a regional climate model. Climate
Dynamics, 33(2-3), 177-186.
Andersen, O.B., S.I. Seneviratne, J. Hinderer, and P. Viterbo, 2005: GRACE-derived
terrestrial water storage depletion associated with the 2003 European heat
wave. Geophysical Research Letters, 32, L18405.
Anderson, C.J., R.W. Arritt, E.S. Takle, Z.T. Pan, W.J. Gutowski, F.O. Otieno, R. da Silva,
D. Caya, J.H. Christensen, D. Luthi, M.A. Gaertner, C. Gallardo, F. Giorgi, S.Y. Hong,
C. Jones, H.M.H. Juang, J.J. Katzfey, W.M. Lapenta, R. Laprise, J.W. Larson, G.E.
Liston, J.L. McGregor, R.A. Pielke, J.O. Roads, and J.A. Taylor, 2003: Hydrological
processes in regional climate model simulations of the central United States
flood of June-July 1993. Journal of Hydrometeorology, 4(3), 584-598.
Andrade, C., H.O. Pires, R. Taborda, and M.C. Freitas, 2007: Projecting future changes
in wave climate and coastal response in Portugal by the end of the 21st century.
Journal of Coastal Research, SI50, 253-257.
Andreadis, K.M. and D.P. Lettenmaier, 2006: Trends in 20th century drought over the
continental United States. Geophysical Research Letters, 33, L10403.
Andreadis, K.M., E.A. Clark, A.W. Wood, A.F. Hamlet, and D.P. Lettenmaier, 2005:
Twentieth-century drought in the conterminous United States. Journal of
Hydrometeorology, 6(6), 985-1001.
Arblaster, J.M., G.A. Meehl, and D.J. Karoly, 2011: Future climate change in the
Southern Hemisphere: Competing effects of ozone and greenhouse gases.
Geophysical Research Letters, 38, L02701.
Arriaga-Ramírez
, S. and T. Cavazos, 2010: Regional trends of daily precipitation
indices in Northwest Mexico and the Southwest United States. Journal of
Geophysical Research – Atmospheres, 115, D14111.
Aryal, S.K., B.C. Bates, E.D. Campbell, Y. Li, M.J. Palmer, and N.R. Viney, 2009:
Characterizing and modeling temporal and spatial trends in rainfall extremes.
Journal of Hydrometeorology, 10(1), 241-253.
Ashfaq, M., Y. Shi, W.W. Tung, R.J. Trapp, X. Gao, S.J. Pal, and N.S. Diffenbaugh, 2009:
Suppression of south Asian summer monsoon precipitation in the 21st century.
Geophysical Research Letters, 36, L01704.
Asokan, S.M., and D. Dutta, 2008: Analysis of water resources in the Mahanadi River
Basin, India under projected climate conditions. Hydrological Processes,
22(18), 3589-3603.
Augustinus, P., 1995: Rock mass strength and the stability of some glacial valley
slopes. Zeitschrift Für Geomorphologie, 39(1), 55-68.
Aunan, K., and B. Romstad, 2008: Strong coasts and vulnerable communities:
Potential implications of accelerated sea-level rise for Norway. Journal of
Coastal Research, 24(2), 403-409.
Bader, J., M.D.S. Mesquita, K.I. Hodges, N. Keenlyside, S. Østerhus, and M. Miles,
2011: A review on Northern Hemisphere sea-ice, storminess and the North
Atlantic Oscillation: Observations and projected changes. Atmospheric
Research, 101(4), 809-834, doi:10.1016/j.atmosres.2011.1004.1007.
Bajracharya, S.R. and P. Mool, 2009: Glaciers, glacial lakes and glacial lake outburst
floods in the Mount Everest region, Nepal. Annals of Glaciology, 50(53), 81-86.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
204
Ballester, J., F. Giorgi, and X. Rodo, 2010: Changes in European temperature
extremes can be predicted from changes in PDF central statistics. Climatic
Change, 98(1-2), 277-284.
Barla, G., F. Dutto, and G. Mortara, 2000: Brenva glacier rock avalanche of 18 January
1997 on the Mount Blanc range, northwest Italy. Landslide News, 13, 2-5.
Barnett, D.N., S.J. Brown, J.M. Murphy, D.M.H. Sexton, and M.J. Webb, 2006:
Quantifying uncertainty in changes in extreme event frequency in response to
doubled CO
2
using a large ensemble of GCM simulations. Climate Dynamics,
26(5), 489-511.
Barnett, T.P., D.W. Pierce, H.G. Hidalgo, C. Bonfils, B.D. Santer, T. Das, G. Bala, A.W.
Wood, T. Nozawa, A.A. Mirin, D.R. Cayan, and M.D. Dettinger, 2008: Human-
induced changes in the hydrology of the western United States. Science,
319(5866), 1080-1083.
Barring, L. and K. Fortuniak, 2009: Multi-indices analysis of southern Scandinavian
storminess 1780-2005 and links to interdecadal variations in the NW Europe-
North Sea region. International Journal of Climatology, 29(3), 373-384.
Barring, L., and H. von Storch, 2004: Scandinavian storminess since about 1800.
Geophysical Research Letters, 31, L20202.
Barring, L., T. Holt, M.L. Linderson, M. Radziejewski, M. Moriondo, and J.P. Palutikof,
2006: Defining dry/wet spells for point observations, observed area averages,
and regional climate model gridboxes in Europe. Climate Research, 31(1), 35-49.
Barriopedro, D., E.M. Fischer, J. Luterbacher, R.M. Trigo, and R. García-Herrera, 2011:
The hot summer of 2010: Redrawing the temperature record map of Europe.
Science, 332(6026), 220-224.
Barros, V., L. Chamorro, G. Coronel, and J. Baez, 2004: The major discharge events in
the Paraguay River: Magnitudes, source regions, and climate forcings. Journal
of Hydrometeorology, 5(6), 1161-1170.
Bartholy, J., and R. Pongracz, 2007: Regional analysis of extreme temperature and
precipitation indices for the Carpathian Basin from 1946 to 2001. Global and
Planetary Change, 57, 83-95.
Bartolini, G., M. Morabito, A. Crisci, D. Grifoni, T. Torrigiani, M. Petralli, G. Maracchi,
and S. Orlandini, 2008: Recent trends in Tuscany (Italy) summer temperature
and indices of extremes. International Journal of Climatology, 28(13), 1751-1760.
Bates, B.C., Z.W. Kundzewics, S. Wu, and J.P. Palutikof, 2008: Climate Change and
Water. Technical Paper of the Intergovernmental Panel on Climate Change. IPCC
Secretariat, Geneva, Switzerland, 210 pp.
Beaulant, A.L., B. Joly, O. Nuissier, S. Somot, V. Ducrocq, A. Joly, F. Sevault, M. Deque,
and D. Ricard, 2011: Statistico-dynamical downscaling for Mediterranean heavy
precipitation. Quarterly Journal of the Royal Meteorological Society, 137(656),
736-748.
Begueria, S., S.M. Vicente-Serrano, and M. Angulo-Martinez, 2010: A multiscalar
global drought dataset: The SPEIbase: A new gridded product for the analysis
of drought variability and impacts. Bulletin of the American Meteorological
Society,
91(10), 1351-1354.
Beirlant, J., Y. Goegebeur, J. Teugels, and J. Segers, 2004: Statistics of Extremes:
Theory and Applications. John Wiley & Sons, Chichester, West Sussex, England,
498 pp.
Beltaos, S. and T. Prowse, 2009: River-ice hydrology in a shrinking cryosphere.
Hydrological Processes, 23(1), 122-144.
Bender, M.A., T.R. Knutson, R.E. Tuleya, J.J. Sirutis, G.A. Vecchi, S.T. Garner, and I.M.
Held, 2010: Modeled impact of anthropogenic warming on the frequency of
intense Atlantic hurricanes. Science, 327(5964), 454-458.
Benestad, R.E., 2001: A comparison between two empirical downscaling strategies.
International Journal of Climatology, 21(13), 1645-1668.
Benestad, R.E., 2003: How often can we expect a record event? Climate Research,
25(1), 3-13.
Benestad, R.E., 2006: Can we expect more extreme precipitation on the monthly
time scale? Journal of Climate, 19(4), 630-637.
Benestad, R.E., 2007: Novel methods for inferring future changes in extreme
rainfall over Northern Europe. Climate Research, 34(3), 195-210.
Benestad, R.E., and J.E. Haugen, 2007: On complex extremes: flood hazards and
combined high spring-time precipitation and temperature in Norway. Climatic
Change, 85(3-4), 381-406.
Benestad, R.E., I. Hanssen-Bauer, and E.J. Forland, 2007: An evaluation of statistical
models for downscaling precipitation and their ability to capture long-term
trends. International Journal of Climatology, 27(5), 649-665.
Bengtsson, L., S. Hagemann, and K.I. Hodges, 2004: Can climate trends be calculated
from reanalysis data? Journal of Geophysical Research – Atmospheres, 109,
D11111.
Bengtsson, L., K.I. Hodges, and E. Roeckner, 2006: Storm tracks and climate change.
Journal of Climate, 19(15), 3518-3543.
Bengtsson, L., K.I. Hodges, M. Esch, N. Keenlyside, L. Kornblueh, J.J. Luo, and T.
Yamagata, 2007: How may tropical cyclones change in a warmer climate?
Tellus Series A – Dynamic Meteorology and Oceanography, 59(4), 539-561.
Bengtsson, L., K.I. Hodges, and N. Keenlyside, 2009: Will extratropical storms
intensify in a warmer climate? Journal of Climate, 22(9), 2276-2301.
Beniston, M., 2004: The 2003 heat wave in Europe: A shape of things to come? An
analysis based on Swiss climatological data and model simulations.
Geophysical Research Letters, 31, L02202.
Beniston, M., 2009a: Decadal-scale changes in the tails of probability distribution
functions of climate variables in Switzerland. International Journal of
Climatology
, 29(10), 1362-1368.
Beniston, M., 2009b: Trends in joint quantiles of temperature and precipitation in
Europe since 1901 and projected for 2100. Geophysical Research Letters, 36,
L07707.
Beniston, M., D.B. Stephenson, O.B. Christensen, C.A.T. Ferro, C. Frei, S. Goyette, K.
Halsnaes, T. Holt, K. Jylha, B. Koffi, J. Palutikof, R. Schoell, T. Semmler, and K.
Woth, 2007: Future extreme events in European climate: an exploration of
regional climate model projections. Climatic Change, 81, 71-95.
Benito, G., A. Diez-Herrero, and M.F. de Villalta, 2003: Magnitude and frequency of
flooding in the Tagus basin (Central Spain) over the last millennium. Climatic
Change, 58(1-2), 171-192.
Benito, G., M. Barriendos, C. Llasat, M. Machado, and V. Thorndycraft, 2005: Impacts
on natural hazards of climatic origin. Flood risk. In: A Preliminary General
Assessment of the Impacts in Spain Due to the Effects of Climate Change
[Moreno, J.M. (ed.)]. Ministry of Environment, Spain, pp. 507-527.
Benito, G., V.R. Thorndycraft, M. Rico, Y. Sanchez-Moya, and A. Sopena, 2008:
Palaeoflood and floodplain records from Spain: Evidence for long-term climate
variability and environmental changes. Geomorphology, 101(1-2), 68-77.
Bergametti, G. and F. Dulac, 1998: Mineral aerosols: Renewed interest for climate
forcing and tropospheric chemistry studies. IGACtivities Newsletter, 11, 13-17.
Bernier, N.B., K.R. Thompson, J. Ou, and H. Ritchie, 2007: Mapping the return periods
of extreme sea levels: Allowing for short sea level records, seasonality, and
climate change. Global and Planetary Change, 57(1-2), 139-150.
Betts, R.A., O. Boucher, M. Collins, P.M. Cox, P.D. Falloon, N. Gedney, D.L. Hemming,
C. Huntingford, C.D. Jones, D.M.W. Sexton, and M.J. Web, 2007: Projected
increase in continental runoff due to plant responses to increasing carbon
dioxide. Nature, 448(7157), 1037-1041.
Bevan, S.L.N., P. R. J. North, W.M.F. Grey, S.O. Los, and S.E. Plummer, 2009: Impact of
atmospheric aerosol from biomass burning on Amazon dry-season drought.
Journal of Geophysical Research – Atmospheres, 114, D09204.
Bhutiyani, M.R., V.S. Kale, and N.J. Pawar, 2008: Changing streamflow patterns in
the rivers of northwestern Himalaya: Implications of global warming in the 20th
century. Current Science, 95(5), 618-626.
Biasutti, M., A.H. Sobel, and S.J. Camargo, 2009: The role of the Sahara Low in
summertime Sahel rainfall variability and change in the CMIP3 models. Journal
of Climate, 22(21), 5755-5771.
Bindoff, N.L., J. Willebrand, V. Artale, A. Cazenave, J.M. Gregory, S. Gulev, K. Hanawa,
C. Le Quéré, S. Levitus, Y. Nojiri, C.K. Shum, L.D. Talley, and A.S. Unnikrishnan,
2007: Observations: Oceanic climate change and sea level. In: Climate Change
2007: The Physical Science Basis. Contribution of Working Group I to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change
[Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor
and H.L. Miller (eds.)]. Cambridge University Press, Cambridge, UK, pp. 385-432.
Bister, M. and K.A. Emanuel, 1998: Dissipative heating and hurricane intensity.
Meteorology and Atmospheric Physics, Vienna, Austria, 65(3-4), 233-240.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
205
Bister, M. and K.A. Emanuel, 2002: Low frequency variability of tropical cyclone
potential intensity 2. Climatology for 1982-1995. Journal of Geophysical
Research – Atmospheres, 107(D22), 4621-4624.
Björnsson, H., 2003: Subglacial lakes and jökulhlaups in Iceland. Global and
Planetary Change, 35(3-4), 255-271.
Blenkinsop, S. and H.J. Fowler, 2007a: Changes in drought frequency, severity and
duration for the British Isles projected by the PRUDENCE regional climate
models. Journal of Hydrology, 342, 50-71.
Blenkinsop, S. and H.J. Fowler, 2007b: Changes in European drought characteristics
projected by the PRUDENCE regional climate models. International Journal of
Climatology, 27(12), 1595-1610.
Blewitt, G., Z. Altamimi, J. Davis, R. Gross, C.-Y. Kuo, F.G. Lemoine, A.W. Moore, R.E.
Neilan, H.-P. Plag, M. Rothacher, C.K. Shum, M.G. Sideris, T. Schöne, P. Tregoning,
and S. Zerbini, 2010: Geodetic observations and global reference frame
contributions to understanding sea-level rise and variability. In: Understanding
Sea-Level Rise and Variability [Church, J.A., P.L. Woodworth, T. Aarup, and W.S.
Wilson (eds.)]. Wiley-Blackwell, Chichester, UK, pp. 256-284.
Boberg, F., P. Berg, P. Thejll, W.J. Gutowski, and J.H. Christensen, 2009a: Improved
confidence in climate change projections of precipitation evaluated using daily
statistics from the PRUDENCE ensemble. Climate Dynamics, 32(7-8), 1097-1106.
Boberg, F., P. Berg, P. Thejll, W.J. Gutowski, and J.H. Christensen, 2009b: Improved
confidence in climate change projections of precipitation further evaluated using
daily statistics from the PRUDENCE ensemble. Climate Dynamics, 35(7-8),
1509-1520.
Boe, J., and L. Terray, 2008: Uncertainties in summer evapotranspiration changes over
Europe and implications for regional climate change. Geophysical Research
Letters, 35, L05702.
Boer, G.J., 2009: Changes in interannual variability and decadal potential predictability
under global warming. Journal of Climate, 22, 3098-3109.
Bollasina, M., and S. Nigam, 2009: Indian Ocean SST, evaporation, and precipitation
during the South Asian summer monsoon in IPCC-AR4 coupled simulations.
Climate Dynamics, 33(7-8), 1017-1032.
Bonan, G.B., 2008: Forests and climate change: Forcing, feedbacks, and the climate
benefit of forests. Science, 320, 1444-1449.
Bondi, F. and L. Salvatori, 2003: The 5-6 May 1998 mudflows in Campania, Italy. In:
Lessons Learnt from Landslide Disasters in Europe [Hervas, J. (ed.)]. European
Commission, Brussels, Belgium, pp. 102.
Bongaerts, P., T. Ridgway, E. Sampayo, and O. Hoegh-Guldberg, 2010: Assessing the
‘deep reef refugia’ hypothesis: focus on Caribbean reefs. Coral Reefs, 29(2),
309-327.
Boo, K.O., W.T. Kwon, and H.J. Baek, 2006: Change of extreme events of temperature
and precipitation over Korea using regional projection of future climate change.
Geophysical Research Letters, 33, L01701.
Brabson
, B.B., D.H. Lister, P.D. Jones, and J.P. Palutikof, 2005: Soil moisture and
predicted spells of extreme temperatures in Britain. Journal of Geophysical
Research – Atmospheres, 110, D05104.
Branstator, G., and F. Selten, 2009: “Modes of Variability” and climate change.
Journal of Climate, 22(10), 2639-2658.
Brayshaw, D.J., B. Hoskins, and M. Blackburn, 2008: The storm-track response to
idealized SST perturbations in an aquaplanet GCM. Journal of the Atmospheric
Sciences, 65(9), 2842-2860.
Brázdil, R., C. Pfister, H. Wanner, H. Von Storch, and J. Luterbacher, 2005: Historical
climatology in Europe - The state of the art. Climatic Change, 70(3), 363-430.
Brázdil, R., Z.W. Kundzewicz, and G. Benito, 2006: Historical hydrology for studying
flood risk in Europe. Hydrological Sciences Journal, 51(5), 739-764.
Brázdil, R., M. Trnka, P. Dobrovolny, K. Chromi, P. Hlavinka, and Z. Zalud, 2009:
Variability of droughts in the Czech Republic, 1881-2006. Theoretical and
Applied Climatology, 97(3-4), 297-315.
Breda, N., and V. Badeau, 2008: Forest tree responses to extreme drought and some
biotic events: Towards a selection according to hazard tolerance? Comptes
Rendus Geoscience, 340(9-10), 651-662.
Brockhaus, P., D. Luthi, and C. Schär, 2008: Aspects of the diurnal cycle in a regional
climate model. Meteorologische Zeitschrift, 17(4), 433-443.
Bromirski, P.D., R.E. Flick, and D.R. Cayan, 2003: Storminess variability along the
California coast: 1858-2000. Journal of Climate, 16(6), 982.
Brooks, H.E., and N. Dotzek, 2008: The spatial distribution of severe convective
storms and an analysis of their secular changes. In: Climate Extremes and
Society [Diaz, H.F., and R. Murnane (eds.)]. Cambridge University Press,
Cambridge, NY, pp. 35-53.
Brovkin, V., T. Raddatz, C.H. Reick, M. Claussen, and V. Gayler, 2009: Global
biogeophysical interactions between forest and climate. Geophysical Research
Letters, 36, L07405.
Brown, J., M. Collins, and A. Tudhope, 2006: Coupled model simulations of mid-
Holocene ENSO and comparisons with coral oxygen isotope record. Advances
in Geosciences, 6, 29-33.
Brown, J.M., A.J. Souza, and J. Wolf, 2010: Surge modelling in the eastern Irish Sea:
present and future storm impact. Ocean Dynamics, 60(2), 227-236.
Brown, R.D., and P.W. Mote, 2009: The response of Northern Hemisphere snow cover
to a changing climate. Journal of Climate, 22(8)
, 2124-2145.
Brown, S.J., J. Caesar, and C.A.T. Ferro, 2008: Global changes in extreme daily
temperature since 1950. Journal of Geophysical Research – Atmospheres, 113,
D05115.
Brunet, M., P.D. Jones, J. Sigró, O. Saladié, E. Aguilar, A. Moberg, P.M. Della-Marta,
D. Lister, A. Walther, and D. López, 2007: Temporal and spatial temperature
variability and change over Spain during 1850-2005. Journal of Geophysical
Research – Atmospheres, 112, D12117.
Bryan, K.R., P.S. Kench, and D.E. Hart, 2008: Multi-decadal coastal change in New
Zealand: Evidence, mechanisms and implications. New Zealand Geographer,
64(2), 117-128.
Büntgen, U., V. Trouet, D. Frank, H.H. Leuschner, D. Friedrichs, J. Luterbacher, and J.
Esper, 2010: Tree-ring indicators of German summer drought over the last
millennium. Quaternary Science Reviews, 29(7-8), 1005-1016.
Burke, E.J., and S.J. Brown, 2008: Evaluating uncertainties in the projection of future
drought. Journal of Hydrometeorology, 9(2), 292-299.
Burke, E.J., S.J. Brown, and N. Christidis, 2006: Modeling the recent evolution of
global drought and projections for the twenty-first century with the Hadley
Centre climate model. Journal of Hydrometeorology, 7(5), 1113-1125.
Burke, E.J., R.H.J. Perry, and S.J. Brown, 2010: An extreme value analysis of UK drought
and projections of change in the future. Journal of Hydrology, 388(1-2),
131-143.
Burkholder, B.A. and D.J. Karoly, 2007: An assessment of US climate variability
using the Climate Extremes Index. In: Proceedings of the Nineteenth
Conference on Climate Variability and Change, San Antonio, TX, 15-18 January
2007, pp. 2B.9.
Burton, A., C.G. Kilsby, H.J. Fowler, P.S.P. Cowpertwait, and P.E. O’Connell, 2008:
RainSim: A spatial-temporal stochastic rainfall modelling system.
Environmental Modelling & Software, 23(12), 1356-1369.
Caesar, J., L. Alexander, and R. Vose, 2006: Large-scale changes in observed daily
maximum and minimum temperatures: Creation and analysis of a new gridded
data set. Journal of Geophysical Research – Atmospheres, 111, D05101.
Caesar, J., L.V. Alexander, B. Trewin, K. Tse-ring, L. Sorany, V. Vuniyayawa, N.
Keosavang, A. Shimana, M.M. Htay, J. Karmacharya, D.A. Jayasinghearachchi, J.
Sakkamart, A. Soares, L.T. Hung, L.T. Thuong, C.T. Hue, N.T.T. Dung, P.V. Hung,
H.D. Cuong, N.M. Cuong, and S. Sirabaha, 2011: Changes in temperature and
precipitation extremes over the Indo-Pacific region from 1971 to 2005.
International Journal of Climatology, 31(6), 791-801.
Callaghan, J., and S. Power, 2011: Variability and decline in the number of severe
tropical cyclones making land-fall over eastern Australia since the late
nineteenth century. Climate Dynamics, 37(3-4), 647-662, doi:10.1007/s00382-
00010-00883-00382.
Camberlin, P., V. Moron, R. Okoola, N. Philippon, and W. Gitau, 2009: Components
of rainy seasons’ variability in Equatorial East Africa: onset, cessation, rainfall
frequency and intensity. Theoretical and Applied Climatology, 98(3-4),
237-249.
Cambers, G., 2009: Caribbean beach changes and climate change adaptation.
Aquatic Ecosystem Health & Management, 12(2), 168-176.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
206
Cameron, D., 2006: An application of the UKCIP02 climate change scenarios to flood
estimation by continuous simulation for a gauged catchment in the northeast
of Scotland, UK (with uncertainty). Journal of Hydrology, 328(1-2), 212-226.
Camilloni, I.A. and V.R. Barros, 2003: Extreme discharge events in the Paraná River
and their climate forcing. Journal of Hydrology, 278(1-4), 94-106.
Campbell, J.D., M.A. Taylor, T.S. Stephenson, R.A. Watson, and F.S. Whyte, 2011:
Future climate of the Caribbean from a regional climate model. International
Journal of Climatology, 31(12), 1866-1878, doi:10.1002/joc.2200.
Camuffo, D., C. Bertolin, M. Barriendos, F. Dominguez-Castro, C. Cocheo, S. Enzi, M.
Sghedoni, A. della Valle, E. Garnier, M.J. Alcoforado, E. Xoplaki, J. Luterbacher, N.
Diodato, M. Maugeri, M.F. Nunes, and R. Rodriguez, 2010: 500-year temperature
reconstruction in the Mediterranean Basin by means of documentary data and
instrumental observations. Climatic Change, 101(1-2), 169-199.
Cane, M.A., 2005: The evolution of El Niño, past and future. Earth and Planetary
Science Letters, 230(3-4), 227-240.
Cao, Z., 2008: Severe hail frequency over Ontario, Canada: Recent trend and variability.
Geophysical Research Letters, 35, L14803.
Caplan-Auerbach, J., and C. Huggel, 2007: Precursory seismicity associated with
frequent, large ice avalanches on Iliamna volcano, Alaska, USA. Journal of
Glaciology, 53(180), 128-140.
Capra, L., 2006: Abrupt climatic changes as triggering mechanisms of massive
volcanic collapses. Journal of Volcanology and Geothermal Research, 155(3-4),
329-333.
Carey, M., 2005: Living and dying with glaciers: people’s historical vulnerability to
avalanches and outburst floods in Peru. Global and Planetary Change, 47(2-4),
122-134.
Carey-Smith, T., S. Dean, J. Vial, and C. Thompson, 2010: Changes in precipitation
extremes for New Zealand: climate model predictions. Weather and Climate,
30, 23-48.
Carter, T.R., R.N. Jones, X. Lu, S. Bhadwal, C. Conde, L.O. Mearns, B.C. O’Neill,
M.D.A. Rounsevell, and M.B. Zurek, 2007: New assessment methods and the
characterisation of future conditions. In: Climate Change 2007. Impacts,
Adaptation and Vulnerability. Contribution of Working Group II to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change [Parry,
M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)].
Cambridge University Press, Cambridge, UK, pp. 133-171.
Cassou, C., L. Terray, and A.S. Phillips, 2005: Tropical Atlantic influence on European
heat waves. Journal of Climate, 18(15), 2805-2811.
Catto, J.L., L.C. Shaffrey, and K.I. Hodges, 2010: Can climate models capture the
structure of extratropical cyclones? Journal of Climate, 23(7), 1621-1635.
Cavazos, T., C. Turrent, and D.P. Lettenmaier, 2008: Extreme precipitation trends
associated with tropical cyclones in the core of the North American monsoon.
Geophysical Research Letters, 35, L21703.
Cayan, D.R., P.D. Bromirski, K. Hayhoe, M. Tyree, M.D. Dettinger, and R.E. Flick, 2008:
Climate change projections of sea level extremes along the California coast.
Climatic Change, 87(S1), 57-73.
Cayan, D.R., T. Das, D.W. Pierce, T.P. Barnett, M. Tyree, and A. Gershunov, 2010: Future
dryness in the southwest US and the hydrology of the early 21st century drought.
Proceedings of the National Academy of Sciences, 107(50), 21271-21276.
Cazenave, A. and W. Llovel, 2010: Contemporary sea level rise. Annual Review of
Marine Science, 2, 145-173.
CCSP, 2008: Weather and Climate Extremes in a Changing Climate. Regions of Focus:
North America, Hawaii, Caribbean, and U.S. Pacific Islands. A Report by the U.S.
Climate Change Science Program and the Subcommittee on Global Change
Research [Karl, T.R., G.A. Meehl, D.M. Christopher, S.J. Hassol, A.M. Waple, and
W.L. Murray (eds.)]. Department of Commerce, NOAA’s National Climatic Data
Center, Washington, DC, 164 pp.
Chambers, L.E. and G.M. Griffiths, 2008: The changing nature of temperature
extremes in Australia and New Zealand. Australian Meteorological Magazine,
57(1), 13-35.
Chan, J.C.L. and M. Xu, 2009: Inter-annual and inter-decadal variations of landfalling
tropical cyclones in East Asia. Part I: Time series analysis. International Journal
of Climatology, 29(9), 1285-1293.
Chang, E.K.M., 2007: Assessing the Increasing Trend in Northern Hemisphere Winter
Storm Track Activity Using Surface Ship Observations and a Statistical Storm
Track Model. Journal of Climate, 20, 5607–5628.
Chang, E.K.M., 2009: Are band-pass variance statistics useful measures of storm
track activity? Re-examining storm track variability associated with the NAO
using multiple storm track measures. Climate Dynamics, 33(2-3), 277-296.
Chang, E.K.M. and Y. Guo, 2007: Is the number of North Atlantic tropical cyclones
significantly underestimated prior to the availability of satellite observations?
Geophysical Research Letters, 34, L14801.
Chauvin, F., J.F. Royer, and M. Deque, 2006: Response of hurricane-type vortices to
global warming as simulated by ARPEGE-Climat at high resolution. Climate
Dynamics, 27(4), 377-399.
Chen, C.-T. and T. Knutson, 2008: On the verification and comparison of extreme
rainfall indices from climate models. Journal of Climate, 21(7), 1605-1621.
Chen, J., F.P. Brissette, and R. Leconte, 2011: Uncertainty of downscaling method in
quantifying the impact of climate change on hydrology. Journal of Hydrology,
401(3-4), 190-202.
Chen, J.L., C.R. Wilson, B.D. Tapley, Z.L. Yang, and G.Y. Niu, 2009: 2005 drought event
in the Amazon River basin as measured by GRACE and estimated by climate
models. Journal of Geophysical Research – Solid Earth, 114, B05404.
Chenoweth, M. and M. Devine, 2008: A document-based 318-year record of tropical
cyclones in the Lesser Antilles, 1690-2007. Geochemistry Geophysics
Geosystems, 9, Q08013.
Cherchi, A., S. Masina, and A. Navarra, 2008: Impact of extreme CO
2
levels on
tropical climate: a CGCM study. Climate Dynamics, 31(7-8), 743-758.
Cherry, N.J., 1988: Comment on long-term associations between wind speeds and
the urban heat-island of Phoenix, Arizona. Journal of Applied Meteorology,
27(7), 878-880.
Chiapello, I., C. Moulin, and J.M. Prospero, 2005: Understanding the long-term
variability of African dust transport across the Atlantic as recorded in both
Barbados surface concentrations and large-scale Total Ozone Mapping
Spectrometer (TOMS) optical thickness. Journal of Geophysical Research –
Atmospheres, 110, D18S10.
Chiarle, M., S. Iannotti, G. Mortara, and P. Deline, 2007: Recent debris flow
occurrences associated with glaciers in the Alps. Global and Planetary Change,
56(1-2), 123-136.
Chini, N., P. Stansby, J. Leake, J. Wolf, J. Roberts-Jones, and J. Lowe, 2010: The impact
of sea level rise and climate change on inshore wave climate: A case study for
East Anglia (UK). Coastal Engineering, 57(11-12), 973-984.
Choi, G., D. Collins, R. Guoyu, B. Trewin, M. Baldi, Y. Fukuda, M. Afzaal, T. Pianmana,
P. Gomboluudev, P.T.T. Huong, N. Lias, W.T. Kwon, K.O. Boo, Y.M. Cha, and
Y. Zhou, 2009: Changes in means and extreme events of temperature and
precipitation in the Asia-Pacific Network region, 1955-2007. International
Journal of Climatology, 29(13), 1956-1975.
Chowdhury, M.R., P.S. Chu, and T. Schroeder, 2007: ENSO and seasonal sea-level
variability – A diagnostic discussion for the US-affiliated Pacific Islands.
Theoretical and Applied Climatology, 88(3-4), 213-224.
Christensen, J.H. and O.B. Christensen, 2003: Climate modeling: Severe summertime
flooding in Europe. Nature, 421, 805-806.
Christensen, J.H. and O.B. Christensen, 2007: A summary of the PRUDENCE model
projections of changes in European climate by the end of this century. Climatic
Change, 81(S1), 7-30.
Christensen, J.H., B. Hewitson, A. Busuioc, A. Chen, X. Gau, I. Held, R. Jones, R. Kolli,
W. Kwon, R. Laprise, V. Magaña Rueda, L. Mearns, C. Menéndez, J. Räisänen, A.
Rinke, A. Sarr, and P. Whetton, 2007: Regional climate projections. In: Climate
Change 2007: The Physical Science Basis. Contribution of Working Group I to
the Fourth Assessment Report of the Intergovernmental Panel on Climate
Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M.
Tignor and H.L. Miller (eds.)]. Cambridge University Press, Cambridge, UK, pp.
847-940.
Christidis, N., P.A. Stott, S. Brown, G. Hegerl, and J. Caesar, 2005: Detection of
changes in temperature extremes during the second half of the 20th century.
Geophysical Research Letters, 32, L20716.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
207
Christidis, N., P.A. Stott, G.S. Jones, H. Shiogama, T. Nozawa, and J. Luterbacher,
2011a: Human activity and anomalously warm seasons in Europe. International
Journal of Climatology, doi:10.1002/joc.2262.
Christidis, N., P.A. Stott, and S.J. Brown, 2011b: The role of human activity in the
recent warming of extremely warm daytime temperatures. Journal of Climate,
24(7), 1922-1930.
Church, J.A. and N.J. White, 2011: Sea-level rise from the late 19th to the early 21st
century. Surveys in Geophysics, 32(4-5), 585-602, doi:10.1007/s10712-10011-
19119-10711.
Church, J.A., J.R. Hunter, K.L. McInnes, and N.J. White, 2006a: Sea-level rise around
the Australian coastline and the changing frequency of extreme sea-level
events. Australian Meteorological Magazine, 55(4), 253-260.
Church, J.A., N.J. White, and J.R. Hunter, 2006b: Sea-level rise at tropical Pacific and
Indian Ocean Islands. Global and Planetary Change, 53(3), 155-168.
Church, J.A., N.J. White, J.R. Hunter, K.L. McInnes, P. J. Cowell , and S.P. O’Farrell, 2008:
Sea-level rise. In: Transitions: Pathways Towards Sustainable Urban
Development in Australia [Newton, E.P. (ed.)]. CSIRO Publishing and Springer
Science, Collingwood, Australia, pp. 691.
Church, J.A., D. Roemmich, C.M. Domingues, J.K. Willis, N.J. White, J.E. Gilson, D.
Stammer, A. Köhl, D.P. Chambers, F.W. Landerer, J. Marotzke, J.M. Gregory, T.
Suzuki, A. Cazenave, and P.-Y. Le Traon, 2010: Ocean temperature and salinity
contributions to global and regional sea-level change. In: Understanding Sea-
Level Rise and Variability [Church, J.A., P.L. Woodworth, T. Aarup, and W.S.
Wilson (eds.)]. Wiley-Blackwell, Chichester, UK, pp. 143-176.
Church, J.A., J.M. Gregory, N.J. White, S.M. Platten, and J.X. Mitrovica, 2011:
Understanding and projecting sea level change. Oceanography, 24(2),
130-143.
Chust, G., A. Borja, P. Liria, I. Galparsoro, M. Marcos, A. Caballero, and R. Castro,
2009: Human impacts overwhelm the effects of sea-level rise on Basque coastal
habitats (N Spain) between 1954 and 2004. Estuarine Coastal and Shelf
Science, 84(4), 453-462.
Ciais, P., M. Reichstein, N. Viovy, A. Granier, J. Ogee, V. Allard, M. Aubinet, N.
Buchmann, C. Bernhofer, A. Carrara, F. Chevallier, N. De Noblet, A. Friend, P.
Friedlingstein, T. Grunwald, B. Heinesch, P. Keronen, A. Knohl, G. Krinner, D.
Loustau, G. Manca, G. Matteucci, F. Miglietta, J. Ourcival, D. Papale, K. Pilegaard,
S. Rambal, G. Seufert, J. Soussana, M. Sanz, E. Schulze, T. Vesala, and R. Valentini,
2005: Europe-wide reduction in primary productivity caused by the heat and
drought in 2003. Nature, 437(7058), 529-533.
Clague, J.J. and S.G. Evans, 2000: A review of catastrophic drainage of moraine-
dammed lakes in British Columbia. Quaternary Science Reviews, 19(17-18),
1763-1783.
Clark, R.T., S.J. Brown, and J.M. Murphy, 2006: Modeling northern hemisphere summer
heat extreme changes and their uncertainties using a physics ensemble of
climate sensitivity experiments. Journal of Climate, 19(17), 4418-4435.
Clark, R.T., J.M. Murphy, and S.J. Brown, 2010: Do global warming targets limit
heatwave risk? Geophysical Research Letters, 37, L17703.
Clarke, G.K.C., 1982: Glacier outburst floods from ‘Hazard Lake,’ Yukon Territory, and
the problem of flood magnitude prediction. Journal of Glaciology, 28(98), 3-21.
Clarke, M.L., and H.M. Rendell, 2009: The impact of North Atlantic storminess on
western European coasts: A review. Quaternary International, 195, 31-41.
Clemmensen, L.B., M. Bjornsen, A. Murray, and K. Pedersen, 2007: Formation of
aeolian dunes on Anholt, Denmark since AD 1560: A record of deforestation and
increased storminess. Sedimentary Geology, 199(3-4), 171-187.
Clow, D.W., 2010: Changes in the timing of snowmelt and streamflow in Colorado:
a response to recent warming. Journal of Climate, 23(9), 2293-2306.
Cobb, K.M., C.D. Charles, H. Cheng, and R.L. Edwards, 2003: El Niño/Southern
Oscillation and tropical Pacific climate during the last millennium. Nature, 424,
271-276.
Coelho, C., R. Silva, F. Veloso-Gomes, and F. Taveira-Pinto, 2009: Potential effects of
climate change on northwest Portuguese coastal zones. ICES Journal of Marine
Science, 66(7), 1497-1507.
Coles, S., 2001: An Introduction to Statistical Modeling of Extreme Values. Springer-
Verlag, Heidelberg, Germany, 208 pp.
Collini, E.A., E.H. Berbery, V. Barros, and M. Pyle, 2008: How does soil moisture
influence the early stages of the South American monsoon? Journal of Climate,
21(2), 195-213.
Collins, M., B.B.B. Booth, G.R. Harris, J.M. Murphy, D.M.H. Sexton, and M.J. Webb,
2006: Towards quantifying uncertainty in transient climate change. Climate
Dynamics, 27(2-3), 127-147.
Collins, M., S.I. An, W.J. Cai, A. Ganachaud, E. Guilyardi, F.F. Jin, M. Jochum, M.
Lengaigne, S. Power, A. Timmermann, G. Vecchi, and A. Wittenberg, 2010: The
impact of global warming on the tropical Pacific ocean and El Nino. Nature
Geoscience, 3(6), 391-397.
Compo, G.P., J.S. Whitaker, and P.D. Sardeshmukh, 2006: Feasibility of a 100-year
reanalysis using only surface pressure data. Bulletin of the American
Meteorological Society, 87(2), 175-190.
Conroy, J.L., A. Restrepo, J.T. Overpeck, M. Seinitz-Kannan, J.E. Cole, M.B. Bush, and
P.A. Colinvaux, 2009: Unprecedented recent warming of surface temperatures
in the eastern tropical Pacific Ocean. Nature Geoscience, 2(1), 46-50.
Conway, D., C.E. Hanson, R. Doherty, and A. Persechino, 2007: GCM simulations of
the Indian Ocean dipole influence on East African rainfall: Present and future.
Geophysical Research Letters, 34, L03705.
Conway, D., A. Persechino, S. Ardoin-Bardin, H. Hamandawana, C. Dieulin, and G.
Mahe, 2009: Rainfall and water resources variability in sub-Saharan Africa
during the twentieth century. Journal of Hydrometeorology, 10(1), 41-59.
Cook, B.I., G.B. Bonan, and S. Levis, 2006: Soil moisture feedbacks to precipitation in
southern Africa. Journal of Climate, 19(17), 4198-4206.
Cook, K.H. and E.K. Vizy, 2006: Coupled model simulations of the west African
monsoon system: Twentieth- and twenty-first-century simulations. Journal of
Climate, 19(15), 3681-3703.
Coppola, E. and F. Giorgi, 2005: Climate change in tropical regions from high-
resolution time-slice AGCM experiments. Quarterly Journal of the Royal
Meteorological Society, 131(612), 3123-3145.
Cordero, E.C. and P.M.D. Forster, 2006: Stratospheric variability and trends in models
used for the IPCC AR4. Atmospheric Chemistry and Physics, 6(12), 5369-5380.
Corti, T., V. Muccione, P. Köllner-Heck, D. Bresch, and S.I. Seneviratne, 2009:
Simulating past droughts and associated building damages in France.
Hydrology and Earth System Sciences, 13(9), 1739-1747.
Costa, A.C., and A. Soares, 2009: Trends in extreme precipitation indices derived from
a daily rainfall database for the South of Portugal. International Journal of
Climatology, 29(13), 1956-1975.
CSIRO, 2007: Climate Change in Australia. CSIRO, Australia, 148 pp.
Cubasch, U., J. Waszkewitz, G. Hegerl, and J. Perlwitz, 1995: Regional climate
changes as simulated in time-slice experiments. Climatic Change, 31(2-4),
273-304.
Cullen, N.J., T. Molg, G. Kaser, K. Hussein, K. Steffen, and D.R. Hardy, 2006:
Kilimanjaro Glaciers: Recent areal extent from satellite data and new
interpretation of observed 20th century retreat rates. Geophysical Research
Letters, 33, L16502.
Cunderlik, J.M., and T.B.M.J. Ouarda, 2009: Trends in the timing and magnitude of
floods in Canada. Journal of Hydrology, 375(3-4), 471-480.
Cunningham, P., and M.J. Reeder, 2009: Severe convective storms initiated by
intense wildfires: Numerical simulations of pyro-convection and pyro-tornado-
genesis. Geophysical Research Letters, 36, L12812.
da Silva, A.E. and L.M.V. de Carvalho, 2007: Large-scale index for South America
Monsoon (LISAM). Atmospheric Science Letters, 8(2), 51-57.
Dai, A., T.T. Qian, K.E. Trenberth, and J.D. Milliman, 2009: Changes in continental
freshwater discharge from 1948 to 2004. Journal of Climate, 22(10),
2773-2792.
Dai, A., 2011: Drought under global warming: a review. Wiley Interdisciplinary
Reviews: Climate Change, 2(1), 45-65.
Dai, A.G., K.E. Trenberth, and T.T. Qian, 2004: A global dataset of Palmer Drought
Severity Index for 1870-2002: Relationship with soil moisture and effects of
surface warming. Journal of Hydrometeorology, 5(6), 1117-1130.
Dairaku, K., S. Emori, and H. Higashi, 2008: Potential changes in extreme events
under global climate change. Journal of Disaster Research, 3(1), 39-50.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
208
Danard, M.B., S.K. Dube, G. Gönnert, A. Munroe, T.S. Murty, P. Chittibabu, A.D. Rao,
and P.C. Sinha, 2004: Storm surges from extra-tropical cyclones. Natural
Hazards, 32(2), 177-190.
Dankers, R., O.B. Christensen, L. Feyen, M. Kalas, and A. de Roo, 2007: Evaluation of
very high-resolution climate model data for simulating flood hazards in the
Upper Danube Basin. Journal of Hydrology, 347(3-4), 319-331.
Dankers, R. and L. Feyen, 2008: Climate change impact on flood hazard in Europe:
An assessment based on high-resolution climate simulations. Journal of
Geophysical Research – Atmospheres, 113, D19105.
Dankers, R. and L. Feyen, 2009: Flood hazard in Europe in an ensemble of regional
climate scenarios. Journal of Geophysical Research – Atmospheres, 114,
D16108.
Dasgupta, S., B. Laplante, C. Meisner, D. Wheeler, and J. Yan, 2009: The impact of sea
level rise on developing countries: a comparative analysis. Climatic Change,
93(3-4), 379-388.
Dawson, J.L., and S.G. Smithers, 2010: Shoreline and beach volume change between
1967 and 2007 at Raine Island, Great Barrier Reef, Australia. Global and
Planetary Change, 72(3), 141-154.
de Elía, R., D. Caya, H. Côté, A. Frigon, S. Biner, M. Giguère, D. Paquin, R. Harvey, and
D. Plummer, 2008: Evaluation of uncertainties in the CRCM-simulated North
American climate. Climate Dynamics, 30(2-3), 113-132.
de Jong, R., S. Bjorck, L. Bjorkman, and L.B. Clemmensen, 2006: Storminess variation
during the last 6500 years as reconstructed from an ombrotrophic peat bog in
Halland, southwest Sweden. Journal of Quaternary Science, 21(8), 905-919.
de Jong, R., K. Schoning, and S. Bjorck, 2007: Increased aeolian activity during
humidity shifts as recorded in a raised bog in south-west Sweden during the
past 1700 years. Climate of the Past, 3(3), 411-422.
Dean, S.M. and P.A. Stott, 2009: The effect of local circulation variability on the
detection and attribution of New Zealand temperature trends. Journal of
Climate, 22(23), 6217-6229.
Debernard, J.B. and L.P. Roed, 2008: Future wind, wave and storm surge climate in
the Northern Seas: a revisit. Tellus Series A – Dynamic Meteorology and
Oceanography, 60(3), 427-438.
Deeming, K.R., B. McGuire, and P. Harrop, 2010: Climate forcing of volcano lateral
collapse: evidence from Mount Etna, Sicily. Philosophical Transactions of the
Royal Society A: Mathematical, Physical and Engineering Sciences, 368(1919),
2559-2577.
Defeo, O., A. McLachlan, D.S. Schoeman, T.A. Schlacher, J. Dugan, A. Jones, M. Lastra,
and F. Scapini, 2009: Threats to sandy beach ecosystems: A review. Estuarine
Coastal and Shelf Science, 81(1)
, 1-12.
DeGaetano, A.T., 2009: Time-dependent changes in extreme-precipitation return-
period amounts in the Continental United States. Journal of Applied
Meteorology and Climatology, 48, 2086-2099.
Delaloye, R., T. Strozzi, C. Lambiel, E. Perruchoud, and H. Raetzo, 2008: Landslide-like
development of rockglaciers detected with ERS-1/2 SAR interferometry. In:
Proceedings of the FRINGE 2007 Symposium at ESRIN [Lacoste, H., and L.
Ouwehand (eds.)], Frascati, Italy, 26-30 November 2007, ESA Communication
Production Office, ESTEC, Noordwijk, The Netherlands, pp. 1-6.
Delgado, J.M., H. Apel, and B. Merz, 2009: Flood trends and variability in the Mekong
River. Hydrology and Earth System Sciences, 6(3), 6691-6719.
Della-Marta, P.M. and H. Wanner, 2006: A method of homogenizing the extremes
and mean of daily temperature measurements. Journal of Climate, 19,
4179-4197.
Della-Marta, P.M., M.R. Haylock, J. Luterbacher, and H. Wanner, 2007a: Doubled
length of western European summer heat waves since 1880. Journal of
Geophysical Research – Atmospheres, 112, D15103.
Della-Marta, P.M., J. Luterbacher, H. von Weissenfluh, E. Xoplaki, M. Brunet, and H.
Wanner, 2007b: Summer heat waves over western Europe 1880-2003, their
relationship to large-scale forcings and predictability. Climate Dynamics,
29(2-3), 251-275.
Della-Marta, P.M., H. Mathis, C. Frei, M.A. Liniger, J. Kleinn, and C. Appenzeller,
2009: The return period of wind storms over Europe. International Journal of
Climatology, 29(3), 437-459.
DeMaria, M., J.A. Knaff, and B.H. Connell, 2001: A tropical cyclone genesis parameter
for the tropical Atlantic. Weather and Forecasting, 16(2), 219-233.
Deo, R.C., J.I. Syktus, C.A. McAlpine, P.J. Lawrence, H.A. McGowan, and S.R. Phinn,
2009: Impact of historical land cover change on daily indices of climate
extremes including droughts in eastern Australia. Geophysical Research Letters,
36, L08705.
Department of Climate Change, 2009: Climate Change Risks to Australia’s
Coast. A First Pass National Assessment. Report published by the Department
of Climate Change, Australian Government, Canberra, Australia, 172 pp.
Déqué, M., S. Somot, E. Sanchez-Gomez, C. Goodess, D. Jacob, G. Lenderink, and O.
Christensen, 2011: The spread amongst ENSEMBLES regional scenarios: regional
climate models, driving general circulation models and interannual variability.
Climate Dynamics, doi:10.1007/s00382-00011-01053-x.
Deser, C., R.A. Tomas, and S. Peng, 2007: The transient atmospheric circulation
response to North Atlantic SST and sea ice anomalies. Journal of Climate,
20(18), 4751-4767.
Di Baldassarre, G., A. Montanari, H. Lins, D. Koutsoyiannis, L. Brandimarte, and G.
Blöschl, 2010: Flood fatalities in Africa: From diagnosis to mitigation.
Geophysical Research Letters, 37, L22402.
Dibike, Y.B., P. Gachon, A. St-Hilaire, T. Ouarda, and V.T.V. Nguyen, 2008: Uncertainty
analysis of statistically downscaled temperature and precipitation regimes in
Northern Canada. Theoretical and Applied Climatology, 91(1-4), 149-170.
Diffenbaugh, N.S. and M. Ashfaq, 2010: Intensification of hot extremes in the
United States. Geophysical Research Letters
, 37(15), L15701.
Diffenbaugh, N.S., J.S. Pal, F. Giorgi, and X. Gao, 2007: Heat stress intensification in
the Mediterranean climate change hotspot. Geophysical Research Letters, 34,
L11706.
Diffenbaugh, N.S., R.J. Trapp, and H. Brooks, 2008: Does global warming influence
tornado activity? Eos Transactions (AGU), 89(53), 553.
Ding, T., W. Qian, and Z. Yan, 2010: Changes in hot days and heat waves in China
during 1961-2007. International Journal of Climatology, 30(10), 1452-1462.
Dirmeyer, P.A., R.D. Koster, and Z. Guo, 2006: Do global models properly represent
the feedback between land and atmosphere? Journal of Hydrometeorology,
7(6), 1177-1198.
Dobrovolny, P., A. Moberg, R. Brazdil, C. Pfister, R. Glaser, R. Wilson, A. van Engelen,
D. Limanowka, A. Kiss, M. Halickova, J. Mackova, D. Riemann, J. Luterbacher, and
R. Bohm, 2010: Monthly, seasonal and annual temperature reconstructions for
Central Europe derived from documentary evidence and instrumental records
since AD 1500. Climatic Change, 101(1-2), 69-107.
Dodet, G., X. Bertin, and R. Taborda, 2010: Wave climate variability in the North-East
Atlantic Ocean over the last six decades. Ocean Modelling, 31(3-4), 120-131.
Dole, R., M. Hoerling, J. Perlwitz, J. Eischeid, P. Pegion, T. Zhang, X.-W. Quan, T. Xu,
and D. Murray, 2011: Was there a basis for anticipating the 2010 Russian heat
wave? Geophysical Research Letters, 38, L06702.
Donat, M.G., G.C. Leckebusch, J.G. Pinto, and U. Ulbrich, 2010a: European storminess
and associated circulation weather types: future changes deduced from a multi-
model ensemble of GCM simulations. Climate Research, 42(1), 27-43.
Donat, M.G., G.C. Leckebusch, J.G. Pinto, and U. Ulbrich, 2010b: Examination of wind
storms over Central Europe with respect to circulation weather types and NAO
phases. International Journal of Climatology, 30(9), 1289-1300.
Donat, M.G., G.C. Leckebusch, S. Wild, and U. Ulbrich, 2011: Future changes of
European winter storm losses and extreme wind speeds in multi-model GCM
and RCM simulations. Natural Hazards and Earth System Sciences, 11(5),
1351-1370.
Dong, B., R. Sutton, and T. Woollings, 2011: Changes of interannual NAO variability
in response to greenhouse gases forcing. Climate Dynamics, 37(7-8), 1621-1641,
doi:10.1007/s00382-00010-00936-00386.
Donnelly, J.P., and J.D. Woodruff, 2007: Intense hurricane activity over the past 5,000
years controlled by El Niño and the West African monsoon. Nature, 447,
465-468.
D’Onofrio, A., J.P. Boulanger, and E.C. Segura, 2010: CHAC: a weather pattern
classification system for regional climate downscaling of daily precipitation.
Climatic Change, 98(3-4), 405-427.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
209
Donohue, R.J., T.R. McVicar, and M.L. Roderick, 2010: Assessing the ability of potential
evaporation formulations to capture the dynamics in evaporative demand within
a changing climate. Journal of Hydrology, 386(1-4), 186-197.
Doser, D.I., K.R. Wiest, and J. Sauber, 2007: Seismicity of the Bering Glacier region
and its relation to tectonic and glacial processes. Tectonophysics, 439(1-4),
119-127.
Douglas, E.M., R.M. Vogel, and C.N. Kroll, 2000: Trends in floods and low flows in the
United States: impact of spatial correlation. Journal of Hydrology, 240(1-2),
90-105.
Dragani, W.C., P.B. Martin, C.G. Simionato, and M.I. Campos, 2010: Are wind wave
heights increasing in south-eastern south American continental shelf between
32°S and 40°S? Continental Shelf Research, 30(5), 481-490.
Dufek, A.S. and T. Ambrizzi, 2008: Precipitation variability in Sao Paulo State, Brazil.
Theoretical and Applied Climatology, 93(3-4), 167-178.
Dufek, A.S., T. Ambrizzi, and R.P. da Rocha, 2008: Are reanalysis data useful for
calculating climate indices over South America? In: Trends and Directions in
Climate Research [Gimeno, L., R. GarciaHerrera, and R.M. Trigo (eds.)]. Vol. 1146,
Annals of the New York Academy of Sciences, New York, NY, USA, pp. 87-104.
Durante, F., and G. Salvadori, 2010: On the construction of multivariate extreme
value models via copulas. Environmetrics, 21(2), 143-161.
Durão, R.M., M.J. Pereira, A.C. Costa, J. Delgado, G. del Barrio, and A. Soares, 2010:
Spatial-temporal dynamics of precipitation extremes in southern Portugal: a
geostatistical assessment study. International Journal of Climatology, 30(10),
1526-1537.
Durre, I., J.M. Wallace, and D.P. Lettenmaier, 2000: Dependence of extreme daily
maximum temperatures on antecedent soil moisture in the contiguous United
States during summer. Journal of Climate, 13(14), 2641-2651.
Durre, I., M.J. Menne, and R.S. Vose, 2008: Strategies for evaluating quality
assurance procedures. Journal of Applied Meteorology and Climatology, 47(6),
1785-1791.
Dussaillant, A., G. Benito, W. Buytaert, P. Carling, C. Meier, and F. Espinoza, 2010:
Repeated glacial-lake outburst floods in Patagonia: an increasing hazard?
Natural Hazards, 54(2), 469-481.
Dutton, J.F., C.J. Poulsen, and J.L. Evans, 2000: The effect of global climate change on
the regions of tropical convection in CSM1. Geophysical Research Letters,
27(19), 3049-3052.
Easterling, D.R., D.M. Anderson, S.J. Cohen, W.J. Gutowski, G.J. Holland, K.E. Kunkel,
T.C. Peterson, R.S. Pulwarty, R.J. Stouffer, and M.F. Wehner, 2008: Measures to
improve our understanding of weather and climate extremes. In: Weather and
Climate Extremes in a Changing Climate. Regions of focus: North America,
Hawaii, Caribbean, and U.S. Pacific Islands.
[Karl, T.R., G.A. Meehl, D.M.
Christopher, S.J. Hassol, A.M. Waple, and W.L. Murray (eds.)]. A Report by the
U.S. Climate Change Science Program and the Subcommittee on Global Change
Research, Washington, DC, pp. 117-126.
Eichler, T., and W. Higgins, 2006: Climatology and ENSO-related variability of North
American extratropical cyclone activity. Journal of Climate, 19(10), 2076-2093.
Elsner, J.B., J.P. Kossin, and T.H. Jagger, 2008: The increasing intensity of the strongest
tropical cyclones. Nature, 455(7209), 92-95.
Eltahir, E.A.B. and P.A.J.F. Yeh, 1999: On the asymmetric response of aquifer
water level to floods and droughts in Illinois. Water Resources Research, 35(4),
1199-1217.
Ely, L.L., Y. Enzel, V.R. Baker, and D.R. Cayan, 1993: A 5000-year record of extreme
floods and climate-change in the southwestern United-States. Science,
262(5132), 410-412.
Emanuel, K.A., 1987: Dependence of hurricane intensity on climate. Nature,
326(61112), 483-485.
Emanuel, K.A., 2000: A statistical analysis of tropical cyclone intensity. Monthly
Weather Review, 128(4), 1139-1152.
Emanuel, K.A., 2007: Environmental factors affecting tropical cyclone power
dissipation. Journal of Climate, 20(22), 5497-5509.
Emanuel, K.A., 2010: Tropical cyclone activity downscaled from NOAA-CIRES
reanalysis, 1908-1958. Journal of Advances in Modeling Earth Systems, 2, 1,
doi:10.3894/JAMES.2010.2.1.
Emanuel, K.A., R. Sundararajan, and J. Williams, 2008: Hurricanes and global warming:
Results from downscaling IPCC AR4 simulations. Bulletin of the American
Meteorological Society, 89(3), 347-367.
Emori, S. and S.J. Brown, 2005: Dynamic and thermodynamic changes in mean and
extreme precipitation under changed climate. Geophysical Research Letters,
32, L17706.
Engelstaedter, S., I. Tegen, and R. Washington, 2006: North African dust emissions
and transport. Earth-Science Reviews, 79(1-2), 73-100.
Essery, R., N. Rutter, J. Pomeroy, R. Baxter, M. Stahli, D. Gustafsson, A. Barr, P. Bartlett,
and K. Elder, 2009: An evaluation of forest snow process simulations. Bulletin of
the American Meteorological Society, 90(8), 1120-1135.
Evan, A.T., D.J. Vimont, A.K. Heidinger, J.P. Kossin, and R. Bennartz, 2009: The role
of aerosols in the evolution of tropical North Atlantic Ocean temperature
anomalies. Science, 324(5928), 778-781.
Evans, S.G., and J.J. Clague, 1998: Rock avalanche from Mount Munday, Waddington
Range, British Columbia, Canada. Landslide News, 11, 23-25.
Favre, A., and A. Gershunov, 2006: Extra-tropical cyclonic/anticyclonic activity in
North-Eastern Pacific and air temperature extremes in Western North America.
Climate Dynamics, 26(6), 617-629.
Favre, A., and A. Gershunov, 2009: North Pacific cyclonic and anticyclonic transients
in a global warming context: possible consequences for Western North American
daily precipitation and temperature extremes. Climate Dynamics, 32(7-8), 969-987.
Fedorov, A. and P. Konstantinov, 2003: Observations of surface dynamics with
thermokarst initiation, Yukechi site, central Yukutia. In: Proceedings of the 8th
International Conference on Permafrost [Philips, M., S.M. Springman, and L.U.
Arenson (eds.)]. Zurich, Switzerland, 21-25 July 2003, Balkema, Netherlands, pp.
239-243.
Fekete, B.M., C.J. Vörösmarty, J.O. Roads, and C.J. Willmott, 2004: Uncertainties in
precipitation and their impacts on runoff estimates. Journal of Climate, 17(2),
294-304.
Feyen, L., and R. Dankers, 2009: Impact of global warming on streamflow drought in
Europe. Journal of Geophysical Research – Atmospheres, 114, D17116.
Findell, K.L. and T.L. Delworth, 2005: A modeling study of dynamic and
thermodynamic mechanisms for summer drying in response to global warming.
Geophysical Research Letters, 32, L16702.
Findell, K.L. and E.A.B. Eltahir, 2003a: Atmospheric controls on soil moisture-boundary
layer interactions. Part I: Framework development. Journal of Hydrometeorology,
4(3), 552-569.
Findell, K.L. and E.A.B. Eltahir, 2003b: Atmospheric controls on soil moisture-boundary
layer interactions. Part II: Feedbacks within the continental United States.
Journal of Hydrometeorology, 4(3), 570-583.
Finnis, J., M.M. Holland, M.C. Serreze, and J.J. Cassano, 2007: Response of Northern
Hemisphere extratropical cyclone activity and associated precipitation to climate
change, as represented by the Community Climate System Model. Journal of
Geophysical Research – Biogeosciences, 112, G04S42.
Fiore, M.M.E., E.E. D’Onofrio, J.L. Pousa, E.J. Schnack, and G.R. Bertola, 2009: Storm
surges and coastal impacts at Mar del Plata, Argentina. Continental Shelf
Research, 29(14), 1643-1649.
Fischer, E.M. and C. Schär, 2009: Future changes in daily summer temperature
variability: driving processes and role for temperature extremes. Climate
Dynamics, 33(7-8), 917-935.
Fischer, E.M., and C. Schär, 2010: Consistent geographical patterns of changes in
high-impact European heatwaves. Nature Geoscience, 3(6), 398-403.
Fischer, E.M., S.I. Seneviratne, D. Lüthi, and C. Schär, 2007a: The contribution of land-
atmosphere coupling to recent European summer heatwaves. Geophysical
Research Letters, 34, L06707.
Fischer, E.M., S.I. Seneviratne, P.L. Vidale, D. Lüthi, and C. Schär, 2007b: Soil moisture-
atmosphere interactions during the 2003 European summer heatwave.
Journal
of Climate, 20(20), 5081-5099.
Fischer, L., H. Eisenbeiss, A. Kääb, C. Huggel, and W. Haeberli, 2011: Monitoring
topographic changes in a periglacial high-mountain face using high-resolution
DTMs, Monte Rosa East Face, Italian Alps. Permafrost and Periglacial Processes,
22(2), 140-152.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
210
Fleig, A.K., L.M. Tallaksen, H. Hisdal, and D.M. Hannah, 2011: Regional hydrological
drought in north-western Europe: linking a new Regional Drought Area Index
with weather types. Hydrological Processes, 25(7), 1163-1179.
Fogt, R.L., J. Perlwitz, A.J. Monaghan, D.H. Bromwich, J.M. Jones, and G.J. Marshall,
2009: Historical SAM variability. Part II: Twentieth-century variability and trends
from reconstructions, observations, and the IPCC AR4 models. Journal of
Climate, 22(20), 5346-5365.
Fontaine, B., P. Roucou, M. Gaetani, and R. Marteau, 2011: Recent changes in
precipitation, ITCZ convection and northern tropical circulation over North
Africa (1979–2007). International Journal of Climatology, 31(5), 633-648.
Forbes, D.L., G.S. Parkes, G.K. Manson, and L.A. Ketch, 2004: Storms and shoreline
retreat in the southern Gulf of St. Lawrence. Marine Geology, 210(1-4), 169-204.
Founda, D., and C. Giannakopoulos, 2009: The exceptionally hot summer of 2007 in
Athens, Greece – A typical summer in the future climate? Global and Planetary
Change, 67(3-4), 227-236.
Fowler, H.J., and M. Ekstrom, 2009: Multi-model ensemble estimates of climate
change impacts on UK seasonal precipitation extremes. International Journal of
Climatology, 29(3), 385-416.
Fowler, H.J., and C.G. Kilsby, 2003: A regional frequency analysis of United Kingdom
extreme rainfall from 1961 to 2000. Intl Journal of Climatology, 23(11), 1313-1334.
Fowler, H.J., and R.L. Wilby, 2010: Detecting changes in seasonal precipitation
extremes using regional climate model projections: Implications for managing
fluvial flood risk. Water Resources Research, 46, W0525.
Fowler, H.J., M. Ekstrom, S. Blenkinsop, and A.P. Smith, 2007a: Estimating change in
extreme European precipitation using a multimodel ensemble. Journal of
Geophysical Research – Atmospheres, 112, D18104.
Fowler, H.J., S. Blenkinsop, and C. Tebaldi, 2007b: Linking climate change modelling
to impacts studies: recent advances in downscaling techniques for hydrological
modelling. International Journal of Climatology, 27(12), 1547-1578.
Francou, B., E. Ramirez, B. Caceres, and J. Mendoza, 2000: Glacier evolution in the
tropical Andes during the last decades of the 20th century: Chacaltaya, Bolivia,
and Antizana, Ecuador. Ambio, 29(7), 416-422.
Frappier, A.B., T. Knutson, K.B. Liu, and K. Emanuel, 2007a: Perspective: Coordinating
paleoclimate research on tropical cyclones with hurricane-climate theory and
modelling. Tellus Series A – Dynamic Meteorology and Oceanography, 59(4),
529-537.
Frappier, A.B., D. Sahagian, S.J. Carpenter, L.A. Gonzalez, and B.R. Frappier, 2007b:
Stalagmite stable isotope record of recent tropical cyclone events. Geology,
35(2), 111-114.
Frederiksen, J.S., and C.S. Frederiksen, 2007: Interdecadal changes in southern
hemisphere winter storm track modes. Tellus Series A – Dynamic Meteorology
and Oceanography
, 59(5), 599-617.
Frei, C. and C. Schär, 2001: Detection of probability of trends in rare events: Theory
and application to heavy precipitation in the Alpine region. Journal of Climate,
14(7), 1568-1584.
Frei, C., J.H. Christensen, M. Deque, D. Jacob, R.G. Jones, and P.L. Vidale, 2003: Daily
precipitation statistics in regional climate models: Evaluation and intercomparison
for the European Alps. Journal of Geophysical Research – Atmospheres, 108, 4124.
Frei, C., R. Schöll, S. Fukutome, J. Schmidli, and P.L. Vidale, 2006: Future change of
precipitation extremes in Europe: Intercomparison of scenarios from regional
climate models. Journal of Geophysical Research – Atmospheres, 111, D06105.
Frey, H., W. Haeberli, A. Linsbauer, C. Huggel, and F. Paul, 2010: A multi-level strategy
for anticipating future glacier lake formation and associated hazard potentials.
Natural Hazards and Earth System Sciences, 10(2), 339-352.
Frich, P., L.V. Alexander, P.M. Della-Marta, B. Gleason, M. Haylock, A.M.G. Klein Tank,
and T. Peterson, 2002: Observed coherent changes in climatic extremes during
the second half of the twentieth century. Climate Research, 19(3), 193-212.
Friedlingstein, P., P. Cox, R. Betts, L. Bopp, W. von Bloh, V. Brovkin, P. Cadule, S.
Doney, M. Eby, I. Fung, G. Bala, J. John, C. Jones, F. Joos, T. Kato, M. Kawamiya,
W. Knorr, K. Lindsay, H.D. Matthews, T. Raddatz, P. Rayner, C. Reick, E. Roeckner,
K.G. Schnitzler, R. Schnur, K. Strassmann, A.J. Weaver, C. Yoshikawa, and N.
Zeng, 2006: Climate-carbon cycle feedback analysis: Results from the C4MIP
model intercomparison. Journal of Climate, 19(14), 3337-3353.
Fu, C.B., S.Y. Wang, Z. Xiong, W.J. Gutowski, D.K. Lee, J.L. McGregor, Y. Sato, H. Kato,
J.W. Kim, and M.S. Suh, 2005: Regional climate model intercomparison project
for Asia. Bulletin of the American Meteorological Society, 86(2), 257-266.
Fujibe, F., N. Yamazaki, and K. Kobayashi, 2006: Long-term changes of heavy
precipitation and dry weather in Japan (1901-2004). Journal of the Meteorological
Society of Japan, 84(6), 1033-1046.
Fujihara, Y., K. Tanaka, T. Watanabe, T. Nagano, and T. Kojiri, 2008: Assessing the
impacts of climate change on the water resources of the Seyhan River Basin in
Turkey: Use of dynamically downscaled data for hydrologic simulations. Journal
of Hydrology, 353(1-2), 33-48.
Funk, C., M.D. Dettinger, J.C. Michaelsen, J.P. Verdin, M.E. Brown, M. Barlow, and A.
Hoell, 2008: Warming of the Indian Ocean threatens eastern and southern
African food security but could be mitigated by agricultural development.
Proceedings of the National Academy of Sciences, 105(32), 11081-11086.
Fyfe, J.C., 2003: Extratropical southern hemisphere cyclones: Harbingers of climate
change? Journal of Climate, 16(17), 2802-2805.
Gallant, A.J.E., and D.J. Karoly, 2010: A combined climate extremes index for the
Australian region. Journal of Climate, 23(23), 6153-6165.
Gallant, A.J.E., K.J. Hennessy, and J. Risbey, 2007: Trends in rainfall indices for six
Australian regions: 1910-2005. Australian Meteorological Magazine, 56(4),
223-239.
Gan, M.A., V.B. Rao, and M.C.L. Moscati, 2006: South American monsoon indices.
Atmospheric Science Letters, 6(4), 219-223.
García, J., M.C. Gallego, A. Serrano, and J. Vaquero, 2007: Trends in block-seasonal
extreme rainfall over the Iberian Peninsula in the second half of the twentieth
century. Journal of Climate, 20(1), 113-130.
Gastineau, G., and B.J. Soden, 2009: Model projected changes of extreme wind
events in response to global warming. Geophysical Research Letters, 36, L10810.
Gavriliev, P.P., and P.V. Efremov, 2003: Effects of cryogenic processes on Yakutian
landscapes under climate warming. In: Permafrost [Philips, M., S.M. Springman,
and L.U. Arenson (eds.)]. Proceedings of the 8th International Conference on
Permafrost, Zurich, Switzerland, 21-25 July 2003, Balkema, The Netherlands, pp.
277-282.
Gedney, N., P.M. Cox, R.A. Betts, O. Boucher, C. Huntingford, and P.A. Stott, 2006:
Detection of a direct carbon dioxide effect in continental river runoff records.
Nature, 439(7078), 835-838.
Geertsema, M., J.J. Clague, J.W. Schwab, and S.G. Evans, 2006: An overview of
recent large catastrophic landslides in northern British Columbia, Canada.
Engineering Geology, 83(1-3), 120-143.
Gehrels, W.R., D.F. Belknap, B.R. Pearce, and B. Gong, 1995: Modeling the contribution
of M(2) tidal amplification to the holocene rise of mean high water in the Gulf
of Maine and the Bay of Fundy. Marine Geology, 124(1-4), 71-85.
Geng, Q. and M. Sugi, 2001: Variability of the North Atlantic cyclone activity in winter
analyzed from NCEP–NCAR reanalysis data. Journal of Climate, 14(18), 3863-3873.
Gerten, D., S. Rost, W. von Bloh, and W. Lucht, 2008: Causes of change in 20th
century global river discharge. Geophysical Research Letters, 35, L20405.
Giannakopoulos, C., P. Le Sager, M. Bindi, M. Moriondo, E. Kostopoulou, and C.M.
Goodess, 2009: Climatic changes and associated impacts in the Mediterranean
resulting from a 2°C global warming. Global and Planetary Change, 68(3),
209-224.
Giannini, A., R. Saravanan, and P. Chang, 2003: Oceanic forcing of Sahel rainfall on
interannual to interdecadal time scales. Science, 302(5647), 1027-1030.
Giannini, A., M. Biasutti, and M.M. Verstraete, 2008: A climate model-based review
of drought in the Sahel: Desertification, the re-greening and climate change.
Global and Planetary Change, 64(3-4), 119-128.
Gibelin, A.L., and M. Deque, 2003: Anthropogenic climate change over the
Mediterranean region simulated by a global variable resolution model. Climate
Dynamics, 20(4), 327-339.
Gillett, N.P. and P.A. Stott, 2009: Attribution of anthropogenic influence on seasonal
sea level pressure. Geophysical Research Letters, 36, L23709.
Gillett
, N.P., R.J. Allan, and T.J. Ansell, 2005: Detection of external influence on sea
level pressure with a multi-model ensemble. Geophysical Research Letters, 32,
L19714.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
211
Gillett, N.P., P.A. Stott, and B.D. Santer, 2008a: Attribution of cyclogenesis region sea
surface temperature change to anthropogenic influence. Geophysical Research
Letters, 35, L09707.
Gillett, N.P., D.A. Stone, P.A. Stott, T. Nozawa, A.Y. Karpechko, G.C. Hegerl, M.F.
Wehner, and P.D. Jones, 2008b: Attribution of polar warming to human influence.
Nature Geoscience, 1, 750-754.
Ginoux, P., J.M. Prospero, O. Torres, and M. Chin, 2004: Long-term simulation of global
dust distribution with the GOCART model: correlation with North Atlantic
Oscillation. Environmental Modelling & Software, 19(2), 113-128.
Giorgi, F., 2006: Climate change hot-spots. Geophysical Research Letters, 33, L08707.
Giorgi, F. and P. Lionello, 2008: Climate change projections for the Mediterranean
region. Global and Planetary Change, 63(2-3), 90-104.
Giorgi, F., C. Jones, and G.R. Asra, 2009: Addressing climate information needs at the
regional level: the CORDEX framework. WMO Bulletin, 58(3), 175-183.
Glade, T., 1998: Establishing the frequency and magnitude of landslide-triggering
rainstorm events in New Zealand. Environmental Geology, 55(2), 160-174.
Gleason, K.L., J.H. Lawrimore, D.H. Levinson, and T.R. Karl, 2008: A revised U.S.
climate extremes index. Journal of Climate, 21(10), 2124-2137.
Gong, S.L., X.Y. Zhang, T.L. Zhao, X.B. Zhang, L.A. Barrie, I.G. McKendry, and C.S. Zhao,
2006: A simulated climatology of Asian dust aerosol and its trans-Pacific
transport. Part II: Interannual variability and climate connections. Journal of
Climate, 19(1), 104-122.
Goodkin, N.F., K.A. Hughen, S.C. Doney, and W.B. Curry, 2008: Increased multidecadal
variability of the North Atlantic Oscillation since 1781. Nature Geoscience, 1,
844-848.
Goubanova, K. and L. Li, 2007: Extremes in temperature and precipitation around
the Mediterranean basin in an ensemble of future climate scenario simulations.
Global and Planetary Change, 57(1-2), 27-42.
Goudie, A.S., 2009: Dust storms: Recent developments. Journal of Environmental
Management, 90(1), 89-94.
Goudie, A.S., and N.J. Middleton, 1992: The changing frequency of dust storms
through time. Climatic Change, 20(3), 197-225.
Grabemann, I., and R. Weisse, 2008: Climate change impact on extreme wave
conditions in the North Sea: an ensemble study. Ocean Dynamics, 58(3-4), 199-212.
Graham
, L.P., J. Andreasson, and B. Carlsson, 2007: Assessing climate change
impacts on hydrology from an ensemble of regional climate models, model
scales and linking methods – a case study on the Lule River basin. Climatic
Change, 81(S1), 293-307.
Graham, N.E. and H.F. Diaz, 2001: Evidence for intensification of North Pacific winter
cyclones since 1948. Bulletin of the American Meteorological Society, 82(9),
1869-1893.
Gravelle, G. and N. Mimura, 2008: Vulnerability assessment of sea-level rise in Viti
Levu, Fiji Islands. Sustainability Science, 3(2), 171-180.
Green, D., L. Alexander, K.L. McInnes, J. Church, N. Nicholls, and W. White, 2009: An
assessment of climate change impacts and adaptation for the Torres Strait
Islands, Australia. Climatic Change, 102(3-4), 405-433.
Greene, A.M., A. Giannini, and S.E. Zebiak, 2009: Drought return times in the Sahel:
A question of attribution. Geophysical Research Letters, 36, L12701.
Greeves, C.Z., V.D. Pope, R.A. Stratton, and G.M. Martin, 2007: Representation of
Northern Hemisphere winter storm tracks in climate models. Climate Dynamics,
28(7-8), 683-702.
Grell, G.A., L. Schade, R. Knoche, A. Pfeiffer, and J. Egger, 2000: Nonhydrostatic
climate simulations of precipitation over complex terrain. Journal of
Geophysical Research – Atmospheres, 105(D24), 29595-29608.
Griffiths, G.M., M.J. Salinger, and I. Leleu, 2003: Trends in extreme daily rainfall
across the South Pacific and relationship to the South Pacific Convergence
Zone. International Journal of Climatology, 23(8), 847-869.
Griffiths, G.M., L.E. Chambers, M.R. Haylock, M.J. Manton, N. Nicholls, H.J. Baek, Y.
Choi, P.M. Della-Marta, A. Gosai, N. Iga, R. Lata, V. Laurent, L. Maitrepierre, H.
Nakamigawa, N. Ouprasitwong, D. Solofa, L. Tahani, D.T. Thuy, L. Tibig, B. Trewin,
K. Vediapan, and P. Zhai, 2005: Change in mean temperature as a predictor of
extreme temperature change in the Asia-Pacific region. International Journal of
Climatology, 25(10), 1301-1330.
Grimm, A.M. and R.G. Tedeschi, 2009: ENSO and extreme rainfall events in South
America. Journal of Climate, 22(7), 1589-1609.
Grimm, A.M., J. Pal, and F. Giorgi, 2007: Connection between spring conditions and
peak summer monsoon rainfall in South America: Role of soil moisture, surface
temperature, and topography in eastern Brazil. Journal of Climate, 20(24),
5929-5945.
Grinsted, A., A.J. Moore, and S. Jevrejeva, 2010: Reconstructing sea level from paleo
and projected temperatures 200 to 2100 AD. Climate Dynamics, 34(4), 461-472.
Groisman, P.Y., and R.W. Knight, 2008: Prolonged dry episodes over the conterminous
United States: New tendencies emerging during the last 40 years. Journal of
Climate, 21(9), 1850-1862.
Gruber, S., and W. Haeberli, 2007: Permafrost in steep bedrock slopes and its
temperature-related destabilization following climate change. Journal of
Geophysical Research – Earth Surface, 112, F02S18.
Gruber, S., M. Hoelzle, and W. Haeberli, 2004: Permafrost thaw and destabilization
of Alpine rock walls in the hot summer of 2003. Geophysical Research Letters,
31, L13504.
Gualdi, S., E. Scoccimarro, and A. Navarra, 2008: Changes in tropical cyclone activity
due to global warming: Results from a high-resolution coupled general circulation
model. Journal of Climate, 21(20), 5204-5228.
Guilyardi, E., A. Wittenberg, A. Fedorov, M. Collins, C.Z. Wang, A. Capotondi, G.J. van
Oldenborgh, and T. Stockdale, 2009: Understanding El Niño in ocean-
atmosphere General Circulation Models: Progress and challenges. Bulletin of
the American Meteorological Society, 90(3), 325-340.
Gulev, S.K. and V. Grigorieva, 2004: Last century changes in ocean wind wave height
from global visual wave data. Geophysical Research Letters, 31, L24302.
Gulev, S.K., O. Zolina, and S. Grigoriev, 2001: Extratropical cyclone variability in the
Northern Hemisphere winter from the NCEP/NCAR reanalysis data. Climate
Dynamics, 17(10), 795-809.
Guo, H., M. Xu, and Q. Hu, 2011: Changes in near-surface wind speed in China:
1969–2005. International Journal of Climatology, 31(2), 349-358.
Gutowski, W.J., L.E. Branscome, and D.A. Stewart, 1992: Life-cycles of moist
baroclinic eddies. Journal of the Atmospheric Sciences, 49(4), 306-319.
Gutowski, W.J., S.G. Decker, R.A. Donavon, Z.T. Pan, R.W. Arritt, and E.S. Takle, 2003:
Temporal-spatial scales of observed and simulated precipitation in central U.S.
climate. Journal of Climate, 16(22), 3841-3847.
Gutowski, W.J., G.C. Hegerl, G.J. Holland, T.R. Knutson, L.O. Mearns, R.J. Stouffer, P.J.
Webster, M.F. Wehner, and F.W. Zwiers, 2008a: Causes of observed changes in
extremes and projections of future changes. In: Weather and Climate Extremes
in a Changing Climate. Regions of Focus: North America, Hawaii, Caribbean,
and U.S. Pacific Islands. [Karl, T.R., G.A. Meehl, D.M. Christopher, S.J. Hassol, A.M.
Waple, and W.L. Murray (eds.)]. A Report by the U.S. Climate Change Science
Program and the Subcommittee on Global Change Research, Washington, DC,
pp. 81-116.
Gutowski, W.J., S.S. Willis, J.C. Patton, B.R.J. Schwedler, R.W. Arritt, and E.S. Takle,
2008b: Changes in extreme, cold-season synoptic precipitation events under
global warming. Geophysical Research Letters, 35, L20710.
Haarsma, R.J., F. Selten, B.V. Hurk, W. Hazeleger, and X.L. Wang, 2009: Drier
Mediterranean soils due to greenhouse warming bring easterly winds over
summertime central Europe. Geophysical Research Letters, 36, L04705.
Haeberli, W., 1983: Frequency and characteristics of glacier floods in the Swiss Alps.
Annals of Glaciology, 4, 85-90.
Haeberli, W., and C.R. Burn, 2002: Natural hazards in forests: glacier and permafrost
effects as related to climate change. In: Environmental Change and Geomorphic
Hazards in Forests, IUFRO Research Series 9 [Sidle, R.C. (ed.)]. CABI Publishing,
Wallingford, UK, pp. 167-202.
Haeberli, W. and M. Funk, 1991: Borehole temperatures at the Colle Gnifetti core-
drilling site (Monte Rosa, Swiss Alps). Journal of Glaciology, 37(125), 37-46.
Haeberli, W. and R. Hohmann, 2008: Climate, glaciers and permafrost in the Swiss
Alps 2050: scenarios, consequences and recommendations. In: Permafrost on a
Warming Planet: Impacts on Ecosystems, Infrastructure and Climate [Kane, D.L.,
and K.M. Hinkel (eds.)]. Proc. of the 9th Intl. Conf. on Permafrost at University of
Alaska, Fairbanks, 9 June - 3 July 2008, Inst. of Northern Engineering, pp. 607-612.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
212
Haeberli, W., M. Wegmann, and D. Vonder Mühll, 1997: Slope stability problems
related to glacier shrinkage and permafrost degradation in the Alps. Eclogae
Geologicae Helvetiae, 90(3), 407-414.
Haeberli, W., A. Kääb, D. Vonder Mühll, and P. Teysseire, 2001: Prevention of outburst
floods from periglacial lakes at Grubengletscher, Valais, Swiss Alps. Journal of
Glaciology, 47(156), 111-122.
Haeberli, W., C. Huggel, A. Kääb, S. Zgraggen-Oswald, A. Polkvoj, I. Galushkin, I.
Zotikov, and N. Osokin, 2004: The Kolka-Karmadon rock/ice slide of 20
September 2002: an extraordinary event of historical dimensions in North
Ossetia, Russian Caucasus. Journal of Glaciology, 50(171), 533-546.
Haerter, J.O. and P. Berg, 2009: Unexpected rise in extreme precipitation caused by
a shift in rain type? Nature Geoscience, 2(6), 372-373.
Haigh, I., R. Nicholls, and N. Wells, 2009: Mean sea level trends around the English
Channel over the 20th century and their wider context. Continental Shelf
Research, 29(17), 2083-2098.
Haigh, I., R. Nicholls, and N. Wells, 2010: Assessing changes in extreme sea levels:
Application to the English Channel, 1900–2006. Continental Shelf Research,
30(9), 1042-1055.
Hall, A., X. Qu, and J.D. Neelin, 2008: Improving predictions of summer climate
change in the United States. Geophysical Research Letters, 35, L01702.
Hampel, A., R. Hetzel, and G. Maniatis, 2010: Response of faults to climate-driven
changes in ice and water volumes on Earth’s surface. Philosophical Transactions
of the Royal Society A: Mathematical, Physical and Engineering Sciences,
368(1919), 2501-2517.
Hanna, E., J. Cappelen, R. Allan, T. Jonsson, F. Le Blancq, T. Lillington, and K. Hickey,
2008: New insights into North European and North Atlantic surface pressure
variability, storminess, and related climatic change since 1830. Journal of
Climate, 21(24), 6739-6766.
Hannaford, J., and T.J. Marsh, 2008: High-flow and flood trends in a network of
undisturbed catchments in the UK. International Journal of Climatology,
28(10), 1325-1338.
Hanson, C.E., J.P. Palutikof, and T.D. Davies, 2004: Objective cyclone climatologies of
the North Atlantic – a comparison between the ECMWF and NCEP Reanalyses.
Climate Dynamics, 22(6-7), 757-769.
Harper, B., T. Hardy, L. Mason, and R. Fryar, 2009: Developments in storm tide modelling
and risk assessment in the Australian region. Natural Hazards, 51(1), 225-238.
Harris, C., L.U. Arenson, H.H. Christiansen, B. Etzelmüller, R. Frauenfelder, S. Gruber,
W. Haeberli, C. Hauck, M. Hölzle, O. Humlum, K. Isaksen, A. Kääb, M.A. Kern-
Lütschg, M. Lehning, N. Matsuoka, J.B. Murton, J. Nötzli, M. Phillips, N. Ross, M.
Seppälä, S.M. Springman, and D. Vonder Mühll, 2009: Permafrost and climate in
Europe: Monitoring and modelling thermal, geomorphological and geotechnical
responses. Earth-Science Reviews, 92(3-4), 117-171.
Haug, G.H., K.A. Hughen, D.M. Sigman, L.C. Peterson, and U. Röhl, 2001: Southward
migration of the Intertropical Convergence Zone through the Holocene.
Science, 293(5533), 1304-1308.
Haugen, J.E. and T. Iversen, 2008: Response in extremes of daily precipitation and
wind from a downscaled multi-model ensemble of anthropogenic global climate
change scenarios. Tellus Series A – Dynamic Meteorology and Oceanography,
60(3), 411-426.
Hawkins, E. and R. Sutton, 2009: The potential to narrow uncertainty in regional climate
predictions. Bulletin of the American Meteorological Society, 90(8), 1095-1107.
Hawkins, E. and R. Sutton, 2011: The potential to narrow uncertainty in projections
of regional precipitation change. Climate Dynamics, 37(1-2), 407-418.
Hay, L.E., M.P. Clark, M. Pagowski, G.H. Leavesley, and W.J. Gutowski, 2006: One-way
coupling of an atmospheric and a hydrologic model in Colorado. Journal of
Hydrometeorology, 7(4), 569-589.
Haylock, M.R. and C.M. Goodess, 2004: Interannual variability of European extreme
winter rainfall and links with mean large-scale circulation. International Journal
of Climatology, 24(6), 759-776.
Haylock, M.R., G.C. Cawley, C. Harpham, R.L. Wilby, and C.M. Goodess, 2006a:
Downscaling heavy precipitation over the United Kingdom: A comparison of
dynamical and statistical methods and their future scenarios. International
Journal of Climatology, 26(10), 1397-1415.
Haylock, M.R., T.C. Peterson, L.M. Alves, T. Ambrizzi, Y.M.T. Anunciacao, J. Baez, V.R.
Barros, M.A. Berlato, M. Bidegain, G. Coronel, V. Corradi, V.J. Garcia, A.M.
Grimm, D. Karoly, J.A. Marengo, M.B. Marino, D.F. Moncunill, D. Nechet, J.
Quintana, E. Rebello, M. Rusticucci, J.L. Santos, I. Trebejo, and L.A. Vincent,
2006b: Trends in total and extreme South American rainfall in 1960-2000 and
links with sea surface temperature. Journal of Climate, 19(8), 1490-1512.
Haylock, M.R., N. Hofstra, A. Tank, E.J. Klok, P.D. Jones, and M. New, 2008: A
European daily high-resolution gridded data set of surface temperature and
precipitation for 1950-2006. Journal of Geophysical Research – Atmospheres,
113(D20119).
Hegerl, G.C., F.W. Zwiers, P.A. Stott, and V.V. Kharin, 2004: Detectability of
anthropogenic changes in annual temperature and precipitation extremes.
Journal of Climate, 17(19), 3683-3700.
Hegerl, G.C., F.W. Zwiers, P. Braconnot, N.P. Gillett, Y. Luo, J.A. Marengo Orsini, N.
Nicholls, J.E. Penner, and P.A. Stott, 2007: Understanding and Attributing
Climate Change. In: Climate Change 2007: The Physical Science Basis.
Contribution of Working Group I to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change [Solomon, S., D. Qin, M. Manning,
Z. Chen, M. Marquis, K.B. Averyt, M. Tignor and H.L. Miller (eds.)]. Cambridge
University Press, Cambridge, UK, pp. 663-745.
Hegerl, G.C., O. Hoegh-Guldberg, G. Casassa, M.P. Hoerling, R.S. Kovats, C. Parmesan,
D.W. Pierce, and P.A. Stott, 2010: Good practice guidance paper on detection
and attribution related to anthropogenic climate change. In: Meeting Report of
the Intergovernmental Panel on Climate Change Expert Meeting on Detection
and Attribution of Anthropogenic Climate Change [Stocker, T.F., C.B. Field, D.
Qin, V. Barros, G.-K. Plattner, M. Tignor, P.M. Midgley, and K.L. Ebi (eds.)]. IPCC
Working Group I Technical Support Unit, University of Bern, Bern, Switzerland.
Hegglin, E. and C. Huggel, 2008: An integrated assessment of vulnerability to glacial
hazards. Mountain Research and Development, 28(3), 299-309.
Heim Jr., R.R., 2002: A review of twentieth-century drought indices used in the
United States. Bulletin of the American Meteorological Society, 83(8), 1149-1165.
Held, I.M. and B.J. Soden, 2006: Robust responses of the hydrological cycle to global
warming.
Journal of Climate, 19(21), 5686-5699.
Hemer, M.A., 2010: Historical trends in Southern Ocean storminess: Long-term
variability of extreme wave heights at Cape Sorell, Tasmania. Geophysical
Research Letters, 37, L18601.
Hemer, M.A., J.A. Church, and J.R. Hunter, 2010: Variability and trends in the
directional wave climate of the Southern Hemisphere. International Journal of
Climatology, 30(4), 475-491.
Hendon, H.H., D.W.J. Thompson, and M.C. Wheeler, 2007: Australian rainfall and
surface temperature variations associated with the Southern Hemisphere annular
mode. Journal of Climate, 20(11), 2452-2467.
Herath, S. and U. Ratnayake, 2004: Monitoring rainfall trends to predict adverse
impacts – a case study from Sri Lanka (1964-1993). Global Environmental
Change – Human and Policy Dimensions, 14(1), 71-79.
Hewitson, B.C. and R.G. Crane, 2006: Consensus between GCM climate change
projections with empirical downscaling: Precipitation downscaling over South
Africa. International Journal of Climatology, 26(10), 1315-1337.
Hidalgo, H.G., T. Das, M.D. Dettinger, D.R. Cayan, D.W. Pierce, T.P. Barnett, G. Bala, A.
Mirin, A.W. Wood, C. Bonfils, B.D. Santer, and T. Nozawa, 2009: Detection and
attribution of streamflow timing changes to climate change in the western
United States. Journal of Climate, 22(13), 3838-3855.
Hinkel, J. and R.J.T. Klein, 2009: Integrating knowledge to assess coastal vulnerability
to sea-level rise: The development of the DIVA tool. Global Environmental
Change – Human and Policy Dimensions, 19(3), 384-395.
Hinzman, L.D., N.D. Bettez, W.R. Bolton, F.S. Chapin, M.B. Dyurgerov, B.G. Fastie, B.
Griffith, R.D. Hollister, A. Hope, H.P. Huntington, A.M. Jenson, G.J. Jia, R.
Jorgenson, D. Kane, D.R. Klein, G. Kofinas, A.H. Lynch, A.H. Lloyd, D. McGuire, F.E.
Nelson, W.C. Oechel, T.E. Osterkamp, C.H. Racine, V.E. Romanovsky, R.S. Stone,
D.A. Stow, M. Sturm, C.E. Tweedies, G.L. Vourlitis, M.D. Walker, D.A. Walker, P.J.
Webber, J.M. Welker, K.S. Winker, and K. Yoshikawa, 2005: Evidence and
implications of recent climate change in northern Alaska and other arctic
regions. Climatic Change, 72(3), 251-298.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
213
Hirabayashi, Y., S. Kanae, K. Motoya, K. Masuda, and P. Doll, 2008a: A 59-year
(1948-2006) global meteorological forcing data set for land surface models.
Part II: Global snowfall estimation. Hydrological Research Letters, 2, 65-69.
Hirabayashi, Y., S. Kanae, S. Emori, T. Oki, and M. Kimoto, 2008b: Global projections
of changing risks of floods and droughts in a changing climate. Hydrological
Sciences Journal, 53(4), 754-772.
Hirsch, M.E., A.T. DeGaetano, and S.J. Colucci, 2001: An East Coast winter storm
climatology. Journal of Climate, 14(5), 882-899.
Hirschi, M., S.I. Seneviratne, V. Alexandrov, F. Boberg, C. Boroneant, O.B. Christensen,
H. Formayer, B. Orlowsky, and P. Stepanek, 2011: Observational evidence for
soil-moisture impact on hot extremes in southeastern Europe. Nature
Geoscience, 4(1), 17-21.
Hodges, K.I., B.J. Hoskins, J. Boyle, and C. Thorncroft, 2003: A comparison of recent
reanalysis datasets using objective feature tracking: Storm tracks and tropical
easterly waves. Monthly Weather Review, 131(9), 2012-2037.
Hoegh-Guldberg, O., 1999: Climate change, coral bleaching and the future of the
world’s coral reefs. Marine and Freshwater Research, 50(8), 839-866.
Hoelzle, M., G. Darms, and S. Suter, 2010: Evidence of accelerated englacial warming
in the Monte Rosa area, Switzerland/Italy. The Cryosphere Discussions, 4(4),
2277-2305.
Hoerling, M., J. Hurrell, J. Eischeid, and A. Phillips, 2006: Detection and attribution of
twentieth-century northern and southern African rainfall change. Journal of
Climate, 19(16), 3989-4008.
Hohenegger, C., P. Brockhaus, and C. Schär, 2008: Towards climate simulations at
cloud-resolving scales. Meteorology Zeitschrift, 17(4), 383-394.
Hohenegger, C., P. Brockhaus, C.S. Bretherton, and C. Schar, 2009: The soil moisture-
precipitation feedback in simulations with explicit and parameterized convection.
Journal of Climate, 22(19), 5003-5020.
Holland, G.J. and P.J. Webster, 2007: Heightened tropical cyclone activity in the North
Atlantic: Natural variability or climate trend? Philosophical Transactions of the
Royal Society A, 365, 2695-2716.
Hope, P.K., W. Drosdowsky, and N. Nicholls, 2006: Shifts in the synoptic systems
influencing southwest Western Australia. Climate Dynamics, 26(7-8), 751-764.
Horton, R., C. Herweijer, C. Rosenzweig, J. Liu, V. Gornitz, and A.C. Ruane, 2008: Sea
level rise projections for current generation CGCMs based on the semi-empirical
method. Geophysical Research Letters, 35, L02715.
Hossain, F., I. Jeyachandran, and R. Pielke Jr, 2009: Have large dams altered extreme
precipitation patterns. EOS Transactions (AGU), 90(48), 453.
Hourdin, F., I. Musat, F. Guichard, P.M. Ruti, F. Favot, M.A. Filiberti, M. Pham, J.Y.
Grandpeix, J. Polcher, P. Marquet, A. Boone, J.P. Lafore, J.L. Redelsperger,
A. Dell’aquila, T.L. Doval, A.K. Traore, and H. Gallee, 2010: AMMA-model
intercomparison project. Bulletin of the American Meteorological Society,
91(1), 95-104.
Huebener, H., U. Cubasch, U. Langematz, T. Spangehl, F. Niehoerster, I. Fast, and M.
Kunze, 2007: Ensemble climate simulations using a fully coupled ocean-
troposphere-stratosphere general circulation model. Philosophical Transactions
of the Royal Society A, 365(1857), 2089-2101.
Huggel, C., 2009: Recent extreme slope failures in glacial environments: effects of
thermal perturbation. Quaternary Science Reviews, 28(11-12), 1119-1130.
Huggel, C., W. Haeberli, A. Kääb, D. Bieri, and S. Richardson, 2004: An assessment
procedure for glacial hazards in the Swiss Alps. Canadian Geotechnical Journal,
41(6), 1068-1083.
Huggel, C., S. Zgraggen-Oswald, W. Haeberli, A. Kääb, A. Polkvoj, I. Galushkin, and
S.G. Evans, 2005: The 2002 rock/ice avalanche at Kolka/Karmadon, Russian
Caucasus: assessment of extraordinary avalanche formation and mobility, and
application of QuickBird satellite imagery. Natural Hazards and Earth System
Sciences, 5(2), 173-187.
Huggel, C., J. Caplan-Auerbach, C.F. Waythomas, and R.L. Wessels, 2007: Monitoring
and modeling ice-rock avalanches from ice-capped volcanoes: A case study of
frequent large avalanches on Iliamna Volcano, Alaska. Journal of Volcanology
and Geothermal Research, 168(1-4), 114-136.
Huggel, C., J. Caplan-Auerbach, and R. Wessels, 2008: Recent extreme avalanches:
Triggered by climate change. Eos Transactions (AGU), 89(47), 469-470.
Huggel, C., N. Salzmann, S. Allen, J. Caplan-Auerbach, L. Fischer, W. Haeberli, C.
Larsen, D. Schneider, and R. Wessels, 2010: Recent and future warm extreme
events and high-mountain slope failures. Philosophical Transactions of the
Royal Society A, 368, 2435-2459.
Hulme, M., R. Doherty, T. Ngara, M. New, and D. Lister, 2001: African climate change:
1900-2100. Climate Research, 17(2), 145-168.
Hundecha, Y. and A. Bardossy, 2008: Statistical downscaling of extreme of daily
precipitation and temperature and construction of their future scenarios.
International Journal of Climatology, 28(5), 589-610.
Hundecha, Y., A. St-Hilaire, T.B.M.J. Ouarda, S. El Adlouni, and P. Gachon, 2008: A
nonstationary extreme value analysis for the assessment of changes in extreme
annual wind speed over the Gulf of St. Lawrence, Canada. Journal of Applied
Meteorology and Climatology, 47(11), 2745-2759.
Hungate, B.A., J.S. Dukes, M.R.L.Y. Shaw, and C.B. Field, 2003: Nitrogen and climate
change. Science, 302(5650), 1512-1513.
Hunter, J., 2010: Estimating sea-level extremes under conditions of uncertain sea-
level rise. Climatic Change,
99(3-4), 331-350.
Huss, M., D. Farinotti, A. Bauder, and M. Funk, 2008: Modelling runoff from highly
glacierized alpine drainage basins in a changing climate. Hydrological
Processes, 22(19), 3888-3902.
Im, E.S. and W.-T. Kwon, 2007: Characteristics of extreme climate sequences over
Korea using a regional climate change scenario. SOLA, 3, 17-20.
Im, E.S., J.B. Ahn, W.T. Kwon, and F. Giorgi, 2008: Multi-decadal scenario simulation
over Korea using a one-way double-nested regional climate model system. Part
2: Future climate projection (2021-2050). Climate Dynamics, 30(2-3), 239-254.
Im, E.S., I.W. Jung, and D.H. Bae, 2011: The temporal and spatial structures of recent and
future trends in extreme indices over Korea from a regional climate projection.
International Journal of Climatology, 31(1), 72-86.
Inoue, J., J.P. Liu, J.O. Pinto, and J.A. Curry, 2006: Intercomparison of Arctic Regional
Climate Models: Modeling clouds and radiation for SHEBA in May 1998. Journal
of Climate, 19(17), 4167-4178.
IPCC, 2007a: Climate Change 2007: The Physical Science Basis. Contribution of
Working Group I to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M.
Marquis, K.B. Averyt, M. Tignor, and H.L. Miller (eds.)]. Cambridge University
Press, Cambridge, UK, 996 pp.
IPCC, 2007b: Summary for Policymakers. In: Climate Change 2007: The Physical
Science Basis. Contribution of Working Group I to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Solomon, S., D. Qin, M.
Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor and H.L. Miller (eds.)].
Cambridge University Press, Cambridge, UK, pp. 1-18.
Irish, J.L. and D.T. Resio, 2010: A hydrodynamics-based surge scale for hurricanes.
Ocean Engineering, 37(1), 69-81.
Iverson, R.M., 2000: Landslide triggering by rain infiltration. Water Resources
Research, 36(7), 1897-1910.
Jakob, D., 2010: Challenges in developing a high-quality surface wind-speed data set
for Australia. Australian Meteorological and Oceanographic Journal, 60(4),
227-236.
Jakob, M. and H. Weatherly, 2003: A hydroclimatic threshold for landslide initiation
on the North Shore Mountains of Vancouver, British Columbia. Geomorphology,
54(3-4), 137-156.
Janicot, S., 2009: A comparison of Indian and African monsoon variability at different
time scales. Comptes Rendus Geoscience, 341(8-9), 575-590.
Jansen, E., J. Overpeck, K.R. Briffa, J.C. Duplessy, F. Joos, V. Masson-Delmotte, V. Olago,
B. Otto-Bliesner, W.R. Peltier, S. Rahmstorf, R. Ramesh, D. Raynaud, D. Rind, O.
Solomina, R. Villalba, and D. Zhang, 2007: Palaeoclimate. In: Climate Change
2007: The Physical Science Basis. Contribution of Working Group I to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change
[Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor
and H.L. Miller (eds.)]. Cambridge University Press, Cambridge, UK, pp. 433-497.
Jiang, T., Z.W. Kundzewicz, and B. Su, 2008: Changes in monthly precipitation and
flood hazard in the Yangtze River Basin, China. International Journal of
Climatology, 28(11), 1471-1481.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
214
Jiang, Y., Y. Luo, Z.C. Zhao, and S.W. Tao, 2010a: Changes in wind speed over China
during 1956-2004. Theoretical and Applied Climatology, 99(3-4), 421-430.
Jiang, Y., Y. Luo, Z.C. Zhao, Y. Shi, Y.L. Xu, and J.H. Zhu, 2010b: Projections of wind
changes for 21st century in China by three regional climate models. Chinese
Geographical Science, 20(3), 226-235.
Johnson, N.C. and S.-P. Xie, 2010: Changes in the sea surface temperature threshold
for tropical convection. Nature Geoscience, 3(12), 842-845.
Joly, M. and A. Voldoire, 2010: Role of the Gulf of Guinea in the inter-annual variability
of the West African monsoon: what do we learn from CMIP3 coupled simulations?
International Journal of Climatology, 30(12), 1843-1856.
Jomelli, V., V.P. Pech, C. Chochillon, and D. Brunstein, 2004: Geomorphic variations of
debris flows and recent climatic change in the French Alps. Climatic Change,
64(1), 77-102.
Jones, B.M., C.D. Arp, M.T. Jorgenson, K.M. Hinkel, J.A. Schmutz, and P.L. Flint, 2009:
Increase in the rate and uniformity of coastline erosion in Arctic Alaska.
Geophysical Research Letters, 36, L03503.
Jones, G.S., P.A. Stott, and N. Christidis, 2008: Human contribution to rapidly increasing
frequency of very warm Northern Hemisphere summers. Journal of Geophysical
Research, 113, D02109.
Jorgenson, M.T., Y.L. Shur, and E.R. Pullman, 2006: Abrupt increase in permafrost
degradation in Arctic Alaska. Geophysical Research Letters, 33, L02503.
Joseph, R. and S. Nigam, 2006: ENSO evolution and teleconnections in IPCC’s
twentieth-century climate simulations: realistic representation? Journal of Climate,
19(17), 4360-4377.
Jull, M. and D. McKenzie, 1996: The effect of deglaciation on mantle melting
beneath Iceland. Journal of Geophysical Research – Solid Earth, 101(B10),
21815-21828.
Jung, M., M. Reichstein, P. Ciais, S.I. Seneviratne, J. Sheffield, M.L. Goulden, G. Bonan,
A. Cescatti, J.Q. Chen, R. de Jeu, A.J. Dolman, W. Eugster, D. Gerten, D. Gianelle,
N. Gobron, J. Heinke, J. Kimball, B.E. Law, L. Montagnani, Q.Z. Mu, B. Mueller, K.
Oleson, D. Papale, A.D. Richardson, O. Roupsard, S. Running, E. Tomelleri, N.
Viovy, U. Weber, C. Williams, E. Wood, S. Zaehle, and K. Zhang, 2010: Recent
decline in the global land evapotranspiration trend due to limited moisture
supply. Nature, 467(7318), 951-954.
Jung, T., S.K. Gulev, I. Rudeva, and V. Soloviov, 2006: Sensitivity of extratropical
cyclone characteristics to horizontal resolution in the ECMWF model. Quarterly
Journal of the Royal Meteorological Society, 132(619), 1839-1857.
Kääb, A., R. Frauenfelder, and I. Roer, 2007: On the response of rockglacier creep to
surface temperature increase. Global and Planetary Change, 56(1-2), 172-187.
Kallis, G., 2008: Droughts. Annual Review of Environment and Resources
, 33, 85-118.
Kamiguchi, K., A. Kitoh, T. Uchiyama, R. Mizuta, and A. Noda, 2006: Changes in
precipitation-based extremes indices due to global warming projected by a
global 20-km-mesh atmospheric model. SOLA, 2, 64-67.
Kanada, S., M. Nakano, and T. Kato, 2010a: Changes in mean atmospheric structures
around Japan during July due to global warming in regional climate
experiments using a cloud system resolving model. Hydrological Research Letters,
4, 11-14.
Kanada, S., M. Nakano, and T. Kato, 2010b: Climatological characteristics of daily
precipitation over Japan in the Kakushin regional climate experiments using a
non-hydrostatic 5-km-mesh model: Comparison with an outer global 20-km-
mesh atmospheric climate model. SOLA, 6, 117-120.
Kang, I.S. and J. Shukla, 2006: Dynamic seasonal prediction and predictability of the
monsoon. In: The Asian Monsoon [Wang, B. (ed.)]. Springer/Praxis, New York,
NY, pp. 585-612.
Karl, T.R., G.A. Meehl, D.M. Christopher, S.J. Hassol, A.M. Waple, and W.L. Murray,
2008: Weather and Climate Extremes in a Changing Climate. Regions of Focus:
North America, Hawaii, Caribbean, and U.S. Pacific Islands. A Report by the U.S.
Climate Change Science Program and the Subcommittee on Global Change
Research, Washington, DC, 164 pp.
Karl, T.R., J.M. Melillo, and T.C. Peterson (eds.), 2009: Global Climate Change Impacts
in the United States. Cambridge University Press, New York, NY, 189 pp.
Karoly, D.J. and Q.G. Wu, 2005: Detection of regional surface temperature trends.
Journal of Climate, 18(21), 4337-4343.
Kaser, G., J.G. Cogley, M.B. Dyurgerov, M.F. Meier, and A. Ohmura, 2006: Mass balance
of glaciers and ice caps: Consensus estimates for 1961-2004. Geophysical
Research Letters, 33, L19501.
Kay, A.L., H.N. Davies, V.A. Bell, and R.G. Jones, 2009: Comparison of uncertainty
sources for climate change impacts: flood frequency in England. Climatic
Change, 92(1-2), 41-63.
Kay, G. and R. Washington, 2008: Future southern African summer rainfall variability
related to a southwest Indian Ocean dipole in HadCM3. Geophysical Research
Letters, 35, L12701.
Keeling, R.F., A. Kortzinger, and N. Gruber, 2010: Ocean deoxygenation in a warming
world. Annual Review of Marine Science, 2, 199-229.
Keenlyside, N.S., M. Latif, J. Jungclaus, L. Kornblueh, and E. Roeckner, 2008:
Advancing decadal-scale climate prediction in the North Atlantic sector. Nature,
453, 84-88.
Kellogg, C.A. and D.W. Griffin, 2006: Aerobiology and the global transport of desert
dust. Trends in Ecology & Evolution, 21(11)
, 638-644.
Kendon, E.J., D.P. Rowell, R.G. Jones, and E. Buonomo, 2008: Robustness of future
changes in local precipitation extremes. Journal of Climate, 21(17), 4280-4297.
Kendon, E.J., D.P. Rowell, and R.G. Jones, 2010: Mechanisms and reliability of future
projected changes in daily precipitation. Climate Dynamics, 35(2-3), 489-509.
Kenyon, J., and G.C. Hegerl, 2008: Influence of modes of climate variability on global
temperature extremes. Journal of Climate, 21(15), 3872-3889.
Kershaw, J.A., J.J. Clague, and S.G. Evans, 2005: Geomorphic and sedimentological
signature of a two-phase outburst flood from moraine-dammed Queen Bess
Lake, British Columbia, Canada. Earth Surface Processes and Landforms, 30(1),
1-25.
Kharin, V., F.W. Zwiers, X. Zhang, and G.C. Hegerl, 2007: Changes in temperature and
precipitation extremes in the IPCC ensemble of global coupled model simulations.
Journal of Climate, 20(8), 1419-1444.
Khon, V.C., I.I. Mokhov, E. Roeckner, and V.A. Semenov, 2007: Regional changes of
precipitation characteristics in Northern Eurasia from simulations with global
climate model. Global and Planetary Change, 57(1-2), 118-123.
Kidston, J., and E.P. Gerber, 2010: Intermodel variability of the poleward shift of the
austral jet stream in the CMIP3 integrations linked to biases in 20th century
climatology. Geophysical Research Letters, 37(9), L09708.
Kiktev, D., D.M.H. Sexton, L. Alexander, and C.K. Folland, 2003: Comparison of
modeled and observed trends in indices of daily climate extremes. Journal of
Climate, 16(22), 3560-3571.
Kiktev, D.M., J. Caesar, and L. Alexander, 2009: Temperature and precipitation
extremes in the second half of the twentieth century from numerical modeling
results and observational data. Izvestiya Atmospheric and Oceanic Physics, 45,
284-293.
Kilsby, C., P. Jones, A. Burton, A. Ford, H. Fowler, C. Harpham, P. James, A. Smith, and
R. Wilby, 2007: A daily weather generator for use in climate change studies.
Environmental Modelling and Software, 22(12), 1705-1719.
Kim, H.J., R.C. Sidle, R.D. Moore, and R. Hudson, 2004: Throughflow variability during
snowmelt in a forested mountain catchment, coastal British Columbia, Canada.
Hydrological Processes, 18(7), 1219-1236.
Kim, H.J., B. Wang, and Q.H. Ding, 2008: The global monsoon variability simulated by
CMIP3 coupled climate models. Journal of Climate, 21(20), 5271-5294.
Kim, S., E. Nakakita, Y. Tachikawa, and K. Takara, 2010: Precipitation changes in Japan
under the A1B climate change scenario. Annual Journal of Hydraulic
Engineering, JSCE, 54, 127-132.
Kim, S.K., W.P. Hong, and Y.M. Kim, 1991: Prediction of rainfall-triggered landslides in
Korea. In:
Proceedings of the 6th International Symposium on Landslides,
Christchurch, New Zealand [Bell, D.H. (ed.)]. Balkema, Rotterdam, The
Netherlands, pp. 989-994.
Kimoto, M., N. Yasutomi, C. Yokoyama, and S. Emori, 2005: Projected changes in
precipitation characteristics around Japan under the global warming. SOLA, 1,
85-88.
King’uyu, S.M., L.A. Ogallo, and E.K. Anyamba, 2000: Recent trends of minimum and
maximum surface temperatures over eastern Africa. Journal of Climate, 13(16),
2876-2886.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
215
Kitoh, A. and T. Uchiyama, 2006: Changes in onset and withdrawal of the East Asian
summer rainy season by multi-model global warming experiments. Journal of
the Meteorological Society of Japan, 84(2), 247-258.
Kitoh, A., T. Ose, K. Kurihara, S. Kusunoki, M. Sugi, and KAKUSHIN Team-3 Modeling
Group, 2009: Projection of changes in future weather extremes using super-
high-resolution global and regional atmospheric models in the KAKUSHIN
Program: Results of preliminary experiments. Hydrological Research Letters, 3,
49-53.
Kitoh, A., S. Kusunoki, and T. Nakaegawa, 2011: Climate change projections over
South America in the late 21st century with the 20 and 60 km mesh
Meteorological Research Institute atmospheric general circulation model (MRI-
AGCM). Journal of Geophysical Research – Atmospheres, 116, D06105.
Kjellström, E. and P. Lind, 2009: Changes in the water budget in the Baltic Sea
drainage basin in future warmer climates as simulated by the regional climate
model RCA3. Boreal Environment Research, 14(1), 114-124.
Kjellström, E., L. Bärring, D. Jacob, R. Jones, G. Lenderink, and C. Schär, 2007:
Modelling daily temperature extremes: recent climate and future changes over
Europe. Climatic Change, 81(S1), 249-265.
Klein Tank, A.M.G., and G.P. Können, 2003: Trends in indices of daily temperature
and precipitation extremes in Europe, 1946-1999. Journal of Climate, 16(22),
3665-3680.
Klein Tank, A.M.G., T.C. Peterson, D.A. Quadir, S. Dorji, X. Zou, H. Tang, K. Santhosh,
U.R. Joshi, A.K. Jaswal, R.K. Kolli, A.B. Sikder, N.R. Deshpande, J.V. Revadekar, K.
Yeleuova, S. Vandasheva, M. Faleyeva, P. Gomboluudev, K.P. Budhathoki, A.
Hussain, M. Afzaal, L. Chandrapala, H. Anvar, D. Amanmurad, V.S. Asanova, P.D.
Jones, M.G. New, and T. Specktorman, 2006: Changes in daily temperature and
precipitation extremes in central and south Asia. Journal of Geophysical
Research, 111, D16105.
Knapp, K.R. and J.P. Kossin, 2007: A new global tropical cyclone data set from ISCCP
B1 geostationary satellite observations. Journal of Applied Remote Sensing, 1,
013505.
Knapp, K.R. and M.C. Kruk, 2010: Quantifying inter-agency differences in tropical
cyclone best track wind speed estimates. Monthly Weather Review, 138(4),
1459-1473.
Kniveton, D.R., R. Layberry, C.J.R. Williams, and M. Peck, 2009: Trends in the start
of the wet season over Africa. International Journal of Climatology, 29(9),
1216-1225.
Knowles, N., M.D. Dettinger, and D.R. Cayan, 2006: Trends in snowfall versus rainfall
in the western United States. Journal of Climate, 19(18), 4545-4559.
Knox, J.C., 2000: Sensitivity of modern and Holocene floods to climate change.
Quaternary Science Reviews, 19(1-5), 439-457.
Knutson, T.R. and R.E. Tuleya, 2004: Impact of CO
2
-induced warming on simulated
hurricane intensity and precipitation: Sensitivity to the choice of climate model
and convective parameterization. Journal of Climate, 17(18), 3477-3495.
Knutson, T.R., T.L. Delworth, K.W. Dixon, I.M. Held, J. Lu, V. Ramaswamy, M.D.
Schwarzkopf, G. Stenchikov, and R.J. Stouffer, 2006: Assessment of twentieth-
century regional surface temperature trends using the GFDL CM2 coupled
models. Journal of Climate, 19(9), 1624-1651.
Knutson, T.R., J.J. Sirutis, S.T. Garner, I.M. Held, and R.E. Tuleya, 2007: Simulation of
the recent multidecadal increase of Atlantic hurricane activity using an 18-km-
grid regional model. Bulletin of the American Meteorological Society, 88(10),
1549-1565.
Knutson, T.R., J.J. Sirutis, S.T. Garner, G.A. Vecchi, and I.M. Held, 2008: Simulated
reduction in Atlantic hurricane frequency under twenty-first- century warming
conditions. Nature Geoscience, 1(6), 359-364.
Knutson, T.R., J.L. McBride, J. Chan, K. Emanuel, G. Holland, C. Landsea, I. Held, J.P.
Kossin, A.K. Srivastava, and M. Sugi, 2010: Tropical cyclones and climate
change. Nature Geoscience, 3(3), 157-163.
Knutti, R., 2010: The end of model democracy? Climatic Change, 102(3-4),
395-404.
Knutti, R., R. Furrer, C. Tebaldi, J. Cermak, and G.A. Meehl, 2010a: Challenges in
combining projections from multiple climate models. Journal of Climate,
23(10), 2739-2758.
Knutti, R., G. Abramowitz, M. Collins, V. Eyring, P.J. Glecker, B. Hewitson, and L.
Mearns, 2010b: Good practice guidance paper on assessing and combining
multi model climate projections. In: Meeting Report of the Intergovernmental
Panel on Climate Change Expert Meeting on Assessing and Combining Multi-
Model Climate Projections [Stocker, T.F., Q. Dahe, G.-K. Plattner, M. Tignor, and
P.M. Midgley (eds.)]. IPCC Working Group I Technical Support Unit, University of
Bern, Bern, Switzerland, pp. 1-13.
Kochel, R.C., and V.R. Baker, 1982: Paleoflood hydrology. Science, 215(4531), 353-361.
Kodama, C. and T. Iwasaki, 2009: Influence of the SST rise on baroclinic instability
wave activity under an aquaplanet condition. Journal of the Atmospheric
Sciences, 66(8), 2272-2287.
Koffi, B. and E. Koffi, 2008: Heat waves across Europe by the end of the 21st century:
multiregional climate simulations. Climate Research, 36(2), 153-168.
Komar, P.D. and J.C. Allan, 2008: Increasing hurricane-generated wave heights along
the US East Coast and their climate controls. Journal of Coastal Research,
24(2), 479-488.
Kont, A., J. Jaagus, R. Aunap, U. Ratas, and R. Rivis, 2008: Implications of sea-level
rise for Estonia. Journal of Coastal Research, 24(2), 423-431.
Koo, G.S., K.O. Boo, and W.T. Kwon, 2009: Projection of temperature over Korea using
an MM5 regional climate simulation. Climate Research, 40(2-3), 241-248.
Kossin, J.P., 2008: Is the North Atlantic hurricane season getting longer? Geophysical
Research Letters, 35, L23705.
Kossin, J.P. and D.J. Vimont, 2007: A more general framework for understanding
Atlantic hurricane variability and trends. Bulletin of the American Meteorology
Society, 88(11), 1767-1781.
Kossin, J.P., K.R. Knapp, D.J. Vimont, R.J. Murnane, and B.A. Harper, 2007: A globally
consistent reanalysis of hurricane variability and trends. Geophysical Research
Letters, 34, L04815.
Kossin, J.P., S.J. Camargo, and M. Sitkowski, 2010: Climate modulation of North
Atlantic hurricane tracks. Journal of Climate, 23(11), 3057-3076.
Koster, R.D. and M.J. Suarez, 2001: Soil moisture memory in climate models. Journal
of Hydrometeorology, 2(6), 558-570.
Koster, R.D., M.J. Suarez, P. Liu, U. Jambor, A. Berg, M. Kistler, R. Reichle, M. Rodell,
and J. Famiglietti, 2004a: Realistic initialization of land surface states: Impacts
on subseasonal forecast skill. Journal of Hydrometeorology, 5(6), 1049-1063.
Koster, R.D., P.A. Dirmeyer, Z.C. Guo, G. Bonan, E. Chan, P. Cox, C.T. Gordon, S. Kanae,
E. Kowalczyk, D. Lawrence, P. Liu, C.H. Lu, S. Malyshev, B. McAvaney, K. Mitchell,
D. Mocko, T. Oki, K. Oleson, A. Pitman, Y.C. Sud, C.M. Taylor, D. Verseghy, R. Vasic,
Y.K. Xue, and T. Yamada, 2004b: Regions of strong coupling between soil
moisture and precipitation. Science, 305, 1138-1140.
Koster, R.D., Z.C. Guo, R.Q. Yang, P.A. Dirmeyer, K. Mitchell, and M.J. Puma, 2009: On
the nature of soil moisture in land surface models. Journal of Climate, 22(16),
4322-4335.
Kotlyakov, V.M., O.V. Rototaeva, and G.A. Nosenko, 2004: The September 2002
Kolka glacier catastrophe in North Ossetia, Russian Federation: evidence and
analysis. Mountain Research and Development, 24(1), 78-83.
Kripalani, R.H., J.H. Oh, A. Kulkarni, S.S. Sabade, and H.S. Chaudhari, 2007: South
Asian summer monsoon precipitation variability: Coupled climate model
simulations and projections under IPCC AR4. Theoretical and Applied
Climatology, 90(3-4), 133-159.
Krishnamurthy, C.K.B., U. Lall, and H.H. Kwon, 2009: Changing frequency and
intensity of rainfall extremes over India from 1951 to 2003. Journal of Climate,
22(18), 4737-4746.
Kruger, A.C., 2006: Observed trends in daily precipitation indices in South Africa:
1910-2004. International Journal of Climatology, 26(15), 2275-2285.
Kubota, H. and J.C.L. Chan, 2009: Interdecadal variability of tropical cyclone landfall
in the Philippines from 1902 to 2005. Geophysical Research Letters, 36, L12802.
Kürbis, K., M. Mudelsee, G. Tetzlaff, and R. Brazdil, 2009: Trends in extremes of
temperature, dewpoint, and precipitation from long instrumental series from
central Europe. Theoretical and Applied Climatology, 98(1-2), 187-195.
Kuglitsch, F.G., A. Toreti, E. Xoplaki, P.M. Della-Marta, J. Luterbacher, and H. Wanner,
2009: Homogenization of daily maximum temperature series in the
Mediterranean. Journal of Geophysical Research – Atmospheres, 114, D15108.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
216
Kuglitsch, F.G., A. Toreti, E. Xoplaki, P.M. Della-Marta, C. Zerefos, M. Türkes, and J.
Luterbacher, 2010: Heat wave changes in the eastern Mediterranean since
1960. Geophysical Research Letters, 37, L04802.
Kuleshov, Y., L. Qi, R. Fawcett, and D. Jones, 2008: On tropical cyclone activity in the
Southern Hemisphere: Trends and the ENSO connection. Geophysical Research
Letters, 35, L14S08.
Kumar, K.R., A.K. Sahai, K.K. Kumar, S.K. Patwardhan, P.K. Mishra, J.V. Revadekar, K.
Kamala, and G.B. Pant, 2006: High-resolution climate change scenarios for India
for the 21st century. Current Science, 90(3), 334-345.
Kundzewicz, Z.W., L.J. Mata, N.W. Arnell, P. Doll, P. Kabat, B. Jimenez, K.A. Miller, T.
Oki, Z. Sen, and I.A. Shiklomanov, 2007: Freshwater resources and their
management. In: Climate Change 2007. Impacts, Adaptation and Vulnerability.
Contribution of Working Group II to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P.
Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 173-210.
Kunkel, K.E., P.D. Bromirski, H.E. Brooks, T. Cavazos, A.V. Douglas, D.R. Easterling, K.A.
Emanuel, P.Y. Groisman, G.J. Holland, T.R. Knutson, J.P. Kossin, P.D. Komar, D.H.
Levinson, and R.L. Smith, 2008: Observed changes in weather and climate
extremes. In: Weather and Climate Extremes in a Changing Climate. Regions of
Focus: North America, Hawaii, Caribbean, and U.S. Pacific Islands. [Karl, T.R.,
G.A. Meehl, D.M. Christopher, S.J. Hassol, A.M. Waple, and W.L. Murray (eds.)]. A
Report by the U.S. Climate Change Science Program and the Subcommittee on
Global Change Research, Washington, DC, pp. 222.
Kunkel, K.E., X.-Z. Liang, and J. Zhu, 2010: Regional climate model projections
and uncertainties of U.S. summer heat waves. Journal of Climate, 23(16),
4447-4458.
Kunz, M., J. Sander, and C. Kottmeier, 2009: Recent trends of thunderstorm and
hailstorm frequency and their relation to atmospheric characteristics in
southwest Germany. International Journal of Climatology, 29(15), 2283-2297.
Kusunoki, S. and R. Mizuta, 2008: Future changes in the Baiu rain band projected by
a 20-km mesh global atmospheric model: sea surface temperature dependence.
SOLA, 4, 85-88.
Kwarteng, A.Y., A.S. Dorvlo, and G.T.V. Kumar, 2009: Analysis of a 27-year rainfall
data (1977-2003) in the Sultanate of Oman. International Journal of
Climatology, 29(4), 605-617.
Kyselý, J., 2009: Trends in heavy precipitation in the Czech Republic over 1961-2005.
International Journal of Climatology, 29(12), 1745-1758.
Kyselý, J. and R. Beranová, 2009: Climate-change effects on extreme precipitation in
central Europe: uncertainties of scenarios based on regional climate models.
Theoretical and Applied Climatology, 95(3-4), 361-374.
Laine, A., M. Kageyama, D. Salas-Melia, G. Ramstein, S. Planton, S. Denvil, and S.
Tyteca, 2009: An energetics study of wintertime Northern Hemisphere storm
tracks under 4 x CO
2
conditions in two ocean-atmosphere coupled models.
Journal of Climate, 22(3), 819-839.
Lambeck, K., C.D. Woodroffe, F. Antonioli, M. Anzidei, W.R. Gehrels, J. Laborel, and
A.J. Wright, 2010: Paleoenvironmental records, geophysical modeling, and
reconstruction of sea-level trends and variability on centennial and longer
timescales. In: Understanding Sea-Level Rise and Variability [Church, J.A., P.L.
Woodworth, T. Aarup, and W.S. Wilson (eds.)]. Wiley-Blackwell, Chichester, UK,
pp. 61-121.
Landsea
, C.W., 1993: A climatology of intense (or major) Atlantic hurricanes.
Monthly Weather Review, 121(6), 1703-1713.
Landsea, C.W., 2007: Counting Atlantic tropical cyclones back to 1900. EOS
Transactions (AGU), 88(18), 197-202.
Landsea, C.W., C. Anderson, N. Charles, G. Clark, J. Dunion, J. Fernandez-Partagas, P.
Hungerford, C. Neumann, and M. Zimmer, 2004: The Atlantic hurricane database
re-analysis project: Documentation for the 1851-1910 alterations and additions
to the HURDAT database. In: Hurricanes and Typhoons: Past, Present and Future
[Murnane, R.J. and K.B. Liu (eds.)]. Columbia University Press, New York, NY, pp.
177-221.
Landsea, C.W., B.A. Harper, K. Hoarau, and J.A. Knaff, 2006: Can we detect trends in
extreme tropical cyclones? Science, 313(5786), 452-454.
Lantz, T.C., and S.V. Kokelj, 2008: Increasing rates of retrogressive thaw slump activity
in the Mackenzie Delta region, N.W.T., Canada. Geophysical Research Letters,
35, L06502.
Laprise, R., R. de Elia, D. Caya, S. Biner, P. Lucas-Picher, E. Diaconescu, M. Leduc, A.
Alexandru, L. Separovic, and C. Canadian Network Regional, 2008: Challenging
some tenets of regional climate modelling. Meteorology and Atmospheric
Physics, 100(1-4), 3-22.
Larsen, C.F., R.J. Motyka, A.A. Arendt, K.A. Echelmeyer, and P.E. Geissler, 2007:
Glacier changes in southeast Alaska and northwest British Columbia and
contribution to sea level rise. Journal of Geophysical Research, 112,
F01007.
Lau, K.M. and H.T. Wu, 2007: Detecting trends in tropical rainfall characteristics,
1973-2003. International Journal of Climatology, 27(8), 979-988.
Lau, K.M., M.K. Kim, and K.M. Kim, 2006: Asian summer monsoon anomalies induced
by aerosol direct forcing: the role of the Tibetan Plateau. Climate Dynamics,
26(7-8), 855-864.
Lau, K.M., Y.P. Zhou, and H.T. Wu, 2008a: Have tropical cyclones been feeding more
extreme rainfall? Journal of Geophysical Research, 113, D23113.
Lau, N.C., A. Leetma, and M.J. Nath, 2008b: Interactions between the responses of
North American climate to El Niño–La Niña and to the secular warming trend
in the Indian-Western Pacific Oceans. Journal of Climate, 21(3), 476-494.
Laurent, C. and S. Parey, 2007: Estimation of 100-year return-period temperatures in
France in a non-stationary climate: Results from observations and IPCC scenarios.
Global and Planetary Change, 57(1-2), 177-188.
Le Cozannet, G., S. Lecacheux, E. Delvallee, N. Desramaut, C. Oliveros, and R.
Pedreros, 2011: Teleconnection pattern influence on sea-wave climate in the
Bay of Biscay. Journal of Climate, 24(3), 641-652.
Le Quesne, C., D.W. Stahle, M.K. Cleaveland, M.D. Therrell, J.C. Aravena, and J.
Barichivich, 2006: Ancient Austrocedrus tree-ring chronologies used to
reconstruct central Chile precipitation variability from A.D. 1200 to 2000.
Journal of Climate, 19(22), 5731-5744.
Le Quesne, C., C. Acuna, J.A. Boninsegna, A. Rivera, and J. Barichivich, 2009: Long-
term glacier variations in the Central Andes of Argentina and Chile, inferred
from historical records and tree-ring reconstructed precipitation.
Palaeogeography, Palaeoclimatology, Palaeoecology, 281(3-4), 334-344.
Leake, J., J. Wolf, J. Lowe, P. Stansby, G. Jacoub, R. Nicholls, M. Mokrech, S. Nicholson-
Cole, M. Walkden, A. Watkinson, and S. Hanson, 2007: Predicted wave climate
for the UK: towards an integrated model of coastal impacts of climate change.
In: Estuarine and Coastal Modeling Congress 2007 [Spaulding, M. (ed.)].
Proceedings of the Tenth International Conference on Estuarine and Coastal
Modeling Congress 2007, Newport, Rhode Island, 5-7 Nov 2007, American
Society of Civil Engineers, pp 393-406.
Leander, R., T.A. Buishand, B.J.J.M. van den Hurk, and M.J.M. de Wit, 2008: Estimated
changes in flood quantiles of the river Meuse from resampling of regional
climate model output. Journal of Hydrology, 351(3-4), 331-343.
Leckebusch, G.C., B. Koffi, U. Ulbrich, J.G. Pinto, T. Spangehl, and S. Zacharias, 2006:
Analysis of frequency and intensity of European winter storm events from a
multi-model perspective, at synoptic and regional scales. Climate Research,
31(1), 59-74.
Leckebusch, G.C., D. Renggli, and U. Ulbrich, 2008: Development and application of
an objective storm severity measure for the Northeast Atlantic region.
Meteorologische Zeitschrift, 17(5), 575-587.
Leclercq, P., J. Oerlemans, and J. Cogley, 2011: Estimating the glacier contribution to
sea-level rise for the period 1800–2005. Surveys in Geophysics, 32(4-5),
519-535, doi:10.1007/s10712-10011-19121-10717.
Lee, T. and M.J. McPhaden, 2010: Increasing intensity of El Niño in the central-
equatorial Pacific. Geophysical Research Letters, 37, L14603.
Legates, D.R., H.F. Lins, and G.J. McCabe, 2005: Comments on “Evidence for global
runoff increase related to climate warming” by Labat et al. Advances in Water
Resources, 28(12), 1310-1315.
Lehmann, A., K. Getzlaff, and J. Harlass, 2011: Detailed assessment of climate
variability in the Baltic Sea area for the period 1958 to 2009. Climate Research,
46(2), 185-196.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
217
Lehner, B., P. Doll, J. Alcamo, T. Henrichs, and F. Kaspar, 2006: Estimating the impact
of global change on flood and drought risks in Europe: A continental, integrated
analysis. Climatic Change, 75(3), 273-299.
Leloup, J., M. Lengaigne, and J.-P. Boulanger, 2008: Twentieth century ENSO
characteristics in the IPCC database. Climate Dynamics, 30(2-3),
227-291.
Lemke, P., J. Ren, R.B. Alley, I. Allison, J. Carrasco, G. Flato, Y. Fujii, G. Kaser, P. Mote,
R.H. Thomas, and T. Zhang, 2007: Observations: Changes in snow, ice and frozen
ground. In: Climate Change 2007: The Physical Science Basis. Contribution of
Working Group I to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M.
Marquis, K.B. Averyt, M. Tignor and H.L. Miller (eds.)]. Cambridge University
Press, Cambridge, UK, and New York, NY, pp. 337-383.
Lenderink, G. and E. Van Meijgaard, 2008: Increase in hourly precipitation extremes
beyond expectations from temperature changes. Nature Geoscience, 1(8),
511-514.
Lenderink, G. and E. van Meijgaard, 2009: Unexpected rise in extreme precipitation
caused by a shift in rain type? Nature Geoscience, 2(6), 373-373.
Lenton, T.M., H. Held, E. Kriegler, J.W. Hall, W. Lucht, S. Rahmstorf, and H.J.
Schellnhuber, 2008: Tipping elements in the Earth’s climate system. Proceedings
of the National Academy of Sciences, 105(6), 1786-1793.
Leslie, L.M., M. Leplastrier, and B.W. Buckley, 2008: Estimating future trends in
severe hailstorms over the Sydney Basin: A climate modelling study.
Atmospheric Research, 87(1), 37-57.
Lewis, S.L., P.M. Brando, O.L. Phillips, G.M.F. van der Heijden, and D. Nepstad, 2011:
The 2010 Amazon Drought. Science, 331(6017), 554.
Lewkowicz, A.G. and C. Harris, 2005: Morphology and geotechnique of active-layer
detachment failures in discontinuous and continuous permafrost, northern
Canada. Geomorphology, 69(1-4), 275-297.
L’Hôte, Y., G. Mahé, B. Somé, and J.P. Triboulet, 2002: Analysis of a Sahelian annual
rainfall index from 1896 to 2000; the drought continues. Hydrological Sciences
Journal, 47(4), 563-572.
Li, H., A. Dai, T. Zhou, and J. Lu, 2008: Responses of East Asian summer monsoon to
historical SST and atmospheric forcing during 1950-2000. Climate Dynamics,
34(4), 501-514.
Li, X.Y., L.Y. Liu, and J.H. Wang, 2004: Wind tunnel simulation of aeolian sandy soil
erodibility under human disturbance. Geomorphology, 59(1-4), 3-11.
Liggins, F., R.A. Betts, and B. McGuire, 2010: Projected future climate changes in the
context of geological and geomorphological hazards. Philosophical
Transactions of the Royal Society A: Mathematical, Physical and Engineering
Sciences, 368(1919), 2347-2367.
Lim, E.-P., and I. Simmonds, 2009: Effect of tropospheric temperature change on the
zonal mean circulation and SH winter extratropical cyclones.
Climate Dynamics,
33(1), 19-32.
Lin, J.L., B.E. Mapes, K.M. Weickmann, G.N. Kiladis, S.D. Schubert, M.J. Suarez, J.T.
Bacmeister, and M.I. Lee, 2008a: North American monsoon and convectively
coupled equatorial waves simulated by IPCC AR4 coupled GCMs. Journal of
Climate, 21(12), 2919-2937.
Lin, J.L., K.M. Weickman, G.N. Kiladis, B.E. Mapes, S.D. Schubert, M.J. Suarez, J.T.
Bacmeister, and M.I. Lee, 2008b: Subseasonal variability associated with Asian
summer monsoon simulated by 14 IPCC AR4 coupled GCMs. Journal of Climate,
21(18), 4541-4567.
Lindström, G. and S. Bergström, 2004: Runoff trends in Sweden 1807-2002.
Hydrological Sciences Journal, 49(1), 69-83.
Lins, H.F. and J.R. Slack, 1999: Streamflow trends in the United States. Geophysical
Research Letters, 26(2), 227-230.
Lionello, P., S. Cogo, M.B. Galati, and A. Sanna, 2008: The Mediterranean surface
wave climate inferred from future scenario simulations. Global and Planetary
Change, 63(2-3), 152-162.
Lipovsky, P., S. Evans, J. Clague, C. Hopkinson, R. Couture, P. Bobrowsky, G. Ekström,
M. Demuth, K. Delaney, N. Roberts, G. Clarke, and A. Schaeffer, 2008: The July
2007 rock and ice avalanches at Mount Steele, St. Elias Mountains, Yukon,
Canada. Landslides, 5(4), 445-455.
Liu, B., M. Xu, and M. Henderson, 2011: Where have all the showers gone? Regional
declines in light precipitation events in China, 1960–2000. International Journal
of Climatology, 31(8), 1177-1191.
Llano, M.P. and O. Penalba, 2011: A climatic analysis of dry sequences in Argentina.
International Journal of Climatology, 31(4), 504-513.
Llasat, M.C., M. Barriendos, A. Barrera, and T. Rigo, 2005: Floods in Catalonia (NE
Spain) since the 14th century. Climatological and meteorological aspects from
historical documentary sources and old instrumental records. Journal of
Hydrology, 313(1-2), 32-47.
Lloyd-Hughes, B., and M.D. Saunders, 2002: A drought climatology for Europe.
International Journal of Climatology, 22(13), 1571-1592.
Lobell, D.B., C.J. Bonfils, L.M. Kueppers, and M.A. Snyder, 2008: Irrigation cooling
effect on temperature and heat index extremes. Geophysical Research Letters,
35, L09705.
Loeptien, U., O. Zolina, S. Gulev, M. Latif, and V. Soloviov, 2008: Cyclone life cycle
characteristics over the Northern Hemisphere in coupled GCMs. Climate
Dynamics, 31(5), 507-532.
Lohmann, U. and J. Feichter, 2007: Global indirect aerosol effects: a review.
Atmospheric Chemistry and Physics, 5(3), 715-737.
Long, Z., W. Perrie, J. Gyakum, R. Laprise, and D. Caya, 2009: Scenario changes in the
climatology of winter midlatitude cyclone activity over eastern North America
and the Northwest Atlantic. Journal of Geophysical Research – Atmospheres,
114, D12111.
Lorenz, D.J., and E.T. DeWeaver, 2007: Tropopause height and zonal wind response
to global warming in the IPCC scenario integrations. Journal of Geophysical
Research – Atmospheres, 112, D10119.
Lorenz, D.J. and D.L. Hartmann, 2003: Eddy-zonal flow feedback in the Northern
Hemisphere winter. Journal of Climate, 16(8), 1212-1227.
Lorenz, R., E.B. Jaeger, and S.I. Seneviratne, 2010: Persistence of heat waves and its
link to soil moisture memory. Geophysical Research Letters, 37, L09703.
Lowe, J.A., P.L. Woodworth, T. Knutson, R.E. McDonald, K.L. McInnes, K. Woth, H. von
Storch, J. Wolf, V. Swail, N.B. Bernier, S. Gulev, K.J. Horsburgh, A.S. Unnikrishnan,
J.R. Hunter, and R. Weisse, 2010: Past and future changes in extreme sea levels
and waves. In: Understanding Sea-Level Rise and Variability [Church, J.A., P.L.
Woodworth, T. Aarup, and W.S. Wilson (eds.)]. Wiley-Blackwell, Chichester, UK,
pp. 326-375.
Lu, J., G.A. Vecchi, and T. Reichler, 2007: Expansion of the Hadley cell under global
warming. Geophysical Research Letters, 34, L06805.
Lugon, R. and M. Stoffel, 2010: Rock-glacier dynamics and magnitude-frequency
relations of debris flows in a high-elevation watershed: Ritigraben, Swiss Alps.
Global and Planetary Change, 73(3-4), 202-210.
Lupikasza, E., 2010: Spatial and temporal variability of extreme precipitation in
Poland in the period 1951–2006. International Journal of Climatology, 30(7),
991-1007.
Luterbacher, J., D. Dietrich, E. Xoplaki, M. Grosjean, and H. Wanner, 2004: European
seasonal and annual temperature variability, trends, and extremes since 1500.
Science, 303(5663 ), 1499-1503.
Lynch, A.H., J.A. Curry, R.D. Brunner, and J.A. Maslanik, 2004: Toward an integrated
assessment of the impacts of extreme wind events on Barrow, Alaska. Bulletin
of the American Meteorological Society, 85(2), 209-221.
Mabuchi, K., Y. Sato, and H. Kida, 2005a: Climatic impact of vegetation change in the
Asian tropical region. Part I: Case of the Northern Hemisphere summer. Journal
of Climate, 18(3), 410-428.
Mabuchi, K., Y. Sato, and H. Kida, 2005b: Climatic impact of vegetation change in
the Asian tropical region. Part II: Case of the Northern Hemisphere winter and
impact on the extratropical circulation. Journal of Climate, 18(3), 429-446.
Mann, M.E. and K.A. Emanuel, 2006: Atlantic hurricane trends linked to climate
change. Eos Transactions (AGU), 87(24), 233-241.
Mann, M.E., K.A. Emanuel, G.J. Holland, and P.J. Webster, 2007a: Atlantic tropical
cyclones revisited. Eos Transactions (AGU), 88, 349-350.
Mann, M.E., T.A. Sabbatelli, and U. Neu, 2007b: Evidence for a modest undercount
bias in early historical Atlantic tropical cyclone counts. Geophysical Research
Letters, 34, L22707.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
218
Mann, M.E., J.D. Woodruff, J.P. Donnelly, and Z. Zhang, 2009: Atlantic hurricanes and
climate over the past 1,500 years. Nature, 460, 880-883.
Manson, G.K., and S.M. Solomon, 2007: Past and future forcing of Beaufort Sea
coastal change. Atmosphere-Ocean, 45(2), 107-122.
Manton, M.J., P.M. Della-Marta, M.R. Haylock, K.J. Hennessy, N. Nicholls, L.E.
Chambers, D.A. Collins, G. Daw, A. Finet, D. Gunawan, K. Inape, H. Isobe, T.S.
Kestin, P. Lefale, C.H. Leyu, T. Lwin, L. Maitrepierre, N. Ouprasitwong, C.M. Page,
J. Pahalad, N. Plummer, M.J. Salinger, R. Suppiah, V.L. Tran, B. Trewin, I. Tibig, and
D. Yee, 2001: Trends in extreme daily rainfall and temperature in Southeast Asia
and the South Pacific: 1961–1998. International Journal of Climatology, 21(3),
269-284.
Maraun, D., T.J. Osborn, and N.P. Gillet, 2008: United Kingdom daily precipitation
intensity: improved early data, error estimates and an update from 2000 to
2006. International Journal of Climatology, 28(6), 833-842.
Maraun, D., F. Wetterhall, A.M. Ireson, R.E. Chandler, E.J. Kendon, M. Widmann, S.
Brienen, H.W. Rust, T. Sauter, M. Themessl, V.K.C. Venema, K.P. Chun, C.M.
Goodess, R.G. Jones, C. Onof, M. Vrac, and I. Thiele-Eich, 2010: Precipitation
downscaling under climate change. Recent developments to bridge the gap
between dynamical models and the end user. Reviews of Geophysics, 48,
RG3003.
Marcos, M., M.N. Tsimplis, and A.G.P. Shaw, 2009: Sea level extremes in southern
Europe. Journal of Geophysical Research – Oceans, 114, C01007.
Marengo, J.A., R. Jones, L.M. Alves, and M.C. Valverde, 2009a: Future change of
temperature and precipitation extremes in South America as derived from the
PRECIS regional climate modeling system. International Journal of Climatology,
29(15), 2241-2255.
Marengo, J.A., M. Rusticucci, O. Penalba, and M. Renom, 2009b: An intercomparison
of observed and simulated extreme rainfall and temperature events during the
last half of the twentieth century: part 2: historical trends. Climatic Change,
98(3-4), 509-529.
Marengo, J.A., J. Tomasella, W. Soares, L. Alves, and C. Nobre, 2011: Extreme climatic
events in the Amazon basin. Theoretical and Applied Climatology,
doi:10.1007/s00704-00011-00465-00701.
Mariotti, A., N. Zeng, J.H. Yoon, V. Artale, A. Navarra, P. Alpert, and L.Z.X. Li, 2008:
Mediterranean water cycle changes: transition to drier 21st century conditions
in observations and CMIP3 simulations. Environmental Research Letters, 3(4),
044001.
Marques, C., A. Rocha, and J. Corte-Real, 2011: Global diagnostic energetics of five
state-of-the-art climate models. Climate Dynamics, 36(9-10), 1767-1794.
Martin, J.H., and R.M. Gordon, 1988: Northeast Pacific iron distributions in relation
to phytoplankton productivity. Deep-Sea Research Part A – Oceanographic
Research Papers, 35(2), 177-196.
Mastin, L.G., 1994: Explosive tephra emissions at Mount St. Helens, 1989–1991: The
violent escape of magmatic gas following storms? Geological Society of
America Bulletin, 106(2), 175-185.
Mastrandrea, M.D., C.B. Field, T.F. Stocker, O. Edenhofer, K.L. Ebi, D.J. Frame, H. Held,
E. Kriegler, K.J. Mach, P.R. Matschoss, G.-K. Plattner, G.W. Yohe, and F.W. Zwiers,
2010: Guidance Note for Lead Authors of the IPCC Fifth Assessment Report on
Consistent Treatment of Uncertainties. Intergovernmental Panel on Climate
Change (IPCC), Geneva, Switzerland.
Matthews, A.J., J. Barclay, S. Carn, G. Thompson, J. Alexander, R. Herd, and C.
Williams, 2002: Rainfall-induced volcanic activity on Montserrat. Geophysical
Research Letters, 29(13), 1644.
Matulla, C., W. Schoener, H. Alexandersson, H. von Storch, and X.L. Wang, 2008:
European storminess: late nineteenth century to present. Climate Dynamics,
31(2-3), 125-130.
Maue, R.N., 2009: Northern Hemisphere tropical cyclone activity. Geophysical
Research Letters, 36, L05805.
May, W., 2008: Potential future changes in the characteristics of daily precipitation
in Europe simulated by the HIRHAM regional climate model. Climate Dynamics,
30(6), 581-603.
McCabe, G.J., and D.M. Wolock, 2002: A step increase in streamflow in the
conterminous United States. Geophysical Research Letters, 29(24), 2185.
McDonald, R., 2011: Understanding the impact of climate change on Northern
Hemisphere extra-tropical cyclones. Climate Dynamics, 37(7-8), 1399-1425,
doi:10.1007/s00382-00010-00916-x.
McGranahan, G., D. Balk, and B. Anderson, 2007: The rising tide: assessing the risks
of climate change and human settlements in low elevation coastal zones.
Environment and Urbanization, 19(1), 17-37.
McGuire, B., 2010: Potential for a hazardous geospheric response to projected future
climate changes. Philosophical Transactions of the Royal Society A, 368(1919),
2317-2345.
McInnes, K.L., K.J.E. Walsh, G.D. Hubbert, and T. Beer, 2003: Impact of sea-level rise
and storm surges on a coastal community. Natural Hazards, 30(2), 187-207.
McInnes, K.L., I. Macadam, and J.G. O’Grady, 2009a: The Effect of Climate Change
on Extreme Sea Levels along Victoria’s Coast. A Project Undertaken for the
Department of Sustainability and Environment, Victoria as part of the ‘Future
Coasts’ Program. CSIRO, Victoria, Australia, 55 pp.
McInnes, K.L., I. Macadam, G.D. Hubbert, and J.G. O’Grady, 2009b: A modelling
approach for estimating the frequency of sea level extremes and the impact of
climate change in southeast Australia. Natural Hazards, 51(1), 115-137.
McInnes, K.L., T.A. Erwin, and J.M. Bathols, 2011: Global Climate Model projected
changes in 10 m wind speed and direction due to anthropogenic climate
change. Atmospheric Science Letters, doi:10.1002/asl.1341.
McKee, T.B., N.J. Doesken, and J. Kleist, 1993: The relationship of drought frequency
and duration to time scales. In: Proceedings of the 8th Conference on Applied
Climatology, Anaheim, California, 17-22 Jan 1993, pp. 179-184.
McKendry, I.G., K.B. Strawbridge, N.T. O’Neill, A.M. Macdonald, P.S.K. Liu, W.R.
Leaitch, K.G. Anlauf, L. Jaegle, T.D. Fairlie, and D.L. Westphal, 2007: Trans-Pacific
transport of Saharan dust to western North America: A case study. Journal of
Geophysical Research – Atmospheres, 112, D01103.
McVicar, T.R., T.G. Van Niel, L.T. Li, M.L. Roderick, D.P. Rayner, L. Ricciardulli, and R.J.
Donohue, 2008: Wind speed climatology and trends for Australia, 1975-2006:
Capturing the stilling phenomenon and comparison with near-surface reanalysis
output. Geophysical Research Letters, 35, L20403.
Mearns, L.O., W. Gutowski, R. Jones, R. Leung, S. McGinnis, A. Nunes, and Y. Qian,
2009: A regional climate change assessment program for North America. Eos
Transactions (AGU), 90(36), 311.
Meehl, G.A. and C. Tebaldi, 2004: More intense, more frequent, and longer lasting
heat waves in the 21st century. Science, 305(5686 ), 994-997.
Meehl, G.A., J.M. Arblaster, and C. Tebaldi, 2005: Understanding future patterns of
increased precipitation intensity in climate model simulations. Geophysical
Research Letters, 32, L18719.
Meehl, G.A., C. Covey, T. Delworth, M. Latif, B. McAvaney, J.F.B. Mitchell, R.J. Stouffer,
and K.E. Taylor, 2007a: The WCRP CMIP3 multimodel dataset – A new era in
climate change research. Bulletin of the American Meteorological Society,
88(9), 1383-1394.
Meehl, G.A., T.F. Stocker, W.D. Collins, P. Friedlingstein, A.T. Gaye, J.M. Gregory, A.
Kitoh, R. Knutti, J.M. Murphy, A. Noda, S.C.B. Raper, I.G. Watterson, A.J. Weaver,
and Z.C. Zhao, 2007b: Global climate projections. In: Climate Change 2007: The
Physical Science Basis. Contribution of Working Group I to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change
[Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor
and H.L. Miller (eds.)]. Cambridge University Press, Cambridge, UK, and New
York, NY, pp. 747-845.
Meehl, G.A., J.M. Arblaster, and W.D. Collins, 2008: Effects of black carbon aerosols
on the Indian monsoon. Journal of Climate, 21(12), 2869-2882.
Meehl, G.A., J.M. Arblaster, K. Matthes, F. Sassi, and H. van Loon, 2009a: Amplifying
the Pacific climate system response to a small 11-year solar cycle forcing.
Science, 325(5944), 1114-1118.
Meehl, G.A., A. Hu, and B.D. Santer, 2009b: The mid-1970s climate shift in the Pacific
and the relative roles of forced versus inherent decadal variability. Journal of
Climate, 22(3), 780-792.
Meehl, G.A., C. Tebaldi, G. Walton, D. Easterling, and L. McDaniel, 2009c: The relative
increase of record high maximum temperatures compared to record low
minimum temperatures in the U.S. Geophysical Research Letters, 36, L23701.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
219
Meinshausen, M., N. Meinshausen, W. Hare, S.C.B. Raper, K. Frieler, R. Knutti, D.J.
Frame, and M.R. Allen, 2009: Greenhouse-gas emission targets for limiting
global warming to 2°C. Nature, 458(7242), 1158-1162.
Menendez, C.G., M. de Castro, J.P. Boulanger, A. D’Onofrio, E. Sanchez, A.A.
Sorensson, J. Blazquez, A. Elizalde, D. Jacob, H. Le Treut, Z.X. Li, M.N. Nunez, N.
Pessacg, S. Pfeiffer, M. Rojas, A. Rolla, P. Samuelsson, S.A. Solman, and C.
Teichmann, 2010: Downscaling extreme month-long anomalies in southern
South America. Climatic Change, 98(3-4), 379-403.
Menendez, M. and P.L. Woodworth, 2010: Changes in extreme high water levels based
on a quasi-global tide-gauge dataset. Journal of Geophysical Research, 115, C10011.
Menendez, M., F.J. Mendez, I.J. Losada, and N.E. Graham, 2008: Variability of
extreme wave heights in the northeast Pacific Ocean based on buoy
measurements. Geophysical Research Letters, 35, L22607.
Menne, M.J., and C.N. Williams Jr., 2009: Homogenization of temperature series via
pairwise comparisons. Journal of Climate, 22(7), 1700-1717.
Merrifield, M.A., Y.L. Firing, and J.J. Marra, 2007: Annual climatologies of extreme
water levels. In: Aha Hulikoa: Extreme Events. Proceedings of the Hawaiian
Winter Workshop, University of Hawaii at Manoa, 23 -26 Jan 2007. SOEST,
University of Hawaii, Manoa, HI, pp. 27-32.
Miller, R., G. Schmidt, and D. Shindell, 2006: Forced annular variations in the 20th
century Intergovernmental Panel on Climate Change Fourth Assessment Report
models. Journal of Geophysical Research, 111, D18101.
Mills, G.A., 2005: A re-examination of the synoptic and mesoscale meteorology of
Ash Wednesday 1983. Australian Meteorological Magazine, 54(1), 35-55.
Milly, P.C.D. and K.A. Dunne, 2011: Hydrologic adjustment of climate-model projections:
The potential pitfall of potential evapotranspiration. Earth Interactions, 15(1), 1-14.
Milly, P.C.D., R.T. Wetherald, K.A. Dunne, and T.L. Delworth, 2002: Increasing risk of
great floods in a changing climate. Nature, 415(6871), 514-517.
Mimura, N., 1999: Vulnerability of island countries in the South Pacific to sea level
rise and climate change. Climate Research, 12(2-3), 137-143.
Min, S.-K., X. Zhang, and F. Zwiers, 2008: Human-induced arctic moistening. Science,
320(5875), 518-520.
Min, S.-K., X. Zhang, F.W. Zwiers, and G.C. Hegerl, 2011: Human contribution to more
intense precipitation extremes. Nature, 470(7334), 378-381.
Min, S.K., and A. Hense, 2007: A Bayesian assessment of climate change using
multimodel ensembles. Part II: Regional and seasonal mean surface temperatures.
Journal of Climate, 20(12), 2769-2790.
Mitrovica, J.X., M.E. Tamisiea, E.R. Ivins, L.L.A. Vermeersen, G.A. Milne, and K.
Lambeck, 2010: Surface mass loading on a dynamic earth: complexity and
contamination in the geodetic analysis of global sea-level trends. In:
Understanding Sea-Level Rise and Variability [Church, J.A., P.L. Woodworth, T.
Aarup, and W.S. Wilson (eds.)]. Wiley-Blackwell, Chichester, UK, pp. 285-325.
Moberg, A., P.D. Jones, D. Lister, A. Walther, M. Brunet, J. Jacobeit, L.V. Alexander, P.M.
Della-Marta, J. Luterbacher, P. Yiou, D. Chen, A.M.G.K. Tank, O. Saladie, J. Sigro, E.
Aguilar, H. Alexandersson, C. Almarza, I. Auer, M. Barriendos, M. Begert, H.
Bergstroem, R. Boehm, C.J. Butler, J. Caesar, A. Drebs, D. Founda, F.-W. Gerstengarbe,
G. Micela, M. Maugeri, H. Osterle, K. Pandzic, M. Petrakis, L. Srnec, R. Tolasz, H.
Tuomenvirta, P.C. Werner, H. Linderholm, A. Philipp, H. Wanner, and E. Xoplaki, 2006:
Indices for daily temperature and precipitation extremes in Europe analyzed for the
period 1901-2000. Journal of Geophysical Research – Atmospheres, 111, D22106.
Moise, A.F., and R. Colman, 2009: Tropical Australia and the Australian Monsoon:
general assessment and projected changes. In: 18th World IMACS Congress and
MODSIM09 International Congress on Modelling and Simulation. Modelling
and Simulation Society of Australia and New Zealand and International
Association for Mathematics and Computers in Simulation, July 2009
[Anderssen, R.S., R.D. Braddock, and L.T.H. Newham (eds.)]. pp. 2042-2048.
Morgenstern, O., H. Akiyoshi, S. Bekki, P. Braesicke, N. Butchart, M.P. Chipperfield, D.
Cugnet, M. Deushi, S.S. Dhomse, R.R. Garcia, A. Gettelman, N.P. Gillett, S.C.
Hardiman, J. Jumelet, D.E. Kinnison, J.F. Lamarque, F. Lott, M. Marchand, M. Michou,
T. Nakamura, D. Olivie, T. Peter, D. Plummer, J.A. Pyle, E. Rozanov, D. Saint-Martin,
J.F. Scinocca, K. Shibata, M. Sigmond, D. Smale, H. Teyssedre, W. Tian, A. Voldoire,
and Y. Yamashita, 2010: Anthropogenic forcing of the Northern Annular Mode in
CCMVal-2 models. Journal of Geophysical Research – Atmospheres, 115, D00M03.
Mori, N., T. Yasuda, H. Mase, T. Tom, and Y. Oku, 2010: Projection of extreme wave
climate change under global warming. Hydrological Research Letters, 4, 15-19.
Mudelsee, M., M. Borngen, G. Tetzlaff, and U. Grunewald, 2003: No upward trends
in the occurrence of extreme floods in central Europe. Nature, 425(6954),
166-169.
Muir-Wood, R., 2000: Deglaciation Seismotectonics: a principal influence on intraplate
seismogenesis at high latitudes. Quaternary Science Reviews, 19(14-15),
1399-1411.
Mullan, B., D. Wratt, S. Dean, M. Hollis, S. Allan, T. Williams, and G. Kenny, 2008:
Climate Change Effects and Impacts Assessment: A Guidance Manual for Local
Government in New Zealand. 2nd ed. Ministry for the Environment, Wellington,
New Zealand, 149 pp.
Müller, W.A. and E. Roeckner, 2008: ENSO teleconnections in projections of future
climate in ECHAM5/MPI-OM. Climate Dynamics, 31(5), 533-549.
Murakami, H., B. Wang, and A. Kitoh, 2011: Future change of western North Pacific
typhoons: Projections by a 20-km-mesh global atmospheric model. Journal of
Climate, 24(4), 1154-1169.
Murphy, J.M., B.B.B. Booth, M. Collins, G.R. Harris, D.M.H. Sexton, and M.J. Webb,
2007: A methodology for probabilistic predictions of regional climate change
from perturbed physics ensembles. Philosophical Transactions of the Royal
Society A, 365(1857), 1993-2028.
Murphy, J.M., D.M. Sexton, G.J. Jenkins, B.B.B. Booth, C.C. Brown, R.T. Clark, M.
Collins, G.R. Harris, E.J. Kendon, R.A. Betts, S.J. Brown, K.A. Humphrey, M.P.
McCarthy, R.E. McDonald, A. Stephens, C. Wallace, R. Warren, R. Wilby, and R.A.
Wood, 2009: UK Climate Projections Science Report: Climate change projections.
Met Office Hadley Centre, Exeter, UK.
Nakaegawa, T. and W. Vergara, 2010: First projection of climatological mean river
discharges in the Magdalena River Basin, Colombia, in a changing climate during
the 21st century. Hydrological Research Letters, 4, 50-54.
Nakamura, M., S. Kanada, Y. Wakazuki, C. Muroi, A. Hashimoto, T. Kato, A. Noda, M.
Yoshizaki, and K. Yasunaga, 2008: Effects of global warming on heavy rainfall
during the Baiu season projected by a cloud-system-resolving model. Journal of
Disaster Research
, 3(1), 15-24.
Nandintsetseg, B., J.S. Greene, and C.E. Goulden, 2007: Trends in extreme daily
precipitation and temperature near lake Hövsgöl, Mongolia. International
Journal of Climatology, 27(3), 341-347.
Narama, C., Y. Shimamura, D. Nakayama, and K. Abdrakhmatov, 2006: Recent
changes of glacier coverage in the western Terskey-Alatoo range, Kyrgyz
Republic, using Corona and Landsat. Annals of Glaciology, 43(1), 223-229.
National Climate Centre, 2009: The Exceptional January-February 2009 Heatwave
in Southeastern Australia. Special Climate Statement 17 (www.bom.gov.au/
climate/current/statements/scs17d.pdf), Bureau of Meteorology, Melbourne,
Australia.
Neill, S.P., J.D. Scourse, G.R. Bigg, and K. Uehara, 2009: Changes in wave climate over
the northwest European shelf seas during the last 12,000 years. Journal of
Geophysical Research – Oceans, 114, C06015.
Neu, U., 2008: Is recent major hurricane activity normal? Nature, 451(7181), E5.
New, M., B. Hewitson, D.B. Stephenson, A. Tsiga, A. Kruger, A. Manhique, B. Gomez,
C.A.S. Coelho, D.N. Masisi, E. Kulunaga, E. Mbambalala, F. Adesina, H. Saleh, J.
Kanyanga, J. Adosi, L. Bulane, L. Fortunata, M.L. Mdoka, and R. Lajoie, 2006:
Evidence of trends in daily climate extremes over southern and west Africa.
Journal of Geophysical Research, 111, D14102.
Newhall, C.G. and R.S. Punongbayan (eds.), 1996: Fire and Mud: Eruptions and
Lahars of Mount Pinatubo, Philippines. University of Washington Press, Seattle,
WA, 1126 pp.
Niall, S. and K. Walsh, 2005: The impact of climate change on hailstorms in
southeastern Australia. International Journal of Climatology, 25(14), 1933-1952.
Nicholls, N., 1995: Long-term climate monitoring and extreme events. Climatic
Change, 31(2-4), 231-245.
Nicholls, N., 2008: Recent trends in the seasonal and temporal behaviour of the El
Niño Southern Oscillation. Geophysical Research Letters, 35, L19703.
Nicholls, N., 2010: Local and remote causes of the southern Australian autumn-
winter rainfall decline, 1958-2007. Climate Dynamics, 34(6), 835-845.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
220
Nicholls, N. and L. Alexander, 2007: Has the climate become more variable or
extreme? Progress 1992-2006. Progress in Physical Geography, 31(1), 77-87.
Nicholls, R.J., 2010: Impacts of and responses to sea-level rise. In: Understanding
Sea-Level Rise and Variability [Church, J.A., P.L. Woodworth, T. Aarup, and W.S.
Wilson (eds.)]. Wiley-Blackwell, Chichester, UK, pp. 17-51.
Nicholls, R.J. and A.C. de la Vega-Leinert, 2008: Implications of sea-level rise for
Europe’s coasts: An introduction. Journal of Coastal Research, 24(2), 285-287.
Nicholls, R.J., P.P. Wong, V.R. Burkett, J.O. Codignotto, J.E. Hay, R.F. McLean, S.
Ragoonaden, and C.D. Woodroffe, 2007: Coastal systems and low-lying areas.
In: Climate Change 2007. Impacts, Adaptation and Vulnerability. Contribution of
Working Group II to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der
Linden, and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK, pp.
315-356.
Nieto-Ferreira, R. and T. Rickenbach, 2011: Regionality of monsoon onset in South
America: a three stage conceptual model. International Journal of Climatology,
31(9), 1309-1321.
Nissen, K.M., G.C. Leckebusch, J.G. Pinto, D. Renggli, S. Ulbrich, and U. Ulbrich, 2010:
Cyclones causing wind storms in the Mediterranean: characteristics, trends and
links to large-scale patterns. Natural Hazards and Earth System Sciences, 10(7),
1379-1391.
Niu, F., G. Cheng, W. Ni, and D. Jin, 2005: Engineering-related slope failure in
permafrost regions of the Qinghai-Tibet Plateau. Cold Regions Science and
Technology, 42(3), 215-225.
Noetzli, J., S. Gruber, T. Kohl, N. Salzmann, and W. Haeberli, 2007: Three-dimensional
distribution and evolution of permafrost temperatures in idealized high-
mountain topography. Journal of Geophysical Research – Earth Surface, 112,
F02S13.
Noetzli, J. and S. Gruber, 2009: Transient thermal effects in Alpine permafrost. The
Cryosphere, 3(1), 85-99.
Nott, J., J. Haig, H. Neil, and D. Gillieson, 2007: Greater frequency variability of
landfalling tropical cyclones at centennial compared to seasonal and decadal
scales. Earth and Planetary Science Letters, 255(3-4), 367-372.
Nott, J., S. Smithers, K. Walsh, and E. Rhodes, 2009: Sand beach ridges record 6000
year history of extreme tropical cyclone activity in northeastern Australia.
Quaternary Science Reviews, 28(15-16), 1511-1520.
Nunez, M.N., S.A. Solman, and M.F. Cabre, 2009: Regional climate change experiments
over southern South America. II: Climate change scenarios in the late twenty-
first century. Climate Dynamics, 32(7-8), 1081-1095.
Nyberg, J., B.A. Malmgren, A. Winter, M.R. Jury, K. Halimeda Kilbourne, and T.M.
Quinn, 2007: Low Atlantic hurricane activity in the 1970s and 1980s compared
to the past 270 years. Nature, 447(7145), 698-701.
O’Gorman, P.A., 2010: Understanding the varied response of the extratropical storm
tracks to climate change. Proceedings of the National Academy of Sciences,
107(45), 19176-19180.
O’Gorman, P.A. and T. Schneider, 2008: Energy of midlatitude transient eddies in
idealized simulations of changed climates. Journal of Climate, 21(22), 5797-5806.
O’Gorman, P.A. and T. Schneider, 2009a: The physical basis for increases in
precipitation extremes in simulations of 21st-century climate change.
Proceedings of the National Academy of Sciences, 106(35), 14773-14777.
O’Gorman, P.A., and T. Schneider, 2009b: Scaling of precipitation extremes over a
wide range of climates simulated with an idealized GCM. Journal of Climate,
22(21), 5676-5685.
Oki, T., T. Nishimura, and P. Dirmeyer, 1999: Assessment of annual runoff from land
surface models using Total Runoff Integrating Pathways (TRIP). Journal of the
Meteorological Society of Japan, 77(1), 235-255.
Okin, G.S., D.A. Gillette, and J.E. Herrick, 2006: Multi-scale controls on and
consequences of aeolian processes in landscape change in arid and semi-arid
environments. Journal of Arid Environments, 65(2), 253-275.
Oouchi, K., J. Yoshimura, H. Yoshimura, R. Mizuta, S. Kusunoki, and A. Noda, 2006:
Tropical cyclone climatology in a global-warming climate as simulated in a 20
km-mesh global atmospheric model: Frequency and wind intensity analyses.
Journal of the Meteorological Society of Japan, 84(2), 259-276.
Oppikofer, T., M. Jaboyedoff, and H.R. Keusen, 2008: Collapse at the eastern Eiger
flank in the Swiss Alps. Nature Geoscience, 1(8), 531-535.
Orlowsky, B. and S.I. Seneviratne, 2011: Global changes in extremes events: Regional
and seasonal dimension. Climatic Change, doi:10.1007/s10584-011-0122-9.
Orsolini, Y. and A. Sorteberg, 2009: Projected changes in Eurasian and Arctic Summer
cyclones under global warming in the Bergen Climate Model. Atmospheric and
Oceanic Science Letters, 2(1), 62.
Osterkamp, T.E., 2005: The recent warming of permafrost in Alaska. Global and
Planetary Change, 49, 187-202.
Osterkamp, T.E., 2007: Characteristics of the recent warming of permafrost in
Alaska. Journal of Geophysical Research, 112, F02S02.
Osterkamp, T.E., M.T. Jorgenson, E.A.G. Schuur, Y.L. Shur, M.Z. Kanevskiy, J.G. Vogel,
and V.E. Tumskoy, 2009: Physical and ecological changes associated with warming
permafrost and thermokarst in Interior Alaska. Permafrost and Periglacial
Processes, 20(3), 235-256.
Osti, R., and S. Egashira, 2009: Hydrodynamic characteristics of the Tam Pokhari
glacial lake outburst flood in the Mt. Everest region, Nepal. Hydrological
Processes, 23(20), 2943-2955.
Otto-Bliesner, B.L., R. Schneider, E.C. Brady, M. Kucera, A. Abe-Ouchi, E. Bard, P.
Braconnot, M. Crucifix, C.D. Hewitt, M. Kageyama, O. Marti, A. Paul, A. Rosell-
Melé, C. Waelbroeck, S.L. Weber, M. Weinelt, and Y. Yu, 2009: A comparison of
PMIP2 model simulations and the MARGO proxy reconstruction for tropical sea
surface temperatures at last glacial maximum. Climate Dynamics
, 32(6), 799-815.
Paciorek, C.J., J.S. Risbey, V. Ventura, and R.D. Rosen, 2002: Multiple indices of
Northern Hemisphere cyclone activity, winters 1949-99. Journal of Climate,
15(13), 1573-1590.
Paeth, H. and H.-P. Thamm, 2007: Regional modelling of future African climate north
of 15 degrees S including greenhouse warming and land degradation. Climatic
Change, 83(3), 401-427.
Pagli, C. and F. Sigmundsson, 2008: Will present day glacier retreat increase volcanic
activity? Stress induced by recent glacier retreat and its effect on magmatism
at the Vatnajökull ice cap, Iceland. Geophysical Research Letters, 35, L09304.
Pall, P., M.R. Allen, and D.A. Stone, 2007: Testing the Clausius-Clapeyron constraint
on changes in extreme precipitation under CO
2
warming. Climate Dynamics,
28(4), 351-363.
Pall, P., T. Aina, D.A. Stone, P.A. Stott, T. Nozawa, A.G.J. Hilbberts, D. Lohmann, and
M.R. Allen, 2011: Anthropogenic greenhouse gas contribution to flood risk in
England and Wales in autumn 2000. Nature, 470(7334), 382-385.
Palmer, W.C., 1965: Meteorological Drought. Report 45, US Weather Bureau,
Washington, DC.
Pan, Z.T., R.W. Arritt, E.S. Takle, W.J. Gutowski, C.J. Anderson, and M. Segal, 2004:
Altered hydrologic feedback in a warming climate introduces a “warming
hole.” Geophysical Research Letters, 31, L17109.
Pattanaik, D.R. and M. Rajeevan, 2010: Variability of extreme rainfall events over
India during southwest monsoon season. Meteorological Applications, 17(1),
88-104.
Paul, F. and W. Haeberli, 2008: Spatial variability of glacier elevation changes in the
Swiss Alps obtained from two digital elevation models. Geophysical Research
Letters, 35, L21502.
Paul, F., A. Kääb, M. Maisch, T. Kellenberger, and W. Haeberli, 2004: Rapid disintegration
of Alpine glaciers observed with satellite data. Geophysical Research Letters,
31, L21402.
Pavan, V., R. Tomozeiu, C. Cacciamani, and M. Di Lorenzo, 2008: Daily precipitation
observations over Emilia-Romagna: mean values and extremes. International
Journal of Climatology, 28(15), 2065-2079.
Peel, M.C. and T.A. McMahon, 2006: Continental runoff: A quality-controlled global
runoff data set. Nature, 444(E14-E15).
Penalba, O.C. and F.A. Robledo, 2010: Spatial and temporal variability of the
frequency of extreme daily rainfall regime in the La Plata Basin during the 20th
century. Climate Change, 98(3-4), 531-550.
Perkins, S.E., A.J. Pitman, and S.A. Sisson, 2009: Smaller projected increases in 20-
year temperature returns over Australia in skill-selected climate models.
Geophysical Research Letters, 36, L06710.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
221
Peters, E., P.J.J.F. Torfs, H.A.J. van Lanen, and G. Bier, 2003: Propagation of drought
through groundwater – a new approach using linear reservoir theory.
Hydrological Processes, 17(15), 3023-3040.
Peterson, T.C. and M.J. Manton, 2008: Monitoring changes in climate extremes – A
tale of international collaboration. Bulletin of the American Meteorological
Society, 89(9), 1266-1271.
Peterson, T.C., M.A. Taylor, R. Demerritte, D.L. Duncombe, S. Burton, F. Thompson, A.
Porter, M. Mercedes, E. Villegas, R.S. Fils, A. Klein Tank, A. Martis, R. Warner, A.
Joyette, W. Mills, L. Alexander, and B. Gleason, 2002: Recent changes in climate
extremes in the Caribbean region. Journal of Geophysical Research –
Atmospheres, 107, 4601.
Peterson, T.C., X. Zhang, M. Brunet-India, and J.L. Vazquez-Aguirre, 2008a: Changes
in North American extremes derived from daily weather data. Journal of
Geophysical Research – Atmospheres, 113, DO7113.
Peterson, T.C., D. Anderson, S.J. Cohen, M. Cortez, R. Murname, C. Parmesan, D.
Phillips, R. Pulwarty, and J. Stone, 2008b: Why weather and climate extremes
matter. In: Weather and Climate Extremes in a Changing Climate. Regions of
Focus: North America, Hawaii, Caribbean, and U.S. Pacific Islands. [Karl, T.R.,
G.A. Meehl, D.M. Christopher, S.J. Hassol, A.M. Waple, and W.L. Murray (eds.)]. A
Report by the U.S. Climate Change Science Program and the Subcommittee on
Global Change Research, Washington, DC, pp. 11-33.
Petoukhov, V. and V.A. Semenov, 2010: A link between reduced Barents-Kara sea ice
and cold winter extremes over northern continents. Journal of Geophysical
Research – Atmospheres, 115, D21111.
Petrow, T. and B. Merz, 2009: Trends in flood magnitude, frequency and seasonality
in Germany in the period 1951-2002. Journal of Hydrology, 371(1-4), 129-141.
Pezza, A.B., and I. Simmonds, 2008: Large-scale factors in tropical and extratropical
cyclone transition and extreme weather events. Trends and Directions in
Climate Research: Annals of the New York Academy of Sciences, 1146, 189-211.
Pezza, A.B., I. Simmonds, and J.A. Renwick, 2007: Southern Hemisphere cyclones and
anticyclones: Recent trends and links with decadal variability in the Pacific
Ocean. International Journal of Climatology, 27(11), 1403-1419.
Pfeffer, W.T., J.T. Harper, and S. O’Neel, 2008: Kinematic constraints on glacier
contributions to 21st-century sea-level rise. Science, 321(5894), 1340-1343.
Philander, S.G., 1990: El Niño, La Niña and the Southern Oscillation. Academic Press,
San Diego, CA, 293 pp.
Piani, C., J.O. Haerter, and E. Coppola, 2010: Statistical bias correction for daily
precipitation in regional climate models over Europe. Theoretical and Applied
Climatology, 99(1-2), 187-192.
Piao, S., P. Friedlingstein, P. Ciais, N. de Noblet-Ducoudré, D. Labat, and S. Zaehle,
2007: Changes in climate and land use have a larger direct impact than rising
CO
2
on global river runoff trends. Proceedings of the National Academy of
Sciences, 104(39), 15242-15247.
Pielke Jr., R.A., J. Gratz, C.W. Landsea, D. Collins, M. Saunders, and R. Musulin, 2008:
Normalized hurricane damages in the United States. Natural Hazards Review,
9(1), 29-42.
Pinto, J.G., T. Spangehl, U. Ulbrich, and P. Speth, 2006: Assessment of winter cyclone
activity in a transient ECHAM4-OPYC3 GHG experiment. Meteorologische
Zeitschrift, 15(3), 279-291.
Pinto, J.G., U. Ulbrich, G.C. Leckebusch, T. Spangehl, M. Reyers, and S. Zacharias,
2007a: Changes in storm track and cyclone activity in three SRES ensemble
experiments with the ECHAM5/MPI-OM1 GCM. Climate Dynamics, 29(2-3),
195-210.
Pinto, J.G., E.L. Frohlich, G.C. Leckebusch, and U. Ulbrich, 2007b: Changing European
storm loss potentials under modified climate conditions according to ensemble
simulations of the ECHAM5/MPI-OM1 GCM. Natural Hazards and Earth System
Sciences, 7(1), 165-175.
Pinto, J.G., S. Zacharias, A.H. Fink, G.C. Leckebusch, and U. Ulbrich, 2009: Factors
contributing to the development of extreme North Atlantic cyclones and their
relationship with the NAO. Climate Dynamics, 32(5), 711-737.
Pinto, J.G., M. Reyers, and U. Ulbrich, 2011: The variable link between PNA and NAO
in observations and in multi-century CGCM simulations. Climate Dynamics,
36(1), 337-354.
Pirazzoli, P.A., and A. Tomasin, 2003: Recent near-surface wind changes in the
central Mediterranean and Adriatic areas. International Journal of Climatology,
23(8), 963-973.
Pitman, A.J., N. de Noblet-Ducoudré, F.T. Cruz, E. Davin, G.B. Bonan, V. Brovkin, C. M.,
D. C., V. Gayler, B.J.J.M. van den Hurk, P.J. Lawrence, M.K. van der Molen, C.
Müller, C.H. Reick, S.I. Seneviratne, B.J. Strengers, and A. Voldoire, 2009:
Uncertainties in climate responses to past land cover change: first results from
the LUCID intercomparison study. Geophysical Research Letters, 36, L14814.
Planton, S., M. Deque, F. Chauvin, and L. Terray, 2008: Expected impacts of climate
change on extreme climate events. Comptes Rendus Geoscience, 340(9-10),
564-574.
Portmann, R.W., S. Solomon, and G.C. Hegerl, 2009: Linkages between climate
change, extreme temperature and precipitation across the United States.
Proceedings of the National Academy of Sciences, 106(18), 7324-7329.
Poulter, B., F. Hattermann, E.D. Hawkins, S. Zaehle, S. Sitch, N. Restrepo-Coupe, U.
Heyder, and W. Cramer, 2010: Robust dynamics of Amazon dieback to climate
change with perturbed ecosystem model parameters. Global Change Biology,
16(9), 2476-2495.
Power, S.B. and I.N. Smith, 2007: Weakening of the Walker Circulation and apparent
dominance of El Niño both reach record levels, but has ENSO really changed?
Geophysical Research Letters, 34, L18702.
Prospero, J.M. and P.J. Lamb, 2003: African droughts and dust transport to the
Caribbean: Climate change implications. Science, 302(5647), 1024-1027.
Prospero, J.M., E. Blades, R. Naidu, and M.C. Lavoie, 2009: Reply to: African dust and
asthma in the Caribbean-medical and statistical perspectives by M.A. Monteil
and R. Antoine. International Journal of Biometeorology, 53(5), 383-385.
Prudhomme, C. and H. Davies, 2009: Assessing uncertainties in climate change
impact analyses on the river flow regimes in the UK. Part 2: future climate.
Climatic Change, 93(1-2), 197-222.
Pruszak, Z., and E. Zawadzka, 2008: Potential implications of sea-level rise for
Poland. Journal of Coastal Research, 24(2), 410-422.
Pryor, S.C., R.J. Barthelmie, and E.S. Riley, 2007: Historical evolution of wind climates
in the USA. Journal of Physics: Conference Series, 75, 012065.
Pryor, S.C., J.A. Howe, and K.E. Kunkel, 2009: How spatially coherent and statistically
robust are temporal changes in extreme precipitation in the contiguous USA?
International Journal of Climatology, 29(1), 31-45.
Purvis, M.J., P.D. Bates, and C.M. Hayes, 2008: A probabilistic methodology to estimate
future coastal flood risk due to sea level rise. Coastal Engineering, 55(12),
1062-1073.
Qian, B., S. Gameda, and H. Hayhoe, 2008: Performance of stochastic weather
generators LARS-WG and AAFC-WG for reproducing daily extremes of diverse
Canadian climates. Climate Research, 37(1), 17-33.
Raff, D.A., T. Pruitt, and L.D. Brekke, 2009: A framework for assessing flood frequency
based on climate projection information. Hydrology and Earth System Sciences,
13(11), 2119-2136.
Rahimzadeh, F., A. Asgari, and E. Fattahi, 2009: Variability of extreme temperature
and precipitation in Iran during recent decades. International Journal of
Climatology, 29(3), 329-343.
Rahmstorf, S., 2007: A semi-empirical approach to projecting future sea-level rise.
Science, 315(5810), 368-370.
Rahmstorf, S., M. Crucifix, A. Ganopolski, H. Goosse, I. Kamenkovich, R. Knutti, G.
Lohmann, R. Marsh, L.A. Mysak, Z.M. Wang, and A.J. Weaver, 2005:
Thermohaline circulation hysteresis: A model intercomparison. Geophysical
Research Letters, 32, L23605.
Raia, A., and I.F.D. Cavalcanti, 2008: The life cycle of the South American monsoon
system. Journal of Climate, 21(23), 6227-6246.
Raible, C.C., 2007: On the relation between extremes of midlatitude cyclones and
the atmospheric circulation using ERA40. Geophysical Research Letters, 34,
L07703.
Raible, C.C., P.M. Della-Marta, C. Schwierz, H. Wernli, and R. Blender, 2008:
Northern hemisphere extratropical cyclones: A comparison of detection and
tracking methods and different reanalyses. Monthly Weather Review,
136(3),
880-897.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
222
Räisänen, J., 2007: How reliable are climate models? Tellus Series A – Dynamic
Meteorology and Oceanography, 59, 2-29.
Raje, D., and P.P. Mujumdar, 2010: Constraining uncertainty in regional hydrologic
impacts of climate change: Nonstationarity in downscaling. Water Resources
Research, 46, W07543.
Rajeevan, M., J. Bhate, and A.K. Jaswal, 2008: Analysis of variability and trends of
extreme rainfall events over India using 104 years of gridded daily rainfall data.
Geophysical Research Letters, 35, L18707.
Rajendran, K., and A. Kitoh, 2008: Indian summer monsoon in future climate
projection by a super high-resolution global model. Current Science, 95(11),
1560-1569.
Ramsay, H.A., and A.H. Sobel, 2011: The effects of relative and absolute sea surface
temperature on tropical cyclone potential intensity using a single column
model. Journal of Climate, 24(1), 183-193.
Ranasinghe, R., and M. Stive, 2009: Rising seas and retreating coastlines. Climatic
Change, 97(3), 465-468.
Ranasinghe, R., R. McLoughlin, A. Short, and G. Symonds, 2004: The Southern
Oscillation Index, wave climate, and beach rotation. Marine Geology, 204(3-4),
273-287.
Randall, D.A., R.A. Wood, S. Bony, R. Colman, T. Fichefet, J. Fyfe, V. Kattsov, A. Pitman,
J. Shukla, J. Srinivasan, R.J. Stouffer, A. Sumi, and K.E. Taylor, 2007: Climate
models and their evaluation. In: Climate Change 2007: The Physical Science
Basis. Contribution of Working Group I to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change [Solomon, S., D. Qin, M. Manning,
Z. Chen, M. Marquis, K.B. Averyt, M. Tignor and H.L. Miller (eds.)]. Cambridge
University Press, Cambridge, UK, and New York, NY, pp. 589-662.
Rappaport, E.N., 2000: Loss of life in the United States associated with recent
Atlantic tropical cyclones. Bulletin of the American Meteorological Society,
81(9), 2065-2073.
Rauthe, M., M. Kunz, and C. Kottmeier, 2010: Changes in wind gust extremes over
Central Europe derived from a small ensemble of high resolution regional
climate models. Meteorologische Zeitschrift, 19(3), 299-312.
Ravanel, L. and P. Deline, 2011: Climate influence on rockfalls in high-Alpine steep
rockwalls: The north side of the Aiguilles de Chamonix (Mont Blanc massif)
since the end of the ‘Little Ice Age’. The Holocene, 21(2), 357-365.
Re, M., and V. Ricardo Barros, 2009: Extreme rainfalls in SE South America. Climatic
Change, 96(1-2), 119-136.
Rebetez, M., O. Dupont, and M. Giroud, 2008: An analysis of the July 2006 heatwave
extent in Europe compared to the record year of 2003. Theoretical and Applied
Climatology, 95(1-2), 1-7.
Redmond, K.T., Y. Enzel, P.K. House, and F. Biondi, 2002: Climate impact on flood
frequency at the decadal to millennial time scales. In: Ancient Floods, Modern
Hazards: Principles and Applications of Paleoflood Hydrology
[House, P.K., R.H.
Webb, V.R. Baker, and D.R. Levish (eds.)]. Vol. 5. American Geophysical Union
Water Science and Applications, AGU, Washington, DC, pp. 21-46.
Regonda, S.K., B. Rajagopalan, M. Clark, and J. Pitlick, 2005: Seasonal cycle shifts in
hydroclimatology over the western United States. Journal of Climate, 18(2),
372-384.
Rehman, S., 2010: Temperature and rainfall variation over Dhahran, Saudi Arabia,
(1970-2006). International Journal of Climatology, 30(3), 445-449.
Reichert, B.K., L. Bengtsson, and J. Oerlemans, 2002: Recent glacier retreat exceeds
internal variability. Journal of Climate, 15(21), 3069-3081.
Reichstein, M., P. Ciais, D. Papale, R. Valentini, S. Running, N. Viovy, W. Cramer, A.
Granier, J. Ogée, V. Allard, M. Aubinet, C. Bernhofer, N. Buchmann, A. Carrara, T.
Grünwald, M. Heimann, B. Heinesch, A. Knohl, W. Kutsch, D. Loustau, G. Manca,
G. Matteucci, F. Miglietta, J.M. Ourcival, K. Pilegaard, J. Pumpanen, S. Rambal, S.
Schaphoff, S. Seufert, J.F. Soussana, M.J. Sanz, T. Vesala, and M. Zhao, 2007:
Reduction of ecosystem productivity and respiration during the European summer
2003 climate anomaly: a joint flux tower, remote sensing and modeling analysis.
Global Change Biology, 13(3), 634-651.
Rein, B., A. Lückge, L. Reinhardt, F. Sirocko, A. Wolf, and W.C. Dullo, 2005: El Niño
variability off Peru during the last 20,000 years. Paleoceanography, 20,
PA4003.
Renard, B., and M. Lang, 2007: Use of a Gaussian copula for multivariate extreme
value analysis: Some case studies in hydrology. Advances in Water Resources,
30(4), 897-912.
Renard, B., M. Lang, P. Bois, A. Dupeyrat, O. Mestre, H. Niel, E. Sauquet, C. Prudhomme,
S. Parey, E. Paquet, L. Neppel, and J. Gailhard, 2008: Regional methods for trend
detection: Assessing field significance and regional consistency. Water
Resources Research, 44, W08419.
Reynolds, J.M., 1998: High-altitude glacial lake hazard assessment and mitigation:
a Himalayan perspective. Geological Society, London, Engineering Geology
Special Publications, 15(1), 25-34.
Reynolds, J.M., A. Dolecki, and C. Portocarrero, 1998: The construction of a drainage
tunnel as part of glacial lake hazard mitigation at Hualcan, Cordillera Blanca,
Peru. Engineering Geology Special Publications, 15(1), 41-48.
Richardson, S.D. and J.M. Reynolds, 2000: An overview of glacial hazards in the
Himalayas. Quaternary International, 65, 31-47.
Rickenmann, D. and M. Zimmermann, 1993: The 1987 debris flows in Switzerland:
documentation and analysis. Geomorphology (Amsterdam), 8(2-3), 175-189.
Rignot, E., I. Velicogna, M.R. van den Broeke, A. Monaghan, and J. Lenaerts, 2011:
Acceleration of the contribution of the Greenland and Antarctic ice sheets to
sea level rise. Geophysical Research Letters,
38(L05503).
Rist, A., and M. Phillips, 2005: First results of investigations on hydrothermal processes
within the active layer above alpine permafrost in steep terrain. Norsk
Geografisk Tidsskrift, 59(2), 177-183.
Robeson, S., 2004: Trends in time-varying percentiles of daily minimum and maximum
temperature over North America. Geophysical Research Letters, 31, L04203.
Robinson, W.A., 2000: A baroclinic mechanism for the eddy feedback on the zonal
index. Journal of the Atmospheric Sciences, 57(3), 415-422.
Robock, A., K.Y. Vinnikov, G. Srinivasan, J.K. Entin, S.E. Hollinger, N.A. Speranskaya, S.
Liu, and A. Namkhai, 2000: The global soil moisture data bank. Bulletin of the
American Meteorological Society, 81(6), 1281-1299.
Robson, A.J., T.K. Jones, D.W. Reed, and A.C. Bayliss, 1998: A study of national trend
and variation in UK floods. International Journal of Climatology, 18(2), 165-182.
Rocha, A., P. Melo-Goncalves, C. Marques, J. Ferreira, and J.M. Castanheira, 2008: High-
frequency precipitation changes in southeastern Africa due to anthropogenic
forcing. International Journal of Climatology, 28(9), 1239-1253.
Rockel, B., and K. Woth, 2007: Extremes of near-surface wind speed over Europe and
their future changes as estimated from an ensemble of RCM simulations
Climatic Change, 81(S1), 267-280.
Rodda, J.C., M.A. Little, H.J.E. Rodda, and P.E. McSharry, 2010: A comparative study
of the magnitude, frequency and distribution of intense rainfall in the United
Kingdom. International Journal of Climatology, 30(12), 1776-1783.
Rodier, J.A., and M. Roche, 1984: World Catalogue of Maximum Observed Floods.
IAHS Pub. No. 143, IAHS Press, Wallingford, UK.
Rodrigo, F.S., 2010: Changes in the probability of extreme daily precipitation
observed from 1951 to 2002 in the Iberian Peninsula. International Journal of
Climatology, 30(10), 1512-1525.
Rodríguez-Puebla, C., A. Encinas, L. García-Casado, and S. Nieto, 2010: Trends in
warm days and cold nights over the Iberian Peninsula: relationships to large-
scale variables. Climatic Change, 100(3), 667-684.
Roer, I., W. Haeberli, and M. Avian, 2008: Observations and considerations on
destabilizing active rock glaciers in the European Alps. In: Permafrost on a
Warming Planet: Impacts on Ecosystems, Infrastructure and Climate [Kane, D.L.
and K.M. Hinkel (eds.)]. Proceedings of the Ninth International Conference on
Permafrost at University of Alaska, Fairbanks, 29 June - 03 July 2008, Institute
of Northern Engineering, pp 1505-1510.
Roesch, A., 2006: Evaluation of surface albedo and snow cover in AR4 coupled
climate models. Journal of Geophysical Research, 111, D15111.
Romanovsky, V.E., D.S. Drozdov, N.G. Oberman, G.V. Malkova, A.L. Kholodov, S.S.
Marchenko, N.G. Moskalenko, D.O. Sergeev, N.G. Ukraintseva, A.A. Abramov,
D.A. Gilichinsky, and A.A. Vasiliev, 2010: Thermal state of permafrost in Russia.
Permafrost and Periglacial Processes, 21(2), 136-155.
Ropelewski, C.F., and M.A. Bell, 2008: Shifts in the statistics of daily rainfall in South
America conditional on ENSO phase. Journal of Climate, 21(5), 849-865.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
223
Rosenbluth, B., H.A. Fuenzalida, and P. Aceituno, 1997: Recent temperature variations
in southern South America. International Journal of Climatology, 17(1), 67-85.
Rosenzweig, C., G. Casassa, D.J. Karoly, A. Imeson, C. Liu, A. Menzel, S. Rawlins, T.L.
Root, B. Seguin, and P. Tryjanowski, 2007: Assessment of Observed Changes and
Responses in Natural and Managed Systems. In: Climate Change 2007. Impacts,
Adaptation and Vulnerability. Contribution of Working Group II to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change [Parry,
M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der Linden, and C.E. Hanson (eds.)].
Cambridge University Press, Cambridge, UK, pp. 79-131.
Rotstayn, L.D., M.D. Keywood, B.W. Forgan, A.J. Gabric, I.E. Galbally, J.L. Gras, A.K.
Luhar, G.H. McTainsh, R.M. Mitchell, and S.A. Young, 2009: Possible impacts of
anthropogenic and natural aerosols on Australian climate: A review.
International Journal of Climatology, 29(4), 461-479.
Rott, H., S.H. Yueh, D.W. Cline, C. Duguay, R. Essery, C. Haas, F. Heliere, M. Kern, G.
Macelloni, E. Malnes, T. Nagler, J. Pulliainen, H. Rebhan, and A. Thompson, 2010:
Cold regions hydrology high-resolution observatory for snow and cold land
processes. IEEE Proceedings, 98(5), 752-765.
Rowell, D.P. and R.G. Jones, 2006: Causes and uncertainty of future summer drying
over Europe. Climate Dynamics, 27(2-3), 281-299.
Ruckstuhl, C. and J.R. Norris, 2009: How do aerosol histories affect solar “dimming”
and “brightening” over Europe? IPCC-AR4 models versus observations. Journal
of Geophysical Research Atmospheres, 114, D00D04.
Rusticucci, M. and M. Barrucand, 2004: Observed trends and changes in temperature
extremes over Argentina. Journal of Climate, 17(20), 4099-4107.
Rusticucci, M. and M. Renom, 2008: Variability and trends in indices of quality-
controlled daily temperature extremes in Uruguay. International Journal of
Climatology, 28(8), 1083-1095.
Rusticucci, M., J. Marengo, O. Penalba, and M. Renom, 2010: An intercomparison of
model-simulated in extreme rainfall and temperature events during the last
half of the twentieth century. Part 1: mean values and variability. Climate
Change, 98(3-4), 493-508.
Ryan, B.F., I.G. Watterson, and J.L. Evans, 1992: Tropical cyclone frequencies inferred
from Gray’s yearly genesis parameter - Validation of GCM tropical climates.
Geophysical Research Letters, 19(18), 1831-1834.
Sailor, D.J., M. Smith, and M. Hart, 2008: Climate change implications for wind power
resources in the Northwest United States. Renewable Energy, 33(11), 2393-2406.
Saito, K., M. Kimoto, T. Zhang, K. Takata, and S. Emori, 2007: Evaluating a high-
resolution climate model: Simulated hydrothermal regimes in frozen ground
regions and their change under the global warming scenario. Journal of
Geophysical Research – Earth Surface, 112, F02S11.
Santer, B.D., T.M.L. Wigley, P.J. Glecker, C. Bonfils, M.F. Wehner, K. AchutaRoa, T.P.
Barnett, J.S. Boyle, W. Brüggemann, M. Fiorino, N. Gillet, J.E. Hansen, P.D. Jones,
S.A. Klein, G.A. Meehl, S.C.B. Raper, R.W. Reynolds, K.E. Taylor, and W.M.
Washington, 2006: Forced and unforced ocean temperature changes in Atlantic
and Pacific tropical cyclogenesis regions. Proceedings of the National Academy
of Sciences, 103(38), 13905-13910.
Santer, B.D., C. Mears, F.J. Wentz, K.E. Taylor, P.J. Gleckler, T.M.L. Wigley, T.P. Barnett,
J.S. Boyle, W. Bruggemann, N.P. Gillett, S.A. Klein, G.A. Meehl, T. Nozawa, D.W.
Pierce, P.A. Stott, W.M. Washington, and M.F. Wehner, 2007: Identification of
human-induced changes in atmospheric moisture content. Proceedings of the
National Academy of Sciences, 104(39), 15248-15253.
Santos, C.A., and J.I.B. Brito, 2007: Análise dos índices de extremos para o semi-
árido do Brasil e suas relações com TSM e IVDN. Revista Brasileira de
Meteorologia, 22(3), 303-312.
Santos, C.A.C., J. Brito, I, B., T.V.R. Rao, and E.A. Meneses, 2009: Tendências dos
Índices de precipitação no Estado do Ceará. Revista Brasileira de Meteorologia,
24(1), 39-47.
Sasaki, W., and H. Toshiyuki, 2007: Interannual variability and predictability of
summertime significant wave heights in the western North Pacific. Journal of
Oceanography, 63(2), 203-213.
Sauber, J.M., and B.F. Molnia, 2004: Glacier ice mass fluctuations and fault instability
in tectonically active Southern Alaska. Global and Planetary Change, 42(1-4),
279-293.
Sauber, J.M., and N.A. Ruppert, 2008: Rapid ice mass loss: does it have an influence
on earthquake occurrence in southern Alaska? In: Active Tectonics and Seismic
Potential of Alaska [Freymueller, J.T., P.J. Haeussler, R. Wesson, and G. Ekström
(eds.)]. American Geophysical Union, Washington, DC.
Scaife, A.A., C.K. Folland, L.V. Alexander, A. Moberg, and J.R. Knight, 2008: European
climate extremes and the North Atlantic Oscillation. Journal of Climate, 21(1),
72-83.
Scaife, A.A., F. Kucharski, C.K. Folland, J. Kinter, S. Bronnimann, D. Fereday, A.M.
Fischer, S. Grainger, E.K. Jin, I.S. Kang, J.R. Knight, S. Kusunoki, N.C. Lau, M.J.
Nath, T. Nakaegawa, P. Pegion, S. Schubert, P. Sporyshev, J. Syktus, J.H. Yoon, N.
Zeng, and T. Zhou, 2009: The CLIVAR C20C project: selected twentieth century
climate events. Climate Dynamics, 33(5), 603-614.
Scaife, A.A., T. Spangehl, D. Fereday, U. Cubasch, U. Langematz, H. Akiyoshi, S. Bekki,
P. Braesicke, N. Butchart, M. Chipperfield, A. Gettelman, S. Hardiman, M.
Michou, E. Rozanov, and T. Shepherd, 2011: Climate change projections and
stratosphere–troposphere interaction. Climate Dynamics, doi:10.1007/s00382-
00011-01080-00387.
Schaefer, K., T. Zhang, L. Bruhwiler, and A.P. Barrett, 2011: Amount and timing of
permafrost carbon release in response to climate warming. Tellus Series B –
Chemical and Physical Meteorology, 63(2), 165-180.
Schär, C., and G. Jendritzky, 2004: Climate change: Hot news from summer 2003.
Nature, 432(7017), 559-560.
Schär, C., P.L. Vidale, D. Lüthi, C. Frei, C. Häberli, M.A. Liniger, and C. Appenzeller,
2004: The role of increasing temperature variability in European summer
heatwaves. Nature, 427(322), 332-336.
Scheffer, M., J. Bascompte, W.A. Brock, V. Brovkin, S.R. Carpenter, V. Dakos, H. Held,
E.H. van Nes, M. Rietkerk, and G. Sugihara, 2009: Early-warning signals for
critical transitions. Nature, 461(7260), 53-59.
Schiefer, E., B. Menounos, and R. Wheate, 2007: Recent volume loss of British
Columbian glaciers, Canada. Geophysical Research Letters
, 34, L16503.
Schmidli, J., and C. Frei, 2005: Trends of heavy precipitation and wet and dry spells
in Switzerland during the 20th century. International Journal of Climatology,
25(6), 753-771.
Schmidli, J., C.M. Goodess, C. Frei, M.R. Haylock, Y. Hundecha, J. Ribalaygua, and T.
Schmith, 2007: Statistical and dynamical downscaling of precipitation: An
evaluation and comparison of scenarios for the European Alps. Journal of
Geophysical Research – Atmospheres, 112, D04105.
Schneidereit, A., R. Blender, K. Fraedrich, and F. Lunkeit, 2007: Icelandic climate and
north Atlantic cyclones in ERA-40 reanalyses. Meteorologische Zeitschrift,
16(1), 17-23.
Schölzel, C., and P. Friederichs, 2008: Multivariate non-normally distributed random
variables in climate research - introduction to the copula approach. Nonlinear
Processes in Geophysics, 15(5), 761-772.
Schubert, S.D., M.J. Suarez, P.J. Pegion, R.D. Koster, and J.T. Bacmeister, 2004: On the
cause of the 1930s Dust Bowl. Science, 303(5665), 1855-1859.
Schubert, S.D., Y. Chang, M.J. Suarez, and P.J. Pegion, 2008a: ENSO and wintertime
extreme precipitation events over the contiguous united states. Journal of
Climate, 21(1), 22-39.
Schubert, S.D., M.J. Suarez, P.J. Pegion, R.D. Koster, and J.T. Bacmeister, 2008b:
Potential predictability of long-term drought and pluvial conditions in the US
Great Plains. Journal of Climate, 21(4), 802-816.
Schubert, S.D., D. Gutzler, H.L. Wang, A. Dai, T. Delworth, C. Deser, K. Findell, R. Fu, W.
Higgins, M. Hoerling, B. Kirtman, R. Koster, A. Kumar, D. Legler, D. Lettenmaier,
B. Lyon, V. Magana, K. Mo, S. Nigam, P. Pegion, A. Phillips, R. Pulwarty, D. Rind,
A. Ruiz-Barradas, J. Schemm, R. Seager, R. Stewart, M. Suarez, J. Syktus, M.F.
Ting, C.Z. Wang, S. Weaver, and N. Zeng, 2009: A US CLIVAR project to assess
and compare the responses of global climate models to drought-related SST
forcing patterns: Overview and results. Journal of Climate, 22(19), 5251-5272.
Scileppi, E., and J.P. Donnelly, 2007: Sedimentary evidence of hurricane strikes in
western Long Island, New York. Geochemistry Geophysics Geosystems, 8, Q06011.
Scott, K.M., J.W. Vallance, N. Kerle, J. Luis Macías, W. Strauch, and G. Devoli, 2005:
Catastrophic precipitation-triggered lahar at Casita volcano, Nicaragua: occurrence,
bulking and transformation. Earth Surface Processes and Landforms, 30(1), 59-79.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
224
Seager, R., M.F. Ting, I. Held, Y. Kushnir, J. Lu, G. Vecchi, H.P. Huang, N. Harnik, A.
Leetmaa, N.C. Lau, C.H. Li, J. Velez, and N. Naik, 2007: Model projections of an
imminent transition to a more arid climate in southwestern North America.
Science, 316(5828), 1181-1184.
Seager, R., A. Tzanova, and J. Nakamura, 2009: Drought in the southeastern United
States: Causes, variability over the last millennium, and the potential for future
hydroclimate change. Journal of Climate, 22(19), 5021-5045.
Seierstad, I.A., D.B. Stephenson, and N.G. Kvamsto, 2007: How useful are
teleconnection patterns for explaining variability in extratropical storminess?
Tellus Series A – Dynamic Meteorology and Oceanography, 59(2), 170-181.
Seleshi, Y. and P. Camberlin, 2006: Recent changes in dry spell and extreme rainfall
events in Ethiopia. Theoretical and Applied Climatology, 83(1-4), 181-191.
Semenov, M.A., 2008: Simulation of extreme weather events by a stochastic weather
generator. Climate Research, 35(3), 203-212.
Semmler, T., S. Varghese, R. McGrath, P. Nolan, S. Wang, P. Lynch, and C. O’Dowd,
2008: Regional model simulation of North Atlantic cyclones: Present climate
and idealized response to increased sea surface temperature. Journal of
Geophysical Research – Atmospheres, 113, D02107.
Sen Roy, S., 2009: A spatial analysis of extreme hourly precipitation patterns in
India. International Journal of Climatology, 29(3), 345-355.
Seneviratne, S.I., D. Lüthi, M. Litschi, and C. Schär, 2006a: Land-atmosphere
coupling and climate change in Europe. Nature, 443(7108), 205-209.
Seneviratne, S.I., R.D. Koster, Z.C. Guo, P.A. Dirmeyer, E. Kowalczyk, D. Lawrence, P.
Liu, C.H. Lu, D. Mocko, K.W. Oleson, and D. Verseghy, 2006b: Soil moisture
memory in AGCM simulations: Analysis of global land-atmosphere coupling
experiment (GLACE) data. Journal of Hydrometeorology, 7(5), 1090-1112.
Seneviratne, S.I., T. Corti, E.L. Davin, M. Hirschi, E. Jaeger, I. Lehner, B. Orlowsky, and
A.J. Teuling, 2010: Investigating soil moisture-climate interactions in a changing
climate: A review. Earth Science Reviews, 99(3-4), 125-161.
Sepp, M., and J. Jaagus, 2011: Changes in the activity and tracks of Arctic cyclones.
Climatic Change, 105(3-4), 577-595.
Shanahan, T.M., J.T. Overpeck, K.J. Anchukaitis, J.W. Beck, J.E. Cole, D.L. Dettman, J.A.
Peck, C.A. Scholz, and J.W. King, 2009: Atlantic forcing of persistent drought in
West Africa. Science, 324(5925), 377-380.
Shao, Y., and C.H. Dong, 2006: A review on East Asian dust storm climate, modelling
and monitoring. Global and Planetary Change, 52(1-4), 1-22.
Shaw, S.B., and S.J. Riha, 2011: Assessing temperature-based PET equations under a
changing climate in temperate, deciduous forests. Hydrological Processes,
25(9), 1466-1478.
Sheffield, J., and E.F. Wood, 2008a: Global trends and variability in soil moisture and
drought characteristics, 1950-2000, from observation-driven simulations of the
terrestrial hydrologic cycle. Journal of Climate, 21(3), 432-458.
Sheffield, J., and E.F. Wood, 2008b: Projected changes in drought occurrence under
future global warming from multi-model, multi-scenario, IPCC AR4 simulations.
Climate Dynamics, 31(1), 79-105.
Sheppard, C., D.J. Dixon, M. Gourlay, A. Sheppard, and R. Payet, 2005: Coral mortality
increases wave energy reaching shores protected by reef flats: Examples from
the Seychelles. Estuarine Coastal and Shelf Science, 64(2-3), 223-234.
Sherwood, S.C., and M. Huber, 2010: An adaptability limit to climate change due
to heat stress. Proceedings of the National Academy of Sciences, 107(21),
9552-9555.
Shiklomanov, A.I., R.B. Lammers, M.A. Rawlins, L.C. Smith, and T.M. Pavelsky,
2007: Temporal and spatial variations in maximum river discharge from a new
Russian data set. Journal of Geophysical Research – Biogeosciences, 112,
G04S53.
Shiklomanov, N.I., D.A. Streletskiy, F.E. Nelson, R.D. Hollister, V.E. Romanovsky, C.E.
Tweedie, J.G. Bockheim, and J. Brown, 2010: Decadal variations of active-layer
thickness in moisture-controlled landscapes, Barrow, Alaska. Journal of
Geophysical Research – Biogeosciences, 115, G00I04.
Shongwe, M.E., G.J. van Oldenborgh, B.J.J.M. van den Hurk, B. de Boer, C.A.S.
Coelho, and M.K. van Aalst, 2009: Projected changes in mean and extreme
precipitation in Africa under global warming. Part I: Southern Africa. Journal of
Climate, 22(13), 3819-3837.
Shongwe, M.E., G.J. van Oldenborgh, B.J.J.M. van den Hurk, B. de Boer, C.A.S.
Coelho, and M.K. van Aalst, 2011: Projected changes in mean and extreme
precipitation in Africa under global warming. Part II: East Africa. Journal of
Climate, 24(14), 3718-3732.
Shukla, J., 2007: Monsoon mysteries. Science, 318(5848), 204-205.
Shulmeister, J., D.T. Rodbell, M.K. Gagan, and G.O. Seltzer, 2006: Inter-hemispheric
linkages in climate change: paleo-perspectives for future climate change.
Climates of the Past, 2(2), 167-185.
Sidle, R.C., and H. Ochiai, 2006: Landslides: processes, prediction, and land use. Water
Resources Monograph 18, American Geophysical Union, Washington, DC, 312 pp.
Sigmundsson, F., V. Pinel, B. Lund, F. Albino, C. Pagli, H. Geirsson, and E. Sturkell,
2010: Climate effects on volcanism: influence on magmatic systems of loading
and unloading from ice mass variations, with examples from Iceland.
Philosophical Transactions of the Royal Society A, 368(1919), 2519-2534.
Sillmann, J. and E. Roeckner, 2008: Indices for extreme events in projections of
anthropogenic climate change. Climatic Change, 86(1-2), 83-104.
Silva, A.G., and P. Azevedo, 2008: Índices de tendências de Mudanças Climáticas no
Estado da Bahia. Engenheiria Ambiental, 5, 141-151.
Silva Dias, M.A.F., S. Rutledge, P. Kabat, P.L. Silva Dias, C. Nobre, G. Fisch, A.J.
Dolman, E. Zipser, M. Garstang, A.O. Manzi, J.D. Fuentes, H.R. Rocha, J. Marengo,
A. Plana-Fattori, L.D.A. Sá, R.C.S. Alvalá, M.O. Andreae, P. Artaxo, R. Gielow, and
L. Gatti, 2002: Cloud and rain processes in a biosphere-atmosphere interaction
context in the Amazon Region. Journal of Geophysical Research – Atmospheres,
107(D20), 8072.
Silvestri, G.E. and C.S. Vera, 2009: Nonstationary impacts of the Southern Annular
Mode on Southern Hemisphere climate. Journal of Climate, 22(22), 6142-6148.
Simmonds, I. and K. Keay, 2002: Surface fluxes of momentum and mechanical energy
over the North Pacific and North Atlantic Oceans. Meteorology and Atmospheric
Physics, 80(1), 1-18.
Simmonds, I., C. Burke, and K. Keay, 2008: Arctic climate change as manifest in
cyclone behavior. Journal of Climate, 21(22), 5777-5796.
Simmons, A.J., K.M. Willett, P.D. Jones, P.W. Thorne, and D.P. Dee, 2010: Low-frequency
variations in surface atmospheric humidity, temperature, and precipitation:
Inferences from reanalyses and monthly gridded observational data sets.
Journal of Geophysical Research – Atmospheres, 115, D01110.
Sinha, A., K.G. Cannariato, L.D. Stott, H. Cheng, R.L. Edwards, M.G. Yadava, R.
Ramesh, and I.B. Singh, 2007: A 900-year (600 to 1500 A. D.) record of the
Indian summer monsoon precipitation from the core monsoon zone of India.
Geophysical Research Letters, 34, L16707.
Sinton, J., K. Grönvold, and K. Sæmundsson, 2005: Postglacial eruptive history of
the Western Volcanic Zone, Iceland. Geochemistry Geophysics Geosystems, 6,
Q12009.
Sjogren, P., 2009: Sand mass accumulation rate as a proxy for wind regimes in the
SW Barents Sea during the past 3 ka. Holocene, 19(4), 591-598.
Small, I., J. van der Meer, and R.E.G., Upshur, 2001: Acting on an environmental
health disaster: The case of the Aral Sea. Environmental Health Perspectives,
109(6), 547-549.
Smith, J.M., M.A. Cialone, T.V. Wamsley, and T.O. McAlpin, 2010: Potential impact of
sea level rise on coastal surges in southeast Louisiana. Ocean Engineering,
37(1), 37-47.
Smith, K., and R. Ward, 1998: Floods. Physical Processes and Human Impacts. John
Wiley, Chichester, UK, 382 pp.
Smith, L.C., 2000: Trends in Russian Arctic river-ice formation and breakup, 1917 to
1994. Physical Geography, 21(1), 46-56.
Smith, L.C., Y. Sheng, G.M. MacDonald, and L.D. Hinzman, 2005: Disappearing Arctic
lakes. Science, 308(5727), 1429.
Smith, S.L., S.A. Wolfe, D.W. Riseborough, and F.M. Nixon, 2009: Active-layer
characteristics and summer climatic indices, Mackenzie Valley, Northwest
Territories, Canada. Permafrost and Periglacial Processes, 20(2), 201-220.
Smith, S.L., V.E. Romanovsky, A.G. Lewkowicz, C.R. Burn, M. Allard, G.D. Clow, K.
Yoshikawa, and J. Throop, 2010: Thermal state of permafrost in North America:
a contribution to the international polar year. Permafrost and Periglacial
Processes, 21(2), 117-135.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
225
Smits, A., A.M.G.K. Tank, and G.P. Konnen, 2005: Trends in storminess over the
Netherlands, 1962-2002. International Journal of Climatology, 25(10), 1331-1344.
Soares, W.R., and J.A. Marengo, 2009: Assessments of moisture fluxes east of the
Andes in South America in a global warming scenario. International Journal of
Climatology, 29(10), 1395-1414.
Son, S.W., L.M. Polvani, D.W. Waugh, H. Akiyoshi, R. Garcia, D. Kinnison, S. Pawson, E.
Rozanov, T.G. Shepherd, and K. Shibata, 2008: The impact of stratospheric ozone
recovery on the Southern Hemisphere westerly jet. Science, 320(5882),
1486-1489.
Son, S.W., E.P. Gerber, J. Perlwitz, L.M. Polvani, N.P. Gillett, K.H. Seo, V. Eyring, T.G.
Shepherd, D. Waugh, H. Akiyoshi, J. Austin, A. Baumgaertner, S. Bekki, P.
Braesicke, C. Bruhl, N. Butchart, M.P. Chipperfield, D. Cugnet, M. Dameris, S.
Dhomse, S. Frith, H. Garny, R. Garcia, S.C. Hardiman, P. Jockel, J.F. Lamarque, E.
Mancini, M. Marchand, M. Michou, T. Nakamura, O. Morgenstern, G. Pitari, D.A.
Plummer, J. Pyle, E. Rozanov, J.F. Scinocca, K. Shibata, D. Smale, H. Teyssedre, W.
Tian, and Y. Yamashita, 2010: Impact of stratospheric ozone on Southern
Hemisphere circulation change: A multimodel assessment. Journal of
Geophysical Research – Atmospheres, 115, D00M07.
Sørensen, L.S., S.B. Simonsen, K. Nielsen, P. Lucas-Picher, G. Spada, G.
Adalgeirsdottir, R. Forsberg, and C.S. Hvidberg, 2011: Mass balance of the
Greenland ice sheet (2003–2008) from ICESat data – the impact of interpolation,
sampling and firn density. The Cryosphere, 5(1), 173-186.
Sorensson, A.A., C.G. Menéndez, R. Ruscica, P. Alexander, P. Samuelsson, and U.
Willén, 2010: Projected precipitation changes in South America: a dynamical
downscaling within CLARIS. Meteorologische Zeitschrift, 19(4), 347-355.
Sorteberg, A. and J.E. Walsh, 2008: Seasonal cyclone variability at 70 degrees N and
its impact on moisture transport into the Arctic. Tellus Series A – Dynamic
Meteorology and Oceanography, 60(3), 570-586.
Sosio, R., G.B. Crosta, and O. Hungr, 2008: Complete dynamic modeling calibration
for the Thurwieser rock avalanche (Italian Central Alps) Engineering Geology,
100(1-2), 11-26.
Stahl, K., H. Hisdal, J. Hannaford, L.M. Tallaksen, H.A.J. van Lanen, E. Sauquet, S.
Demuth, M. Fendekova, and J. Jódar, 2010: Streamflow trends in Europe:
evidence from a dataset of near-natural catchments. Hydrology and Earth
System Sciences Discussion, 7(4), 5769-5804.
Stephenson, D.B., H.F. Diaz, and R.J. Murnane, 2008: Definition, diagnosis, and
origin of extreme weather and climate events. In: Climate Extremes and Society
[Murnane, R.J. and H.F. Diaz (eds.)]. Cambridge University Press, Cambridge, UK,
pp. 11-23.
Stephenson, T.S., C.M. Goodess, M.R. Haylock, A.A. Chen, and M.A. Taylor, 2008:
Detecting inhomogeneities in Caribbean and adjacent Caribbean temperature
data using sea-surface temperatures. Journal of Geophysical Research –
Atmospheres, 113, D21116.
Sterl, A., C. Severijns, H. Dijkstra, W. Hazeleger, G.J. van Oldenborgh, M. van den
Broeke, G. Burgers, B. van den Hurk, P.J. van Leeuwen, and P. van Velthoven,
2008: When can we expect extremely high surface temperatures? Geophysical
Research Letters, 35, L14703.
Sterl
, A., H. van den Brink, H. de Vries, R. Haarsma, and E. van Meijgaard, 2009: An
ensemble study of extreme storm surge related water levels in the North Sea in
a changing climate. Ocean Science, 5(3), 369-378.
Sterr, H., 2008: Assessment of vulnerability and adaptation to sea-level rise for the
coastal zone of Germany. Journal of Coastal Research, 24(2), 380-393.
Stewart, I.S., J. Sauber, and J. Rose, 2000: Glacio-seismotectonics: ice sheets, crustal
deformation and seismicity. Quaternary Science Reviews, 19(14-15), 1367-1389.
Stive, M.J.F., Z.B. Wang, and C. Lakhan, 2003: Morphodynamic modeling of tidal
basins and coastal inlets. In: Advances in Coastal Modeling [Lakhan, V.C. (ed.)].
Elsevier, Amsterdam, The Netherlands, pp. 367-392.
Stoffel, M., I. Lièvre, D. Conus, M.A. Grichting, H. Raetzo, H.W. Gärtner, and M.
Monbaron, 2005: 400 years of debris-flow activity and triggering weather
conditions: Ritigraben, Valais, Switzerland. Arctic, Antarctic, and Alpine
Research, 37(3), 387-395.
Stott, P.A., D.A. Stone, and M.R. Allen, 2004: Human contribution to the European
heatwave of 2003. Nature, 432(7017), 610-614.
Stott, P.A., N.P. Gillett, G.C. Hegerl, D. Karoly, D. Stone, X. Zhang, and F.W. Zwiers,
2010: Detection and attribution of climate change: a regional perspective. Wiley
Interdisciplinary Reviews: Climate Change, 1(2), 192-211.
Su, B., Z.W. Kundzewicz, and T. Jiang, 2009: Simulation of extreme precipitation over
the Yangtze River Basin using Wakeby distribution. Theoretical and Applied
Climatology, 96(3-4), 209-219.
Sugahara, S., R.P. da Rocha, and R. Silveira, 2009: Non-stationary frequency analysis
of extreme daily rainfall in Sao Paulo, Brazil. International Journal of
Climatology, 29(9), 1339-1349.
Sugi, M., A. Noda, and N. Sato, 2002: Influence of the global warming on tropical
cyclone climatology: An experiment with the JMA global model. Journal of the
Meteorological Society of Japan, 80(2), 249-272.
Sugi, M., H. Murakami, and J. Yoshimura, 2009: A reduction in global tropical cyclone
frequency due to global warming. SOLA, 5, 164-167.
Sugiyama, M., H. Shiogama, and S. Emori, 2010: Precipitation extreme changes
exceeding moisture content increases in MIROC and IPCC climate models.
Proceedings of the National Academy of Sciences, 107(2), 571-575.
Suppiah, R., K. Hennessy, P.H. Whetton, K. McInnes, I. Macadam, J. Bathols, J.
Ricketts, and C.M. Page, 2007: Australian climate change projections derived
from simulations performed for the IPCC 4th Assessment Report. Australian
Meteorological Magazine, 56(3), 131-152.
Suter
, S., M. Laternser, W. Haeberli, R. Frauenfelder, and M. Hoelzle, 2001: Cold firn
and ice of high-altitude glaciers in the Alps: measurements and distribution
modelling. Journal of Glaciology, 47(156), 85-96.
Sutton, R.T., B. Dong, and J.M. Gregory, 2007: Land/sea warming ratio in response
to climate change: IPCC AR4 model results and comparison with observations.
Geophysical Research Letters, 34, L02701.
Suursaar, U., and T. Kullas, 2009: Decadal variations in wave heights off Cape Kelba,
Saaremaa Island, and their relationships with changes in wind climate.
Oceanologia, 51(1), 39-61.
Svensson, C., and D.A. Jones, 2002: Dependence between extreme sea surge, river
flow and precipitation in eastern Britain. International Journal of Climatology,
22(10), 1149-1168.
Swanson, K.L., 2008: Non-locality of Atlantic tropical cyclone intensities.
Geochemistry Geophysics Geosystems, 9, Q04V01.
Syvitski, J.P.M., A.J. Kettner, I. Overeem, E.W.H. Hutton, M.T. Hannon, G.R.
Brakenridge, J. Day, C. Vorosmarty, Y. Saito, L. Giosan, and R.J. Nicholls, 2009:
Sinking deltas due to human activities. Nature Geoscience, 2(10), 681-686.
Tallaksen, L.M., H. Hisdal, and H.A.J. Van Lanen, 2009: Space-time modelling of
catchment scale drought characteristics. Journal of Hydrology, 375(3-4), 363-372.
Tamura, T., K. Horaguchi, Y. Saito, L.N. Van, M. Tateishi, K.O.T. Thi, F. Nanayama, and
K. Watanabe, 2010: Monsoon-influenced variations in morphology and sediment
of a mesotidal beach on the Mekong River delta coast. Geomorphology, 116(1-2),
11-23.
Tanaka, H.L., N. Ishizaki, and D. Nohara, 2005: Intercomparison of the intensities and
trends of Hadley, Walker and monsoon circulations in the global warming
projections. SOLA, 1, 77-80.
Taschetto, A.S., C.C. Ummenhofer, A. Sen Gupta, and M.H. England, 2009: The effect
of anomalous warming in the central Pacific on the Australian monsoon.
Geophysical Research Letters, 36, L12704.
Taye, M.T., V. Ntegeka, N.P. Ogiramoi, and P. Willems, 2011: Assessment of climate
change impact on hydrological extremes in two source regions of the Nile River
Basin. Hydrology and Earth System Sciences, 15(1), 209-222.
Taylor, A.E., K. Wang, S.L. Smith, M.M. Burgess, and A.S. Judge, 2006: Canadian Arctic
Permafrost Observatories: Detecting contemporary climate change through
inversion of subsurface temperature time series. Journal of Geophysical
Research – Solid Earth, 111, B02411.
Tebaldi, C. and R. Knutti, 2007: The use of the multi-model ensemble in probabilistic
climate projections. Philosophical Transactions of the Royal Society A,
365(1857), 2053-2075.
Tebaldi, C. and B. Sanso, 2009: Joint projections of temperature and precipitation
change from multiple climate models: a hierarchical Bayesian approach. Journal
of the Royal Statistical Society Series A – Statistics in Society, 172(1), 83-106.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
226
Tebaldi, C., K. Hayhoe, J.M. Arblaster, and G.A. Meehl, 2006: Going to the extremes.
An intercomparison of model-simulated historical and future changes in
extreme events. Climatic Change, 79(3-4), 185-211.
Teng, H., W.M. Washington, and G.A. Meehl, 2008: Interannual variations and future
change of wintertime extratropical cyclone activity over North America in
CCSM3. Climate Dynamics, 30(7-8), 673-686.
Teuling, A.J., M. Hirschi, A. Ohmura, M. Wild, M. Reichstein, P. Ciais, N. Buchmann, C.
Ammann, L. Montagnani, A.D. Richardson, G. Wohlfahrt, and S.I. Seneviratne,
2009: A regional perspective on trends in continental evaporation. Geophysical
Research Letters, 36, L02404.
Teuling, A.J., S.I. Seneviratne, R. Stockli, M. Reichstein, E. Moors, P. Ciais, S. Luyssaert,
B. van den Hurk, C. Ammann, C. Bernhofer, E. Dellwik, D. Gianelle, B. Gielen, T.
Grunwald, K. Klumpp, L. Montagnani, C. Moureaux, M. Sottocornola, and G.
Wohlfahrt, 2010: Contrasting response of European forest and grassland energy
exchange to heatwaves. Nature Geoscience, 3(10), 722-727.
Textor, C., M. Schulz, S. Guibert, S. Kinne, Y. Balkanski, S. Bauer, T. Berntsen, T. Berglen,
O. Boucher, M. Chin, F. Dentener, T. Diehl, R. Easter, H. Feichter, D. Fillmore, S.
Ghan, P. Ginoux, S. Gong, J.E. Kristjansson, M. Krol, A. Lauer, J.F. Lamarque, X.
Liu, V. Montanaro, G. Myhre, J. Penner, G. Pitari, S. Reddy, O. Seland, P. Stier, T.
Takemura, and X. Tie, 2006: Analysis and quantification of the diversities of
aerosol life cycles within AeroCom. Atmospheric Chemistry and Physics, 6(7),
1777-1813.
Themeßl, M.J., A. Gobiet, and A. Leuprecht, 2011: Empirical-statistical
downscaling and error correction of daily precipitation from regional climate
models. International Journal of Climatology, 31(10), 1530-1544, doi:
10.1002/joc.2168.
Thodsen, H., 2007: The influence of climate change on stream flow in Danish rivers.
Journal of Hydrology, 333(2-4), 226-238.
Thomas, D.S.G., M. Knight, and G.F.S. Wiggs, 2005: Remobilization of southern
African desert dune systems by twenty-first century global warming. Nature,
435(7046), 1218-1221.
Thompson, L.G., E. Mosley-Thompson, H. Brecher, M. Davis, B. Leon, D. Les, P.N.
Lin, T. Mashiotta, and K. Mountain, 2006: Abrupt tropical climate change: Past
and present. Proceedings of the National Academy of Sciences, 103(28),
10536-10543.
Tietsche, S., D. Notz, J.H. Jungclaus, and J. Marotzke, 2011: Recovery mechanisms of
Arctic summer sea ice. Geophysical Research Letters, 38, L02707.
Timbal, B., P. Hope, and S. Charles, 2008: Evaluating the consistency between
statistically downscaled and global dynamical model climate change projections.
Journal of Climate, 21(22), 6052-6059.
Timbal, B., E. Fernandez, and Z. Li, 2009: Generalization of a statistical downscaling
model to provide local climate change projections for Australia. Environmental
Modelling & Software, 24(3), 341-358.
Timmermann, A., S. McGregor, and F.F. Jin, 2010: Wind effects on past and future
regional sea level trends in the southern Indo-Pacific. Journal of Climate,
23(16), 4429-4437.
Tolika, K., P. Maheras, and I. Tegoulias, 2009: Extreme temperatures in Greece
during 2007: Could this be a “return to the future”? Geophysical Research
Letters, 36, L10813.
Tomassini, L. and D. Jacob, 2009: Spatial analysis of trends in extreme precipitation
events in high-resolution climate model results and observations for Germany.
Journal of Geophysical Research – Atmospheres, 114, D12113.
Toreti, A. and F. Desiato, 2008: Changes in temperature extremes over Italy in the last
44 years. International Journal of Climatology, 28(6), 733-745.
Toreti, A., E. Xoplaki, D. Maraun, F.G. Kuglitsch, H. Wanner, and J. Luterbacher, 2010:
Characterisation of extreme winter precipitation in Mediterranean coastal sites
and associated anomalous atmospheric circulation patterns. Natural Hazards
and Earth System Sciences, 10(5), 1037-1050.
Trenberth, K.E., D. Stepaniak, J. Hurrell, and M. Fiorino, 2001: Quality of reanalyses
in the tropics. Journal of Climate, 14(7), 1499-1510.
Trenberth, K.E., J. Fasullo, and L. Smith, 2005: Trends and variability in
column-integrated atmospheric water vapor. Climate Dynamics, 24(7-8),
741-758.
Trenberth, K.E., P.D. Jones, P. Ambenje, R. Bojariu, D. Easterling, A. Klein Tank, D.
Parker, F. Rahimzadeh, J.A. Renwick, M. Rusticucci, B. Solden, and P. Zhai, 2007:
Observations: Surface and atmospheric climate change. In: Climate Change
2007: The Physical Science Basis. Contribution of Working Group I to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change
[Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor
and H.L. Miller (eds.)]. Cambridge University Press, Cambridge, UK, and New
York, NY, pp. 235-336.
Trewin, B. and H. Vermont, 2010: Changes in the frequency of record temperatures
in Australia, 1957-2009. Australian Meteorological and Oceanographic Journal,
60(2), 113-119.
Trigo, I.F., 2006: Climatology and interannual variability of storm-tracks in the Euro-
Atlantic sector: a comparison between ERA-40 and NCEP/NCAR reanalyses.
Climate Dynamics, 26(2-3), 127-143.
Trömel, S. and C.D. Schönwiese, 2005: A generalized method of time series
decomposition into significant components including probability assessments
of extreme events and application to observed German precipitation data.
Meteorologische Zeitschrift, 14, 417-427.
Trouet, V., J. Esper, N.E. Graham, A. Baker, J.D. Scourse, and D.C. Frank, 2009:
Persistent positive North Atlantic Oscillation mode dominated the medieval
climate anomaly. Science, 324(5923), 78-80.
Tsimplis, M.N., M. Marcos, and S. Somot, 2008: 21st century Mediterranean sea level
rise: Steric and atmospheric pressure contributions from a regional model.
Global and Planetary Change, 63(2-3), 105-111.
Tu, M., M.J. Hall, P.J.M. de Laat, and M.J.M. de Wit, 2005: Extreme floods in the Meuse
river over the past century: aggravated by land-use changes? Physics and
Chemistry of the Earth, 30(4-5), 267-276.
Tudhope, A.W., C.P. Chilcott, M.T. McCulloch, E.R. Cook, J. Chappell, R.M. Ellam, D.W.
Lea, J.M. Lough, and G.B. Shimmield, 2001: Variability in the El Niño-Southern
Oscillation through a glacial-interglacial cycle. Science, 291(5508), 1511-1517.
Tuffen, H., 2010: How will melting of ice affect volcanic hazards in the twenty-first
century? Philosophical Transactions of the Royal Society A,
368(1919), 2535-2558.
Turner, A.G. and J.M. Slingo, 2009: Subseasonal extremes of precipitation and active-
break cycles of the Indian summer monsoon in a climate-change scenario.
Quarterly Journal of the Royal Meteorological Society, 135(640), 549-567.
Turner, J., S.R. Colwell, G.J. Marshall, T.A. Lachlan-Cope, A.M. Carleton, P.D. Jones, V.
Lagun, P.A. Reid, and S. Iagovkina, 2005: Antarctic climate change during the
last 50 years. International Journal of Climatology, 25(3), 279-294.
Turpeinen, H., A. Hampel, T. Karow, and G. Maniatis, 2008: Effect of ice sheet growth
and melting on the slip evolution of thrust faults. Earth and Planetary Science
Letters, 269(1-2), 230-241.
Uchiyama, T., R. Mizuta, K. Kamiguchi, A. Kitoh, and A. Noda, 2006: Changes in
temperature-based extremes indices due to global warming projected by a
global 20-km-mesh atmospheric model. SOLA, 2, 68-71.
Ulbrich, U., J.G. Pinto, H. Kupfer, G.C. Leckebusch, T. Spangehl, and M. Reyers, 2008:
Changing northern hemisphere storm tracks in an ensemble of IPCC climate
change simulations. Journal of Climate, 21(8), 1669-1679.
Ulbrich, U., G.C. Leckebusch, and J.G. Pinto, 2009: Extra-tropical cyclones in the
present and future climate: a review. Theoretical and Applied Climatology,
96(1-2), 117-131.
Ullmann, A., P.A. Pirazzoli, and A. Tomasin, 2007: Sea surges in Camargue: Trends
over the 20th century. Continental Shelf Research, 27(7), 922-934.
Ullmann, A., P.A. Pirazzoli, and V. Moron, 2008: Sea surges around the Gulf of Lions
and atmospheric conditions. Global and Planetary Change, 63(2-3), 203-214.
Ummenhofer, C.C., A. Sen Gupta, M.J. Pook, and M.H. England, 2008: Anomalous
rainfall over southwest Western Australia forced by Indian Ocean sea surface
temperatures. Journal of Climate, 21(19), 5113-5134.
Ummenhofer, C.C., A. Sen Gupta, M.H. England, and C.J.C. Reason, 2009a:
Contributions of Indian Ocean sea surface temperatures to enhanced East
African rainfall. Journal of Climate, 22(4), 993-1013.
Ummenhofer, C.C., A. Sen Gupta, A.S. Taschetto, and M.H. England, 2009b:
Modulation of Australian precipitation by meridional gradients in East Indian
Ocean sea surface temperature. Journal of Climate, 22(21), 5597-5610.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
227
Ummenhofer, C.C., M.H. England, G.A. Meyers, P.C. McIntosh, M.J. Pook, J.S. Risbey,
A. Sen Gupta, and A.S. Taschetto, 2009c: What causes Southeast Australia’s
worst droughts? Geophysical Research Letters, 36, L04706.
Vallée, S. and S. Payette, 2007: Collapse of permafrost mounds along a subarctic river
over the last 100years (northern Québec). Geomorphology, 90(1-2), 162-170.
Vallve, M.B., and J. Martin-Vide, 1998: Secular climatic oscillations as indicated by
catastrophic floods in the Spanish Mediterranean coastal area (14th-19th
centuries). Climatic Change, 38(4), 473-491.
Van den Brink, H.W., G.P. Konnen, J.D. Opsteegh, G.J. Van Oldenborgh, and G.
Burgers, 2005: Estimating return periods of extreme events from ECMWF
seasonal forecast ensembles. International Journal of Climatology, 25(10),
1345-1354.
van den Hurk, B.J.J.M., and E. van Meijgaard, 2010: Diagnosing land–atmosphere
interaction from a Regional Climate Model simulation over West Africa. Journal
of Hydrometeorology, 11(2), 467-481.
van der Linden, P. and J.F.B. Mitchell (eds.), 2009: ENSEMBLES: Climate Change and
its Impacts: Summary of research and results from the ENSEMBLES project. Met
Office Hadley Centre, Exeter, UK, 160 pp.
van der Schrier, G., K.R. Briffa, T.J. Osborn, and E.R. Cook, 2006a: Summer moisture
availability across North America. Journal of Geophysical Research –
Atmospheres, 111, D11102.
van der Schrier, G., K.R. Briffa, P.D. Jones, and T.J. Osborn, 2006b: Summer moisture
variability across Europe. Journal of Climate, 19(12), 2818-2834.
van Oldenborgh, G.J., S. Philip, and M. Collins, 2005: El Niño in a changing climate:
a multi-model study. Ocean Science, 1(2), 267-298.
van Oldenborgh, G.J., S. Drijfhout, A. van Ulden, R. Haarsma, A. Sterl, C. Severijns,
W. Hazeleger, and H. Dijkstra, 2009: Western Europe is warming much faster
than expected. Climate of the Past, 5(1), 1-12.
van Pelt, S.C., P. Kabat, H.W. ter Maat, B.J.J.M. van den Hurk, and A.H. Weerts, 2009:
Discharge simulations performed with a hydrological model using bias corrected
regional climate model input. Hydrology and Earth System Sciences, 13(12),
2387-2397.
Vautard, R., P. Yiou, F. D’Andrea, N. de Noblet, N. Viovy, C. Cassou, J. Polcher, P. Ciais,
M. Kageyama, and Y. Fan, 2007: Summertime European heat and drought
waves induced by wintertime Mediterranean rainfall deficit. Geophysical
Research Letters, 34, L07711.
Vautard, R., J. Cattiaux, P. Yiou, J.-N. Thepaut, and P. Ciais, 2010: Northern
Hemisphere atmospheric stilling partly attributed to an increase in surface
roughness. Nature Geoscience, 3(11), 756-761.
Vecchi, G.A. and T.R. Knutson, 2008: On estimates of historical North Atlantic tropical
cyclone activity. Journal of Climate, 21(14)
, 3580-3600.
Vecchi, G.A. and B.J. Soden, 2007a: Effect of remote sea surface temperature change
on tropical cyclone potential intensity. Nature, 450(7172), 1066-1070.
Vecchi, G.A. and B.J. Soden, 2007b: Global warming and the weakening of the
tropical circulation. Journal of Climate, 20(17), 4316-4340.
Vecchi, G.A. and B.J. Soden, 2007c: Increased tropical Atlantic wind shear in model
projections of global warming. Geophysical Research Letters, 34, L08702.
Vecchi, G.A., and A.T. Wittenberg, 2010: El Niño and our future climate: where do we
stand? Climate Change, 1(2), 260-270.
Vecchi, G.A., B.J. Soden, A.T. Wittenberg, I.M. Held, A. Leetmaa, and M.J. Harrison, 2006:
Weakening of tropical Pacific atmospheric circulation due to anthropogenic
forcing. Nature, 441, 73-76.
Vecchi, G.A., K.L. Swanson, and B.J. Soden, 2008: Whither hurricane activity. Science,
322(5902), 687-689.
Velicogna, I., 2009: Increasing rates of ice mass loss from the Greenland and
Antarctic ice sheets revealed by GRACE. Geophysical Research Letters, 36,
L19503.
Vera, C. and G. Silvestri, 2009: Precipitation interannual variability in South America
from the WCRP-CMIP3 multi-model dataset. Climate Dynamics, 32(7-8),
1003-1014.
Vera, C., G. Silvestri, B. Liebmann, and P. Gonzalez, 2006: Climate change scenarios
for seasonal precipitation in South America from IPCC-AR4 models. Geophysical
Research Letters, 33, L13707.
Vermeer, M. and S. Rahmstorf, 2009: Global sea level linked to global temperature.
Proceedings of the National Academy of Sciences, 106(51), 21527-21532.
Vicente-Serrano, S.M. and J.I. Lopez-Moreno, 2008: Nonstationary influence of the
North Atlantic Oscillation on European precipitation. Journal of Geophysical
Research – Atmospheres, 113, D20120.
Vicente-Serrano, S.M., S. Begueria, and J.I. Lopez-Moreno, 2010: A multiscalar
drought index sensitive to global warming: The standardized precipitation
evapotranspiration index. Journal of Climate, 23(7), 1696-1718.
Vidal, J.P., E. Martin, L. Franchisteguy, F. Habets, J.M. Soubeyroux, M. Blanchard, and
M. Baillon, 2010: Multilevel and multiscale drought reanalysis over France with
the Safran-Isba-Modcou hydrometeorological suite. Hydrology and Earth
System Sciences, 14(3), 459-478.
Vilibic, I. and J. Sepic, 2010: Long-term variability and trends of sea level storminess
and extremes in European Seas. Global and Planetary Change,
71(1-2), 1-12.
Villarini, G., F. Serinaldi, J.A. Smith, and W.F. Krajewski, 2009: On the stationarity of
annual flood peaks in the continental United States during the 20th century.
Water Resources Research, 45, W08417.
Vincent, C., E. Le Meur, D. Six, M. Funk, M. Hoelzle, and S. Preunkert, 2007: Very high-
elevation Mont Blanc glaciated areas not affected by the 20th century climate
change. Journal of Geophysical Research – Atmospheres, 112, D09120.
Vincent, C., S. Auclair, and E.L.e. Meur, 2010: Outburst flood hazard for glacier-
dammed Lac de Rochemelon, France. Journal of Glaciology, 56(195), 91-100.
Vincent, L.A. and E. Mekis, 2006: Changes in daily and extreme temperature and
precipitation indices for Canada over the 20th century. Atmosphere-Ocean,
44(2), 177-193.
Vincent, L.A., X. Zhang, B.R. Bonsal, and W.D. Hogg, 2002: Homogenization of daily
temperatures over Canada. Journal of Climate, 15(11), 1322-1334.
Vincent, L.A., T.C. Peterson, V.R. Barros, M.B. Marino, M. Rusticucci, G. Carrasco, E.
Ramirez, L.M. Alves, T. Ambrizzi, M.A. Berlato, A.M. Grimm, J.A. Marengo, J.
Molion, D.F. Moncunill, E. Rebello, Y.M.T. Anunciação, J. Quintana, J.L. Santos, J.
Baez, G. Coronel, J. Garcia, I. Trebejo, M. Bidegain, M.R. Haylock, and D. Karoly,
2005: Observed trends in indices of daily temperature extremes in South
America 1960-2000. Journal of Climate, 18(23), 5011-5023.
Vinchon, C., S. Aubie, Y. Balouin, L. Closset, M. Garcin, D. Idier, and C. Mallet, 2009:
Anticipate response of climate change on coastal risks at regional scale in
Aquitaine and Languedoc Roussillon (France). Ocean & Coastal Management,
52(1), 47-56.
Vinnikov, K.Y., A. Robock, N.A. Speranskaya, and A. Schlosser, 1996: Scales of
temporal and spatial variability of midlatitude soil moisture. Journal of
Geophysical Research – Atmospheres, 101(D3), 7163-7174.
Voldoire, A., and J.F. Royer, 2004: Tropical deforestation and climate variability.
Climate Dynamics, 22(8), 857-874.
Vose, R.S., D.R. Easterling, and B. Gleason, 2005: Maximum and minimum temperature
trends for the globe: an update through 2004. Geophysical Research Letters,
32, L23822.
Vrac, M., M. Stein, and K. Hayhoe, 2007: Statistical downscaling of precipitation
through nonhomogeneous stochastic weather typing. Climate Research, 34(3),
169-184.
Vuichard, D. and M. Zimmermann, 1987: The 1985 catastrophic drainage of a
moraine-dammed lake, Khumbu Himal, Nepal: cause and consequences.
Mountain Research and Development, 7(2), 91-110.
Wakazuki, Y., M. Nakamura, S. Kanada, and C. Muroi, 2008: Climatological
reproducibility evaluation and future climate projection of extreme precipitation
events in the Baiu season using a high-resolution non-hydrostatic RCM in
comparison with an AGCM. Journal of the Meteorological Society of Japan,
86(6), 951-967.
Walkden, M. and M. Dickson, 2008: Equilibrium erosion of soft rock shores with a
shallow or absent beach under increased sea level rise. Marine Geology,
251(1-2), 75-84.
Walsh, J., 2005: Cryosphere and hydrology. In: Arctic Climate Impact Assessment.
Cambridge Univesity Press, Cambridge, UK, pp. 183-242.
Wan, H., X.L. Wang, and V.R. Swail, 2010: Homogenization and trend analysis of
Canadian near-surface wind speeds. Journal of Climate, 23(5), 1209-1225.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
228
Wang, A.H., T.J. Bohn, S.P. Mahanama, R.D. Koster, and D.P. Lettenmaier, 2009:
Multimodel ensemble reconstruction of drought over the Continental United
States. Journal of Climate, 22(10), 2694-2712.
Wang, B. and Q.H. Ding, 2006: Changes in global monsoon precipitation over the
past 56 years. Geophysical Research Letters, 33, L06711.
Wang, B., Q.H. Ding, X.H. Fu, I.S. Kang, K. Jin, J. Shukla, and F. Doblas-Reyes, 2005:
Fundamental challenge in simulation and prediction of summer monsoon rainfall.
Geophysical Research Letters, 32, L15711.
Wang, G. and H.H. Hendon, 2007: Sensitivity of Australian rainfall to inter-El Niño
variations. Journal of Climate, 20(16), 4211-4226.
Wang, G.L., 2005: Agricultural drought in a future climate: results from 15 global
climate models participating in the IPCC 4th assessment. Climate Dynamics,
25(7-8), 739-753.
Wang, J. and X. Zhang, 2008: Downscaling and projection of winter extreme daily
precipitation over North America. Journal of Climate, 21(5), 923-937.
Wang, S., R. McGrath, J. Hanafin, P. Lynch, T. Semmler, and P. Nolan, 2008: The impact
of climate change on storm surges over Irish waters. Ocean Modelling, 25(1-2),
83-94.
Wang, X., P. Zhai, and C. Wang, 2009: Variations in extratropical cyclone activity in
northern East Asia. Advances in Atmospheric Sciences, 26(3), 471-479.
Wang, X.L., V.R. Swail, and F.W. Zwiers, 2006: Climatology and changes of extratropical
cyclone activity: Comparison of ERA-40 with NCEP-NCAR reanalysis for 1958-
2001. Journal of Climate, 19(13), 3145-3166.
Wang, X.L., V.R. Swail, F.W. Zwiers, X. Zhang, and Y. Feng, 2009a: Detection of external
influence on trends of atmospheric storminess and northern oceans wave
heights. Climate Dynamics, 32(2-3), 189-203.
Wang, X.L., F.W. Zwiers, V.R. Swail, and Y. Feng, 2009b: Trends and variability of
storminess in the Northeast Atlantic region, 1874-2007. Climate Dynamics,
33(7-8), 1179-1195.
Wang, X.M., Y. Yang, Z.B. Dong, and C.X. Zhang, 2009: Responses of dune activity
and desertification in China to global warming in the twenty-first century.
Global and Planetary Change, 67(3-4), 167-185.
Wang, Y.Q., L.R. Leung, J.L. McGregor, D.K. Lee, W.C. Wang, Y.H. Ding, and F. Kimura,
2004: Regional climate modeling: Progress, challenges, and prospects. Journal
of the Meteorological Society of Japan, 82(6), 1599-1628.
Wanner, H., J. Beer, J. Butikofer, T.J. Crowley, U. Cubasch, J. Fluckiger, H. Goosse, M.
Grosjean, F. Joos, J.O. Kaplan, M. Kuttel, S.A. Muller, I.C. Prentice, O. Solomina,
T.F. Stocker, P. Tarasov, M. Wagner, and M. Widmann, 2008: Mid- to Late Holocene
climate change: an overview. Quaternary Science Reviews, 27(19-20)
, 1791-1828.
Watterson, I.G., 2006: The intensity of precipitation during extratropical cyclones in
global warming simulations: a link to cyclone intensity? Tellus Series A –
Dynamic Meteorology and Oceanography, 58(1), 82-97.
Webb, A.P. and P. Kench, 2010: The dynamic response of reef islands to sea level rise:
Evidence from multi-decadal analysis of island change in the central Pacific.
Global and Planetary Change, 72(3), 234-246.
Webster, P.J., 2008: Myanmar’s deadly daffodil. Nature Geoscience, 1, 488-490.
Webster, P.J., G.J. Holland, J.A. Curry, and H.R. Chang, 2005: Changes in tropical
cyclone number, duration, and intensity in a warming environment. Science,
309(5742), 1844-1846.
Wehner, M.F., G. Bala, P. Duffy, A.A. Mirin, and R. Romano, 2010: Towards direct
simulation of future tropical cyclone statistics in a high-resolution global
atmospheric model. Advances in Meteorology, 2010, Article 915303.
Weisse, R. and H. Günther, 2007: Wave climate and long-term changes for the
Southern North Sea obtained from a high-resolution hindcast 1958-2002.
Ocean Dynamics, 57(3), 161-172.
Weisse, R., H. Von Storch, and F. Feser, 2005: Northeast Atlantic and North Sea
storminess as simulated by a regional climate model during 1958-2001 and
comparison with observations. Journal of Climate, 18(3), 465-479.
Wells, N., S. Goddard, and M.J. Hayes, 2004: A self-calibrating Palmer Drought
Severity Index. Journal of Climate, 17(12), 2335-2351.
Weng, H., S.K. Behera, and T. Yamagata, 2009: Anomalous winter climate conditions
in the Pacific Rim during recent El Niño Modoki and El Niño events. Climate
Dynamics, 32(5), 663-674.
Wernli, H., S. Dirren, M.A. Liniger, and M. Zillig, 2002: Dynamical aspects of the life
cycle of the winter storm ‘Lothar’ (24-26 December 1999). Quarterly Journal of
the Royal Meteorological Society, 128(580), 405-429.
Wettstein, J.J., and J.M. Wallace, 2010: Observed patterns of month-to-month storm-
track variability and their relationship to the background flow. Journal of the
Atmospheric Sciences, 67(5), 1420-1437.
Whitney, F.A., H.J. Freeland, and M. Robert, 2007: Persistently declining oxygen levels
in the interior waters of the eastern subarctic Pacific. Progress in
Oceanography, 75(2), 179-199.
Wieczorek, G.F., T. Glade, M. Jakob, and O. Hungr, 2005: Climatic factors influencing
occurrence of debris flows. In: Debris-Flow Hazards and Related Phenomena
[Jakob, M. and O. Hungr (eds.)]. Springer, Berlin, Germany, pp. 325-362.
Wiggs, G.F.S., A.J. Baird, and R.J. Atherton, 2004: The dynamic effects of moisture on
the entrainment and transport of sand by wind. Geomorphology, 59(1-4), 13-30.
Wilby, R.L., K.J. Beven, and N.S. Reynard, 2008: Climate change and fluvial flood risk
in the UK: more of the same? Hydrological Processes, 22(14), 2511-2523.
Willett, K.M., N.P. Gillett, P.D. Jones, and P.W. Thorne, 2007: Attribution of observed
surface humidity changes to human influence. Nature, 449(7163), 710-712.
Willett, K.M., P.D. Jones, N.P. Gillett, and P.W. Thorne, 2008: Recent changes in
surface humidity: Development of the HadCRUH dataset. Journal of Climate,
21(20), 5364-5383.
Willett, K.M., P.D. Jones, P.W. Thorne, and N.P. Gillett, 2010: A comparison of large
scale changes in surface humidity over land in observations and CMIP3 general
circulation models. Environmental Research Letters, 5(2), 025210.
Wilson, P., J. McGourty, and M.D. Bateman, 2004: Mid- to late-Holocene coastal dune
event stratigraphy for the north coast of Northern Ireland. Holocene, 14(3),
406-416.
Wing, A.A., A.H. Sobel, and S.J. Camargo, 2007: Relationship between the potential
and actual intensities of tropical cyclones on interannual time scales.
Geophysical Research Letters, 34, L08810.
Wittenberg, A.T., 2009: Are historical records sufficient to constrain ENSO simulations?
Geophysical Research Letters, 36, L12702.
Wong, S., A.E. Dessler, N.M. Mahowald, P.R. Colarco, and A. da Silva, 2008: Long-term
variability in Saharan dust transport and its link to North Atlantic sea surface
temperature. Geophysical Research Letters, 35, L07812.
Woodroffe, C.D., 2008: Reef-island topography and the vulnerability of atolls to
sea-level rise. Global and Planetary Change, 62(1-2), 77-96.
Woodruff, J.D., J.P. Donnelly, K.A. Emanuel, and P. Lane, 2008a: Assessing sedimentary
records of paleohurricane activity using modeled hurricane climatology.
Geochemistry Geophysics Geosystems, 9, Q09V10.
Woodruff, J.D., J.P. Donnelly, D. Mohrig, and W.R. Geyer, 2008b: Reconstructing
relative flooding intensities responsible for hurricane-induced deposits from
Laguna Playa Grande, Vieques, Puerto Rico. Geology, 36(5), 391-394.
Woodworth, P.L. and D.L. Blackman, 2004: Evidence for systematic changes in
extreme high waters since the mid-1970s. Journal of Climate, 17(6), 1190-1197.
Woodworth, P.L., M. Menendez, and W.R. Gehrels, 2011: Evidence for century-
timescale acceleration in mean sea levels and for recent changes in extreme
sea levels. Surveys in Geophysics, 32(4-5), 603-618, doi: 10.1007/s10712-
10011-19112-10718.
Woollings, T.J., B.J. Hoskins, M. Blackburn, and P. Berrisford, 2008: A new Rossby
wave-breaking interpretation of the North Atlantic Oscillation. Journal of
Atmospheric Science, 65(2), 609-626.
Woollings
, T.J., A. Hannachi, B. Hoskins, and A. Turner, 2010: A regime view of the
North Atlantic Oscillation and its response to anthropogenic forcing. Journal of
Climate, 23(6), 1291-1307.
Wu, Q., and T. Zhang, 2010: Changes in active layer thickness over the Qinghai-
Tibetan Plateau from 1995-2007. Journal of Geophysical Research, 115,
D09107.
Wu, Y., M. Ting, R. Seager, H.-P. Huang, and M. Cane, 2011: Changes in storm tracks
and energy transports in a warmer climate simulated by the GFDL CM2.1
model. Climate Dynamics, 37(1-2), 53-72.
Xie, B., Q. Zhang, and Y. Wang, 2008: Trends in hail in China during 1960–2005.
Geophysical Research Letters, 35, L13801.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
229
Xie, S.-P., C. Deser, G.A. Vecchi, J. Ma, H. Teng, and A.T. Wittenberg, 2010: Global
warming pattern formation: Sea surface temperature and rainfall. Journal of
Climate, 23(4), 966-986.
Xin, W., L. Shiyin, G. Wanqin, and X. Junli, 2008: Assessment and simulation of glacier
lake outburst floods for Longbasaba and Pida Lakes, China. Mountain Research
and Development, 28(3), 310-317.
Xoplaki, E., J.F. Gonzalez-Rouco, J. Luterbacher, and H. Wanner, 2003: Mediterranean
summer air temperature variability and its connection to the large-scale
atmospheric circulation and SSTs. Climate Dynamics, 20(7-8), 723-739.
Xu, M., C.P. Chang, C.B. Fu, Y. Qi, A. Robock, D. Robinson, and H.M. Zhang, 2006:
Steady decline of east Asian monsoon winds, 1969-2000: Evidence from direct
ground measurements of wind speed. Journal of Geophysical Research –
Atmospheres, 111, D24111.
Yao, C., S. Yang, W. Qian, Z. Lin, and M. Wen, 2008: Regional summer precipitation
events in Asia and their changes in the past decades. Journal of Geophysical
Research – Atmospheres, 113, D17107.
Yeh, S.W., J.S. Kug, B. Dewitte, M.H. Kwon, B.P. Kirtman, and F.F. Jin, 2009: El Niño in
a changing climate. Nature, 461, 511-514.
Yin, J., S.M. Griffies, and R.J. Stouffer, 2010: Spatial variability of sea level rise in
twenty-first century projections. Journal of Climate, 23(17), 4585-4607.
Yiou, P., P. Ribereau, P. Naveau, M. Nogaj, and R. Brazdil, 2006: Statistical analysis of
floods in Bohemia (Czech Republic) since 1825. Hydrological Sciences Journal,
51(5), 930-945.
Yoshimura, J., M. Sugi, and A. Noda, 2006: Influence of greenhouse warming on
tropical cyclone frequency. Journal of the Meteorological Society of Japan,
84(2), 405-428.
You, Q., S. Kang, E. Aguilar, and Y. Yan, 2008: Changes in daily climate extremes in the
eastern and central Tibetan Plateau during 1961-2005. Journal of Geophysical
Research – Atmospheres, 113, D07101.
You, Q., S. Kang, E. Aguilar, N. Pepin, W.A. Flügel, Y. Yan, Y. Xu, Y. Zhang, and J. Huang,
2011: Changes in daily climate extremes in China and their connection to the
large scale atmospheric circulation during 1961-2003. Climate Dynamics,
36(11-12), 2399-2417.
Young, I.R., S. Zieger, and A.V. Babanin, 2011: Global trends in wind speed and wave
height. Science, 332(6028), 451-455.
Yu, J.Y. and H.Y. Kao, 2007: Decadal changes of ENSO persistence barrier in SST and
ocean heat content indices: 1958-2001. Journal of Geophysical Research –
Atmospheres, 112, D13106.
Yu, K.F., J.X. Zhao, Q. Shi, and Q.S. Meng, 2009: Reconstruction of storm/tsunami
records over the last 4000 years using transported coral blocks and lagoon
sediments in the southern South China Sea. Quaternary International
, 195(1-2),
128-137.
Zahn, M. and H. von Storch, 2010: Decreased frequency of North Atlantic polar lows
associated with future climate warming. Nature, 467(7313), 309-312.
Zemp, M., W. Haeberli, M. Hoelzle, and F. Paul, 2006: Alpine glaciers to disappear
within decades? Geophysical Research Letters, 33, L13504.
Zemp, M., W. Haeberli, and J. Eamer, 2007: Glaciers and ice caps. In: Global Outlook
for Ice and Snow. UNEP/GRID, Arendal, Norway, pp. 115-152.
Zhai, P.M., X. Zhang, H. Wan, and X.H. Pan, 2005: Trends in total precipitation and
frequency of daily precipitation extremes over China. Journal of Climate, 18(7),
1096-1108.
Zhang, K.Q., B.C. Douglas, and S.P. Leatherman, 2004: Global warming and coastal
erosion. Climatic Change, 64(1-2), 41-58.
Zhang, Q., Y. Guan, and H. Yang, 2008: ENSO amplitude change in observation and
coupled models. Advances in Atmospheric Sciences, 25(3), 361-366.
Zhang, R. and T.L. Delworth, 2009: A new method for attributing climate variations
over the Atlantic Hurricane Basin’s main development region. Geophysical
Research Letters, 36, L06701.
Zhang, T., 2005: Influence of the seasonal snow cover on the ground thermal regime:
An overview. Reviews of Geophysics, 43, RG4002.
Zhang, T., R.G. Barry, K. Knowles, J.A. Heginbottom, and J. Brown, 1999: Statistics and
characteristics of permafrost and ground ice distribution in the Northern
Hemisphere. Polar Geography, 23(2), 147-169.
Zhang, T., O.W. Frauenfeld, M.C. Serreze, A. Etringer, C. Oelke, J. McCreight, R.G.
Barry, D. Gilichinsky, D. Yang, H. Ye, F. Ling, and S. Chudinova, 2005: Spatial and
temporal variability of active layer thickness over the Russian Arctic drainage
basin. Journal of Geophysical Research, 110, D16101.
Zhang, X., K.D. Harvey, W.D. Hogg, and T.R. Yuzyk, 2001: Trends in Canadian
streamflow. Water Resources Research, 37(4), 987-998.
Zhang, X., F.W. Zwiers, and G. Li, 2004: Monte Carlo experiments on the detection of
trends in extreme values. Journal of Climate, 17(10), 1945-1952.
Zhang, X., E. Aguilar, S. Sensoy, H. Melkonyan, U. Tagiyeva, N. Ahmed, N. Kutaladze,
F. Rahimzadeh, A. Taghipour, T.H. Hantosh, P. Albert, M. Semawi, M.K. Ali, M.H.S.
Al-Shabibi, Z. Al-Oulan, T. Zatari, I.A. Khelet, S. Hamoud, R. Sagir, M. Demircan,
M. Eken, M. Adiguzel, L. Alexander, T.C. Peterson, and T. Wallis, 2005: Trends in
Middle East climate extreme indices from 1950 to 2003. Journal of Geophysical
Research – Atmospheres, 110, D22104.
Zhang, X., F.W. Zwiers, G.C. Hegerl, F.H. Lambert, N.P. Gillett, S. Solomon, P.A. Stott,
and T. Nozawa, 2007: Detection of human influence on twentieth-century
precipitation trends. Nature, 448(7152), 461-465.
Zhang, X., J. Wang, F.W. Zwiers, P. Ya, and P.Y. Groisman, 2010: The influence of large
scale climate variability on winter maximum daily precipitation over North
America. Journal of Climate, 23(11), 2902-2915.
Zhang, X.D., J.E. Walsh, J. Zhang, U.S. Bhatt, and M. Ikeda, 2004: Climatology and
interannual variability of arctic cyclone activity: 1948-2002. Journal of Climate,
17(12), 2300-2317.
Zhang, X.D., A. Sorteberg, J. Zhang, R. Gerdes, and J.C. Comiso, 2008: Recent radical
shifts of atmospheric circulations and rapid changes in Arctic climate system.
Geophysical Research Letters, 35, L22701.
Zhang, X.Y., S.L. Gong, T.L. Zhao, R. Arimoto, Y.Q. Wang, and Z.J. Zhou, 2003: Sources
of Asian dust and role of climate change versus desertification in Asian dust
emission. Geophysical Research Letters, 30(24), 2272.
Zhang, Y.Q., C.M. Liu, Y.H. Tang, and Y.H. Yang, 2007: Trends in pan evaporation and
reference and actual evapotranspiration across the Tibetan Plateau. Journal of
Geophysical Research – Atmospheres, 112, D12110.
Zhao, C.S., X. Dabu, and Y. Li, 2004: Relationship between climatic factors and dust
storm frequency in Inner Mongolia of China. Geophysical Research Letters, 31,
L01103.
Zhao, L., Q. Wu, S.S. Marchenko, and N. Sharkhuu, 2010: Thermal state of permafrost
and active layer in Central Asia during the international polar year. Permafrost
and Periglacial Processes, 21(2), 198-207.
Zhao, M., I.M. Held, S.J. Lin, and G.A. Vecchi, 2009: Simulations of global hurricane
climatology, interannual variability, and response to global warming using a
50-km resolution GCM. Journal of Climate, 22(24), 6653-6678.
Zhong, D.C., 1999: The dynamic changes and trends of modern desert in China.
Advance in Earth Sciences, 14(3), 229-234 (in Chinese).
Zhou, T.J., L.X. Zhang, and H.M. Li, 2008a: Changes in global land monsoon area and
total rainfall accumulation over the last half century. Geophysical Research
Letters, 35, L16707.
Zhou, T.J., R.C. Yu, H.M. Li, and B. Wang, 2008b: Ocean forcing to changes in global
monsoon precipitation over the recent half-century. Journal of Climate, 21(15),
3833-3852.
Zhou, Z.J. and G.C. Zhang, 2003: Typical severe dust storms in northern China during
1954-2002. Chinese Science Bulletin, 48(21), 2366-2370.
Zimmermann, M., P. Mani, and H. Romang, 1997: Magnitude-frequency aspects of
alpine debris flows. Eclogae Geologicae Helvetiae, 90(3), 415-420.
Zin, W.Z.W., S. Jamaludin, S.M. Deni, and A.A. Jemain, 2009: Recent changes in
extreme rainfall events in Peninsular Malaysia: 1971-2005. Theoretical and
Applied Climatology, 99(3-4), 303-314.
Zolina, O., C. Simmer, A. Kapala, S. Bachner, S. Gulev, and H. Maechel, 2008:
Seasonally dependent changes of precipitation extremes over Germany since
1950 from a very dense observational network. Journal of Geophysical
Research – Atmospheres, 113, D06110.
Zolina, O., C. Simmer, K. Belyaev, A. Kapala, and S. Gulev, 2009: Improving estimates
of heavy and extreme precipitation using daily records from European rain
gauges. Journal of Hydrometeorology, 10(3), 701-716.
Chapter 3 Changes in Climate Extremes and their Impacts on the Natural Physical Environment
230
Zolina, O., C. Simmer, S.K. Gulev, and S. Kollet, 2010: Changing structure of European
precipitation: Longer wet periods leading to more abundant rainfalls.
Geophysical Research Letters, 37, L06704.
Zorita, E., T.F. Stocker, and H. von Storch, 2008: How unusual is the recent series of
warm years? Geophysical Research Letters, 35, L24706.
Zou, X., L.V. Alexander, D. Parker, and J. Caesar, 2006: Variations in severe storms over
China. Geophysical Research Letters, 33, L17701.
Zwiers, F.W., X. Zhang, and Y. Feng, 2011: Anthropogenic influence on long return
period daily temperature extremes at regional scales. Journal of Climate, 24(3),
881-892.
Chapter 3Changes in Climate Extremes and their Impacts on the Natural Physical Environment
231
Coordinating Lead Authors:
John Handmer (Australia), Yasushi Honda (Japan), Zbigniew W. Kundzewicz (Poland/Germany)
Lead Authors:
Nigel Arnell (UK), Gerardo Benito (Spain), Jerry Hatfield (USA), Ismail Fadl Mohamed (Sudan),
Pascal Peduzzi (Switzerland), Shaohong Wu (China), Boris Sherstyukov (Russia), Kiyoshi Takahashi
(Japan), Zheng Yan (China)
Review Editors:
Sebastian Vicuna (Chile), Avelino Suarez (Cuba)
Contributing Authors:
Amjad Abdulla (Maldives), Laurens M. Bouwer (Netherlands), John Campbell (New Zealand),
Masahiro Hashizume (Japan), Fred Hattermann (Germany), Robert Heilmayr (USA), Adriana Keating
(Australia), Monique Ladds (Australia), Katharine J. Mach (USA), Michael D. Mastrandrea (USA),
Reinhard Mechler (Germany), Carlos Nobre (Brazil), Apurva Sanghi (World Bank), James Screen
(Australia), Joel Smith (USA), Adonis Velegrakis (Greece), Walter Vergara (World Bank), Anya M. Waite
(Australia), Jason Westrich (USA), Joshua Whittaker (Australia), Yin Yunhe (China), Hiroya Yamano
(Japan)
This chapter should be cited as:
Handmer, J., Y. Honda, Z.W. Kundzewicz, N. Arnell, G. Benito, J. Hatfield, I.F. Mohamed, P. Peduzzi, S. Wu, B. Sherstyukov,
K. Takahashi, and Z. Yan, 2012: Changes in impacts of climate extremes: human systems and ecosystems. In: Managing the
Risks of Extreme Events and Disasters to Advance Climate Change Adaptation [Field, C.B., V. Barros, T.F. Stocker, D. Qin,
D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. A Special
Report of Working Groups I and II of the Intergovernmental Panel on Climate Change (IPCC). Cambridge University Press,
Cambridge, UK, and New York, NY, USA, pp. 231-290.
4
Changes in Impacts
of Climate Extremes: Human
Systems and Ecosystems
Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
232
Executive Summary .................................................................................................................................234
4.1. Introduction..............................................................................................................................237
4.2. Climatic Extremes in Natural and Socioeconomic Systems.....................................................237
4.2.1. How Do Climate Extremes Impact on Humans and Ecosystems?....................................................................................237
4.2.2. Complex Interactions among Climate Events, Exposure, and Vulnerability ....................................................................238
4.3. System- and Sector-Based Aspects of Vulnerability, Exposure, and Impacts...........................239
4.3.1. Introduction......................................................................................................................................................................239
4.3.2. Water................................................................................................................................................................................241
4.3.3. Ecosystems .......................................................................................................................................................................244
4.3.3.1. Heat Waves......................................................
..............................................................................................
...................................244
4.3.3.2. Drought ............................................................................................................................................................................................246
4.3.3.3. Floods ...............................................................................................................................................................................................246
4.3.3.4. Other Events.....................................................................................................................................................................................246
4.3.4. Food Systems and Food Security ...
....
..............................................................................................................................246
4.3.5. Human Settlements, Infrastructure, and Tourism.............................................................................................................247
4.3.5.1. Human Settlements .....................................................
............................................................................................
.........................247
4.3.5.2. Infrastructure ....................................................................................................................................................................................248
4.3.5.3. Tourism.............................................................................................................................................................................................250
4.3.6. Human Health, Well-Being, and Security...
....
...................................................................................................................251
4.4. Regionally Based Aspects of Vulnerability, Exposure, and Impacts.........................................252
4.4.1. Introduction......................................................................................................................................................................252
4.4.2 Africa ................................................................................................................................................................................253
4.4.2.1. Introduction ......................................................
..............................................................................................
..................................253
4.4.2.2. Droughts and Heat Waves ................................................................................................................................................................253
4.4.2.3. Extreme Rainfall Events and Floods..................................................................................................................................................253
4.4.2.4. Dust Storms ......................................................................................................................................................................................254
4.4.3. Asia...
....
............................................................................................................................................................................254
4.4.3.1. Tropical Cyclones (Typhoons or Hurricanes)......................................................................................................................................254
4.4.3.2.
Flooding............................................................................................................................................................................................254
4.4.3.3. Temperature Extremes......................................................................................................................................................................255
4.4.3.4. Droughts...........................................................................................................................................................................................255
4.4.3.5. Wildfires ...........................................................................................................................................................................................255
4.4.4. Central and South America ...
....
.......................................................................................................................................255
4.4.4.1. Extreme Rainfalls in South America.............................................................................................................................................
.....255
4.4.4.2. Wildfires ...........................................................................................................................................................................................255
4.4.4.3. Regional Costs..................................................................................................................................................................................256
4.4.5. Europe ...
....
.......................................................................................................................................................................256
4.4.5.1. Introduction ....................................................................................................................................................
..................................256
4.4.5.2. Heat Waves.......................................................................................................................................................................................256
4.4.5.3. Droughts and Wildfires .....................................................................................................................................................................256
4.4.5.4. Coastal Flooding...............................................................................................................................................................................256
4.4.5.5. Gale Winds .......................................................................................................................................................................................257
4.4.5.6. Flooding............................................................................................................................................................................................258
Chapter 4
Table of Contents
233
4.4.5.7. Landslides.........................................................................................................................................................................................258
4.4.5.8. Snow.................................................................................................................................................................................................258
4.4.6. North America ...
....
...........................................................................................................................................................258
4.4.6.1. Introduction ...................................................................................................................................................
...................................258
4.4.6.2. Heat Waves.......................................................................................................................................................................................258
4.4.6.3. Drought and Wildfire ........................................................................................................................................................................259
4.4.6.4. Inland Flooding.................................................................................................................................................................................259
4.4.6.5. Coastal Storms and Flooding............................................................................................................................................................259
4.4.7. Oceania...
....
......................................................................................................................................................................260
4.4.7.1. Introduction ...................................................................................................................................................
...................................260
4.4.7.2. Temperature Extremes......................................................................................................................................................................260
4.4.7.3. Droughts...........................................................................................................................................................................................260
4.4.7.4. Wildfire.............................................................................................................................................................................................261
4.4.7.5. Intense Precipitation and Floods ......................................................................................................................................................261
4.4.7.6. Storm Surges ....................................................................................................................................................................................261
4.4.8. Open Oceans ...
....
.............................................................................................................................................................261
4.4.9. Polar Regions ...................................................................................................................................................................261
4.4.9.1. Introduction .....................................................
..............................................................................................
...................................261
4.4.9.2. Warming Cryosphere ........................................................................................................................................................................262
4.4.9.3. Floods ...............................................................................................................................................................................................262
4.4.9.4. Coastal Erosion.................................................................................................................................................................................263
4.4.10. Small Island States...
....
....................................................................................................................................................263
4.5. Costs of Climate Extremes and Disasters ................................................................................264
4.5.1. Framing the Costs of Extremes and Disasters .................................................................................................................264
4.5.2. Extreme Events, Impacts, and Development....................................................................................................................265
4.5.3. Methodologies for Evaluating Impact and Adaptation Costs of Extreme Events and Disasters ....................................266
4.5.3.1. Methods and Tools for Costing Impacts.....................................................
......................................................................................
.266
4.5.3.2. Methods and Tools for Evaluating the Costs of Adaptation..............................................................................................................267
4.5.3.3. Attribution of Impacts to Climate Change: Observations and Limitations........................................................................................268
4.5.4. Assessment of Impact Costs ...
....
.....................................................................................................................................269
4.5.4.1. Estimates of Global and Regional Costs of Disasters .......................................................................................................................269
4.5.4.2.
Potential Trends in Key Extreme Impacts ..........................................................................................................................................271
4.5.5. Assessment of Adaptation Costs ...
....
..............................................................................................................................273
4.5.6. Uncertainty in Assessing the Economic Costs of Extremes and Disasters ......................................................................274
References ...............................................................................................................................................275
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
234
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
Extreme impacts can result from extreme weather and climate events, but can also occur without extreme
events. This chapter examines two broad categories of impacts on human and ecological systems, both of which are
influenced by changes in climate, vulnerability, and exposure: first, the chapter primarily focuses on impacts that
result from extreme weather and climate events, and second, it also considers extreme impacts that are triggered by
less-than-extreme weather or climate events. These two categories of impacts are examined across sectors, systems,
and regions. Extreme events can have positive as well as negative impacts on ecosystems and human activities.
Economic losses from weather- and climate-related disasters have increased, but with large spatial and
interannual variability (high confidence, based on high agreement, medium evidence). Global weather- and
climate-related disaster losses reported over the last few decades reflect mainly monetized direct damages to assets,
and are unequally distributed. Estimates of annual losses have ranged since 1980 from a few US$ billion to above
200 billion (in 2010 dollars), with the highest value for 2005 (the year of Hurricane Katrina). In the period 2000 to
2008, Asia experienced the highest number of weather- and climate-related disasters. The Americas suffered the most
economic loss, accounting for the highest proportion (54.6%) of total loss, followed by Asia (27.5%) and Europe
(15.9%). Africa accounted for only 0.6% of global economic losses. Loss estimates are lower bound estimates because
many impacts, such as loss of human lives, cultural heritage, and ecosystem services, are difficult to value and monetize,
and thus they are poorly reflected in estimates of losses. Impacts on the informal or undocumented economy, as well
as indirect effects, can be very important in some areas and sectors, but are generally not counted in reported esti-
mates of losses. [4.5.1, 4.5.3.3, 4.5.4.1]
Economic, including insured, disaster losses associated with weather, climate, and geophysical events are
higher in developed countries. Fatality rates and economic losses expressed as a proportion of gross
domestic product (GDP) are higher in developing countries (high confidence). During the period from 1970 to
2008, over 95% of deaths from natural disasters occurred in developing countries. Middle-income countries with
rapidly expanding asset bases have borne the largest burden. During the period from 2001 to 2006, losses amounted
to about 1% of GDP for middle-income countries, while this ratio has been about 0.3% of GDP for low-income coun-
tries and less than 0.1% of GDP for high-income countries, based on limited evidence. In small exposed countries, par-
ticularly small island developing states, losses expressed as a percentage of GDP have been particularly high, exceed-
ing 1% in many cases and 8% in the most extreme cases, averaged over both disaster and non-disaster years for the
period from 1970 to 2010. [4.5.2, 4.5.4.1]
Increasing exposure of people and economic assets has been the major cause of long-term increases in
economic losses from weather- and climate-related disasters (high confidence). Long-term trends in
economic disaster losses adjusted for wealth and population increases have not been attributed to climate
change, but a role for climate change has not been excluded (high agreement, medium evidence). These
conclusions are subject to a number of limitations in studies to date. Vulnerability is a key factor in disaster losses, yet
not well accounted. Other limitations are: (i) data availability, as most data are available for standard economic sec-
tors in developed countries; and (ii) type of hazards studied, as most studies focus on cyclones, where confidence in
observed trends and attribution of changes to human influence is low. The second conclusion is subject to additional
limitations: the processes used to adjust loss data over time, and record length. [4.5.3.3]
Settlement patterns, urbanization, and changes in socioeconomic conditions have all influenced observed
trends in exposure and vulnerability to climate extremes (high confidence). Settlements concentrate the
exposure of humans, their assets, and their activities. The most vulnerable populations include urban poor in informal
settlements, refugees, internally displaced people, and those living in marginal areas. Population growth is also a driv-
er of changing exposure and vulnerability. [4.2.1, 4.2.2, 4.3.5.1]
In much of the developed world, societies are aging and hence can be more vulnerable to climate
extremes, such as heat waves. For example, Europe currently has an aging population, with a higher population
Executive Summary
235
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
density and lower birth rate than any other continent. Nonetheless, exposure to climate extremes in Europe has
increased whereas vulnerability has decreased as a result of implementation of policy, regulations, risk prevention, and
risk management. Urban heat islands pose an additional risk to urban inhabitants, most affecting the elderly, ill, and
socially isolated. [4.3.5.1, 4.3.6, 4.4.5]
Transportation, infrastructure, water, and tourism are sectors sensitive to climate extremes. Transport
infrastructure is vulnerable to extremes in temperature, precipitation/river floods, and storm surges, which can lead to
damage in road, rail, airports, and ports, and electricity transmission infrastructure is also vulnerable to extreme storm
events. The tourism sector is sensitive to climate, given that climate is the principal driver of global seasonality in
tourism demand. [4.3.5.2, 4.3.5.3]
Agriculture is also an economic sector exposed and vulnerable to climate extremes. The economies of many
developing countries rely heavily on agriculture, dominated by small-scale and subsistence farming, and livelihoods in
this sector are especially exposed to climate extremes. Droughts in Africa, especially since the end of the 1960s, have
impacted agriculture, with substantial famine resulting. [4.3.4, 4.4.2]
Coastal settlements in both developed and developing countries are exposed and vulnerable to climate
extremes. For example, the major factor increasing the vulnerability and exposure of North America to hurricanes is
the growth in population and increase in property values, particularly along the Gulf and Atlantic coasts of the United
States. Small island states are particularly vulnerable to climate extremes, especially where urban centers and/or island
infrastructure predominate in coastal locations. Asia’s mega-deltas are also exposed to extreme events such as flood-
ing and have vulnerable populations in expanding urban areas. Mountain settlements are also exposed and vulnerable
to climate extremes. [4.3.5.1, 4.4.3, 4.4.6, 4.4.9, 4.4.10]
In many regions, the main drivers of future increases in economic losses due to some climate extremes will
be socioeconomic in nature (medium confidence, based on medium agreement, limited evidence). The
frequency and intensity of extreme weather and climate events are only one factor that affects risks, but few studies
have specifically quantified the effects of changes in population, exposure of people and assets, and vulnerability as
determinants of loss. However, these studies generally underline the important role of projected changes (increases) in
population and capital at risk. Additionally, some researchers argue that poorer developing countries and smaller
economies are more likely to suffer more from future disasters than developed countries, especially in relation to
extreme impacts. [4.5.2, 4.5.4.2]
Increases in exposure will result in higher direct economic losses from tropical cyclones. Losses will
depend on future changes in tropical cyclone frequency and intensity (high confidence). Overall losses due to
extratropical cyclones will also increase, with possible decreases or no change in some areas (medium confidence).
Although future flood losses in many locations will increase in the absence of additional protection measures (high
agreement, medium evidence), the size of the estimated change is highly variable, depending on location, climate
scenarios used, and methods used to assess impacts on river flow and flood occurrence. [4.5.4.2]
Extreme events will have greater impacts on sectors with closer links to climate, such as water, agriculture
and food security, forestry, health, and tourism. For example, while it is not currently possible to reliably project
specific changes at the catchment scale, there is high confidence that changes in climate have the potential to serious-
ly affect water management systems. However, climate change is in many instances only one of the drivers of future
changes in supply reliability, and is not necessarily the most important driver at the local scale. The impacts of changes
in flood characteristics are also highly dependent on how climate changes in the future, and as noted in Section 3.5.2,
there is low confidence in projected changes in flood magnitude or frequency. However, based on the available
literature, there is high confidence that, in some places, climate change has the potential to substantially affect flood
losses. Climate-related extremes are also expected to produce large impacts on infrastructure, although detailed
236
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
analysis of potential and projected damages are limited to a few countries, infrastructure types, and sectors.
[4.3.2, 4.3.5.2]
Estimates of adaptation costs to climate change exhibit a large range and relate to different assessment
periods. For 2030, the estimated global cost ranges from US$ 48 to 171 billion per year (in 2005 US$) with recent
estimates for developing countries broadly amounting to the average of this range with annual costs of up to US$ 100
billion. Confidence in individual estimates is low because the estimates are derived from only three relatively
independent studies. These studies have not explicitly separated costs of adapting to changes in climate extremes
from other climate change impacts, do not include costs incurred by all sectors, and are based on extrapolations of
bottom-up assessments and on top-down analysis lacking site-specificity. [4.5.3, 4.5.5, 4.5.6]
237
4.1. Introduction
Chapter 3 evaluates observed and projected changes in the frequency,
intensity, spatial extent, and duration of extreme weather and climate
events. This physical basis provides a picture of climate change and
extreme events. But it does not by itself indicate the impacts experienced
by humans or ecosystems. For example, for some sectors and groups of
people, severe impacts may result from relatively minor weather and
climate events. To understand impacts triggered by weather and climate
events, the exposure and vulnerability of humans and ecological systems
need to be examined. The emphasis of this chapter is on negative
impacts, in line with this report’s focus on managing the risks of extreme
events and disasters. Weather and climate events, however, can and often
do have positive impacts for some people and ecosystems.
In this chapter, two different types of impacts on human and ecological
systems are examined: (i) impacts of extreme weather and climate
events; and (ii) extreme impacts triggered by less-than-extreme weather
or climate events (in combination with non-climatic factors, such as
high exposure and/or vulnerability). Where data are available, impacts
are examined from sectoral and regional perspectives. Throughout this
chapter, the term ‘climate extremes’ will be used to refer in brief to
‘extreme weather and extreme climate events,’ as defined in the
Glossary and discussed more extensively in Section 3.1.2.
Activities undertaken as disaster risk reduction may also act as adaptation
to trends in climate extremes resulting from climate change, and they
may thereby act to reduce impacts. Strategies to reduce risk from one
type of climate extreme may act to increase or decrease the risk from
another. In writing this chapter, we have not considered these issues as
subsequent chapters are dedicated to adaptation. Here, impacts are
assessed without discussion of the specific possible adaptation or disaster
risk reduction strategies or policies evaluated in subsequent chapters.
Examination of trends in impacts and disasters highlights the difficulties
in attributing trends in weather- and climate-related disasters to climate
change. Trends in exposure and vulnerability and their relationship with
climate extremes are discussed. The chapter then examines system- and
sector-based aspects of vulnerability, exposure, and impacts, both
observed and projected. The same issues are examined regionally before
the chapter concludes with a section on the costs of weather- and
climate-related impacts, disasters, and adaptation.
4.2. Climatic Extremes in Natural
and Socioeconomic Systems
4.2.1. How Do Climate Extremes Impact
on Humans and Ecosystems?
The impacts of weather and climate extremes are largely determined by
exposure and vulnerability. This is occurring in a context where all three
components – exposure, vulnerability, and climate – are highly dynamic
and subject to continuous change. Some changes in exposure and
vulnerability can be considered as adaptive actions. For example, migration
away from high-hazard areas (see Chapter 1 and the Glossary for a
definition of the term ‘hazard’) reduces exposure and the chance of
disaster and is also an adaptation to increasing risk from climate
extremes (Adger et al., 2001; Dodman and Satterthwaite, 2008; Revi,
2008). Similar adaptive actions are reflected in changes in building
regulations and livelihoods, among many other examples.
Extreme impacts on humans and ecosystems can be conceptualized as
‘disasters’ or ‘emergencies.’ Many contemporary definitions emphasize
either that a disaster results when the impact is such that local capacity
to cope is exceeded or such that it severely disrupts normal activities.
There is a significant literature on the definitional issues, which include
factors of scale and irreversibility (Quarantelli, 1998; Handmer and
Dovers, 2007). Disasters result from impacts that require both exposure
to the climate event and a susceptibility to harm by what is exposed.
Impacts can include major destruction of assets and disruption to
economic sectors, loss of human lives, mental health effects, or loss and
impacts on plants, animals, and ecosystem services. The Glossary provides
the definition of disaster used in this chapter.
Exposure can be conceptualized as the presence of human and ecosystem
tangible and intangible assets and activities (including services) in areas
affected by climate extremes (see Sections 1.1.2 and 2.2 and the
Glossary for definitional discussion). Without exposure there is no
impact. Temporal and spatial scales are also important. Exposure can be
more or less permanent; for example, exposure can be increased by
people visiting an area or decreased by evacuation of people and
livestock after a warning. As human activity and settlements expand
into an exposed area, more people will be subject to and affected by
local climatic hazards. Population growth is predominantly in developing
countries (Peduzzi et al., 2011; UNISDR, 2011). Newly occupied areas
around or in urban areas were previously left vacant because they are
prone to the occurrence of climatic hazards (Handmer and Dovers,
2007; Satterthwaite et al., 2007; Wilbanks et al., 2007), for example with
movement of squatters to and development of informal settlements in
areas prone to flooding (Huq et al., 2007) and landslides (Anderson et
al., 2007). ‘Informal settlements’ are characterized by an absence of
involvement by government in planning, building, or infrastructure and
lack of secure tenure. In addition, there are affluent individuals pursuing
environmental amenity through coastal canal estates, riverside, and
bush locations, which are often at greater risk from floods and fires
(Handmer and Dovers, 2007).
Exposure is a necessary but not sufficient condition for impacts. For
exposed areas to be subjected to significant impacts from a weather or
climate event there must be vulnerability. Vulnerability is composed of
(i) susceptibility of what is exposed to harm (loss or damage) from the
event, and (ii) its capacity to recover (Cutter and Emrich, 2006; see
Sections 1.1.2 and 2.2 and the Glossary). Vulnerability is defined here as
in the Glossary as the propensity or predisposition to be adversely
affected. For example, those whose livelihoods are weather-dependent
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
238
or whose housing offers limited protection from weather events will be
particularly susceptible to harm (Dodman and Satterthwaite, 2008).
Others with limited capacity to recover include those with limited
personal resources for recovery or with no access to external resources
such as insurance or aid after an event, and those with limited personal
support networks (Handmer and Dovers, 2007). Knowledge, health, and
access to services of all kinds including emergency services and political
support help reduce both key aspects of vulnerability.
The concept of ‘resilience’ (developed in an ecological context by Holling,
1978; in a broad social sustainability context by Handmer and Dovers,
2007; and by Adger, 2000; Folke et al., 2002; see also the Glossary)
emphasizes the positive components of resistance or adaptability in the
face of an event and ability to cope and recover. This concept of
‘resilience’ is often seen as a positive way of expressing a similar
concept to that contained in the term ‘vulnerability’ (Handmer, 2003).
Refugees, internally displaced people, and those driven into marginal
areas as a result of violence can be dramatic examples of people
vulnerable to the negative effects of weather and climate events, cut off
from coping mechanisms and support networks (Handmer and Dovers,
2007). Reasons for the increase in vulnerability associated with warfare
include destruction or abandonment of infrastructure (e.g., transport,
communications, health, and education) and shelter, redirection of
resources from social to military purposes, collapse of trade and
commerce, abandonment of subsistence farmlands, lawlessness, and
disruption of social networks (Levy and Sidel, 2000; Collier et al., 2003).
The proliferation of weapons and minefields, the absence of basic
health and education, and collapse of livelihoods can ensure that the
effects of war on vulnerability to disasters are long lasting, although
some also benefit (Korf, 2004). These areas are also characterized by an
exodus of trained people and an absence of inward investment.
Many ecosystems are dependent on climate extremes for reproduction
(e.g., through fire and floods), disease control, and in many cases
general ecosystem health (e.g., fires or windstorms allowing new
growth to replace old). How such extreme events interact with other
trends and circumstances can be critical to the outcome. For example,
floods that would normally be essential to river gum reproduction may
carry disease and water weeds (Rogers and Ralph, 2010).
Climate extremes can cause substantial mortality of individual species
and contribute to determining which species exist in ecosystems
(Parmesan et al., 2000). For example, drought plays an important role in
forest dynamics, as a major influence on the mortality of trees (Villalba
and Veblen, 1997; Breshears and Allen, 2002; Breshears et al., 2005).
4.2.2. Complex Interactions among Climate Events,
Exposure, and Vulnerability
There exist complex interactions between different climatic and non-
climatic hazards, exposure, and vulnerability that have the potential of
triggering complex, scale-dependent impacts. Anthropogenic changes in
atmospheric systems are influencing changes in many climatic variables
and the corresponding physical impacts (see Chapter 3). However, the
impacts that climatic extremes have on humans and ecosystems
(including those altered by humans) depend also on several other non-
climatic factors (Adger, 2006). This section will explore these factors,
drawing on examples of flooding and drought.
Changes in socioeconomic status are a key component of exposure; in
particular, population growth is a major driver behind changing exposure
and vulnerability (Downton et al., 2005; Barredo, 2009). In many
regions, people have been encroaching into flood-prone areas where
effective flood protection is not assured, due to human pressure and lack
of more suitable and available land (McGranahan et al., 2007; Douglas
et al., 2008). Urbanization, often driven by rural poverty, drives such
migration (Douglas et al., 2008). In these areas, both population and
wealth are accumulating, thereby increasing the flood damage potential.
In many developed countries, population and wealth accumulation also
occur in hazard-prone areas for reasons of lifestyle and/or lower cost
(e.g., Radeloff et al., 2005). Here, a tension between climate change
adaptation and development is seen; living in these areas without
appropriate adaptation may be maladaptive from a climate change
perspective, but this may be a risk people are willing to take, or a risk
over which they have limited choice, considering their economic
circumstances (Wisner et al., 2004). Furthermore, there is often a
deficient risk perception present, stemming from an unjustified faith in
the level of safety provided by flood protection systems and dikes in
particular (Grothmann and Patt, 2005) (e.g., 2005 Hurricane Katrina in
New Orleans).
Economic development and land use change can also lead to changes
in natural systems. Land cover changes induce changes in rainfall-runoff
patterns, which can impact on flood intensity and frequency (e.g.,
Kundzewicz and Schellnhuber, 2004). Deforestation, urbanization,
reduction of wetlands, and river regulation (e.g., channel straightening,
shortening, and embankments) change the percentage of precipitation
becoming runoff by reducing the available water storage capacity (Few,
2003; Douglas et al., 2008). The proportion of impervious areas (e.g.,
roofs, yards, roads, pavements, parking lots, etc.) and the value of the
runoff coefficient are increased. As a result, water runs off faster to
rivers or the sea, and the flow hydrograph has a higher peak and a
shorter time-to-peak (Cheng and Wang, 2002; Few, 2003; Douglas et al.,
2008), reducing the time available for warnings and emergency action.
In mountainous areas, developments extending into hilly slopes are
potentially endangered by landslides and debris flows triggered by
intense rains. These changes have resulted in rain that is less extreme
leading to serious impacts (Crozier, 2010).
Similarly, the socioeconomic impacts of droughts may arise from the
interaction between natural conditions and human water use, which
can be conceptualized as a combination of supply and demand factors.
Human activities (such as over-cultivation, overgrazing, deforestation)
have exacerbated desertification of vulnerable areas in Africa and Asia,
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
239
where soil and bio-productive resources became permanently degraded
(Dregne, 1986). An extreme example of a human-made, pronounced
hydrological drought comes from the Aral Sea basin in Central Asia. Due
to excessive and non-sustainable water withdrawals from the
tributaries (Syr Darya and Amu Darya), their inflow into the Aral Sea has
shrunk in volume by some 75% (Micklin, 2007; Rodell et al., 2009)
resulting in severe economic and ecological impacts.
The changing impacts of climate extremes on sectors, such as water and
food, depend not only on changes in the characteristics of climate-
related variables relevant to a given sector, but also on sector-relevant
non-climatic stressors, management characteristics (including
organizational and institutional aspects), and adaptive capacity
(Kundzewicz, 2003).
There also may be increasing risks from possible interactions of hazards
(Cruz, 2005; see Sections 3.1.3 and 3.1.4 for discussion of interactions
and feedbacks). One hazard may influence other hazards or exacerbate
their effects, also with dependence on scale (Buzna et al., 2006). For
instance, temperature rise can lead to permafrost thaw, reduced slope
stability, and damage to buildings. Another example is that intense
precipitation can lead to flash flood, landslides, and infrastructure damage,
for example, collapse of bridges, roads, and buildings, and interruption
of power and water supplies. In the Philippines, two typhoons hitting
the south of Luzon Island in 2004 caused a significant flood disaster as
well as landslides on the island, leading to 900 fatalities (Pulhin et al.,
2010). It is worthwhile to note that cascading system failures (e.g.,
among infrastructure) can happen rapidly and over large areas due to
their interdependent nature.
4.3. System- and Sector-Based Aspects
of Vulnerability, Exposure, and Impacts
4.3.1. Introduction
In this subsection, studies evaluating impacts and risks of extreme
events are surveyed for major affected sectors and systems. Sectors and
systems considered here include water; ecosystems; food systems and
food security; human settlements, infrastructure, and tourism; and
human health, well-being, and security. Impacts of climate extremes are
determined by the climate extremes themselves as well as by exposure
and vulnerability. Climate extremes, exposure, and vulnerability are
characterized by uncertainty and continuous change, and shifts in any
of these components of risk will have implications for the impacts of
extreme events. Generally, there is limited literature on the potential
future impacts of extreme events; most literature analyzes current
impacts of extreme events. This focus may result in part from incomplete
knowledge and uncertainties regarding future changes in some extreme
events (see, for example, Section 3.2.3 and Tables 3-1 and 3-3) as well
as from uncertainties regarding future exposure and vulnerabilities.
Nonetheless, understanding current impacts can be important for
decisionmakers preparing for future risks. Analyses of both observed
and projected impacts due to extreme climate and weather events are
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
Box 4-1 | Evolution of Climate, Exposure, and Vulnerability – The Melbourne Fires, 7 February 2009
The fires in the Australian state of Victoria, on 7 February 2009, demonstrate the evolution of risk through the relationships between the
weather- and climate-related phenomena of a decade-long drought, record extreme heat, and record low humidity of 5% (Karoly, 2010;
Trewin and Vermont, 2010) interacting with rapidly increasing exposure. Together the climate phenomena created the conditions for
major uncontrollable wildfires (Victorian Bushfires Royal Commission, 2010).
The long antecedent drought, record heat, and a 35-day period with no rain immediately before the fires turned areas normally seen as
low to medium wildfire risk into very dry high-risk locations. A rapidly expanding urban-bush interface and valuable infrastructure (Berry,
2003; Burnley and Murphy, 2004; Costello, 2007, 2009) provided the values exposed and the potential for extreme impacts that was
realized with the loss of 173 lives and considerable tangible and intangible damage. There was a mixture of natural and human sources
of ignition, showing that human agency can trigger such fires and extreme impacts.
Many people were not well-prepared physically or psychologically for the fires, and this influenced the level of loss and damage they
incurred. Levels of physical and mental health also affected people’s vulnerability. Many individuals with ongoing medical conditions,
special needs because of their age, or other impairments struggled to cope with the extreme heat and were reliant on others to respond
safely (Handmer et al., 2010). However, capacity to recover in a general sense was high for humans and human activities through
insurance, government support, private donations, and nongovernmental organizations (NGOs) and was variable for the affected bush
with some species and ecosystems benefitting (Lindenmayer et al., 2010; Banks et al., 2011; see also Case Study 9.2.2).
Chapter 3 details projected changes in climate extremes for this region that could increase fire risk, in particular warm temperature
extremes, heat waves, and dryness (see Table 3-3 for summary).
240
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
Box 4-2 | Observed and Projected Trends in Human Exposure: Tropical Cyclones and Floods
The International loss databases with global coverage such as EM-DAT, NatCat, and Sigma (maintained by the Centre for the
Epidemiology of Disasters, Munich Re, and Swiss Re, respectively) present an increase in reported disasters through time. Although the
number of reported tropical cyclone disasters, for example, has increased from a yearly average of 21.7 during the 1970s to 63 during
the 2000s (see Table 4-1), one should not simply conclude that the number of disasters is increasing due to climate change. There are
four factors that may individually or together explain this increase: improved access to information, higher population exposure, higher
vulnerability, and higher frequency and/or intensity of hazards (Dao and Peduzzi, 2004; Peduzzi et al., 2009). Due to uncertainties in the
significance of the role of each of these four possible factors (especially regarding improved access to information), a vulnerability and
risk trend analysis cannot be performed based on reported losses (e.g., from EM-DAT or Munich Re). To better understand this trend,
international loss databases would have to be standardized.
Here for both tropical cyclones and floods, we overview a method for better understanding these factors through calculation of past
trends and future projections of human exposure at regional and global scales. Changes in population size strongly influence changes in
exposure to hazards. It is estimated that currently about 1.15
billion people live in tropical cyclone-prone areas. The physical
exposure (yearly average number of people exposed) to tropical
cyclones is estimated to have increased from approximately
73 million in 1970 to approximately 123 million in 2010
(Figure 4-1; Peduzzi et al., 2011). The number of times that
countries are hit by tropical cyclones per year is relatively
steady (between 140 and 155 countries per year
1
on average;
see Table 4-1 (UNISDR, 2011).
In most oceans, the frequency of tropical cyclones is likely to
decrease or remain unchanged while mean tropical cyclone
Number of tropical cyclones
as identified in best track
data (average per year)
Number of countries hit by
tropical cyclones as detected
by satellite (average per year)
Number of disasters triggered
by tropical cyclones as
reported by EM-DAT (average
per year)
Reported disasters as a
percentage of number of
countries hit by tropical
cyclones
88.4 88.2 87.2 86.5
142.1 144.0 155.0 146.3
21.7 37.5 50.6 63.0
15% 26% 33% 43%
1970 - 79 1980 - 89 1990 - 99 2000 - 09
Table 4-1 | Trend in tropical cyclone disasters reported versus tropical cyclones detected
by satellite during the last four decades. The reported disasters as a percentage of the
number of countries hit by tropical cyclones increased three-fold. Note that ‘best track
data’ generally comprise four-times daily estimates of tropical cyclone intensity and
position; these data are based on post-season reprocessing of data that were collected
operationally during each storm’s lifetime. Source: UNISDR, 2011.
Figure 4-1 | Average physical exposure to tropical cyclones assuming constant hazard (in thousands of people per year). Data from Peduzzi et al., 2011.
in 2030
in 1970
Circles are proportional
to the number of persons affected
4,870
30
100
50
100
ISLANDS
Indian Ocean, Pacific Ocean,
Caribbean and other Islands
4,
1
ISLAND
S
68,000
125,950
NORTH
AMERICA
CENTRAL AND
SOUTH AMERICA
AFRICA
ASIA
AUSTRALIA
NEW ZEALAND
2,610
500
2,280
1,910
3,490
Average Physical Exposure
to Tropical Cyclones
Assuming Constant Hazard
in thousands of people per year
____________
1
This is the number of intersections between countries and tropical cyclones.
One cyclone can affect several countries, but also many tropical cyclones
occur only over the oceans.
Continued next page
241
thus assessed. Sections 4.3.2 to 4.3.6, building on an understanding of
exposure and vulnerability, evaluate knowledge of current and future
risks of extreme events by sectors and systems.
4.3.2. Water
Past and future changes in exposure and vulnerability to climate
extremes in the water sector are driven by both changes in the volume,
timing, and quality of available water and changes in the property, lives,
and systems that use the water resource or that are exposed to water-
related hazards (Aggarwal and Singh, 2010). With a constant resource
or physical hazard, there are two opposing drivers of change in exposure
and vulnerability. On the one hand, vulnerability increases as more
demands are placed on the resource (due to increased water consumption,
for example, or increased discharge of polluting effluent) or exposure
increases as more property, assets, and lives encounter flooding. On the
other hand, vulnerability is reduced as measures are implemented to
improve the management of resources and hazards and to enhance the
ability to recover from extreme events. For example, enhancing water
supplies, improving effluent treatment, and employing flood management
measures (including the provision of insurance or disaster relief) would
all lead to reductions in vulnerability in the water sector. Such measures
have been widely implemented, and the runoff regime of many rivers
has been considerably altered (Vörösmarty, 2002). The change in exposure
and vulnerability in any place is a function of the relationship between
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
maximum wind speed is likely to increase (see Section 3.4.4). Figure 4-1 provides the modeled change in human exposure at constant
hazard (without forecast of the influence of climate change on the hazard). It shows that the average number of people exposed to
tropical cyclones per year globally would increase by 11.6% from 2010 to 2030 from population growth only. In relative terms, Africa
has the largest percentage increase in physical exposure to tropical cyclones. In absolute terms, Asia has more than 90% of the global
population exposed to tropical cyclones.
In terms of exposure to flooding, about 800 million people are currently living in flood-prone areas, and about 70 million people currently
living in flood-prone areas are, on average, exposed to floods each year (UNISDR, 2011). Given the lack of complete datasets on past
flood events, and the uncertainty associated with projected trends in future flood frequencies and magnitudes (see Section 3.5.2), it is
difficult to estimate future flood hazards. However, using population increase in the flood-prone area, it is possible to look at trends in
the number of people exposed per year on average at constant hazard (UNISDR, 2011). Figure 4-2 shows that population growth will
continue to increase exposure to floods. Due to model constraints, areas north of 60°N and south of 60°S, as well as catchments smaller
than 1,000 km
2
(typically small islands) are not modeled. The data provided in Figure 4-2 correspond to river flooding.
A number of factors underlie increases in impacts from floods and cyclones. However, trends in the population exposed to these hazards
are an important factor. Population projections in tropical cyclone areas and flood-prone areas imply that impacts will almost certainly
continue to increase based on this factor alone.
29,780
77,640
NORTH
AMERICA
CENTRAL AND
SOUTH AMERICA
AFRICA
ASIA
AUSTRALIA
NEW ZEALAND
640
850
3,640
Average Physical
Exposure to Floods Assuming
Constant Hazard
in thousands of people per year
CARIBBEAN*
EUROPE
in thousand
ds o
in 2030
in 1970
Circles are proportional
to the number of persons affected
1,190
550
1,320
30
60
*Only catchments bigger than 1,000 km
2
were included in this analysis. Therefore, only the largest islands in the Caribbean are covered.
70
180
1,650
1,870
Figure 4-2 | Average physical exposure to floods assuming constant hazard (in thousands of people per year). Data from Peduzzi et al., 2011.
242
these two opposing drivers, which also interact. Flood or water
management measures may reduce vulnerability in the short term, but
increased security may generate more development and ultimately lead
to increased exposure and vulnerability.
Extreme events considered in this section can threaten the ability of the
water supply ‘system’ (from highly managed systems with multiple
sources to a single rural well) to supply water to users. This may be
because a surplus of water affects the operation of systems, but more
typically results from a shortage of water relative to demands – a
drought. Water supply shortages may be triggered by a shortage of river
flows and groundwater, deterioration in water quality, an increase in
demand, or an increase in vulnerability to water shortage. There is
medium confidence that since the 1950s some regions of the world
have experienced more intense and longer droughts, in particular in
southern Europe and West Africa (see Section 3.5.1), but it is not possible
to attribute trends in the human impact of drought directly or just to
these climatic trends because of the simultaneous change in the other
drivers of drought impact.
There is medium confidence that the projected duration and intensity of
hydrological drought will increase in some regions with climate change
(Section 3.5.1), but other factors leading to a reduction in river flows or
groundwater recharge are changes in agricultural land cover and
upstream interventions. A deterioration in water quality may be driven
by climate change (as shown for example by Delpla et al., 2009;
Whitehead et al., 2009; Park et al., 2010), change in land cover, or
upstream human interventions. An increase in demand may be driven by
demographic, economic, technological, or cultural drivers as well as by
climate change (see Section 2.5). An increase in vulnerability to water
shortage may be caused, for example, by increasing reliance on specific
sources or volumes of supply, or changes in the availability of alternatives.
Indicators of hydrological and water resources drought impact include
lost production (of irrigated crops, industrial products, and energy), the
cost of alternative or replacement water sources, and altered human
well-being, alongside consequences for freshwater ecosystems (impacts
of meteorological and agricultural droughts on production of rain-fed
crops are summarized in Section 4.3.4).
Few studies have so far been published on the effect of climate change
on the impacts of drought in water resources terms at the local catchment
scale. Virtually all of these have looked at water system supply reliability
during a drought, or the change in the yield expected with a given
reliability, rather than indicators such as lost production, cost, or well-
being. Changes in the reliability of a given yield, or yield with a given
reliability, of course vary with local hydrological and water management
circumstances, the details of the climate scenarios used, and other
drivers of drought risk. Some studies show large potential reductions in
supply reliability due to climate change that challenge existing water
management systems (e.g., Fowler et al., 2003; Kim et al., 2009; Takara
et al., 2009; Vanham et al., 2009); some show relatively small reductions
that can be managed – albeit at increased cost – by existing systems
(e.g., Fowler et al., 2007), and some show that under some scenarios the
reliability of supply increases (e.g., Kim and Kaluarachchi, 2009; Li et al.,
2010). While it is not currently possible to reliably project specific
changes at the catchment scale, there is high confidence that changes
in climate have the potential to seriously affect water management
systems. However, climate change is in many instances only one of the
drivers of future changes in supply reliability, and is not necessarily the
most important driver at the local scale. MacDonald et al. (2009), for
example, demonstrate that the future reliability of small-scale rural water
sources in Africa is largely determined by local demands, biological
aspects of water quality, or access constraints, rather than changes in
regional recharge, because domestic supply requires only 3-10 mm of
recharge per year. However, they noted that up to 90 million people in
low rainfall areas (200-500 mm) would be at risk if rainfall reduces to
the point at which groundwater resources become nonrenewable.
There have been several continental- or global-scale assessments of
potential change in hydrometeorological drought indicators (see
Section 3.5.1), but relatively few on measures of water resources
drought or drought impacts. This is because these impacts are very
dependent on context. One published large-scale assessment (Lehner et
al., 2006) used a generalized drought deficit volume indicator, calculated
by comparing simulated river flows with estimated withdrawals for
municipal, industrial, and agricultural uses. The indicator was calculated
across Europe, using climate change projections from two climate
models and assuming changes in withdrawals over time. They showed
substantial changes in the return period of the drought deficit volume,
comparing the 100-year return period for the 1961-1990 period with
projections for the 2070s (Figure 4-3). Across large parts of Europe, the
1961-1990 100-year drought deficit volume is projected to have a
return period of less than 10 years by the 2070s. Lehner et al. (2006)
also demonstrated that this projected pattern of change was generally
driven by changes in climate, rather than the projected changes in
withdrawals of water (Figure 4-3). In southern and western Europe,
changing withdrawals alone only are projected to increase deficit
volumes by less than 5%, whereas the combined effect of changing
withdrawals and climate change is projected to increase deficit volumes
by at least 10%, and frequently by more than 25%. In eastern Europe,
increasing withdrawals are projected to increase drought deficit
volumes by over 5%, and more than 10% across large areas, but this is
offset under both climate scenarios by increasing runoff.
Climate change has the potential to change river flood characteristics
through changing the volume and timing of precipitation, by altering the
proportions of precipitation falling as snow and rain, and to a lesser
extent, by changing evaporation and hence accumulated soil moisture
deficits. However, there is considerable uncertainty in the magnitude,
frequency, and direction of change in flood characteristics (Section
3.5.2). Changes in catchment surface characteristics (such as land
cover), floodplain storage, and the river network can also lead to
changes in the physical characteristics of river floods (e.g., along the
Rhine: Bronstert et al., 2007). The impacts of extreme flood events
include direct effects on livelihoods, property, health, production, and
communication, together with indirect effects of these consequences
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
243
through the wider economy. There have, however, been very few studies
that have looked explicitly at the human impacts of changes in flood
frequency, rather than at changes in flood frequencies and magnitudes.
One study has so far looked at changes in the area inundated by floods
with defined return periods (Veijalainen et al., 2010), showing that the
relationship between change in flood magnitude and flood extent
depended strongly on local topographic conditions.
An early study in the United States (Choi and Fisher, 2003) constructed
regression relationships between annual flood loss and socioeconomic
and climate drivers, concluding that a 1% increase in average annual
precipitation would, other things being equal, lead to an increase in
annual national flood loss of around 6.5%. However, the conclusions
are highly dependent on the regression methodology used, and the
spatial scale of analysis. More sophisticated analyses combine estimates
of current and future damage potential (as represented by a damage-
magnitude relationship) with estimates of current and future flood
frequency curves to estimate event damages and average annual damages
(sometimes termed expected annual damage). For example, Mokrech et
al. (2008) estimated damages caused by the current 10- and 75-year
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
2070s
ECHAM4
2070s water use
and HadCM3
climate
2070s water use
without climate
change
2070s
HadCM3
Future return period [years] of
droughts with an intensity of today’s
100-year events:
Future change [%] in intensity
(deficit volume) of 100-year
droughts:
more frequent
no change
no change
less frequent
decrease increase
100<704010>
05
10 25 ><
Figure 4-3 | Change in indicators of water resources drought across Europe by the 2070s. (top): projected changes in the return period of the
1961–1990 100-year drought deficit volume for the 2070s, with change in river flows and withdrawals for two climate models, ECHAM4 and HadCM3;
(bottom): projected changes in the intensity (deficit volume) of 100-year droughts with changing withdrawals for the 2070s, with climate change (left,
with HadCM3 climate projections) and without climate change (right). Source: Lehner et al., 2006.
244
events in two regions of England, combining fluvial and coastal flooding.
The two main conclusions from their work were as follows. First, the
percentage change in cost was greater for the rarer event than the more
frequent event. Second, the absolute value of impacts, and therefore the
percentage change from current impacts, was found to be highly
dependent on the assumed socioeconomic change. In one region, event
damage varied, in monetary terms, between four and five times across
socioeconomic scenarios. An even wider range in estimated average
annual damage was found in the UK Foresight Future Flooding and
Coastal Defence project (Evans et al., 2004; Hall et al., 2005), which
calculated average annual damage in 2080 of £1.5 billion, £5 billion, and
£21 billion under similar climate scenarios but different socioeconomic
futures (current average annual damage was estimated at £1 billion).
The Foresight project represented the effect of climate change on flood
frequency by altering the shape of the flood frequency curve using
precipitation outputs from climate models and rainfall-runoff models for
a sample of UK catchments. The EU-funded Projection of Economic
impacts of climate change in Sectors of the European Union based on
boTtom-up Analysis (PESETA) project (Ciscar, 2009; Feyen et al., 2009)
used a hydrological model to simulate river flows, flooded areas, and flood
frequency curves from climate scenarios derived from regional climate
models, but – in contrast to the UK Foresight project – assumed no
change in economic development in flood-prone areas. Figure 4-4
summarizes estimated changes in the average annual number of people
flooded and average annual damage, by European region (Ciscar, 2009).
There are strong regional variations in impact, with particularly large
projected increases in both number of people flooded and economic
damage (over 200%) in central and Eastern Europe, while in parts of
North Eastern Europe, average annual flood damages decrease.
At the global scale, two studies have estimated the numbers of people
affected by increases (or decreases) in flood hazard. Kleinen and Petschel-
Held (2007) calculated the percentage of population living in river basins
where the return period of the current 50-year event becomes shorter,
for three climate models and a range of increases in global mean
temperature. With an increase in global mean temperature of 2°C (above
late 20th-century temperatures), between (approximately) 5 and 27%
of the world’s population would live in river basins where the current
50-year return period flood occurs at least twice as frequently. Hirabayashi
and Kanae (2009) used a different metric, counting each year the number
of people living in grid cells where the flood peak exceeded the (current)
100-year magnitude, using runoff as simulated by a high-resolution
climate model fed through a river routing model. Beyond 2060, they
found that at least 300 million people could be affected by substantial
flooding even in years with relatively low flooding, with of the order of
twice as many being flooded in flood-rich years (note that they used
only one climate scenario with one climate model). This compares with
a current range (using the same index) of between 20 and 300 million
people. The largest part of the projected increase is due to increases in
the occurrence of floods, rather than increases in population.
The impacts of changes in flood characteristics are highly dependent on
how climate changes in the future, and as noted in Section 3.5.2, there
is low confidence in projections of changes in flood magnitude or
frequency. However, based on the available literature, there is high
confidence that, in some places, climate change has the potential to
substantially affect flood losses.
4.3.3. Ecosystems
Available information shows that high temperature extremes (i.e., heat
wave), drought, and floods substantially affect ecosystems. Increasing
gaps and overall contraction of the distribution range for species
habitat could result from increases in the frequency of large-scale
disturbances due to extreme weather and climate events (Opdam and
Wascher, 2004). Fischlin et al. (2007), from assessment of 19 studies,
found that 20 to 30% of studied plant and animal species may be at an
increased risk of extinction if warming exceeds 2 to 3°C above the
preindustrial level. Changes due to climate extremes could also entail
shifts of ecosystems to less-desired states (Scheffer et al., 2001; Chapin
et al., 2004; Folke et al., 2004) through, for example, the exceedance of
critical temperature thresholds, with potential loss of ecosystem services
dependent on the previous state (Reid et al., 2005; see also Fischlin et
al., 2007).
4.3.3.1. Heat Waves
Heat waves can directly impact ecosystems by, for example, constraining
carbon and nitrogen cycling and reducing water availability, with the
result of potentially decreasing production or even causing species
mortality.
Warming can decrease net ecosystem carbon dioxide (CO
2
) exchange
by inducing drought that suppresses net primary productivity. More
frequent warm years may lead to a sustained decrease in CO
2
uptake by
terrestrial ecosystems (Arnone et al., 2008). Extreme temperature
conditions can shift forest ecosystems from being a net carbon sink to
being a net carbon source. For example, tall-grass prairie net ecosystem
CO
2
exchange levels decreased in both an extreme warming year
(2003) and the following year in grassland monoliths from central
Oklahoma, United States (Arnone et al., 2008). A 30% reduction in gross
primary productivity together with decreased ecosystem respiration
over Europe during the heat wave in 2003 resulted in a strong net
source of CO
2
(0.5 Pg C yr
-1
) to the atmosphere and reversed the effect
of four years of net ecosystem carbon sequestration. Such a reduction
in Europe’s primary productivity is unprecedented during the last century
(Ciais et al., 2005).
Impacts are determined not only by the magnitude of warming but also
by organisms’ physiological sensitivity to that warming and by their
ability to compensate behaviorally and physiologically. For example,
warming may affect tropical forest lizards’ physiological performance in
summer, as well as their ability to compete with warm-adapted, open-
habitat competitors (Huey et al., 2009). Projected increases in maximum
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
245
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
Figure 4-4 | Impact of climate change by 2071–2100 on flood risk in Europe. Note that the numbers assume no change in population or development in flood-prone
areas. As illustrated in the legend on the right of each panel, projections are given for two Special Report on Emissions Scenarios (SRES) scenarios (A2 and B2) and for
two global climate models (HadAM3h and ECHAM4). Projected mean temperature increase in the European region for the period 2071–2100 compared with
1961–1990 is indicated for each scenario and model combination. (top): For each region, baseline simulated population affected over 1961–1990 (thousands per year)
and expected population affected (thousands per year) for 2071–2100 for each scenario and model combination. (bottom): For each region, baseline simulated
economic damage over 1961–1990 (million € per year, 2006 prices) and expected economic damage (million € per year, 2006 prices) for 2071–2100 for each scenario
and model combination. Data from Ciscar, 2009.
BRITISH ISLES
NORTH
EUROPE
NO
ORTH
O
BRITIS
ITISH ISLE
S
SOUTH
EUROPE
Number of people affected
per year (in thousands)
Baseline
Population
affected
during
1961-1990
period
0
50
100
150
200
2.5°C (B2)
3.9°C (A2)
4.1°C (B2)
5.4°C (A2)
Model
HadAM3h
Period
2071
2100
Model
ECHAM4
Period
2071
2100
BRITISH ISLES
NORTH
EUROPE
NORTHERN
CENTRAL EUROPE
NORTHERN
CENTRAL EUROPE
SOUTHERN
CENTRAL EUROPE
SOUTHERN
CENTRAL EUROPE
SOUTH
EUROPE
Expected economic damage
in million Euros per year (2006 prices)
Model
HadAM3h
Period
2071
2100
Model
ECHAM4
Period
2071
2100
Baseline
Economic
damage
during
1961-1990
period
2.5°C (B2)
3.9°C (A2)
4.1°C (B2)
5.4°C (A2)
0
1 000
2 000
3 000
4 000
5 000
6 000
246
air temperatures may increase evaporative water requirements in birds,
thus influencing survival during extreme heat events (McKechnie and
Wolf, 2010). Heat waves could also cause increased likelihood of
catastrophic avian mortality events (McKechnie and Wolf, 2010).
4.3.3.2. Drought
A rapid, drought-induced die-off of overstory woody plants at a
subcontinental scale was triggered by the 2000–2003 drought in
southwestern North America. Following 15 months of diminished soil
water content, more than 90% of the dominant tree species, Pinus edulis,
died. Limited observations indicate that die-off was more extensive than
during the previous drought of the 1950s, also affecting wetter sites
within the tree species’ distribution (Breshears et al., 2005). Regional-
scale piñon pine mortality was observed following an extended drought
(2000–2004) in northern New Mexico (Rich et al., 2008). Dominant
plant species from diverse habitat types (i.e., riparian, chaparral, and
low-to high-elevation forests) exhibited significant mortality during a
drought in the southwestern United States; average mortality among
dominant species was 3.3 to 41.4% (Gitlin et al., 2006).
Evergreen coniferous species mortality caused by the coupling of drought
and higher temperatures from winter to spring has been observed in the
Republic of Korea (Lim et al., 2010). In 1998, 2002, 2007, and 2009, years
of high winter-spring temperatures and lower precipitation, P. densiflora
and P. koraiensis were affected by droughts, with many dying in the
crown layer, while deciduous species survived. Similarly, Abies koreana,
an endemic species in Korea, at high elevation has declined following a
rise in winter temperatures since the late 1990s (Lim et al., 2010). Beech
crown condition was observed to decline following severe drought in
1976 (Power, 1994), 1989 (Innes, 1992), and 1990 (Stribley and
Ashmore, 2002). Similarly, the percentage of moderately or severely
damaged trees displayed an upward trend after the 1989 drought in
Central Italy, especially for P. pinea and Fagus sylvatica (Bussotti et al.,
1995). As final examples, defoliation and mortality in Scots pine observed
in each year during 1996 to 2002 was related to the precipitation deficit
and hot conditions of the previous year in the largest inner-alpine
valley of Switzerland (Valais) (Rebetez and Dobbertin, 2004), and both
gross primary production and total ecosystem respiration decreased in
2003 in many regions of Europe (Granier et al., 2007).
In a shallow temperate southern European estuary, the Mondego Estuary
in Portugal, the severe drought in 2004–2005 was responsible for
spatial shifts in the estuary’s zooplankton community, with an increase
in abundance and diversity during the period of low freshwater flow
(Marques et al., 2007).
4.3.3.3. Floods
Floods also impact ecosystems. Floods can cause population- and
community-level changes superimposed on a background of more
gradual trends (Thibault and Brown, 2008). As an example, an extreme
flood event affected a desert rodent community (that had been
monitored for 30 years) by inducing a large mortality rate, eliminating
the advantage of previously dominant species, resetting long-term
population and community trends, altering competitive and meta-
population dynamics, and reorganizing the community (Thibault and
Brown, 2008).
4.3.3.4. Other Events
Other events, such as hurricanes and storms, can also impact ecosystems.
Hurricanes can cause widespread mortality of wild birds, and their
aftermath may cause declines due to the birds’ loss of resources
required for foraging and breeding (Wiley and Wunderle, 1994). Winter
storms can also impact forest ecosystems, particularly in pre-alpine and
alpine areas (Faccio, 2003; Schelhaas et al., 2003; Fuhrer et al., 2006). In
addition, saltmarshes, mangroves, and coral reefs can be vulnerable to
climate extremes (e.g., Bertness and Ewanchuk, 2002; Hughes et al.,
2003; Fischlin et al., 2007).
4.3.4. Food Systems and Food Security
Food systems and food security can be affected by extreme events that
impair food production and food storage and delivery systems (food
logistics). Impacts transmitted through an increase in the price of food
can be especially challenging for the urban poor in developing countries
(FAO, 2008). Global food price increases are borne disproportionally by
low-income countries, where people spend more of their income on
food (OECD-FAO, 2008).
When agricultural production is not consumed where it is produced, it
must be transported and often processed and stored. This process
involves complex interdependent supply chains exposed to multiple
hazards. At every step of the process, transport and associated
infrastructure such as roads, railways, bridges, warehouses, airports,
ports, and tunnels can be at risk of direct damage from climate
events, making the processing and delivery chain as a whole at risk of
disruption resulting from damage or blockages at any point in the
chain.
The economies of many developing countries rely heavily on agriculture,
dominated by small-scale and subsistence farming. People’s livelihoods
in this sector are especially exposed to weather extremes (Easterling
and Apps, 2005; Easterling et al., 2007). Subsistence farmers can be
severely impacted by climate and weather events. For example, the
majority of households produce maize in many African countries, but
only a modest proportion sells it – the great majority eat all they
produce. In Kenya for example, nearly all households grow maize, but
only 36% sell it, with 20% accounting for the majority of sales (FAO,
2009). Both such famers and their governments have limited capacity
for recovery (Easterling and Apps, 2005).
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
247
Evidence that the current warming trends around the world have
already begun to impact agriculture is reported by Lobell et al. (2011). They
show that crop yields have already declined due to warmer conditions
compared to the expected yields without warming. Both Schlenker and
Roberts (2009) and Muller et al. (2011), after their evaluation of projected
temperature effects on crops in the United States and Africa, concluded
that climate change would have negative impacts on crop yields. These
effects were based on temperature trends and an expected increase in
the probability of extremes during the growing season; however, there
is also the potential occurrence of extreme events after the crop is
grown, which could affect harvest and grain quality. Fallon and Betts
(2010) stated that increasing flooding and drought risks could affect
agricultural production and require the adoption of robust management
practices to offset these negative impacts. Their analysis for Europe
showed a probable increase in crop productivity in northern regions
but a decrease in the southern regions, leading to a greater disparity in
production.
In a recent evaluation of high temperature as a component of climate
trends, Battisti and Naylor (2009) concluded that future growing season
temperatures are very likely to exceed the most extreme temperatures
observed from 1900 to 2006, for both tropical and subtropical regions,
with substantial potential implications for food systems around the world.
The effects of temperature extremes on a number of different crop
species have been summarized in Hatfield et al. (2011). Many crops are
especially sensitive to extreme temperatures that occur just prior to or
during the critical pollination phase of crop growth (Wheeler et al., 2000;
Hatfield et al., 2008, 2011). Crop sensitivity and ability to compensate
during later improved weather will depend on the length of time for
anthesis in each crop.
Extreme temperatures can negatively impact grain yield (Kim et al.,
1996; Prasad et al., 2006). For example, Tian et al. (2010) observed in rice
that high temperatures (>35°C) coupled with high humidity and low
wind speed caused panicle temperatures to be as much as 4°C higher
than air temperature, inducing floret sterility. Impacts of temperature
extremes may not be limited to daytime events. Mohammed and Tarpley
(2009) observed that rice yields were reduced by 90% when night
temperatures were increased from 27 to 32°C. An additional impact of
extremes has been found in the quality of the grain. Kettlewell et al.
(1999) found that wheat quality in the United Kingdom was related to
the North Atlantic Oscillation and probably caused by variation in
rainfall during the grain-filling period. In a more recent study, Hurkman
et al. (2009) observed that high-temperature events during grain-filling
of wheat altered the protein content of the grain, and these responses
were dependent upon whether the exposure was imposed early or
midway through the grain-filling period. Skylas et al. (2002) observed that
high temperature during grain-filling was one of the most significant
factors affecting both yield and flour quality in wheat.
Drought causes yield variation, and an example from Europe demonstrates
that historical yield records show that drought has been a primary cause
of interannual yield variation (Hlavinka et al., 2009; Hatfield, 2010).
Water supply for agricultural production will be critical to sustain
production and even more important to provide the increase in food
production required to sustain the world’s growing population. With
glaciers retreating due to global warming and El Niño episodes, the
Andean region faces increasing threats to its water supply (Mark and
Seltzer, 2003; Cadier et al., 2007). With precipitation limited to only a few
months of the year, melt from glaciers is the only significant source of
water during the dry season (Mark and Seltzer, 2003). Glacier recession
reduces the buffering role of glaciers, hence inducing more floods during
the rainy season and more water shortages during the dry season.
Cadier et al. (2007) found that warm anomalies of the El Niño-Southern
Oscillation (ENSO) corresponded to an increase in melting four months
later.
Food security is linked to our ability to adapt agricultural systems to
extreme events using our understanding of the complex system of
production, logistics, utilization of the produce, and the socioeconomic
structure of the community. The spatial variability and context sensitivity
of each of these factors points to the value of downscaled scenarios of
climate change and extreme events.
4.3.5. Human Settlements, Infrastructure, and Tourism
4.3.5.1. Human Settlements
Settlements concentrate the exposure of humans, their assets, and their
activities. In the case of very large cities, these concentrations can
represent a significant proportion of national wealth and may result in
additional forms of vulnerability (Mitchell, 1998). Flooding, landslides,
storms, heat waves, and wildfires have produced historically important
damages in human settlements, and the characteristics of these events
and their underlying climate drivers are projected to change (see
Chapter 3; Kovats and Akhtar, 2008; Satterthwaite, 2008). The
concentration of economic assets and people creates the possibility
of large impacts, but also the capacity for recovery (Cutter et al.,
2008). Coastal settlements are especially at risk with sea level rise
and changes in coastal storm activity (see Sections 3.4.4 and 3.4.5
and Case Study 9.2.8).
At very high risk of impacts are the urban poor in informal settlements
(Satterthwaite, 2008; Douglas, 2009). Worldwide, about one billion
people live in informal settlements, and informal settlements are
growing faster than formal settlements (UN-HABITAT, 2008; UNISDR,
2011). Informal settlements are also found in developed countries; for
example, there are about 50 million people in such areas in Europe
(UNECE, 2009). Occupants of informal settlements are typically more
exposed to climate events with no or limited hazard-reducing
infrastructure. The vulnerability is high due to very low-quality housing
and limited capacity to cope due to a lack of assets, insurance, and
marginal livelihoods, with less state support and limited legal protection
(Dodman and Satterthwaite, 2008).
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
248
The number and size of coastal settlements and their associated
infrastructure have increased significantly over recent decades
(McGranahan et al., 2007; Hanson et al., 2011; see also Case Study
9.2.8). In many cases these settlements have affected the ability of
natural coastal systems to respond effectively to extreme climate events
by, for example, removing the protection provided by sand dunes and
mangroves. Small island states, particularly small island developing
states (see Case Study 9.2.9), may face substantial impacts from climate
change-related extremes.
Urbanization exacerbates the negative effects of flooding through
greatly increased runoff concentration, peak, and volume, the increased
occupation of flood plains, and often inadequate drainage planning
(McGranahan et al., 2007; Douglas et al., 2008). These urbanization
issues are universal but often at their worst in informal settlements,
which are generally the most exposed to flooding and usually do not
have the capacity to deal with the issues (Hardoy et al., 2001). Flooding
regularly disrupts cities, and urban food production can be severely
affected by flooding, undermining local food security in poor communities
(Douglas, 2009; Aggarwal and Singh, 2010). A further concern for low-
and middle-income cities as a result of flooding, particularly in developing
countries, is human waste, as most of these cities are not served by
proper water services such as sewers, drains, or solid waste collection
services (Hardoy et al., 2001).
Slope failure can affect settlements in tropical mountainous areas,
particularly in deforested areas (e.g., Vanacker et al., 2003)and hilly
areas (Loveridge et al., 2010), and especially following heavy prolonged
rain (e.g., see Case Study 9.2.5). Informal settlements are often exposed
to potential slope failure as they are often located on unstable land with
no engineering or drainage works (Alexander, 2005; Anderson et al.,
2007). Informal settlements have been disproportionately badly impacted
by landslides in Colombia and Venezuela in the past (e.g., Takahashi et
al., 2001; Ojeda and Donnelly, 2006) and were similarly affected in 2010
during unusual heavy rains associated with the La Niña weather
phenomenon (NCDC, 2011). Densely settled regions in the Alps (Crosta
et al., 2004) and Himalayas have been similarly impacted (Petley et al.,
2007).
Cities can substantially increase local temperatures and reduce
temperature drop at night (e.g., see Case Study 9.2.1). This is the urban
heat island effect resulting from the large amount of heat-absorbing
material, building characteristics, and emissions of anthropogenic heat
from air conditioning units and vehicles (e.g., Rizwan et al., 2008; for a
critical review of heat island research, see Stewart, 2011). Heat waves
combined with urban heat islands (Basara et al., 2010; Tan et al., 2010)
can result in large death tolls with the elderly, the unwell, the socially
isolated, and outdoor workers (Maloney and Forbes, 2011) being
especially vulnerable, although acclimatization and heat health-warning
systems can substantially reduce excess deaths (Fouillet et al., 2008).
Heat waves thus pose a future challenge for major cities (e.g., Endlicher
et al., 2008; Bacciniet al., 2011; for London, Wilby, 2003). In urban areas,
heat waves also have negative effects on air quality and the number of
days with high pollutants, ground level ozone, and suspended particle
concentrations (Casimiro and Calheiros, 2002; Sanderson et al., 2003;
Langner et al., 2005).
The largest impacts from coastal inundation due to sea level rise (and/or
relative sea level rise) in low-elevation coastal zones (i.e., coastal areas
with an elevation less than 10 m above present mean sea level; see
McGranahan et al., 2007) are thought to be associated with extreme
sea levels due to tropical and extratropical storms (e.g., Ebersole et al.,
2010; Mozumder et al., 2011) that will be superimposed upon the long-
term sea level rise (e.g., Frazier et al., 2010). An increase in the mean
maximum wind speed of tropical cyclones is likely over the 21st century,
but possibly not in all ocean basins (see Table 3-1). The destructive
potential of tropical cyclones may increase in some regions as a result
of this projected increase in intensity of mean maximum wind speed
and tropical cyclone-related rainfall rates (see Section 3.4.4). Storms
generally result in considerable disruption and local destruction, but
cyclones and their associated storm surges have in some cases caused
very substantial destruction in modern cities (e.g., New Orleans and
Darwin; see also Case Study 9.2.5). The impacts are considered to be
more severe for large urban centers built on deltas and small island
states (McGranahan et al., 2007; Love et al., 2010; Wardekker et al.,
2010), particularly for those at the low end of the international income
distribution (Dasgupta et al., 2009). The details of exposure will be
controlled by the natural or human-induced characteristics of the system,
for example, the occurrence/distribution of protecting barrier islands
and/or coastal wetlands that may attenuate surges (see, e.g., Irish et al.,
2010; Wamsley et al., 2010) or changes such as land reclamation (Guo
et al., 2009). Recent studies (Nicholls et al., 2008; Hanson et al., 2011)
have assessed the asset exposure of port cities with more than one
million inhabitants (in 2005). They demonstrated that large populations
are already exposed to coastal inundation (~40 million people or 0.6%
of the global population) by a 1-in-100-year extreme event, while the
total value of exposed assets was estimated at US$ 3,000 billion (~ 5%
of the global GDP in 2005). By the 2070s, population exposure was
estimated to triple, whereas asset exposure could grow tenfold to some
US$ 35,000 billion; these estimates, however, do not account for the
potential construction of effective coastal protection schemes (see also
Dawson et al., 2005), with the exposure growth rate being more rapid in
developing countries (e.g., Adamo, 2010). Lenton et al. (2009) estimated a
substantial increase in the exposure of coastal populations to inundation
(see Figure 4-5).
4.3.5.2. Infrastructure
Weather- and climate-related extremes are expected to produce large
impacts on infrastructure, although detailed analyses of potential and
projected damages are limited to a few countries (e.g., Australia,
Canada, the United States; Holper et al., 2007), infrastructure types
(e.g., power lines), and sectors (e.g., transport, tourism). Inadequate
infrastructure design may increase the impacts of climate and weather
extremes, and some infrastructure may become inadequate where climate
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
249
change alters the frequency and severity of extremes, for example, an
increase in heavy rainfalls may affect the capacity and maintenance of
storm water, drainage, and sewerage infrastructure (Douglas et al., 2008).
In some infrastructure, secondary risks in case of extreme weather may
cause additional hazards (e.g., extreme rainfall can damage dams). The
same is true for industrial and mining installations containing hazardous
substances (e.g., heavy rainfall is the main cause of tailings dam failure,
accounting for 25% of incidents worldwide and 35% in Europe; Rico et
al., 2008).
In many parts of the world, including Central Asia and parts of Europe,
aging infrastructure, high operating costs, low responsiveness to
customers, and poor access to capital markets may limit the operability
of sewerage systems (Evans and Webster, 2008). Moreover, most urban
centers in sub-Saharan Africa and in Asia have no sewers (Hardoy et al.,
2001). Current problems of pollution and flooding will be exacerbated
by an increase in climatic and weather extremes (e.g., intense rainfall;
see Table 3-3 for projected regional changes).
Major settlements are dependent on lengthy infrastructure networks for
water, power, telecommunications, transport, and trade, which are exposed
to a wide range of extreme events (e.g., heavy precipitation and snow,
gale winds). Modern logistics systems are intended to minimize slack
and redundancies and as a result are particularly vulnerable to disruption
by extreme events (Love et al., 2010).
Transport infrastructure is vulnerable to extremes in temperature,
precipitation/river floods, and storm surges, which can lead to damage
in roads, rail, airports, and ports. Impacts on coastal infrastructure, on
services, and particularly on ports, key nodes of international supply
chains, are expected (e.g., Oh and Reuveny, 2010). This may have far-
reaching implications for international trade, as more than 80% of global
trade in goods (by volume) is carried by sea (UNCTAD, 2009). All coastal
modes of transportation are considered vulnerable, but exposure and
impacts will vary, for example, by region, mode of transportation, location/
elevation, and condition of transport infrastructure (NRC, 2008; UNCTAD,
2009). Coastal inundation due to storm surges and river floods can affect
terminals, intermodal facilities, freight villages, storage areas, and cargo
and disrupt intermodal supply chains and transport connectivity (see
Figure 4-6). These effects would be of particular concern to small island
states, whose transportation facilities are mostly located in low-elevation
coastal zones (UNCTAD, 2009; for further examples, see Love et al., 2010).
Regarding road infrastructure, Meyer (2008) pointed to bridges and
culverts as vulnerable elements in areas with projected increases in
heavy precipitation. Moreover, the lifetime of these rigid structures is
longer than average road surfaces and they are costly to repair or
replace. Increased temperatures could reduce the lifetime of asphalt on
road surfaces (Meizhu et al., 2010). Extreme temperature may cause
expansion and increased movement of concrete joints, protective
cladding, coatings, and sealants on bridges and airport infrastructure,
impose stresses in the steel in bridges, and disrupt rail travel (e.g., Arkell
and Darch, 2006). Nevertheless, roads and railways are typically replaced
every 20 years and can accommodate climate change at the time of
replacement (Meyer, 2008).
Electricity transmission infrastructure is also vulnerable to extreme
storm events, particularly wind and lightning, and in some cases heat
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
Figure 4-5 | For low-elevation coastal areas, current and future (2050) population exposure to inundation in the case of the 1-in-100-year extreme storm for sea level rise of
0.15 m and for sea level rise of 0.50 m due to the partial melting of the Greenland and West Antarctic Ice Sheets. Data from Lenton et al., 2009.
Population exposed
in 2050 in millions
0.50m SLR
0.15m SLR
Current population
exposed
6.2
11.7
3.8
60.2
2.32.7
Height of columns represents the number of exposed persons.
5.8
2.8
47.8
9.6
4.6
4.8
1.8
82.7
16.4
7.4
8.9
NORTH
AMERICA
SOUTH
AMERICA
EUROPE
AFRICA
ASIA
OCEANIA
5.6
250
waves (McGregor et al., 2007). The passage of the Lothar and Martin
storms across France in 1999 caused the greatest devastation to an
electricity supply network ever seen in a developed country, as 120
high-voltage transmission pylons were toppled, and 36 high-tension
transmission lines (one-quarter of the total lines in France) were lost
(Abraham et al., 2000). Severe droughts may also affect the supply of
cooling water to power plants, disrupting the ongoing supply of power
(see Box 4-4; Rübbelke and Vögele, 2011).
Buildings and urban facilities may be vulnerable to increasing frequency
of heavy precipitation events (see Section 3.3.2). Those close to the
coast are particularly at risk when storm surges are combined with sea
level rise. In commercial buildings, vulnerable elements are lightweight
roofs commonly used for warehouses, causing water spoilage to stored
goods and equipment. During the Lothar and Martin storms, the most
vulnerable public facilities were schools, particularly those built in the
1960s and 1970s and during the 1990s with the use of lightweight
architectural elements of metal, plastic, and glass in walls and roofs
(Abraham et al., 2000).
4.3.5.3. Tourism
The tourism sector is highly sensitive to climate, since climate is the
principal driver of global seasonality in tourism demand (Lise and Tol,
2002; Becken and Hay, 2007). Approximately 10% of global GDP is
spent on recreation and tourism, constituting a major source of income
and foreign currency in many developing countries (Berrittella et al.,
2006). Extreme events may play an important role in tourist decisions
(e.g., Hein et al., 2009; Yu et al., 2009).
There are three broad categories of impacts of climate extremes that
can affect tourism destinations, competitiveness, and sustainability (Scott
et al., 2008): (1) direct impacts on tourist infrastructure (hotels, access
roads, etc.), on operating costs (heating/cooling, snowmaking, irrigation,
food and water supply, evacuation, and insurance costs), on emergency
preparedness requirements, and on business disruption (e.g., sun-and-sea
or winter sports holidays); (2) indirect environmental change impacts
of extreme events on biodiversity and landscape change (e.g., coastal
erosion), which may negatively affect the quality and attractiveness of
tourism destinations; and (3) tourism-adverse perception of particular
touristic regions after occurrence of the extreme event itself. For example,
adverse weather conditions or the occurrence of an extreme event can
reduce a touristic region’s popularity among tourists during the following
season.
Apart from extreme events, large impacts on some tourist destinations
may be produced by medium-term projected climate change effects (e.g.,
Bigano et al., 2008). Salinization of the groundwater resources due to
sea level rise, land reclamation, and overexploitation of coastal aquifers
(e.g., Alpa, 2009) as well as changing weather extreme patterns (Hein
et al., 2009) will pose additional stresses for the industry. Nevertheless,
the potential impacts on the tourist industry will depend also on
tourists’ perceptions of the coastal destinations (e.g., of destinations
experiencing beach erosion) that, however, cannot be easily predicted
(Buzinde et al., 2009). Capacity to recover is related to the degree of
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
Figure 4-6 | Freight-handling port facilities at risk from storm surge of 5.5 and 7 m on the US Gulf Coast. Adapted from CCSP, 2008.
Galveston
Pascagoula
Gulfport/
Biloxi
Freeport
South Louisiana
Port Arthur
Texas City
Plaquemines
Mobile
Houston
Beaumont
New Orleans
Baton Rouge
Lake Charles
Galveston
Pascagoula
Gulfport/
Biloxi
Freeport
South Louisiana
Port Arthur
Texas City
Plaquemines
Red River
Mobile
Houston
Beaumont
New Orleans
Baton Rouge
Lake Charles
Mississippi River
Mobile River
Texas
Louisiana
Mississippi
Alabama
Florida
0 50 100 150 20025
Miles
States
Study Area Counties
Rivers
USACE Cargo Ports
18 to 23 feet
Below 18 feet
Elevation
251
dependence on tourism, with diversified economies being more robust
(Ehmer and Heymann, 2008). However, low-lying coastal areas and areas
currently on the edge of the snow limit may have limited alternatives.
Some ski resorts will be able to adapt using snowmaking, which has
become an integral component of the ski industry in Europe (Elsasser
and Bürki, 2002), although at the expense of high water and energy
consumption.
In some regions, the main impact of extreme events on tourism is decline
in revenue, with loss of livelihoods for those working in the sector
(Hamilton et al., 2005; Scott et al., 2008; Hein et al., 2009). Quantitative
regional climate projections of the frequency or magnitude of certain
weather and climate extremes (e.g., heat waves and droughts; see, for
example, Table 3-3) inform qualitative understanding of regional impacts
on tourism activities (see Box 4-3). The vulnerable hotspot regions in
terms of extreme impacts of climate change on tourism include the
Mediterranean, Caribbean, small islands of the Indian and Pacific
Oceans, and Australia and New Zealand (Scott et al., 2008). Direct and
indirect effects of extremes in these regions will vary greatly with
location (Gössling and Hall, 2006a,b; Wilbanks et al., 2007).
Box 4-3 points out a number of potential climate extreme impacts on
tourism regions and activities.
4.3.6. Human Health, Well-Being, and Security
Climate extremes, such as heat waves, floods, droughts, and cyclones,
influence human health, well-being, and security.
Heat waves have affected developed countries, as exemplified by the
2003 European heat wave (see Case Study 9.2.1 and Box 4-4). In much
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
Box 4-3 | Regional Examples of Potential Impacts of Climate Extremes on Tourism
Tropics
Projections indicate a likely increase in mean maximum wind speed (but not in all basins) and in tropical cyclone-related rainfall rates
(see Table 3-1). In the Caribbean, tourist activities may be reduced where beaches erode with sea level rise and where coral is bleached,
impacting snorkelers and divers (Uyarra et al., 2005).
Small island states are dependent on tourism, and the tourism infrastructure that lies on the coast is threatened by climate change
(Berrittella et al., 2006). Sea level rise in the 20th century – with an average rate of 1.7 mm yr
-1
and a significantly higher rate
(3.1 mm yr
-1
) for the period 1993–2003 (Bindoff et al., 2007) – poses risks for many touristic resorts of small islands in the Pacific and
Indian Oceans (Becken and Hay, 2007; Scott et al., 2008).
Alpine Regions
Warming temperatures will raise the snowline elevation (Elsasser and Bürki, 2002; Scott et al., 2006). In Switzerland, only 44% of ski
resorts are projected to be above the ‘snow-reliable’ altitude (snow for 100 days per season) by approximately 2030, as opposed to 85%
today (Elsasser and Bürki, 2002). In Austria, 83% of ski resorts are currently snow-reliable but an increase in temperature of 1 and 2°C is
projected to reduce this number to 67 and 50%, respectively (Abegg et al., 2007). Ski season simulations show that snowmaking
technology can maintain snow-reliable conditions in Austria until the 2040s (A1B) to the 2050s (B1), but by the end of the century the
required production in snow volume is projected to increase by up to 330% (Steiger, 2010). This artificial snow production will increase
vulnerability to water shortage and local water conflicts, in particular in the French Alps (EEA, 2009).
Mediterranean Countries
More frequent heat waves and tropical nights (>20°C) in summer (see Table 3-3) may lead to exceedance of comfortable temperature
levels and reduce the touristic flow by 2060 (Hein et al., 2009). Tourism occupancy may increase during spring and autumn and decrease
in summer (Perry, 2003; Esteban-Talaya et al., 2005). Northern European countries are expected to become relatively more attractive,
closing their gap with the currently popular southern European countries (Hamilton et al., 2005).
There are major regional gaps in understanding how climate change may affect the natural and cultural resources in Africa and South
America, preventing further insight on corresponding impacts for tourism activities (Scott et al., 2008).
In many regions, some types of tourism will benefit from, or be unaffected by, climate extremes (Scott et al., 2008). When an area is
impacted directly by an extreme event, tourists will often go to another destination with the result that one area’s loss becomes another’s
gain. The impacted area may also gain in the longer term through the provision of new infrastructure. City and cultural tourism is
generally seen as relatively unaffected by climate and weather events (Scott et al., 2008).
252
of the developed world, societies are aging and hence can be more
sensitive to climate extremes, such as heat waves (Hennessy et al.,
2007). Heat extremes can claim casualties even in tropical countries,
where people are acclimatized to the hot climate; McMichael et al.
(2008) evaluated the relation between daily temperature and mortality
in middle- and low-income countries, and reported that higher mortality
was observed on very hot days in most of the cities, including tropical
cities, such as Bangkok, Thailand; Delhi, India; and Salvador, Brazil.
Floods can cause deaths and injuries and can be followed by infectious
diseases (such as diarrhea) and malnutrition due to crop damage (see
Section 4.4.2.3). In Dhaka, Bangladesh, the severe flood in 1998 was
associated with an increase in diarrhea during and after the flood, and
the risk of non-cholera diarrhea was higher among those from a lower
socioeconomic group and not using tap water (Hashizume et al., 2008).
Floods may also lead to a geographical shift of malaria epidemic
regions by changing breeding sites for vector mosquitoes. Outbreaks of
malaria were associated with changes in habitat after the 1991 floods in
Costa Rica’s Atlantic region (Saenz et al., 1995; for another example, see
Case Study 9.2.6). Malaria epidemics can also occur when people with
little immunity move into endemic regions, although the displacement
of large populations has rarely occurred as a result of acute natural
disasters (Toole, 1997).
Drought can affect water security, as well as food security through
reduction of agricultural production (MacDonald, 2010), and it can be a
factor contributing to human-ignited forest fires, which can lead to
widespread deforestation and carbon emissions (D’Almeida et al., 2007;
van der Werf et al., 2008; Field et al., 2009; Phillips et al., 2009; Costa
and Pires, 2010). Also, drought can increase or decrease the prevalence
of mosquito-borne infectious diseases such as malaria, depending on the
local conditions (Githeko et al., 2000), and is associated with meningitis
(Molesworth et al., 2003). Studies indicate that there is a climate signal
in forest fires throughout the American West and Canada and that there
is a projected increase in severe wildfires in many areas (Gillett et al.,
2004; Westerling et al., 2006; Westerling and Bryant, 2008). As described
by McMichael et al. (2003a), the direct effects of fire on human health
can include burns and smoke inhalation, with indirect health impacts
potentially resulting from loss of vegetation on slopes, increased soil
erosion, and resulting increased risk of landslides.
Evaluation of how impacts of climate extremes affect human health tend
to focus on the direct, immediate effects of the event, using parameters
that are often easier to obtain and quantify like death statistics or
hospitalizations. These direct observable outcomes are used to
demonstrate the extremity of an event and as a comparison metric to
measure against other extreme events. However, indirect health impacts
are not often reported, because they are one step removed from the
event. Because indirect impacts are hard to monitor and are often
temporally separated from the event, they are effectively removed from
the cause-and-effect linkage to that event. Examples of indirect health
impacts from extreme weather events include illnesses or injury resulting
from disruption of human infrastructure built to deal with basic needs like
medical services; exposure to infectious or toxic agents after an extreme
event like cyclones or flooding (Schmid et al., 2005); stress, anxiety, and
mental illness after evacuation or geographical displacement (Fritze et
al., 2008) as well as increased susceptibility to infection (Yee et al.,
2007); and disruption of socioeconomic structures and food production
that leads to increases of malnutrition that might not manifest until
months after an extreme event (Haines et al., 2006; McMichael et al.,
2006). Indirect health impacts are therefore a potentially large but
under-examined outcome of extreme weather events that lead to a
substantial underestimation of the total health burden.
There is a growing body of evidence that the mental health impact from
extreme events is substantial (Neria et al., 2008; Berry et al., 2010).
Often overshadowed by the physical health outcomes of an event, the
psychological effects can be long lasting and can affect a large portion of
a population (Morrissey and Reser, 2007). An extreme event may affect
mental health directly from acute traumatic stress from an event, with
common outcomes of anxiety and depression. It can also have indirect
impacts during the recovery period associated with the stress and
challenges of loss, disruption, and displacement. Furthermore, indirect
mental health impacts could even affect individuals not directly associated
with an event, like grieving friends and family of those who die from an
event or the rescue and aid workers who suffer post-traumatic stress
disorder (PTSD) after their aid efforts. Long-term mental health impacts
are not often adequately monitored, but the body of research conducted
after natural disasters in the past three decades suggests that the burden
of PTSD among persons exposed to disasters is substantial (Neria et al.,
2008). A range of other stress-related problems such as grief, depression,
anxiety disorders, somatoform disorders, and drug and alcohol abuse
(Fritze et al., 2008) have lasting effects, long after the causative event.
There remain large limitations in evaluating health impacts of climate
change. The largest research gap is a lack of information on impact
outcomes themselves in developing countries in general. This includes the
mortality/morbidity data and information on other contributing factors
such as nutritional status or access to safe water and medical facilities.
4.4. Regionally Based Aspects of
Vulnerability, Exposure, and Impacts
4.4.1. Introduction
The regional subsections presented here discuss the impacts of extreme
weather and climate events within the context of other issues and
trends. Regional perspective, in social and economic dimensions, is
important especially since decisionmaking often has a strong regional
context. For a comprehensive assessment of observed and projected
regional changes in climate extremes, see Sections 3.3 to 3.5 and
Tables 3-2 and 3-3.
For various climate extremes, the following aspects are considered on a
regional basis: exposure of humans and their activities to given climate
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
253
extremes; the vulnerability of what is exposed to the climate extreme;
and the resulting impacts. The individual sections below are structured
as is most logical for the trends relevant to each region.
4.4.2. Africa
4.4.2.1. Introduction
Climate extremes exert a significant control on the day-to-day economic
development of Africa, particularly in traditional rain-fed agriculture
and pastoralism, and water resources, at all scales
. Floods and droughts
can cause major human and environmental impacts on and disruptions
to the economies of African countries, thus exacerbating vulnerability
(AMCEN/UNEP, 2002; Scholes and Biggs, 2004; Washington et al., 2004;
Thornton et al., 2006). There is still limited scientific information available
on observed frequency and projections of many extreme events in Africa
(e.g., see Tables 3-2 and 3-3), despite frequent reporting of such events,
including their impacts.
Agriculture as an economic sector is most vulnerable and most exposed
to climate extremes in Africa. It contributes approximately 50% to
Africa’s total export value and approximately 21% of its total GDP
(Mendelsohn et al., 2000; PACJA, 2009). In particular, with an inefficient
agriculture industry, sub-Saharan Africa is extremely vulnerable to climate
extremes. This vulnerability is exacerbated by poor health, education, and
governance standards (Brooks et al., 2005). Reid et al. (2007) project
climate impacts on Namibia’s natural resources that would cause annual
losses of 1 to 6% of GDP, of which livestock production, traditional
agriculture, and fishing are expected to be hardest hit, with a combined
loss of US$ 461 to 2,045 million per year by 2050.
4.4.2.2. Droughts and Heat Waves
An overall increase in dryness in Africa has been observed (medium
confidence), with prolonged Sahel drought, but regional variability is
observed (see Table 3-2). Droughts have affected the Sahel, the Horn of
Africa, and Southern Africa particularly since the end of the 1960s
(Richard et al., 2001; L’Hôte et al., 2002; Brooks, 2004; Christensen et al.,
2007; Trenberth et al., 2007). One of the main consequences of multi-
year drought periods is severe famine, such as the one associated with
the drought in the Sahel in the 1980s, causing many casualties and
important socioeconomic losses. The people in Africa who live in
drought-prone areas are vulnerable to the direct impacts of droughts
(e.g., famine, death of cattle, soil salinization), as well as indirect
impacts (e.g., illnesses such as cholera and malaria) (Few et al., 2004).
The water sector is strongly influenced by, and sensitive to, periods of
prolonged drought conditions in a continent with limited water storage
infrastructure. Natural water reservoirs such as lakes experience
amarkedinterannual water level fluctuationrelated to rainfall interannual
variability (Nicholson et al., 2000; Verschuren et al., 2000).
Large changes in hydrology and water resources linked to climate
variability have led to water stress conditions in human and ecological
systems in a number of African countries (Schulze et al., 2001; New, 2002;
Legesse et al., 2003; Eriksen et al., 2005; de Wit and Stankiewicz, 2006;
Nkomo and Bernard, 2006). Twenty-five percent of the contemporary
African population has limited water availability and thus constitutes a
drought-sensitive population, whereas 69% of the population experiences
relative water abundance (Vörösmarty et al., 2005). Even for this latter
part of the population, however, relative abundance does not necessarily
correspond to access to safe drinking water and sanitation, and this
effective reduction of the quantity of freshwater available for human use
negatively affects vulnerability. Despite the considerable improvements
in access to freshwater in the 1990s, only about 62% of the African
population had access to improved water supplies in 2000 (WHO/
UNICEF, 2000). As water demand increases, the population exposed to
different drought conditions (agricultural, climate, urban) is expected to
increase as well.
Increasing drought risk may cause a decline in tourism, fisheries, and
cropping (UNWTO, 2003). This could reduce the revenue available to
governments, enterprises, and individuals, and hence further deteriorate
the capacity for adaptation investment. For example, the 2003-2004
drought cost the Namibian Government N$ 275 million (US$ 43-48
million) in provision of emergency relief (Reid et al., 2007). Cameroon’s
economy is highly dependent on rain-fed agriculture; a 14% reduction in
rainfall is projected to cause significant losses, of up to US$ 4.56 billion
(Molua and Lambi, 2006).
4.4.2.3. Extreme Rainfall Events and Floods
There are inconsistent patterns of change in heavy precipitation in
Africa and partial lack of data; hence there is low confidence in
observed precipitation trends (see Table 3-2). Heavy precipitation may
induce landslides and debris flows in tropical mountain regions (Thomas
and Thorp, 2003) with potential impacts for human settlements. In the
arid and semi-arid areas of countries of the Horn of Africa, extreme
rainfall events are often associated with a higher risk of the vector and
epidemic diseases of malaria, dengue fever, cholera, Rift Valley fever,
and hantavirus pulmonary syndrome (Anyamba et al., 2006; McMichael
et al., 2006).
The periods of extreme rainfall and recurrent floods seem to correlate
with the El Niño phase (Reason and Kiebel, 2004; Reason et al., 2005;
Washington and Preston, 2006; Christensen et al., 2007) of ENSO events
(e.g., 1982-1983, 1997-1998, 2006-2007). When such events occur,
important economic and human losses result. In 2000, floods in
Mozambique (see Case Study 9.2.6), particularly along the valleys of the
rivers Limpopo, Save, and Zambezi, resulted in 700 reported deaths and
about half a million homeless. The floods had a devastating effect on
livelihoods, destroying agricultural crops, disrupting electricity supplies,
and demolishing basic infrastructure (Osman-Elasha et al., 2006).
However, floods can be highly beneficial in African drylands (e.g.,
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
254
Sahara and Namib Deserts) since the floodwaters infiltrate and recharge
alluvial aquifers along ephemeral river pathways, extending water
availability to dry seasons and drought years (Morin et al., 2009; Benito
et al., 2010), and supporting riparian systems and human communities
(e.g., Walvis Bay in Namibia).
Damage to African port cities from flooding, storm surge, and high winds
might increase due to climate change. For instance, it is indicated that in
Alexandria, US$ 563.28 billion worth of assets could suffer damage
or be lost because of coastal flooding alone by 2070 (Nicholls et al.,
2008).
4.4.2.4. Dust Storms
Atmospheric dust is a major element of the Saharan and Sahelian
environments. The Sahara Desert is the world’s largest source of
airborne mineral dust, which is transported over large distances,
traversing northern Africa and adjacent regions and depositing dust on
other continents (Osman-Elasha et al., 2006; Moulin et al., 1997). Dust
storms have negative impacts on agriculture, health, and structures.
They erode fertile soil; uproot young plants; bury water canals, homes,
and properties; and cause respiratory problems. Meningitis transmission
is associated with dust in semi-arid conditions and overcrowded living
conditions. The frequency of dust events has increased in the Sahel
zone, but studies of observations and in particular studies of projections
of dust activity are limited (see Section 3.5.8).
4.4.3. Asia
Asia includes mega-deltas, which are susceptible to extreme impacts due
to a combination of the following factors: high-hazard rivers, coastal
flooding, and increased population exposure from expanding urban
areas with large proportions of high vulnerability groups (Nicholls et al.,
2007). Asia can also expect changes in the frequency and magnitude of
extreme weather and climate events, such as heat waves and heavy
precipitation (see, e.g., Table 3-3). Such changes may have ramifications
not only for physical and natural systems but also for human systems.
4.4.3.1. Tropical Cyclones (Typhoons or Hurricanes)
Damage due to storm surge is sensitive to any changes in the magnitude
of tropical cyclones (Xiao and Xiao, 2010). For example, changes in
storm surge and associated damage were projected for the inner parts
of three major bays (Tokyo, Ise, and Osaka) in Japan (Suzuki, 2009). The
projections were based on calculations of inundations for different sea
levels and different strengths of typhoons, using a spatial model with
information on topography and levees. The research indicated that a
typhoon that is 1.3 times as strong as the design standard with a sea
level rise of 60 cm would cause damage costs of about US$ 3, 40, and
27 billion, respectively, in the investigated bays.
Awareness, improved governance, and development are essential in
coping with extreme tropical cyclone and typhoon events in developing
Asian countries (Cruz et al., 2007). For example, two cyclones in the
Indian Ocean (Sidr and Nargis) of similar magnitude and strength
caused a significantly different number of fatalities. A comparison is
presented in Case Study 9.2.5.
For the period from 1983 to 2006, the direct economic losses in China
increased, but there is no trend if the losses are normalized by annual
total GDP and GDP per capita, suggesting Chinese economic development
contributed to the upward trend. This hypothesis is consistent with data
on tropical cyclone casualties, which showed no significant trend over
the 24 years (Zhang et al., 2009). Similarly, normalized losses from
typhoons on the Indian southeast coast since 1977 show no increases
(Raghavan and Rajesh, 2003).
4.4.3.2. Flooding
The geographical distribution of flood risk is heavily concentrated in
India, Bangladesh, and China, causing high human and material losses
(Brouwer et al., 2007; Dash et al., 2007; Shen et al., 2008). Regarding
the occurrence of the extreme events themselves, different flooding
trends have been detected and projected in various catchments, but the
evidence for broader regional trends is limited (see Section 3.5.2).
In July 2005, severe flooding occurred in Mumbai, India, after 944 mm
of rain fell in a 24-hour period (Kshirsagar et al., 2006). The consequent
flooding affected households, even in more affluent neighborhoods.
Poor urban drainage systems in many parts of India can be easily
blocked. Ranger et al. (2011) analyzed risk from heavy rainfall in the city of
Mumbai, concluding that total losses (direct plus indirect) for a 1-in-100
year event could triple by the 2080s compared with the present
(increasing from US$ 700 to 2,305 million), and that adaptation could
help reduce future damages.
As noted in the final report for the Ministry of Environment and Forest
(2005) of the People’s Republic of Bangladesh, flooding in Bangladesh
is a normal, frequently recurrent, phenomenon. Bangladesh experiences
four types of floods: flash floods from the overflowing of hilly rivers; rain
floods due to poor drainage; monsoon floods in the flood plains of
major rivers; and coastal floods following storm surge. In a normal year,
river spills and drainage congestions cause inundation of 20 to 25% of
the country’s area. Inundation areas for 10-, 50-, and 100-year floods
constitute 37, 52, and 60% of the country’s area, respectively. In 1987,
1988, and 1998, floods inundated more than 60% of the country.
The 1998 flood alone led to 1,100 deaths, caused inundation of nearly
100,000 km
2
, left 30 million people homeless, and substantially
damaged infrastructure.
There have been increases in flood impacts associated with changes in
surrounding environments. Flooding has increased over the past few
decades in the Poyang Lake, South China, due to levee construction
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
255
protecting a large rural population (Shankman et al., 2006). Such levees
reduce the area for floodwater storage, leading to higher lake stages
during the summer flood season and then levee failures. The most
extreme floods occurred during or immediately following El Niño events
(Shankman et al., 2006). Fengqing et al. (2005) analyzed losses from
flooding in the Xinjiang autonomous region of China, and found an
increase that seems to be linked to changes in rainfall and flash floods
since 1987.
Heavy rainfall and flooding also affect environmental health in urban
areas because surface water can be quickly contaminated. Urban poor
populations in low- and middle-income countries can experience higher
rates of infectious disease after floods, such as cholera, cryptosporidiosis,
and typhoid fever (Kovats and Akhtar, 2008).
4.4.3.3. Temperature Extremes
Increases in warm days/nights and heat wave duration, frequency,
and/or intensity are observed and projected in Asia (see Tables 3-2 and
3-3), with adverse impacts on both human and natural systems. In 2002,
a heat wave was reported to have killed 622 people in the southern
Indian state of Andhra Pradesh. Persons living in informal settlements
and structures are more exposed to high temperatures (Kovats and
Akhtar, 2008).
Agriculture is also affected directly by temperature extremes. For example,
rice, the staple food in many parts of Asia, is adversely affected by
extremely high temperature, especially prior to or during critical pollination
phases (see Section 4.3.4).
4.4.3.4. Droughts
Asia has a long history of drought, which has been linked with other
climate extremes. Spatially varying trends have been observed during
the second half of the 20th century, with increasing dryness noted in
some areas, particularly in East Asia (see Table 3-2), adversely affecting
socioeconomic, agricultural, and environmental conditions. Drought
causes water shortages, crop failures, starvation, and wildfire.
In Southeast Asia, El Niño is associated with comparatively dry
conditions: 93% of droughts in Indonesia between 1830 and 1953
occurred during El Niño years (Quinn et al., 1978). In four El Niño years
between 1973 and 1992, the average annual rainfall amounted to only
around 67% of the 20-year average in two major rice growing areas in
Java, Indonesia, causing a yield decline of approximately 50% (Amien
et al., 1996).
During drought, severe water scarcity results from one of, or a
combination of, the following mechanisms: insufficient precipitation;
high evapotranspiration; and over-exploitation of water resources
(Bhuiyan et al., 2006).
About 15% (23 million ha) of Asian rice areas experience frequent yield
loss due to drought (Widawsky and O’Toole, 1990). The problem is
particularly pertinent to eastern India, where the area of drought-prone
fields exceeds more than 10 million ha (Pandey et al., 2000). Even when
the total rainfall is adequate, shortages during critical periods reduce
yield (Kumar et al., 2007). Lowland rice production in the Mekong
region is generally reduced because crops are cultivated under rain-fed
conditions, rather than irrigated, and often exposed to drought. In
Cambodia, severe drought that affects grain yield mostly occurs late in
the growing season, and longer-duration genotypes are more likely to
encounter drought during grain filling (Tsubo et al., 2009).
Asian wetlands provide resources to people in inundation areas, who
are susceptible to droughts. For achieving the benefits from fertilization
for inundation agriculture in Cambodia, wide areas along the rivers
need to be flooded (Kazama et al., 2009). Flood protection in this area
needs to consider this benefit of inundation.
4.4.3.5. Wildfires
Grassland fire disaster is a critical problem in China (Su and Liu, 2004;
Zhang et al., 2006), especially in northwestern and northeastern China
due to expansive territory and complex physiognomy. Statistical analysis
of historical grassland fire disaster data has suggested a gradual
increase in grassland fire disasters with economic development and
population growth in 12 northern provinces of China between 1991 and
2006 (Liu et al., 2006).
In tropical Asia, although humans are igniting the fires, droughts are
predisposing factors for fire occurrence (Field et al., 2009). Drought
episodes, forest fires, drainage of rice fields, and oil palm plantations are
drying peatlands, which are then more susceptible to fires (van der Werf
et al., 2008). Peatland fires are an important issue given the difficulties
of extinguishing them and their potential effects on climate.
4.4.4. Central and South America
4.4.4.1. Extreme Rainfalls in South America
Extreme rainfall episodes have caused disasters in parts of South
America, with hundreds to thousands of fatalities in mudslides and
landslides, as typified, for example, by the December 1999 incident in
Venezuela (Lyon, 2003). However, there is low to medium confidence in
observed (Table 3-2) and in projected (Table 3-3) changes in heavy
precipitation in the region.
4.4.4.2. Wildfires
There is a low to medium confidence in projections of trends in dryness
in South America (see Table 3-3). Magrin et al. (2007) indicated that
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
256
more frequent wildfires are probable (an increase in frequency of 60%
for a temperature increase of 3°C) in much of South America. In most of
central and northern Mexico, the semi-arid vegetation could be replaced
by the vegetation of arid regions (Villers and Trejo, 2004). Due to the
interrelated nature of forest fires, deforestation, drought, and climate
change, isolating one of the processes fails to describe the complexity
of the interconnected whole.
4.4.4.3. Regional Costs
Climatic disasters account for the majority of natural disasters in Central
America, with most of its territory located in tropical and equatorial areas.
Low-lying states are especially vulnerable to hurricanes and tropical
storms. In October 1998, Hurricane Mitch, one of the most powerful
hurricanes of the tropical Atlantic Basin of the 20th century, caused
direct and indirect damages to Honduras of US$ 5 billion, equivalent to
95% of Honduras’ 1998 GDP (Cardemil et al., 2000). Some literature
indicates that hurricane losses, when corrected for population and
wealth in Latin America and the Caribbean, have not increased since the
1940s (Pielke Jr. et al., 2003); and that increasing population and assets
at risk are the main reason for increasing impacts.
4.4.5. Europe
4.4.5.1. Introduction
This section assesses vulnerability and exposure to climate extremes in
Europe, evaluating observed and projected impacts, disasters, and risks.
Europe has a higher population density and lower birth rate than any
other continent. It currently has an aging population; life expectancy
is high and increasing, and child mortality is low and decreasing
(Eurostat, 2010). European exposure to climate- and weather-related
hazards has increased whereas vulnerability has decreased as a result
of implementation of policy, regulations, and risk prevention and
management (EEA, 2008; UNISDR, 2009).
4.4.5.2. Heat Waves
Summer heat waves have increased in frequency and duration in most
of Europe (Section 3.3.1 and Table 3-2) and have affected vulnerable
segments of European society. During the 2003 heat wave, several tens
of thousands of additional heat-related deaths were recorded (see Case
Study 9.2.1 and Box 4-4). Urban heat islands pose an additional risk to
urban inhabitants. Those most affected are the elderly, ill, and socially
isolated (Kunst et al., 1993; Laschewski and Jendritzky, 2002; see
Case Study 9.2.1). There are mounting concerns about increasing heat
intensity in major European cities (Wilby, 2003) because of the large
population that inhabits urban areas. Building characteristics, emissions
of anthropogenic heat from air conditioning units and vehicles, as well
as lack of green open areas in some parts of the cities, may exacerbate
heat load during heat waves (e.g., Stedman, 2004; Wilby, 2007).
However, as high summer temperatures and urban heat waves become
more common, populations are able to adapt to such ‘expected’
temperature conditions, decreasing mortality during subsequent heat
waves (Fouillet et al., 2008).
4.4.5.3. Droughts and Wildfires
Drought risk is a function of the frequency, severity, and spatial and
temporal extent of dry spells and of the vulnerability and exposure of a
population and its economic activity (Lehner et al., 2006). In
Mediterranean countries, droughts can lead to economic damages
larger than floods or earthquakes (e.g., the drought in Spain in 1990
affected 6 million people and caused material losses of US$ 4.5 billion;
after CRED, 2010). The most severe human consequences of droughts
are often found in semiarid regions where water availability is already
low under normal conditions, water demand is close to, or exceeds,
natural availability, and/or society lacks the capacity to mitigate or
adapt to drought (Iglesias et al., 2009). Direct drought impacts affect
all forms of water supply (municipal, industrial, and agricultural).
Other sectors and systems affected by drought occurrence are
hydropower generation, tourism, forestry, and terrestrial and aquatic
ecosystems.
Forest fire danger (length of season, frequency, and severity) depends
on the occurrence of drought. There is medium confidence in observed
changes in drought in Europe (Table 3-2). Projections indicate increasing
dryness in central Europe and the Mediterranean, with no major change
in Northern Europe (medium confidence) (see Table 3-3). In the
Mediterranean, an increase in dryness may lead to increased dominance
of shrubs over trees (Mouillot et al., 2002); however, it does not translate
directly into increased fire occurrence or changes in vegetation (Thonicke
and Cramer, 2006). Analysis of post-fire forest resilience contributes to
identifying ‘risk hotspots’ where post-fire management measures
should be applied as a priority (Arianoutsou et al., 2011).
4.4.5.4. Coastal Flooding
Coastal flooding is an important natural disaster, since many Europeans
live near the coasts. Storm surges can be activated as a result of wind-
driven waves and winter storms (Smith et al., 2000), whereas long-term
processes are linked to global mean sea level rise (Woodworth et al.,
2005). Locations currently experiencing adverse impacts such as coastal
erosion and inundation will continue to do so in the future (see Section
3.5.5). Expected sea level rise is projected to have impacts on Europe’s
coastal areas including land loss, groundwater and soil salinization, and
damage to property and infrastructure (Devoy, 2008). Hinkel et al.
(2010) found that the total monetary damage in coastal areas of the
Member Countries of the European Union caused by flooding, salinity
intrusion, land erosion, and migration is projected to rise without
adaptation by 2100 to roughly € 17 billion per year under the A2 and
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
257
B1 emission scenarios. The Netherlands is an example of a country that
is highly susceptible to both sea level rise and coastal flooding, with
damage costs relative to GDP of up to 0.3% of GDP under the A2
scenario (Hinkel et al., 2010). By 2100, adaptation can reduce the
number of people flooded by two orders of magnitude and the total
damage costs by a factor of seven to nine (Hinkel et al., 2010).
4.4.5.5. Gale Winds
Storms have been one of the most important climate hazards for the
insurance industry in Europe (Munich Re NatCatSERVICE data cited in
EEA, 2008). In the most severe extratropical windstorm month,
December 1999, when three events struck Europe (Anatol – December 3,
Denmark; Lothar – December 26, France, Germany, and Switzerland;
and Martin – December 28, France, Spain, and Italy), insured damage
was in excess of US$ 12 billion (Schwierz et al., 2010). Typical economic
losses were generated by gale winds via effects on electrical distribution
systems, transportation, and communication lines; by damage to
vulnerable elements of buildings (e.g., lightweight roofs); and by trees
falling on houses. Some researchers have found no contribution from
climate change to trends in the economic losses from floods in Europe
since the 1970s (Barredo, 2009). Some studies have found evidence of
increasing damages to forests in Sweden and Switzerland (Nilsson et
al., 2004; Usbeck et al., 2010). Still other studies assert that increases in
forest disturbances in Europe are mostly due to changes in forest
management (e.g., Schelhaas et al., 2003).
There is medium confidence in projected poleward shifts of mid-latitude
storm tracks but low confidence in detailed regional projections (see
Section 3.4.5). According to a study by Swiss Re (2009), if by the end of
this century once-in-a-millennium storm surge events strike northern
Europe every 30 years, this could potentially result in a disproportionate
increase in annual expected losses from a current € 0.6 to 2.6 billion by
end of the century. Similar results are obtained from global and regional
climate models run under the IPCC SRES A1B emission scenario (Donat
et al., 2010). Adaptation to the changing wind climate may reduce by
half the estimated losses (Leckebusch et al., 2007; Donat et al., 2010),
indicating that adaptation through adequate sea defenses and the
management of residual risk is beneficial.
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
Box 4-4 | Extraordinary Heat Wave in Europe, Summer 2003
The extraordinarily severe heat wave over large parts of the European continent in the summer of 2003 produced record-breaking
temperatures particularly during June and August (Beniston, 2004; Schär et al., 2004). Average summer (June to August) temperatures
were by up to five standard deviations above the long-term mean, implying that this was an extremely unusual event (Schär and
Jendritzky, 2004). Regional climate model simulations suggest the 2003 heat wave bears resemblance to summer temperatures in the
late 21st century under the A2 scenario (Beniston, 2004).
Electricity demand increased with the high heat levels. Additionally, drought conditions created stress on health, water supplies, food
storage, and energy systems; for example, reduced river flows reduced the cooling efficiency of thermal power plants (conventional and
nuclear), and six power plants were shut down completely (Létard et al., 2004). Many major rivers (e.g., the Po, Rhine, Loire, and Danube)
were at record low levels, resulting in disruption of inland navigation and irrigation, as well as power plant cooling (Beniston and Díaz,
2004; Zebisch et al., 2005). In France, electricity became scarce, construction productivity fell, and the cold storage systems of approximately
20 to 30% of all food-related establishments were found to be inadequate (Létard et al., 2004). The (uninsured) economic losses for the
agriculture sector in the European Union were estimated at € 13 billion (Sénat, 2004). A record drop in crop yield of 36% occurred in
Italy for maize grown in the Po valley, where extremely high temperatures prevailed (Ciais et al., 2005). The hot and dry conditions led to
many very large wildfires. Glacier melting in the Alps prevented even lower river flows in the Danube and Rhine (Fink et al., 2004).
Health and health service-related impacts of the heat wave were dramatic, with excess deaths of about 35,000 (Kosatsky, 2005). Elderly
people were among those most affected (WHO, 2003; Borrell et al., 2006; Kovats and Ebi, 2006), but deaths were also associated with
housing and social conditions, for example, being socially isolated or living on the top floor (Vandentorren et al., 2006). The high mortality
during the 2003 heat wave marked an inflexion point in public awareness of the dangers of high temperatures, conducive to increasing
the preventive measures set up by health institutions and authorities (Koppe et al., 2004; Pascal et al., 2006).
During the July 2006 heat wave, about 2,000 excess deaths occurred in France (Rey et al., 2007). The excess mortality during the 2006
heat wave was markedly lower than that predicted by Fouillet et al. (2008) based on the quantitative association between temperature
and mortality observed during 1975-2003. Fouillet et al. (2008) interpreted this mortality reduction (~4,400 deaths) as a decrease in the
population’s vulnerability to heat, together with increased awareness of the risk related to extreme temperatures, preventive measures,
and the warning system established after the 2003 heat wave.
258
4.4.5.6. Flooding
Flooding is the most frequent natural disaster in Europe (EEA, 2008).
Economic losses from flood hazards in Europe have increased considerably
over previous decades (Lugeri et al., 2010), and increasing exposure of
people and economic assets is probably the major cause of the long-
term changes in economic disaster losses (Barredo, 2009). Exposure is
influenced by socioeconomic development, urbanization, and infrastructure
construction on flood-prone areas. Large flood impacts have been caused
by a few individual flood events (e.g., the 1997 floods in Poland and
Czech Republic, the 2002 floods in much of Europe, and the 2007 summer
floods in the United Kingdom). The projected increase in frequency and
intensity of heavy precipitation over large parts of Europe (Table 3-3) may
increase the probability of flash floods, which pose the highest risk of
fatality (EEA, 2004). Particularly vulnerable are new urban developments
and tourist facilities, such as camping and recreation areas (e.g., a large
flash flood in 1996 in the Spanish Pyrenees, conveying a large amount
of water and debris to a camping site, resulted in 86 fatalities; Benito et
al., 1998). Apart from new developed urban areas, linear infrastructure,
such as roads, railroads, and underground rails with inadequate drainage,
will probably suffer flood damage (DEFRA, 2004; Arkell and Darch, 2006).
Increased runoff volumes may increase risk of dam failure (small water
reservoirs and tailings dams) with high environmental and socioeconomic
damages as evidenced by historical records (Rico et al., 2008).
In glaciated areas of Europe, glacial lake outburst floods, although
infrequent, have the potential to produce immense socioeconomic and
environmental impacts. Glacial lakes dammed by young, unstable, and
unconsolidated moraines, and lakes in contact with the active ice body
of a glacier, increase the potential of triggering an event (e.g., Huggel et
al., 2004). Intense lake level and dam stability monitoring on most glacial
lakes in Europe helps prevent major breach catastrophes. In case of
flooding, major impacts are expected on infrastructure and settlements
even at long distances downstream from the hazard source area
(Haeberli et al., 2001; Huggel et al., 2004).
4.4.5.7. Landslides
There is a general lack of information on trends in landslide activity, and
for regions with reasonably well-established databases (e.g., Switzerland),
significant trends have not been found in the number of events and
impacts (Hilker et al., 2009). Reactivation of large movements usually
occurs in areas with groundwater flow and river erosion. In southern
Europe the risk is reduced through revegetation on scree slopes, which
enhances cohesion and slope stability coupled with improved hazard
mitigation (Corominas, 2005; Clarke and Rendell, 2006).
4.4.5.8. Snow
Snow avalanches are an ever-present hazard with the potential for loss
of life, property damage, and disruption of transportation. Due to an
increased use of mountainous areas for recreation and tourism, there is
increased exposure for the population leading to an increased rate of
mortality due to snow avalanches. During the period 1983 to 2003,
avalanche fatalities have averaged about 25 per year in Switzerland
(McClung and Schaerer, 2006). In economic terms, direct losses related to
avalanches are small (Voigt et al., 2010), although short-term reactions by
tourists may result in a reduction in overnight stays one year after a
disaster (Nöthiger and Elsasser, 2004). Increased winter precipitation
may result in higher than average snow depth or duration of snow
cover, which could contribute to avalanche formation (Schneebeli et al.,
1997). Climate change impacts on snow cover also include decreases in
its duration, depth, and extent and a possible altitudinal shift of the
snow/rain limit (Beniston et al., 2003), with adverse consequences to
winter tourism. Increased avalanche occurrence would have a negative
impact on humans (loss of life and infrastructure) but could have a
positive result in mountain forests due to higher biodiversity within the
affected areas (Bebi et al., 2009).
4.4.6. North America
4.4.6.1. Introduction
North America (Canada, Mexico, and the United States) is relatively well
developed, although differentiation in living standards exists across and
within countries. This differentiation in adaptive capacity, combined
with a decentralized and essentially reactive response capability,
underlies the region’s vulnerability (Field et al., 2007). Furthermore,
population trends within the region have increased vulnerability by
heightening exposure of people and property in areas that are affected
by extreme events. For example, population in coastline regions of the
Gulf of Mexico region in the United States increased by 150% from
1960 to 2008, while total US population increased by 70% (U.S. Census
Bureau, 2010).
4.4.6.2. Heat Waves
For North America, there is medium confidence in observations (Table 3-2)
and high confidence in projections (Table 3-3) of increasing trends in
heat wave frequency and duration.
Heat waves have impacts on many sectors, most notably on human health,
agriculture, forestry and natural ecosystems, and energy infrastructure.
One of the most significant concerns is human health, in particular
mortality and morbidity. In 2006 in California, at least 140 deaths and
more than 1,000 hospitalizations were recorded during a severe heat
wave (CDHS, 2007; Knowlton et al., 2008). In 1995 in Chicago, more
than 700 people died during a severe heat wave. Following that 1995
event, the city developed a series of response measures through an
extreme heat program.In 1999, the city experienced another extreme
heat event but far fewer lives were lost. While conditions in the 1999
eventwere somewhat less severe, the city’s response measures were
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
259
also credited with contributing to the lower mortality (Palecki et al.,
2001).
While heat waves are projected to increase in intensity and duration
(Table 3-3), their net effect on human health is uncertain, largely because
of uncertainties about the structure of cities in the future, adaptation
measures, and access to cooling (Ebi and Meehl, 2007). Many cities
have installed heat watch warning systems. Several studies show that
the sensitivity of the population of large US cities to extreme heat
events has been declining over time (e.g., Davis et al., 2003; Kalkstein
et al., 2011).
Heat waves have other effects. There is increased likelihood of disruption
of electricity supplies during heat waves (Wilbanks et al., 2008). Air
quality can be reduced, particularly if stagnant high-pressure systems
increase in frequency and intensity (Wang and Angell, 1999).
Additionally, extreme heat can reduce yields of grain crops such as corn
and increase stress on livestock (Karl et al., 2009).
4.4.6.3. Drought and Wildfire
There is medium confidence in an overall slight decrease in dryness
since 1950 across the continent, with regional variability (Table 3-2).
For some regions of North America, there is medium confidence in
projections of increasing dryness (Table 3-3).
Droughts are currently the third most costly category of natural disaster
in the United States (Carter et al., 2008). The effects of drought include
reduced water quantity and quality, lower streamflows, decreased crop
production, ecosystem shifts, and increased wildfire risk. The severity of
impacts of drought is related to the exposure and vulnerability of
affected regions.
From 2000 to 2010, excluding 2003, crop losses accounted for nearly all
direct damages resulting from US droughts (NWS, 2011). Similarly,
drought has had regular recurring impacts on agricultural activities in
Northern Mexico (Endfield and Tejedo, 2006). In addition to impacts on
crops and pastures, droughts have been identified as causes of regional-
scale ecosystem shifts throughout southwestern North America (Allen
and Breshears, 1998; Breshears et al., 2005; Rehfeldt et al., 2006).
Drought also has multiple indirect impacts in North America, although
they are more difficult to quantify. Droughts pose a risk to North
American power supplies due to associated reliance on sufficient water
supplies for hydropower generation and cooling of nuclear, coal, and
natural gas generation facilities (Goldstein, 2003; Wilbanks et al., 2008).
Studies of water availability in heavily contested reservoir systems such
as the Colorado River Basin indicate that climate change is projected to
reduce states’ abilities to meet existing agreements (Christensen et al.,
2004). The effects of climate change on the reliability of the water
supply have been thoroughly explored by Barnett and Pierce (2008,
2009).
Additionally, droughts and dry conditions more generally have been
linked to increases in wildfire activity in North America. Westerling et al.
(2006) found that wildfire activity in the western United States increased
substantially in the late 20th century and that the increase is caused by
higher temperatures and earlier snowmelt. Similarly, increases in wildfire
activity in Alaska from 1950 to 2003 have been linked to increased
temperatures (Karl et al., 2009). Anthropogenic warming was identified
as a contributor to increases in Canadian wildfires (Gillett et al., 2004).
In Canada, forest fires are responsible for one-third of all particulate
emissions, leading to heightened incidence of respiratory and cardiac
illnesses as well as mortality (Rittmaster et al., 2006). Wildfires not only
cause direct mortality, but the air pollution produces increases in eye
and respiratory illnesses (Ebi et al., 2008). The principal economic costs
of wildfires include timber losses, property destruction, fire suppression,
and reductions in the tourism sector (Butry et al., 2001; Morton et al., 2003).
4.4.6.4. Inland Flooding
There has been a likely increase in heavy precipitation in many areas of
North America since 1950 (Table 3-2), with projections suggesting
further increases in heavy precipitation in some regions (Table 3-3).
Flooding and heavy precipitation events have a variety of significant
direct and indirect human health impacts (Ebi et al., 2008). Heavy
precipitation events are strongly correlated with the outbreak of
waterborne illnesses in the United States – 51% of waterborne disease
outbreaks were preceded by precipitation events in the top decile
(Curriero et al., 2001). In addition, heavy precipitation events have been
linked to North American outbreaks of vector-borne diseases such as
Hantavirus and plague (Engelthaler et al., 1999; Parmenter et al., 1999;
Hjelle and Glass, 2000).
Beyond direct destruction of property, flooding has important negative
impacts on a variety of economic sectors including transportation and
agriculture. Heavy precipitation and field flooding in agricultural systems
delays spring planting, increases soil compaction, and causes crop losses
through anoxia and root diseases; variation in precipitation is responsible
for the majority of the crop losses (Mendelsohn, 2007). In 1993, heavy
precipitation flooded 8.2 million acres (~3.3 million ha) of American
Midwest soybean and corn fields, leading to a 50% decrease in corn yields
in Iowa, Minnesota, and Missouri, and a 20 to 30% decrease in Illinois,
Indiana, and Wisconsin (Changnon, 1996). Furthermore, flood impacts
include temporary damage or permanent destruction of infrastructure
for most modes of transportation (Zimmerman and Faris, 2010). For
example, heavy precipitation events are a very costly weather condition
facing US rail transportation (Changnon, 2006).
4.4.6.5. Coastal Storms and Flooding
Global observed and projected changes in coastal storms and flooding
are complex. Since 1950, there has been a likely increase in extreme sea
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
260
level, related to trends in mean sea level. With upward trends in sea
level very likely to continue (Section 3.5.3), there is high confidence that
locations currently experiencing coastal erosion and inundation will
continue to do so in the future (Section 3.5.5).
North America is exposed to coastal storms, and in particular, hurricanes.
2005 was a particularly severe year with 14 hurricanes (out of 27 named
storms) in the Atlantic (NCDC, 2005). There were more than 2,000 deaths
during 2005 (Karl et al., 2009) and widespread destruction on the Gulf
Coast and in New Orleans in particular. Property damages exceeded
US$ 100 billion (Beven et al., 2008; Pielke Jr. et al., 2008). Hurricanes
Katrina and Rita destroyed more than 100 oil and gas platforms in the
Gulf and damaged 558 pipelines, halted all oil and gas production in the
Gulf, and disrupted 20% of US refining capacity (Karl et al., 2009). It is
reported that the direct overall losses of Hurricane Katrina were about
US$ 138 billion in 2007 dollars (Spranger, 2008). However, 2005 may be
an outlier for a variety of reasons – the year saw storms of higher than
average frequency, with greater than average intensity, which made
more frequent landfall, including in the most vulnerable region of the
country (Nordhaus, 2010). The major factor increasing the vulnerability
and exposure of North America to hurricanes is the growth in population
(see, e.g., Pielke Jr. et al., 2008) and increase in property values, particularly
along the Gulf and Atlantic coasts of the United States. While some of
this increase has been offset by adaptation and improved building
codes, Nordhaus (2010) suggests the ratio of hurricane damages to
national GDP has increased by 1.5% per year over the past half-century.
However, the choice of start and end dates influences this figure.
Future sea level rise and potential increases in storm surge could
increase inundation and property damage in coastal areas. Hoffman et
al. (2010) assumed no acceleration in the current rate of sea level rise
through 2030 and found that property damage from hurricanes would
increase by 20%. Frey et al. (2010) simulated the combined effects of
sea level rise and more powerful hurricanes on storm surge in southern
Texas in the 2080s. They found that the area inundated by storm surge
could increase from 6-25% to 60-230% across scenarios evaluated. No
adaptation measures were assumed in either study. Globally, uncertainties
associated with changes in tropical and extratropical cyclones mean
that a general assessment of the projected effects of storminess on
future storm surge is not currently possible (Section 3.5.3).
4.4.7. Oceania
The region of Oceania consists of Australia, New Zealand, and several
small island states that are considered separately in Section 4.4.10.
4.4.7.1. Introduction
Extreme events have severe impacts in both Australia and New Zealand.
In Australia, weather- and climate-related events cause around 87% of
economic damage due to natural disasters (storms, floods, cyclones,
earthquakes, fires, and landslides; BTE, 2001). In New Zealand, floods
and droughts are the most costly climate disasters (Hennessy et al.,
2007). Economic damage from extreme weather is projected to increase
and provide challenges for adaptation (Hennessy et al., 2007).
Observed and projected trends in temperature and precipitation
extremes for the region are extensively covered in Chapter 3 (e.g.,
Tables 3-2 and 3-3). ENSO is a strong driver of climate variability in this
region (see Section 3.4.2).
4.4.7.2. Temperature Extremes
During the Eastern Australian heat wave, in February 2004, temperatures
reached 48.5°C in western New South Wales. About two-thirds of
continental Australia recorded maximum temperatures over 39°C. Due
to heat-related stresses, the Queensland ambulance service recorded a
53% increase in ambulance call-outs (Steffen et al., 2006). A week-long
heat wave in Victoria in 2009 corresponded with a sharp increase in
deaths in the state. For the week of the heat wave a total of 606 deaths
were expected and there were a total of 980 deaths, representing a
62% increase (DHS, 2009).
An increase in heat-related deaths is projected given a warming climate
(Hennessy et al., 2007). In Australian temperate cities, the number of
deaths is projected to more than double in 2020 from 1,115 per year at
present and to increase to between 4,300 and 6,300 per year by 2050
for all emission scenarios, including demographic change (McMichael et
al., 2003b). In Auckland and Christchurch, a total of 14 heat-related deaths
occur per year in people aged over 65, but this number is projected to
rise approximately two-, three-, and six-fold for warming of 1, 2, and 3°C,
respectively (McMichael et al., 2003b). An aging society in Australia and
New Zealand would amplify these figures. For example, it has been
projected that, by 2100, the Australian annual death rate in people aged
over 65 would increase from a 1999 baseline of 82 per 100,000 to a
range of between 131 and 246 per 100,000 in 2100 for the scenarios
examined (SRES B2 and A2, with stabilization of atmospheric CO
2
at
450 ppm; Woodruff et al., 2005). In Australia, cities with a temperate
climate are expected to experience more heat-related deaths than those
with a tropical climate (McMichael et al., 2003b).
4.4.7.3. Droughts
There is a complex pattern of observed and projected changes in dryness
over the region, with increasing dryness in some areas, and decreasing
dryness or inconsistent signals in others (Tables 3-2 and 3-3). However,
several high-impact drought events have been recorded (OCDESC,
2007).
In Australia, the damages due to the droughts of 1982-1983, 1991-1995,
and 2002-2003 were US$ 2.3, 3.8, and 7.6 billion, respectively (Hennessy
et al., 2007). Droughts have a negative impact on water security in the
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
261
Murray-Darling Basin in Australia, as it accounts for most of the water
for irrigated crops and pastures in the country.
New Zealand has a high level of economic dependence on agriculture,
and drought can cause significant disruption for this industry. The
1997-1998 El Niño resulted in severe drought conditions across large areas
of New Zealand with losses estimated at NZ$ 750 million (2006 values)
or 0.9% of GDP (OCDESC, 2007). Severe drought in two consecutive
summers, 2007-2009, affected a large area of New Zealand and caused
on-farm net income to drop by NZ$ 1.9 billion (Butcher and Ford, 2009).
Drought conditions also have a serious impact on electricity production in
New Zealand where around two-thirds of supply is from hydroelectricity
and low precipitation periods result in increased use of fossil fuel for
electricity generation, a maladaptation to climate change. Auckland,
New Zealand’s largest city, suffered from significant water shortages in
the early 1990s, but has since established a pipeline to the Waikato
River to guarantee supply (OCDESC, 2007).
Climate change may cause land use change in southern Australia.
Cropping could become non-viable at the dry margins if rainfall
substantially decreases, even though yield increases from elevated CO
2
partly offset this effect (Luo et al., 2003).
4.4.7.4. Wildfire
Wildfires around Canberra in January 2003 caused AUS$ 400 million
damage (Lavorel and Steffen, 2004), with about 500 houses destroyed,
four people killed, and hundreds injured. Three of the city’s four water
storage reservoirs were contaminated for several months by sediment-
laden runoff (Hennessy et al., 2007). The 2009 fire in the state of Victoria
caused immense damage (see Box 4-1 and Case Study 9.2.2).
An increase in fire danger in Australia is associated with a reduced
interval between fire events, increased fire intensity, a decrease in fire
extinguishments, and faster fire spread (Hennessy et al., 2007). In
southeast Australia, the frequency of very high and extreme fire danger
days is expected to rise 15 to 70% by 2050 (Hennessy et al., 2006). By
the 2080s, the number of days with very high and extreme fire danger
are projected to increase by 10 to 50% in eastern areas of New Zealand,
the Bay of Plenty, Wellington, and Nelson regions (Pearce et al., 2005),
with even higher increases (up to 60%) in some western areas. In both
Australia and New Zealand, the fire season length is expected to be
extended, with the window of opportunity for fuel reduction burning
shifting toward winter (Hennessy et al., 2007).
4.4.7.5. Intense Precipitation and Floods
There has been a likely decrease in heavy precipitation in many parts of
southern Australia and New Zealand (Table 3-2), while there is generally
low to medium confidence in projections due to a lack of consistency
between models (Table 3-3).
Floods are New Zealand’s most frequently experienced hazard (OCDESC,
2007) affecting both agricultural and urban areas. Being long and narrow,
New Zealand is characterized by small river catchments and accordingly
shorter time-to-peak and shorter flood warning times, posing a difficult
preparedness challenge. Projected increases in heavy precipitation
events across most parts of New Zealand (Table 3-3) is expected to
cause greater erosion of land surfaces, more landslides, and a decrease
in the protection afforded by levees (Hennessy et al., 2007).
4.4.7.6. Storm Surges
Over 80% of the Australian population lives in the coastal zone, and
outside of the major capital cities is also where the largest population
growth occurs (Harvey and Caton, 2003; ABS, 2010). Over 500,000
addresses are within 3 km of the coast and less than 5 m above sea
level (Chen and McAneney, 2006). As a result of being so close to sea
level, the risk of inundation from sea level rise and large storm surges
increases with climate change (Hennessy et al., 2007). The risk of a
1-in-100 year storm surge in Cairns is expected to more than double by
2050 (McInnes et al., 2003). Projected changes in coastal hazards from
sea level rise and storm surge are also an issue for New Zealand (e.g.,
Ministry for the Environment, 2008).
4.4.8. Open Oceans
The ocean’s huge mass in comparison to the atmosphere gives it a crucial
role in global heat budgets and chemical budgets. Possible extreme
impacts can be triggered by (1) warming of the surface ocean, with a
major cascade of physical effects, (2) ocean acidification induced by
increases in atmospheric CO
2
, and (3) reduction in oxygen concentration
in the ocean due to a temperature-driven change in gas solubility and
physical impacts from (1). All have potentially nonlinear multiplicative
impacts on biodiversity and ecosystem function, and each may increase
the vulnerability of ocean systems, triggering an extreme impact
(Kaplan et al., 2010; Griffith et al., 2011). Surface warming of the oceans
can itself directly impact biodiversity by slowing or preventing growth in
temperature-sensitive species. One of the most well-known biological
impacts of warming is coral bleaching, but ocean acidification can also
affect coral growth rates (Bongaerts et al., 2010). The seasonal sea ice
cycle affects biological habitats. Such species of Arctic mammals as
polar bears, seals, and walruses depend on sea ice for habitat, hunting,
feeding, and breeding. Declining sea ice can decrease polar bear numbers
(Stirling and Parkinson, 2006).
4.4.9. Polar Regions
4.4.9.1. Introduction
The polar regions consist of the Arctic and the Antarctic, including
associated water bodies. The Arctic region consists of a vast treeless
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
262
permafrost territory (parts of northern Europe, northern Asia, and North
America, and several islands including Greenland). Delimitation of the
Arctic may differ according to different disciplinary and political definitions
(ACIA, 2004). Population density in the polar regions is low, so that
impacts of climate change and extremes on humans may not be as
noticeable as elsewhere throughout the world. The territory of the
Russian Arctic is more populated than other polar regions, hence
impacts of climate change are most noticeable there as they affect
human activities. Specific impacts of climate extremes on the natural
physical environment in polar regions are discussed in Section 3.5.7.
4.4.9.2. Warming Cryosphere
Polar regions have experienced significant warming in recent decades.
Warming has been most pronounced across the Arctic Ocean Basin and
along the Antarctic Peninsula, with significant decreases in the extent
and seasonal duration of sea ice, while in contrast, temperatures over
mainland Antarctica have not warmed over recent decades (Lemke et
al., 2007; Trenberth et al., 2007). Sea ice serves as primary habitat for
marine organisms central to the food webs of these regions. Changes
in the timing and extent of sea ice can impose temporal and spatial
mismatches between energy requirements and food availability for
many higher trophic levels, leading to decreased reproductive success,
lower abundances, and changes in distribution (Moline et al., 2008).
Warming in the Arctic may be leading to a shift of vegetation zones
(e.g., Sturm et al., 2001; Tape et al., 2006; Truong et al., 2006), bringing
wide-ranging impacts and changes in species diversity and distribution.
In the Russian North, the seasonal soil thawing depth has increased
overall over the past four decades (Sherstyukov, 2009). As frozen
ground thaws, many existing buildings, roads, pipelines, airports, and
industrial facilities are destabilized. In the 1990s, the number of
damaged buildings increased by 42 to 90% in comparison with the
1980s in the north of western Siberia (Anisimov and Belolutskaya,
2002; Anisimov and Lavrov, 2004). Arctic infrastructure faces increased
risks of damage due to changes in the cryosphere, particularly the loss
of permafrost and land-fast sea ice (SWIPA, 2011).
An apartment building collapsed in the upper part of the Kolyma River
Basin, and over 300 buildings were severely damaged in Yakutsk as a
result of retreating permafrost (Anisimov and Belolutskaya, 2002;
Anisimov and Lavrov, 2004). Changes in permafrost damage the
foundations of buildings and disrupt the operation of vital infrastructure
in human settlements (Anisimov et al., 2004; see also Case Study 9.2.10).
Transport options and access to resources are altered by differences in
the distribution and seasonal occurrence of snow, water, ice, and
permafrost in the Arctic. This affects both daily living and commercial
activities (SWIPA, 2011).
In conditions of land impassability, frozen rivers are often used as
transport ways. In the conditions of climate warming, rivers freeze later
and melt earlier than before, and the duration of operation of transport
routes to the far north of Russia decreases with the increase in air
temperature in winter and spring (Mirvis, 1999).
Ice cover does not allow ship navigation. Navigation in the Arctic Ocean
is only possible during the ice-free period off the northern coasts of
Eurasia and North America. During periods of low ice concentration,
ships navigate toward ice-free passages, away from multi-year ice that has
accumulated over several years. Regional warming provides favorable
conditions for sea transport going through the Northern Sea Route
along the Eurasian coasts and through the Northwest Passage in the
north of Canada and Alaska (ACIA, 2004).
In September 2007, when the Arctic Sea ice area was extremely low, the
Northwest Passage was opened up. In Russia, this enabled service to
ports of the Arctic region and remote northern regions (import of fuel,
equipment, food, timber, and export of timber, oil, and gas). However,
owing to deglaciation in Greenland, New Land, and Northern Land, the
number of icebergs may increase, creating navigation hazards
(Roshydromet, 2005, 2008; Rignot et al., 2010, Straneo et al., 2010).
4.4.9.3. Floods
From the mid-1960s to the beginning of the 1990s, winter runoff in the
three largest rivers of Siberia (Yenisei, Lena, and Ob; jointly contributing
approximately 70% of the global river runoff to the Arctic Ocean)
increased by 165 km
3
(Savelieva et al., 2004).
Rivers in Arctic Russia experience floods, but their frequency, stage, and
incidence are different across the region, depending on flood formation
conditions. Floods on the Siberian rivers can be produced by a high
peak of the spring flood, by rare heavy rain, or by a combination of
snow and rain, as well as by ice jams, hanging dams, and combinations
of factors (Semyonov and Korshunov, 2006).
Maximum river discharge was found to decrease from the mid-20th
century through 1980 in Western Siberia and the Far East (except for
the Yenisei and the Lena rivers). However, since 1980, maximum
streamflow values began to increase over much of Russia (Semyonov
and Korshunov, 2006).
Snowmelt and rain continue to be the most frequent cause of hazardous
floods on the rivers in the Russian Arctic (85% of all hazardous floods
in the past 15 years). Hazardous floods produced by ice jams and wind
tides make up 10 and 5% of the total number of hazardous floods,
respectively. For the early 21st century, Pomeranets (2005) suggests
that the probability of catastrophic wind tide-related floods and ice
jam-related floods increased. The damage from floods depends not only
on their level, but also on the duration of exposure. On average, a flood
lasts 5 to 10 days, but sometimes high water marks have been recorded
to persist longer, for example, for 20 days or more (Semyonov and
Korshunov, 2006).
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
263
4.4.9.4. Coastal Erosion
Coastal erosion is a significant problem in the Arctic, where coastlines
are highly variable due to environmental forcing (wind, waves, sea level
changes, sea ice, etc.), geology, permafrost, and other elements
(Rachold et al., 2005). For example, the amount of coastal erosion along
a 60 km stretch of Alaska’s Beaufort Sea doubled between 2002 and
2007. Jones et al. (2009) considered contributing factors to be melting
sea ice, increasing summer sea surface temperature, sea level rise, and
increases in storm power and associated stronger ocean waves.
Increasing coastal retreat will have further ramifications for Arctic
landscapes, including losses in freshwater and terrestrial wildlife habitats,
in subsistence grounds for local communities, and in disappearing
cultural sites, as well as adverse impacts on coastal villages and towns.
In addition, oil test wells may be impacted (Jones et al., 2009). Coastal
erosion has also become a problem for residents of Inupiat and on the
island of Sarichev (Russian Federation) (Revich, 2008).
Permafrost degradation along the coast of the Kara Sea may lead to
intensified coastal erosion, driving the coastline back by up to 2 to 4 m
per year (Anisimov and Belolutskaya, 2002; Anisimov and Lavrov, 2004).
Coastline retreat poses considerable risks for coastal population centers
in Yamal and Taymyr and other littoral lowland areas.
4.4.10. Small Island States
Small Island States (SIS) in the Pacific, Indian, and Atlantic Oceans have
been identified as being among the most vulnerable to climate change
and climate extremes (e.g., UNFCCC, 1992; DSD, 1994; UNISDR, 2005).
In the light of current experience and model-based projections, SIS, with
high vulnerability and low adaptive capacity, have substantial future
risks (Mimura et al., 2007). Smallness renders island countries at risk of
high proportionate losses when impacted by a climate extreme (Pelling
and Uitto, 2001; see also Case Study 9.2.9 and Box 3-4).
Sea level rise could lead to a reduction in island size (FitzGerald et al.,
2008). Island infrastructure, including international airports, roads, and
capital cities, tends to predominate in coastal locations (Hess et al.,
2008). Sea level rise exacerbates inundation, erosion, and other coastal
hazards; threatens vital infrastructure, settlements, and facilities; and
thus compromises the socioeconomic well-being of island communities
and states (Hess et al., 2008).
In 2005, regionally averaged temperatures were the warmest in the
western Caribbean for more than 150 years (Eakin et al., 2010). These
extreme temperatures caused the most severe coral bleaching ever
recorded in the Caribbean: more than 80% of the corals surveyed were
bleached, and at many sites more than 40% died. Recovery from such
large-scale coral mortality is influenced by the extent to which coral
reef health has been compromised and the frequency and severity of
subsequent stresses to the system.
Since the early 1950s, when the quality of disaster monitoring and
reporting improved in the Pacific Islands region, there has been a
general increasing trend in the number of disasters reported annually
(Hay and Mimura, 2010).
Pacific Island Countries and Territories (PICs) exhibit a variety of
characteristics rendering generalization difficult (see Table 4-2;
Campbell, 2006). One form of PICs is large inter-plate boundary islands
formed by subduction and found in the southwest Pacific Ocean. These
may be compared to the Oceanic (or intra-plate) islands which were, or
are being, formed over ‘hot spots’ in the Earth’s mantle into volcanic
high islands. Some of these are still being formed and some are heavily
eroded with steep slopes and barrier reefs. Another form of PICs is
atolls that consist of coral built on submerging former volcanic high
islands, through raised limestone islands (former atolls stranded above
contemporary sea levels). Each island type has specific characteristics in
relation to disaster risk reduction, with atolls being particularly vulnerable
to tropical cyclones, where storm surges can completely inundate them
and there is no high ground to which people may escape. In contrast,
the inter-plate islands are characterized by large river systems and fertile
flood plains in addition to deltas, both of which tend to be heavily
populated. Fatalities in many of the worst weather- and climate-related
disasters in the region have been mostly from river flooding (AusAID,
2005). Raised atolls are often saved from the storm surge effects of
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
90,5 x 244,4
Island Type Exposure to climate risks
Plate-Boundary Islands
Large land area
High elevations
High biodiversity
Well-developed soils
River flood plains
Orographic rainfall
These islands are located in the western
Pacific. River flooding is more likely to be a
problem than in other island types. In Papua
New Guinea, high elevations expose areas to
frost (extreme during El Niño). Most major
settlements are on the coast and exposed to
storm damage.
Intra-Plate (Oceanic) Islands
Volcanic High Islands
Steep slopes
Different stages of erosion
Barrier reefs
Relatively small land area
Less well-developed river systems
Orographic rainfall
Because of size these islands have substantial
exposure to tropical cyclones, which cause
the most damage in coastal areas and
catchments. Streams and rivers are subject to
flash flooding. Most islands are exposed to
drought. Barrier reefs may ameliorate storm
surge and tsunamis.
Atolls
Very small land areas
Very low elevations
No or minimal soil
Small islets surround a lagoon
Shore platform on windward side
Larger islets on windward side
No surface (fresh) water
Ghyben Herzberg (freshwater) lens
Convectional rainfall
These islands are exposed to storm surge,
‘king’ tides, and high waves. They are
exposed to freshwater shortages and drought.
Freshwater limitations may lead to health
problems.
Raised Limestone Islands
Steep outer slopes
Concave inner basin
Sharp karst topography
Narrow coastal plains
No surface water
No or minimal soil
Depending on height these islands may be
exposed to storm surges and wave damage
during cyclones and storms. They are exposed
to freshwater shortages and drought.
Freshwater problems may lead to health
problems.
Table 4-2 | Pacific Island type and exposure to climate extremes. Adapted from
Campbell, 2006.
264
tropical cyclones, but during Cyclone Heta that struck Niue in 2004, the
cliffs were unable to provide protection.
Drought is a hazard of considerable importance in SIS. In particular,
atolls have very limited water resources, being dependent on their
freshwater lens, whose thickness decreases with sea level rise (e.g.,
Kundzewicz et al., 2007), floating above sea water in the pervious coral,
and is replenished by convectional rainfall. During drought events,
water shortages in SIS become acute on atolls in particular, resulting in
rationing in some cases (Campbell, 2006).
The main impacts from climatic extremes in PICs are damage to
structures, infrastructure, and crops during tropical cyclones and crop
damage and water supply shortages during drought events. On atolls,
salinization of the freshwater lens and garden areas is a serious problem
following storm surges, high wave events, and ‘king’ tides (Campbell,
2006).
4.5. Costs of Climate Extremes and Disasters
The following section focuses on the economic costs imposed by climate
extremes and disasters on humans, societies, and ecosystems and the costs
of adapting to the impacts. Cost estimates are composed of observed
and projected economic impacts, including economic losses, future
trends in extreme events and disasters in key regions, and the costs of
adaptation. The section stands at the interface between chapters, using
the conceptual framework of Chapters 1 and 2 and the scientific
foundation of Chapter 3 and earlier subsections in this chapter, and
leading into the following Chapters 5 through 9.
4.5.1. Framing the Costs of Extremes and Disasters
The economic costs associated with climate extremes and disasters can be
subdivided into impact or damage costs (or simply losses) and adaptation
costs. Costs arise due to economic, social, and environmental impacts of
a climate extreme or disaster and adaptation to those impacts in key
sectors. Residual damage costs are the impact and damage costs after
all desirable and practical adaptation actions have been implemented.
Conceptually, comparing costs of adaptation with damages before
adaptation and residual damages can help in assessing the economic
efficiency of adaptation (Parry et al., 2009).
The impact of climate extremes and disasters on economies, societies,
and ecosystems can be measured as the damage costs and losses of
economic assets or stocks, as well as consequential indirect effects on
economic flows, such as on GDP or consumption. In line with general
definitions in Chapters 1 and 2, economic disaster risk may be defined
as a probability distribution indicating potential economic damage costs
and associated return periods. The cost categories of direct, indirect, and
intangible are rarely fully exclusive, and items or activities can have
elements in all categories.
Direct damage costs or losses are often defined as those that are a
direct consequence of the weather or climate event (e.g., floods,
windstorms, or droughts). They
refer to the costing of the physical
impacts of climate extremes and disasters – on the lives and health of
directly affected persons; on all types of tangible assets, including
private dwellings, and agricultural, commercial, and industrial stocks and
facilities; on infrastructure (e.g., transport facilities such as roads, bridges,
and ports, energy and water supply lines, and telecommunications); on
public facilities (e.g., hospitals, schools); and on natural resources
(ECLAC, 2003; World Bank, 2010).
Indirect impact costs generally arise due to the disruption of the flows
of goods and services (and therefore economic activity) because of a
disaster, and are sometimes termed consequential or secondary impacts
as the losses typically flow from the direct impact of a climate event
(
ECLAC, 2003; World Bank, 2010). Indirect damages may be caused by
the direct damages to physical infrastructure or sources of livelihoods,
or because reconstruction pulls resources away from production.
Indirect damages include additional costs incurred from the need to use
alternative and potentially inferior means of production and/or distribution
of normal goods and services (
Cavallo and Noy, 2010). For example,
electricity transmission lines may be destroyed by wind, a direct impact,
causing a key source of employment to cease operation, putting many
people out of work, and in turn creating other problems that can be
classified as indirect impacts. These impacts can emerge later in the
affected location, as well as outside the directly affected location (Pelling
et al., 2002; ECLAC, 2003; Cavallo and Noy, 2010). Indirect impacts
include both negative and positive factors – for example, transport
disruption, mental illness or bereavement resulting from disaster shock,
rehabilitation, health costs, and reconstruction and disaster-proof
investment, which can include changes in employment in a disaster-hit
area (due to reconstruction and other recovery activity) or additional
demand for goods produced outside of a disaster-affected area (ECLAC,
2003; World Bank, 2010). As another example, long-running droughts
can induce indirect losses such as local economic decline, out-migration,
famine, the partial collapse of irrigation areas, or loss of livelihoods
dependent on hydroelectricity or rain-fed agriculture. It is important to
note that impacts on the informal or undocumented economy may be
very important in some areas and sectors, but are generally not counted
in reported estimates of losses.
Many impacts, such as loss of human lives, cultural heritage, and
ecosystem services, are difficult to measure as they are not normally
given monetary values or bought and sold, and thus they are also poorly
reflected in estimates of losses. These items are often referred to as
intangibles in contrast to tangibles such as tradable assets, structures,
and infrastructure (
Handmer et al., 2002; Pelling et al., 2002; Benson and
Clay, 2003; ECLAC, 2003; Cavallo and Noy, 2010; World Bank, 2010).
Adaptation costs are those associated with adaptation and facilitation
in terms of planning (e.g., developing appropriate processes including
key stakeholders), actual adaptation (e.g., risk prevention, preparedness,
and risk financing), reactive adaptation (e.g., emergency disaster responses,
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
265
rehabilitation, and reconstruction), and finally the implementation of
adaptation measures (including transition costs) (Smit et al., 2001; also
see the Glossary). The benefits of adaptation can generally be assessed
as the value of avoided impacts and damages as well as the co-benefits
generated by the implementation of adaptation measures (Smit et al.,
2001). The value of all avoidable damage can be taken as the gross (or
theoretically maximum) benefit of adaptation and risk management,
which may be feasible to adapt to but not necessarily economically efficient
(Pearce et al., 1996; Tol, 2001; Parry et al., 2009). The adaptation deficit
is identified as the gap between current and optimal levels of adaptation
to climate change (Burton and May, 2004). However, it is difficult to
assess the optimal adaptation level due to the uncertainties inherent in
climate scenarios, the future patterns of exposure and vulnerability to
climate events, and debates over methodological issues such as discount
rates. In addition, as social values and technologies change, what is
considered avoidable also changes, adding additional uncertainty to
future projections.
4.5.2 Extreme Events, Impacts, and Development
The relationship between socioeconomic development and disasters,
including those triggered by climatic events, has been explored by a
number of researchers over the last few years using statistical
techniques and numerical modeling approaches. It has been suggested
that natural disasters exert adverse impacts on the pace and nature of
economic development (Benson and Clay, 1998, 2003; Kellenberg and
Mobarak, 2008). (The ‘poverty trap’ created by disasters is discussed in
Chapter 8.) A growing literature has emerged that identifies these
important adverse macroeconomic and developmental impacts of natural
disasters (Cuny, 1983; Otero and Marti, 1995; Benson and Clay, 1998,
2000, 2003, 2004; Charveriat, 2000; Crowards, 2000; ECLAC, 2003;
Mechler, 2004; Raddatz, 2009; Noy, 2009; Okuyama and Sahin, 2009;
Cavallo and Noy, 2010). Yet, confidence in the adverse economic
impacts of natural disasters is only medium, as, although the bulk of
studies identify negative effects of disasters on shorter-term economic
growth (up to three years after an event), others find positive effects
(Albala-Bertrand, 1993; Skidmore and Toya, 2002; Caselli and Malhotra,
2004; see Section 4.2). Differences can be partly explained by the lack
of a robust counterfactual in some studies (e.g., what would GDP have
been if a disaster had not occurred?), failure to account for the informal
sector, varying ways of accounting for insurance and aid flows, different
patterns of impacts resulting from, for example, earthquakes versus
floods, and the fact that national accounting does not record the
destruction of assets, but reports relief and reconstruction as additions
to GDP (World Bank and UN, 2010). In terms of longer-run economic
growth (beyond three years after events), there are mixed findings with
the exception of very severe disasters, which have been found to set
back development (World Bank and UN, 2010).
In terms of the nexus between development and disaster vulnerability,
researchers argue that poorer developing countries and smaller
economies are more likely to suffer more from future disasters than
developed countries, especially in relation to extreme impacts
(Hallegatte et al., 2007; Heger et al., 2008; Hallegatte and Dumas, 2009;
Loayza et al., 2009; Raddatz, 2009). In general, the observed or modeled
relationship between development and disaster impacts indicates that
a wealthier country is better equipped to manage the consequences of
extreme events by reducing the risk of impacts and by managing the
impacts when they occur. This is due (inter alia) to higher income levels,
more governance capacity, higher levels of expertise, amassed climate-
proof investments, and improved insurance systems that can act to
transfer costs in space and time (Wildavsky, 1988; Albala-Bertrand,
1993; Burton et al., 1993; Tol and Leek, 1999; Mechler, 2004;
Rasmussen, 2004; Brooks et al., 2005; Kahn, 2005; Toya and Skidmore,
2007; Raschky, 2008; Noy, 2009). While the countries with highest
income account for most of the total economic and insured losses from
disasters (Swiss Re, 2010), in developing countries there are higher
fatality rates and the impacts consume a greater proportion of GDP. This
in turn imposes a greater burden on governments and individuals in
developing countries. For example, during the period from 1970 to 2008
over 95% of deaths from natural disasters occurred in developing
countries (Cavallo and Noy, 2010; CRED, 2010). From 1975 to 2007,
Organisation for Economic Co-operation and Development (OECD)
countries accounted for 71.2% of global total economic losses from
tropical cyclones, but only suffered 0.13% of estimated annual loss of
GDP (UNISDR, 2009).
There is general consensus that, as compared to developed countries,
developing countries are more economically vulnerable to climate
extremes largely because: (i) developing countries have less resilient
economies that depend more on natural capital and climate-sensitive
activities (cropping, fishing, etc.; Parry et al., 2007); (ii) they are often
poorly prepared to deal with the climate variability and physical hazards
they currently face (World Bank, 2000); (iii) more damages are caused
by maladaptation due to the absence of financing, information, and
techniques in risk management, as well as weak governance systems;
(iv) there is generally little consideration of climate-proof investment in
regions with a fast-growing population and asset stocks (such as in coastal
areas) (IPCC, 2001; Nicholls et al., 2008); (v) there is an adaptation deficit
resulting from the low level of economic development (World Bank,
2007) and a lack of ability to transfer costs through insurance and fiscal
mechanisms; and vi) they have large informal sectors. However, in some
cases like Hurricane Katrina in New Orleans, United States, developed
countries also suffer severe disasters because of social vulnerability and
inadequate disaster protection (Birch and Wachter, 2006; Cutter and
Finch, 2008).
While some literature has found that the relationship between income
and some natural disaster consequences is nonlinear (Kellenberg and
Mobarak, 2008; Patt et al., 2010), much empirical evidence supports a
negative relationship between the relative share of GDP and fatalities,
with fatalities from hydrometeorological extreme events falling with
rising level of income (Kahn, 2005; Toya and Skidmore, 2007; World
Bank and UN, 2010). Some emerging developing countries, such as China,
India, and Thailand, are projected to face increased future exposure to
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
266
extremes, especially in highly urbanized areas, as a result of the rapid
urbanization and economic growth in those countries (Bouwer et al.,
2007; Nicholls et al., 2008).
It should also be noted that in a small country, a disaster can directly
affect much of the country and therefore the magnitude of losses and
recovery demands can be extremely high relative to GDP and public
financial resources (Mechler, 2004). This is particularly the case in the
event of multiple and/or consecutive disasters in short periods. For
example, in Fiji, consecutive natural disasters have resulted in reduced
national GDP as well as decreased socioeconomic development as
captured by the Human Development Index (HDI) (Lal, 2010). In Mexico,
natural disasters resulted in the HDI regressing by approximately two
years and in an increase in poverty levels (Rodriguez-Oreggia et al., 2010).
Patt et al. (2010) indicated that vulnerability in the least-developed
countries will rise most quickly, which implies an urgent need for
international assistance.
Costs and impacts not only vary among developing and developed
countries, but also between and within countries, regions, local areas,
sectors, systems, and individuals due to the heterogeneity of vulnerability
and resilience (see Chapter 2). Some individuals, sectors, and systems
would be less affected, or may even benefit, while other individuals,
sectors, and systems may suffer significant losses in the same event.
In general, the poorest and those who are socially or economically
marginalized will be the most at risk in terms of being exposed and
vulnerable (Wisner et al., 2004). For example, women and children are
found to be more vulnerable to disasters in many countries, with larger
disasters having an especially unequal impact (Neumayer and Plümper,
2007).
4.5.3. Methodologies for Evaluating Impact and
Adaptation Costs of Extreme Events and Disasters
4.5.3.1. Methods and Tools for Costing Impacts
Direct, tangible impacts are comparatively easy to measure, but costing
approaches are not necessarily standardized and assessments are often
incomplete, which can make aggregation and comparability across the
literature difficult. In some countries, flood impact assessment has long
been standardized, for example, in Britain and parts of the United States
(e.g., Handmer et al., 2002). Intangible losses can generally be estimated
using valuation techniques such as loss of life/morbidity (usually
estimated using value of statistical life benchmarks), replacement value,
benefits transfer, contingent evaluation, travel cost, hedonic pricing
methods, and so on (there is a vast literature on this subject, e.g.,
Handmer et al., 2002; Carson et al., 2003; Pagiola et al., 2004; Ready
and Navrud, 2006; TEEB, 2009). Yet, assessing the intangible impacts of
extremes and disasters in the social, cultural, and environmental fields
is more difficult, and there is little agreement on methodologies (Albala-
Bertrand, 1993; Tol, 1995; Hall et al., 2003; Huigen and Jens, 2006;
Schmidt et al., 2009).
Studies and reports on the economic impacts of extremes, such as
insurance or post-disaster reports, have mostly focused on direct and
tangible losses, such as on impacts on produced capital and economic
activity. Intangibles such as loss of life and impacts on the natural
environment are generally not considered using monetary metrics (Parry
et al., 2009). Loss of life due to natural disasters, including future
changes, is accounted for in some studies (e.g., BTE, 2001; Jonkman,
2007; Jonkman et al., 2008; Maaskant et al., 2009). Estimates of impacts
that account for tangibles and intangibles are expected to be much
larger than those that consider tangible impacts only (Handmer et al.,
2002; Parry et al., 2009). Potential impacts include all direct, indirect,
and intangible costs, including the losses from public goods and natural
capital (in particular ecosystem services), as well as the longer-term
economic impact of disasters. Indirect impacts and intangible impacts
can outweigh those of direct impacts. There will therefore often be a
large gap between potential impacts and the estimates from studies
that consider only direct impacts.
Indirect economic loss assessment methodologies exist but produce
uncertain and method-dependent results. Such assessments at
national, regional, and global levels fall into two categories: a ‘top
down’ approach that uses models of the whole economy under
study, and a bottom-up or partial equilibrium approach that identifies
and values changes in specific parts of an economy (van der Veen,
2004).
The top-down approach is grounded in macroeconomics under which the
economy is described as an ensemble of interacting economic sectors.
Most studies have focused on impact assessment remodeling actual
events in the past and aim to estimate the various, often hidden follow-
on impacts of disasters (e.g., Ellson et al., 1984; Yezer and Rubin, 1987;
Guimaraes et al., 1993; West and Lenze, 1994; Brookshire et al., 1997;
Hallegatte et al., 2007; Rose 2007). Existing macroeconomic or top-
down approaches utilize a range of models such as the Input-Output,
Social Accounting Matrix multiplier, Computable General Equilibrium
models, economic growth frameworks, and simultaneous-equation
econometric models. These models attempt to capture the impact of the
extreme event as it is felt throughout the whole economy. Only a few
models have aimed at representing extremes in a risk-based framework
in order to assess the potential impacts of events and their probabilities
using a stochastic approach, which is desirable given the fact that extreme
events are non-normally distributed and the tails of the distribution
matter (Freeman et al., 2002; Mechler, 2004; Hallegatte and Ghil, 2007;
Hallegatte, 2008).
The bottom-up approach, derived from microeconomics, scales up data
from sectors at the regional or local level to aggregate an assessment
of disaster costs and impacts (see van der Veen, 2004). The bottom-up
approach to disaster impact assessment attempts to evaluate the impact
of an actual or potential disaster on consumers’ willingness to pay (or
willingness to accept). This approach values direct loss of or damage to
property, as well as that of the interruption to the economy, impacts on
health and well-being, and impacts on environmental amenity and
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
267
ecosystem services. In short, it attempts to value the impact of the
disaster on society.
Overall, measuring the many effects of disasters is problematic, prone to
both overestimation (for example, double counting) and underestimation
(because it is difficult to value loss of life or damage to the environment).
Both over- and underestimation can be issues in different parts of the same
impact assessment, for example, ecological and quality of life impacts may
be ignored, while double counting occurs in the measurement of indirect
impacts. As discussed earlier in this section, most large-scale estimates
leave out significant areas of cost and are therefore underestimates.
Biases also affect the accuracy of estimates; for example, the prospect
of aid may create incentives to inflate losses. How disaster impacts are
evaluated depends on numerous factors, such as the types of impacts
being evaluated, the objective of the evaluation, the spatial and temporal
scale under consideration, and importantly, the information, expertise,
and data available. In practice, the great majority of post-disaster
impact assessments are undertaken pragmatically using whatever data
and expertise are available. Many studies utilize both partial and general
equilibrium analysis in an ‘integrated assessment’ that attempts to capture
both the bottom-up and the economy-wide impacts of disasters (Ciscar,
2009; World Bank, 2010).
4.5.3.2. Methods and Tools for Evaluating the Costs of Adaptation
Over the last few years, a wide range of methodologies using different
metrics, time periods, and assumptions has been developed and applied
for assessing adaptation costs and benefits. However, much of the
literature remains focused on gradual changes such as sea level rise and
effects on agriculture (IPCC, 2007). Extreme events are generally
represented in an ad hoc manner using add-on damage functions based
on averages of past impacts and contingent on gradual temperature
increase (see comment in Nordhaus and Boyer, 2000). In a review of
existing literature, Markandya and Watkiss (2009) identify the following
types of analyses: investment and financial flows; impact assessments
(scenario-based assessments); vulnerability assessments; adaptation
assessments; risk management assessments; economic integrated
assessment models; multi-criteria analysis; computable general
equilibrium models; cost-benefit analysis; cost effectiveness analysis;
and portfolio/real options analysis.
Global and regional assessments of adaptation costs, the focus of this
section, have essentially used two approaches: (1) determining the
pure financial costs, that is, outlays necessary for specific adaptation
interventions (known as investment and financial flow analyses); and
(2) economic costs involving estimating the wider overall costs and
benefits to society and comparing this to mitigation, often using
Integrated Assessment Models (IAMs). The IAM approach leads to a
broader estimate of costs (and benefits) over long time scales, but
requires detailed models of the economies under study (UNFCCC,
2007). One way of measuring the costs of adaptation involves first
establishing a baseline development path (for a country or all countries)
with no climate change, and then altering the baseline to take into
account the impacts of climate change (World Bank, 2010). Then the
potential effects of various adaptation strategies on development or
growth can be examined. Adaptation cost estimates are based on
various assumptions about the baseline scenario and the effectiveness
of adaptation measures. The difference between these assumptions
makes it very difficult to compare or aggregate results (Yohe et al.,
1995, 1996; West et al., 2001).
An example illustrating methodological challenges comes from agriculture,
where estimates have been made using various assumptions about
adaptation behavior (Schneider et al., 2000). These assumptions about
behavior range from the farmers who do not react to observed changes in
climate conditions (especially in studies that use crop yield sensitivity to
weather variability) (Deschênes and Greenstone, 2007; Lobell et al., 2008;
Schlenker and Lobell, 2010), to the introduction of selected adaptation
measures within crop yield models (Rosenzweig and Parry, 1994), to the
assumption of ‘perfect’ adaptation – that is, farmers have complete or
‘perfect’ knowledge and apply that knowledge in ways that ensure
outcomes align exactly with theoretical predictions (Kurukulasuriya and
Mendelsohn, 2008a,b; Seo and Mendelsohn, 2008). Realistic assessments
fall between these extremes, and a realistic representation of future
adaptation patterns depends on the in-due-time detection of the climate
change signal (Schneider et al., 2000; Hallegatte, 2009); the inertia in
adoption of new technologies (Reilly and Schimmelpfennig, 2000); the
existence of price signals (Fankhauser et al., 1999); and assessments of
plausible behavior by farmers.
Cost-benefit analysis (CBA) is an established tool for determining the
economic efficiency of development interventions. CBA compares the
costs of conducting such projects with their benefits and calculates the
net benefits or economic efficiency (Benson and Twigg, 2004). Ideally CBA
accounts for all costs and benefits to society including environmental
impacts, not just financial impacts on individual businesses. All costs
and benefits are monetized so that tradeoffs can be compared with a
common measure. The fact that intangibles and other items that are
difficult to value are often left out is one of the major criticisms of the
approach (Gowdy, 2007). In the case of disaster risk reduction (DRR)
and adaptation interventions, CBA weighs the costs of the DRR project
against the disaster damage costs avoided. While the benefits created
by development interventions are the additional benefits due to, for
example, improvements in physical or social infrastructure, in DRR the
benefits are mostly the avoided or reduced potential damages and losses
(Smyth et al., 2004). The net benefit can be calculated in terms of net
present value, the rate of return, or the benefit-cost ratio. OECD
countries such as the United Kingdom and the United States, as well as
international financial institutions such as the World Bank, Asian
Development Bank, and Inter-American Development Bank, have used
CBA for evaluating disaster risk management (DRM) in the context of
development assistance (Venton and Venton, 2004; Ghesquiere et al.,
2006) and use it routinely for assessing engineering DRM strategies
domestically. CBA can be, and has been, applied at any level from the
global to local (see Kramer, 1995; Benson and Twigg, 2004; Venton and
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
268
Venton, 2004; UNFCCC, 2007; Mechler, 2008). Because the chance of
occurrence of a disaster event can be expressed as a probability, it
follows that the benefits of reducing the impact of that event can be
expressed in probabilistic terms. Costs and benefits should be calculated
by multiplying probability by consequences; this leads to risk estimates
that account for hazard intensity and frequency, vulnerability, and
exposure (Smyth et al., 2004; Ghesquiere et al., 2006).
National-level studies of adaptation effectiveness in the European
Union, the United Kingdom, Finland, and The Netherlands, as well as in
a larger number of developing countries using the National Adaptation
Programme of Action approach, have been conducted or are underway
(Ministry of Agriculture and Forestry, 2005; DEFRA, 2006; Lemmen et al.,
2008; de Bruin et al., 2009; Parry et al., 2009). Yet the evidence base on
the economic aspects including economic efficiency of adaptation
remains limited and fragmented (Adger et al., 2007; Moench et al.,
2007; Agrawala and Fankhauser, 2008; Parry et al., 2009). As noted at
the start of Section 4.5.3.2, many adaptation studies focus on gradual
change, especially for agriculture. Those studies considering extreme
events, and finding or reporting net benefits over a number of key
options (Agrawala and Fankhauser, 2008; Parry et al., 2009), do so by
treating extreme events similarly to gradual onset phenomena and
using deterministic impact metrics, which is problematic for disaster
risk. A recent, risk-focused study (ECA, 2009) concentrating on national
and sub-national levels went so far as to suggest an adaptation cost
curve, which organizes relevant adaptation options around their cost-
benefit ratios. However, given available data including future projections
of risk and the effectiveness of options, this is probably at most heuristic
rather than a basis for policy.
There are several complexities and uncertainties inherent in the estimates
required for a CBA of DRR. As these are compounded by climate change,
CBAs utility in evaluating adaptation may be reduced. These include
difficulties in handling intangibles and, as is particularly important for
extremes, in the discounting of future impacts; CBA does not account
for the distribution of costs and benefits or the associated equity issues.
Moench et al. (2007) argue that CBA is most useful as a decision support
tool that helps the policymaker categorize, organize, assess, and present
information on the costs and benefits of a potential project, rather than
give a definite answer. Overall, the applicability of rigorous CBAs for
evaluations of adaptation is thus limited based on limited evidence and
medium agreement.
4.5.3.3. Attribution of Impacts to Climate Change:
Observations and Limitations
Attribution of the impacts of climate change can be defined and used in a
way that parallels the well-developed applications for the physical climate
system (IPCC, 2010). Detection is the process of demonstrating that a
system affected by climate has changed in some defined statistical
sense, without providing a reason for that change. Attribution is
the process of establishing the most probable causes, natural or
anthropogenic, for the detected change with some defined level of
confidence.
The IPCC Working Group II Fourth Assessment Report found, with very
high confidence, that observational evidence shows that biological
systems on all continents and in most oceans are already being affected
by recent climate changes, particularly regional temperature increases
(Rosenzweig et al., 2007).
Attribution of changes in individual weather and climate events to
anthropogenic forcing is complicated because any such event might
have occurred by chance in an unmodified climate as a result of natural
climate variability (see FAQ 3.2). An approach that addresses this problem
is to look at the likelihood of such an event occurring, rather than the
occurrence of the event itself (Stone and Allen, 2005). For example,
human-induced changes in mean temperature have been shown to
increase the likelihood of extreme heat waves (Meehl and Tebaldi, 2004;
Stott et al., 2004). For a large region of continental Europe, Stott et al.
(2004) showed that anthropogenic climate change very likely doubled
the probability of surpassing a mean summer temperature not exceeded
since advent of the instrumental record in 1851, but which was by the
2003 event in Europe. More recent work provides further support for
such a linkage (Barriopedro et al., 2011; see Section 3.3.1).
Most published studies on the attribution of impacts of extremes to
natural and anthropogenic climate change have focused on long-term
records of disaster losses, or examine the likelihood of the event occurring.
Most published effort has gone into the analysis of long-term disaster
loss records.
There is high confidence, based on high agreement and medium
evidence, that economic losses from weather- and climate-related
disasters have increased (Cutter and Emrich, 2005; Peduzzi et al., 2009,
2011; UNISDR, 2009; Mechler and Kundzewicz, 2010; Swiss Re 2010;
Munich Re, 2011). A key question concerns whether trends in such
losses, or losses from specific events, can be attributed to climate
change. In this context, changes in losses over time need to be
controlled for exposure and vulnerability. Most studies of long-term
disaster loss records attribute these increases in losses to increasing
exposure of people and assets in at-risk areas (Miller et al., 2008;
Bouwer, 2011), and to underlying societal trends – demographic,
economic, political, and social – that shape vulnerability to impacts
(Pielke Jr. et al., 2005; Bouwer et al., 2007). Some authors suggest that
a (natural or anthropogenic) climate change signal can be found in the
records of disaster losses (e.g., Mills, 2005; Höppe and Grimm, 2009),
but their work is in the nature of reviews and commentary rather than
empirical research. Attempts have been made to normalize loss records
for changes in exposure and wealth. There is medium evidence and high
agreement that long-term trends in normalized losses have not been
attributed to natural or anthropogenic climate change (Choi and Fisher,
2003; Crompton and McAneney, 2008; Miller et al., 2008; Neumayer
and Barthel, 2011). The evidence is medium because of the issues set
out toward the end of this section.
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
269
The statement about the absence of trends in impacts attributable
to natural or anthropogenic climate change holds for tropical and
extratropical storms and tornados (Boruff et al., 2003; Pielke Jr. et al.,
2003, 2008; Raghavan and Rajesh, 2003; Miller et al 2008; Schmidt et al.,
2009; Zhang et al., 2009; see also Box 4-2). Most studies related increases
found in normalized hurricane losses in the United States since the
1970s (Miller et al., 2008; Schmidt et al., 2009; Nordhaus, 2010) to the
natural variability observed since that time (Miller et al., 2008; Pielke Jr. et
al., 2008). Bouwer and Botzen (2011) demonstrated that other normalized
records of total economic and insured losses for the same series of
hurricanes exhibit no significant trends in losses since 1900.
The absence of an attributable climate change signal in losses also holds
for flood losses (Pielke Jr. and Downton, 2000; Downton et al., 2005;
Barredo, 2009; Hilker et al., 2009), although some studies did find recent
increases in flood losses related in part to changes in intense rainfall
events (Fengqing et al., 2005; Chang et al., 2009). For precipitation-
related events (intense rainfall, hail, and flash floods), the picture is
more diverse. Some studies suggest an increase in damages related to
a changing incidence in extreme precipitation (Changnon, 2001, 2009),
although no trends were found for normalized losses from flash floods
and landslides in Switzerland (Hilker et al., 2009). Similarly, a study of
normalized damages from bushfires in Australia also shows that increases
are due to increasing exposure and wealth (Crompton et al., 2010).
Increasing exposure of people and economic assets has been the major
cause of long-term increases in economic losses from weather- and
climate-related disasters (high confidence). The attribution of economic
disaster losses is subject to a number of limitations in studies to date:
data availability (most data are available for standard economic sectors
in developed countries); type of hazards studied (most studies focus on
cyclones, where confidence in observed trends and attribution of
changes to human influence is low; Section 3.4.4); and the processes
used to normalize loss data over time. Different studies use different
approaches to normalization, and most normalization approaches take
account of changes in exposure of people and assets, but use only limited,
if any, measures of vulnerability trends, which is questionable. Different
approaches are also used to handle variations in the quality and
completeness of data on impacts over time. Finding a trend or ‘signal’
in a system characterized by large variability or ‘noise’ is difficult and
requires lengthy records. These are all areas of potential weakness in
the methods and conclusions of longitudinal loss studies and more
empirical and conceptual efforts are needed. Nevertheless, the results of
the studies mentioned above are strengthened as they show similar
results, although they have applied different data sets and methodologies.
A general area of uncertainty in the studies concerns the impacts of
weather and climate events on the livelihoods and people of informal
settlements and economic sectors, especially in developing countries.
Some one billion people live in informal settlements (UNISDR, 2011),
and over half the economy in some developing countries is informal
(Schneider et al., 2010). These impacts have not been systematically
documented, with the result that they are largely excluded from both
longitudinal impact analysis and attribution to defined weather
episodes.
Another general area of uncertainty comes from confounding factors
that can be identified but are difficult to quantify, and relates to the
usual assumption of constant vulnerability in studies of loss trends.
These include factors that would be expected to increase resilience
(Chapters 2 and 5 of this report) and thereby mask the influence of
climate change, and those that could act to increase the impact of
climate change. Those that could mask the effects of change include
gradual improvements in warnings and emergency management (Adger
et al., 2005), building regulations (Crichton, 2007), and changing
lifestyles (such as the use of air conditioning), and the almost instant
media coverage of any major weather extreme that may help reduce
losses. In the other direction are changes that may be increasing risk,
such as the movement of people in many countries to coastal areas
prone to cyclones (Pompe and Rinehart, 2008) and sea level rise.
4.5.4. Assessment of Impact Costs
Much work has been conducted on the analysis of direct economic losses
from natural disasters. The examples mentioned below mainly focus on
national and regional economic losses from particular climate extremes
and disasters, and also discuss uncertainty issues related to the assessment
of economic impacts.
4.5.4.1. Estimates of Global and Regional Costs of Disasters
Observed trends in extreme impacts: Data on global weather- and
climate-related disaster losses reported since the 1960s reflect mainly
monetized direct damages to assets, and are unequally distributed.
Estimates of annual losses have ranged since 1980 from a few US$
billion to above 200 billion (in 2010 dollars) for 2005 (the year of
Hurricane Katrina) (UNISDR, 2009; Swiss Re 2010; Munich Re, 2011).
These estimates do not include indirect and intangible losses.
On a global scale, annual material damage from large weather and
climate events has been found to have increased eight-fold between the
1960s and the 1990s, while the insured damage has been found to have
increased by 17-fold in the same interval, in inflation-adjusted monetary
units (Mechler and Kundzewicz, 2010). Between 1980 and 2004, the
total costs of extreme weather events totaled US$ 1.4 trillion, of which
only one-quarter was insured (Mills, 2005). Material damages caused by
natural disasters, mostly weather- and water-related, have increased
more rapidly than population or economic growth, so that these factors
alone may not fully explain the observed increase in damage. The loss
of life has been brought down considerably (Mills, 2005; UNISDR, 2011).
Developing regions are vulnerable both because of exposure to
weather- and climate-related extremes and their status as developing
economies. However, disaster impacts are unevenly distributed by type of
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
270
disaster, region, country, and the exposure and vulnerability of different
communities and sectors.
Percentage of direct economic losses by regions: The concentration of
information on disaster risk generally is skewed toward developed
countries and the Northern Hemisphere (World Bank and UN, 2010).
Some global databases, however, do allow a regional breakdown of
disaster impacts. The unequal distribution of the human impact of natural
disasters is reflected in the number of disasters and losses across regions
(Figure 4-7). In the period 2000 to 2008, Asia experienced the highest
number of weather- and climate-related disasters. The Americas suffered
the most economic loss, accounting for the highest proportion (54.6%)
of total loss, followed by Asia (27.5%) and Europe (15.9%). Africa
accounted for only 0.6% of global economic losses, but economic damages
from natural disasters are underreported in these data compared to
other regions (Vos et al., 2010). Although reporting biases exist, they are
judged to provide robust evidence of the regional distribution of the
number of disasters and of direct economic losses for this recent period
2000 to 2008, and there is high agreement regarding this distribution
among different databases collected by independent organizations
(Guha-Sapir et al., 2011; Munich Re, 2011; Swiss Re, 2011).
Damage losses in percentage of GDP by regions: The relative economic
burden in terms of direct loss expressed as a percentage of GDP has been
substantially higher for developing states. Middle-income countries
with rapidly expanding asset bases have borne the largest burden,
where during the period from 2001 to 2006 losses amounted to about
1% of GDP, while this ratio has been about 0.3% of GDP for low-income
countries and less than 0.1% of GDP for high-income countries, based
on limited evidence (Cummins and Mahul, 2009). In small exposed
countries, particularly small island developing states, these wealth losses
expressed as a percentage of GDP and averaged over both disaster and
non-disaster years can be considerably higher, exceeding 1% in many
cases and 8% in the most extreme cases over the period from 1970 to
2010 (World Bank and UN, 2010), and individual events may consume
more than the annual GDP (McKenzie et al., 2005). This indicates a far
higher vulnerability of the economic infrastructure in developing
countries (Cavallo and Noy 2009; UNISDR, 2009).
Increasing weather- and climate-related disasters: The number of reported
weather- and climate-related disasters and their direct financial costs
have increased over the past decades. Figure 4-8 illustrates an increasing
trend (coupled with large interannual variability) in losses based on
data for large weather-and climate-related disasters over the period
1980 to 2010, for which data have been gathered consistently and
systematically (see Neumayer and Barthel, 2011).
This increase in affected population and direct economic losses is also
coupled with the increasing numbers of reported weather- and climate-
related disasters (UNISDR, 2009; Munich Re, 2011; Swiss Re 2011).
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
0.5
9
18
60
55
22.82
13
13.17
45.28
87
1
8
13
136
17
32
58
13
48
1.19
AMERICAS
EUROPE
AFRICA
ASIA
OCEANIA
Number of disasters
Meteorological
Climatological
Hydrological
Damages
Height of columns represents the number of disasters or damages in billion dollars.
Figure 4-7 | Weather- and climate-related disaster occurrence and regional average impacts from 2000 to 2008. The number of climatological (e.g., extreme temperature, drought,
wildfire), meteorological (e.g., storm), and hydrological (e.g., flood, landslides) disasters is given for each region, along with damages (2009 US$ billion). Data from Vos et al., 2010.
271
These statistics imply the increasing cost of such disasters to society,
regardless of cause. It is also important to note that the number of
weather- and climate-related disasters has increased more rapidly than
losses from non-weather disasters (Mills, 2005; Munich Re, 2011; Swiss
Re, 2011). This could indicate a change in climate extremes, but there
are other possible explanations (Bouwer, 2011). Drought and flood
losses may have grown due to a number of non-climatic factors, such as
increasing water withdrawals effectively exacerbating the impact of
droughts, decrease in storage capacity in catchments (urbanization,
deforestation, sealing surfaces, channelization) adversely affecting both
flood and drought preparedness, increase in runoff coefficients, and
growing settlements in floodplains around urban areas (see Section
4.2.2; Field et al., 2009).
4.5.4.2. Potential Trends in Key Extreme Impacts
As indicated in Sections 3.3 to 3.5 and Tables 3-1 and 3-3, climate
extremes may have different trends in the future; some such as heat
waves are projected to increase over most areas in length, frequency,
and intensity, while projected changes in some other extremes are given
with less confidence. However, uncertainty is a key aspect of disaster/
climate change trend analysis due to attribution issues discussed above,
incomparability of methods, changes in exposure and vulnerability over
time, and other non-climatic factors such as mitigation and adaptation.
A challenge is ensuring that the projections of losses from future
changes in extreme events are examined not for current populations
and economies, but for scenarios of possible future socioeconomic
development. See Box 4-2 for a discussion of this with respect to
cyclones.
It is more likely than not that the frequency of the most intense tropical
cyclones will increase substantially in some ocean basins (Section 3.4.4).
Many studies have investigated impacts from tropical cyclones (e.g.,
ABI, 2005a, 2009; Hallegatte, 2007; Pielke Jr., 2007; Narita et al., 2009;
Bender et al., 2010; Nordhaus, 2010; Crompton et al., 2011). Table 4-3
presents the projected percentage increase in direct economic losses from
tropical cyclones from a number of these studies, scaled to the year 2040
relative to a common baseline (year 2000). There is high confidence that
increases in exposure will result in higher direct economic losses from
tropical cyclones and that losses will also depend on future changes in
tropical cyclone frequency and intensity. One study, building on global
climate model results from Bender et al. (2010), found that to attribute
increased losses to increased tropical cyclone activity in the United
States with a high degree of certainty would take another 260 years of
records, due to the high natural variability of storms and their impacts
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
0
50
100
150
200
250
US$ billions
Overall Losses in 2010 Values
Of Which Insured in 2010 Values
0
0
0
0
0
O
v
e
r
all
L
osses
in 2
0
1
0
V
alues
Of Which Insured in 2010 Values
2010200520001995199019851980
Figure 4-8 | The overall losses and insured losses from weather- and climate-related disasters worldwide (in 2010 US$). These data for weather- and climate-related ‘great’ and
‘devastating’ natural catastrophes are plotted without inclusion of losses from geophysical events. A catastrophe in this data set is considered ‘great’ if the number of fatalities
exceeds 2,000, the number of homeless exceeds 200,000, the country’s GDP is severely hit, and/or the country is dependent on international aid. A catastrophe is considered
‘devastating’ if the number of fatalities exceeds 500 and/or the overall loss exceeds US$ 650 million (in 2010 values). Data from Munich Re, 2011.
272
(Crompton et al., 2011). See Section 4.5.3.3 on attribution and the use
of a risk-based approach to cope with this issue. Other studies have
investigated impacts from increases in the frequency and intensity of
extratropical cyclones at high latitudes (Dorland et al., 1999; ABI,
2005a, 2009; Narita et al., 2010; Schwierz et al., 2010; Donat et al.,
2011). In general there is medium confidence that increases in losses
due to extratropical cyclones will occur with climate change, with
possible decreases or no change in some areas. Projected increases
generally are slightly lower than increases in tropical cyclone losses (see
Table 4-3). Patt et al. (2010) projected future losses due to weather- and
climate-related extremes in least-developed countries.
Many studies have addressed future economic losses from river floods,
most of which are focused on Europe, including the United Kingdom
(Hall et al., 2003, 2005; ABI, 2009), Spain (Feyen et al., 2009), and The
Netherlands (Bouwer et al., 2010) (see Table 4-3). Maaskant et al.
(2009) is one of the few studies that addresses future loss of life from
flooding, and projects up to a four-fold increase in potential flood
victims in The Netherlands by the year 2040, when population growth
is accounted for. Some studies are available on future coastal flood risks
(Hall et al., 2005; Mokrech et al., 2008; Nicholls et al., 2008; Dawson et
al., 2009; Hallegatte et al., 2010). Although future flood losses in many
locations will increase in the absence of additional protection measures
(high agreement, medium evidence), the size of the estimated change is
highly variable, depending on location, climate scenarios used, and
methods used to assess impacts on river flow and flood occurrence (see
Table 4-3 for a comparison of some regional studies) (Bouwer, 2010).
Some studies have addressed economic losses from other types of
weather extremes, often smaller-scale compared to river floods and
cyclones. These include hail damage, for which mixed results are found:
McMaster (1999) and Niall and Walsh (2005) found no significant
effect on hailstorm losses for Australia, while Botzen et al. (2010) find a
significant increase (up to 200% by 2050) for damages in the agricultural
sector in The Netherlands, although the approaches used vary
considerably. Rosenzweig et al. (2002) report on a possible doubling
of losses to crops due to excess soil moisture caused by more intense
rainfall. Hoes (2007), Hoes and Schuurmans (2006), and Hoes et al.
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
Pielke (2007) Tropical storm Atlantic 58 1,365 417
30
Nordhaus (2010) Tropical storm United States 12 92 47
Narita et al. (2009) Tropical storm Global 23 130 46
Hallegatte (2007) Tropical storm United States - - 22
ABI (2005a,b) Tropical storm United States, Caribbean 19 46 32
ABI (2005a,b) Tropical storm Japan 20 45 30
ABI (2009) Tropical storm China 9 19 14
Schmidt et al. (2009) Tropical storm United States - - 9
Bender et al. (2010) Tropical storm United States -27 36 14
Narita et al. (2010) Extra-tropical storm High latitude -11 62 22
15
Schwierz et al. (2010) Extra-tropical storm Europe 6 25 16
Leckebusch et al. (2007) Extra-tropical storm United Kingdom, Germany -6 32 11
ABI (2005a,b) Extra-tropical storm Europe - - 14
ABI (2009) Extra-tropical storm United Kingdom -33 67 15
Dorland et al. (1999) Extra-tropical storm Netherlands 80 160 120
Bouwer et al. (2010) River flooding Netherlands 46 201 124
65
Feyen et al. (2009) River flooding Europe - - 83
ABI (2009) River flooding United Kingdom 3 11 7
Feyen et al. (2009) River flooding Spain (Madrid) - - 36
Schreider et al. (2000) Local flooding Australia 67 514 361
Hoes (2007) Local flooding Netherlands 16 70 47
Pielke (2007) Tropical storm Atlantic 164 545 355
Schmidt et al. (2009) Tropical storm United States - - 240
Dorland et al. (1999) Extra-tropical storm Netherlands 12 93 50
Bouwer et al. (2010) River flooding Netherlands 35 172 104
Feyen et al. (2009) River flooding Spain (Mad) - - 349
Hoes (2007) Local flooding Netherlands -4 72 29
172
A. Impact of projected climate change
B. Impact of projected exposure change
Study
Hazard type
Region
Min
Max
Mean Median
Estimated loss change [%] in 2040
Study
Hazard type
Region
Min
Max
Mean Median
Estimated loss change [%] in 2040
Table 4-3 | Estimated change in disaster losses in 2040 under projected climate change and exposure change, relative to 2000, from 21 impact studies including median estimates
by type of weather hazard. Source: Bouwer, 2010.
273
(2005) estimated increases in damages due to extreme rainfall in The
Netherlands by mid-century.
It is well known that the frequency and intensity of extreme weather
and climate events are only one factor that affects risks, as changes in
population, exposure of people and assets, and vulnerability determine
loss potentials (see Sections 4.2 to 4.4). Few studies have specifically
quantified these factors. However, the ones that do generally underline
the important role of projected changes (increases) in population and
capital at risk. Some studies indicate that the expected changes in
exposure are much larger than the effects of climate change (see Table
4-3), which is particularly true for tropical and extratropical storms
(Pielke Jr., 2007; Feyen et al., 2009; Schmidt et al., 2009). Other studies
show that the effect of increasing exposure is about as large as the effect
of climate change (Hall et al., 2003; Maaskant et al., 2009; Bouwer et al.,
2010), or estimate that these are generally smaller (Dorland et al., 1999;
Hoes, 2007). There is therefore medium confidence that, for some climate
extremes in many regions, the main driver for future increasing losses
in many regions will be socioeconomic in nature (based on medium
agreement and limited evidence). Finally, many studies underline that
both factors need to be taken into account, as the factors do in fact
amplify each other, and therefore need to be studied jointly when
expected losses from climate change are concerned (Hall et al., 2003;
Bouwer et al., 2007, 2010; Pielke Jr., 2007; Feyen et al., 2009).
4.5.5. Assessment of Adaptation Costs
The World Bank (2006) estimated the cost of climate-proofing foreign
direct investments, gross domestic investments, and Official Development
Assistance, which was taken up and modified by Stern (2007), Oxfam
(2007), and UNDP (2007). The second source of adaptation cost estimates
is UNFCCC (2007), which calculated the value of existing and planned
investment and financial flows required for the international community
to effectively and appropriately respond to climate change impacts. The
third source is World Bank (2010), which also conducted a number of
country-level studies to complement the global assessment, following
UNFCCC (2007), but aimed at improving upon this by assessing the
climate-proofing of existing and new infrastructure, using more precise
unit cost estimates and including the costs of maintenance as well as
those of port upgrading and the risks from sea level rise and storm
surges. Also, the investment in education necessary to neutralize
impacts of extreme weather is calculated. Estimates of costs to adapt to
climate change (rather than simply to extremes and disasters), which
have mostly been made for developing countries, exhibit a large range
and relate to different assessment periods (such as today, 2015, or 2030).
For 2030, the estimated global cost from UNFCCC (2007) ranges from
US$ 48 to 171 billion per year for developed and developing countries,
and US$ 28 to 67 billion per year for developing countries (in 2005
dollars). Recent estimates from World Bank (2010) for developing
countries lead to higher projected costs and broadly amount to the
average of this range with annual costs of up to US$ 100 billion (in
2005 US$) (see Table 4-4). Confidence in individual global estimates is
low because, as mentioned above and discussed by Parry et al. (2009),
the estimates are derived from only three relatively independent studies,
which explains the seeming convergence of the estimates in latter
studies. As well, Parry et al. (2009) consider the estimates a significant
underestimation by at least a factor of two to three and possibly more
if the costs incurred by other sectors were included, such as ecosystem
services, energy, manufacturing, retailing, and tourism. The adaptation
cost estimates are also based mostly on low levels of investment due to
an existing adaptation deficit in many regions. Unavoidable residual
damages remain absent from these analyses.
In terms of regional costs and as reported in the World Bank (2010)
study, the largest absolute adaptation costs would arise in East Asia and
the Pacific, followed by the Latin American and Caribbean region as well
as sub-Saharan Africa. This pattern held for the two scenarios assessed
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
Study
Results
(billion US$ yr
-1
)
Time Frame and
Coverage
Sectors Methodology and Comment
World Bank (2006)
9–41
1
Present, developing
countries
Unspecified
Cost of climate-proofing foreign direct investments, gross domestic
investments, and Official Development Assistance
Stern (2007) 4–37
1
Present, developing
countries
Unspecified Update of World Bank (2006)
Oxfam (2007) >50
1
Present, developing
countries
Unspecified
World Bank (2006) plus extrapolation of cost estimates from National
Adaptation Programmes of Action and NGO projects
UNDP (2007) 86–109
2
In 2015, developing
countries
Unspecified
World Bank (2006) plus costing of targets for adapting poverty reduction
programs and strengthening disaster response systems
UNFCCC (2007)
48–171
(28–67 for
developing
countries)
2
In 2030, developed
and developing
countries
Agriculture, forestry, and fisheries; water
supply; human health; coastal zones;
infrastructure; ecosystems (but no estimate
for 2030 for ecosystem adaptation)
Additional investment and financial flows needed for adaptation in 2030
World Bank (2010)
70–100
2
Annual from 2010
to 2050, developing
countries
Agriculture, forestry, and fisheries; water
supply and flood protection; human health;
coastal zones; infrastructure; extreme
weather events
Impact costs linked to adaptation costs, improvement upon UNFCCC
(2007): climate-proofing existing and new infrastructure, more precise
unit cost, inclusion of cost of maintenance and port upgrading, risks
from sea level rise and storm surges, riverine flood protection, education
investment to neutralize impacts of extreme weather events
Notes: 1. in 2000 US$; 2. in 2005 US$.
Table 4-4 | Estimates of global costs of adaptation to climate change. Source: Extended based on Agrawala and Fankhauser (2008) and Parry et al. (2009).
274
in the study, which were a scenario with the most precipitation (‘wet’)
and one with the least precipitation (‘dry’) among all scenarios chosen
for the study, which employ socioeconomic driver information from
IPCC’s SRES A2 scenario (see Table 4-5).
Taking Africa as an example, based on various estimates the potential
additional costs of adaptation investment range from US$ 3 to 10 billion
per year by 2030 (UNFCCC, 2007; PACJA, 2009). However, this could be
also an underestimate considering the desirability of improving Africa’s
resilience to climate extremes as well as the flows of international
humanitarian aid in the aftermath of disasters.
4.5.6. Uncertainty in Assessing the
Economic Costs of Extremes and Disasters
Upon reviewing the estimates to date, the costing of weather- and
climate-related disasters and estimating adaptation costs is still
preliminary, incomplete, and subject to a number of assumptions with
the result that there is considerable uncertainty (Agrawala and
Fankhauser, 2008; Parry et al., 2009). This is largely due to modeling
uncertainties in climate change and damage estimates, limited data
availability, and methodological shortcomings in analyzing disaster
damage statistics. Such costing is further limited by the interaction
between numerous adaptation options and assumptions about future
exposure and vulnerabilities, social preferences, and technology, as well
as levels of resilience in specific societies. Additionally the following
challenges can be identified.
Risk assessment methods: Technical challenges remain in developing
robust risk assessment and damage costing methods. Study results can
vary significantly between top-down and bottom-up approaches. Risk-
based approaches are utilized for assessing and projecting disaster risk
(Jones, 2004; Carter et al., 2007), for which input from both climate and
social scenarios is required. All climatic phenomena are subject to the
limitation that historically based relationships between damages and
disasters cannot be used with confidence to deduce future risk of
extreme events under changing characteristics of frequency and intensity
(UNDP, 2004). Yet climate models are today challenged when reproducing
spatially explicit climate extremes, due to coarse resolution and physical
understanding of the relevant process, as well as challenges in modeling
low-probability, high-impact events (see Section 3.2.3). Therefore,
projections of future extreme event risk involve uncertainties that can
limit understanding of sudden onset risk, such as flood risk. Future
socioeconomic development is also inherently uncertain. A uniform set
of assumptions can help to provide a coherent global picture and
comparison and extrapolation between regions.
Data availability and consistency: Lack of data and robust information
increases the uncertainty of costing when scaling up to global levels from
a very limited (and often very local) evidence base. There are double-
counting problems and issues of incompatibility between types of
impacts in the process of multi-sectoral and cross-scale analyses,
especially for the efforts to add both market and non-market values (e.g.,
ecosystem services) (Downton and Pielke Jr., 2005; Pielke Jr. et al., 2008;
Parry et al., 2009). Moreover the full impacts of weather- and climate-
related extremes in developing countries are not fully understood, and
a lack of comprehensive studies on damage, adaptation, and residual
costs indicates that the full costs are underestimated.
Information on future vulnerability: Apart from climate change,
vulnerability and exposure will also change over time, and the
interaction of these aspects should be considered (see, e.g., Hallegatte,
2008; Hochrainer and Mechler, 2011). This has been recognized and
assessments of climate change impacts, vulnerability, and risk are
changing in focus, leading to more integration across questions. While
initial studies focused on an analysis of the problem, the field proceeded
to assess potential impacts and risks, and now more recently started to
combine such assessments with the consideration of specific risk
management methods (Carter et al., 2007).
Some studies have suggested incorporating an analysis of the ongoing
or chronic economic impact of disasters into the adaptation planning
process (Freeman, 2000). A fuller assessment of disaster cost at varying
spatial and temporal scales and costs related to impacts on human,
social, built, and natural capital, and their associated services at different
levels can set the stage for comparisons of post-disaster development
strategies. This would make disaster risk reduction planning and
preparedness investment more cost-effective (Gaddis et al., 2007). For
example, there is consensus on the important role of ecosystems in risk
reduction and well-being, which would make the value of ecosystem
services an integral part of key policy decisions associated with adaptation
(Tallis and Kareiva, 2006; Costanza and Farley, 2007).
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
Wet 25.7 12.6 21.3 3.6 17.1 17.1 97.5
Dry
17.7 6.5 14.5 2.4 14.6 13.8 69.6
East Asia &
Pacific
Europe &
Central Asia
Latin America &
Caribbean
Middle East &
North America
sub-Saharan
Africa
TotalSouth AsiaScenario/Region
Table 4-5 | Range of regionalized annual costs of adaptation for wet and dry scenarios (in 2005 US$ billion). Reflecting the full range of estimated costs, the wet scenario costs
do not include benefits from climate change while the dry scenario costs include benefits from climate change within and across countries. Source: World Bank, 2010.
275
References
A digital library of non-journal-based literature cited in this chapter that
may not be readily available to the public has been compiled as part of
the IPCC review and drafting process, and can be accessed via either the
IPCC Secretariat or IPCC Working Group II web sites.
Abegg, B., S. Agrawala, F. Crick, and A. Montfalcon, 2007: Climate change impacts
and adaptation in winter tourism. In: Climate Change in the European Alps:
Adapting Winter Tourism and Natural Hazards Management [S. Agrawala (ed.)].
Organisation for Economic Co-operation and Development, Paris, France, pp.
25-60.
ABI, 2005a: Financial Risks of Climate Change: Summary Report. Association of
British Insurers, London, UK, 129 pp.
ABI, 2005b: Financial Risks of Climate Change: Technical Annexes. Climate Risk
Management Ltd. for the Association of British Insurers, London, UK, 125 pp.
ABI, 2009. The Financial Risk of Climate Change. Research Paper No. 19, Association
of British Insurers, London, UK, 107 pp.
Abraham, J., F. Bendimerad, A. Berger, A. Boissonnade, O. Collignon, E. Couchmann,
F. Grandjean, S. McKay, C. Miller, C. Mortgat, R. Muir-Wood, B. Page, T. Shah, S.
Smith, P. Wiart, and Y. Xien, 2000: Windstorms Lothar and Martin. December 26-
28, 1999. Risk Management Solutions, Newark, CA, 20 pp.
ABS, 2010: Regional Population Growth, Australia, 2008-09.Australian Bureau of
Statistics, Canberra, Australia, abs.gov.au/ausstats/abs@.nsf/Products/3218.0~
2008-09~Main+Features~Main+Features?OpenDocument#PARALINK5.
ACIA, 2004: Impacts of a Warming Arctic. Overview Report of the Arctic Climate
Impact Assessment, Cambridge University Press, Cambridge, UK, 139 pp.
Adamo, S.B., 2010: Environmental migration and cities in the context of global
environmental change. Current Opinion in Environmental Sustainability, 2(3),
161-165, doi:10.1016/j.cosust.2010.06.005.
Adger, W.N., 2000. Social and ecological resilience: Are they related? Progress in
Human Geography, 24, 347-364.
Adger, W.N., 2006: Vulnerability. Global Environmental Change, 16(3), 268-281.
Adger, W.N., T.A. Benjaminsen, K. Brown, and H. Svarstad, 2001: Advancing a
political ecology of global environmental discourses. Development and Change,
32(4), 681-715. doi:10.1111/1467-7660.00222.
Adger, W.N., T.P. Hughes, C. Folke, S.R. Carpenter, and J. Rockstrom, 2005: Social-
ecological resilience to coastal disasters. Science, 309(5737), 1036-1039.
Adger, W.N., S. Agrawala, M.M.Q. Mirza, C. Conde, K. O’Brien, J. Pulhin, R. Pulwarty,
B. Smit, and K. Takahashi, 2007: Assessment of adaptation practices, options,
constraints and capacity. In: Climate Change 2007: Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani,
J.P. Palutikof, P.J. van der Linden and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 717-743.
Aggarwal, P.K. and A.K. Singh, 2010: Implications of global climatic change on
water and food security. In: Global Change: Impacts on Water and Food Security
[Ringler, C., A.K. Biswas, and S. Kline (eds.)]. Springer-Verlag, Berlin and
Heidelberg, Germany, pp. 49-63.
Agrawala, S. and S. Fankhauser (eds.), 2008: Economic Aspects of Adaptation to
Climate Change: Costs, Benefits and Policy Instruments. Organisation for
Economic Co-operation and Development, Paris, France, 133 pp.
Albala-Bertrand, J., 1993: Political Economy of Large Natural Disasters. Oxford
University Press, New York, NY, 272 pp.
Alexander, D., 2005: Vulnerability to landslides. In:
Landslide Hazards and Risk
[Glade, T., M. Anderson, and M. Crozier (eds.)]. Wiley & Sons, Sussex, UK,
pp.175-198.
Allen, C.D. and D.D. Breshears, 1998: Drought-induced shift of a forest–woodland
ecotone: Rapid landscape response to climate variation. Proceedings of the
National Academy of Sciences, 95(25), 14839 -14842.
Alpa, B., 2009: Vulnerability of Turkish coasts to accelerated sea-level rise.
Geomorphology, 107, 58-63.
AMCEN/UNEP, 2002: Africa environment outlook: past, present and future
perspectives. United Nations Environmental Programme and the African
Ministerial Conference on the Environment, Earthprint, Stevenage, UK, 422 pp.
Amien, I., P. Rejekiningrum, A. Pramudia, and E. Susanti, 1996: Effects of interannual
climate variability and climate change on rice yields in Java, Indonesia. Water,
Air &Soil Pollution, 92, 29-39.
Anderson, M.G., L. Holcombe, and J.-P. Renaud, 2007: Assessing slope stability in
unplanned settlements in developing countries. Journal of Environmental
Management, 85(1), 101-111.
Anisimov, O.A. and M.A. Belolutskaya, 2002: Assessment of the impact of climate
change and permafrost degradation on infrastructure in northern regions of
Russia. Meteorology and Hydrology, 6, 15-22.
Anisimov, O.A. and S.A. Lavrov, 2004: Global warming and permafrost melting:
assessment of risks for energy-sector industrial facilities. Tekhnologii TEK, 3, 78-
83.
Anisimov, O.A., A.A. Velichko, P.F. Demchenko, A.V. Yeliseyev, I.I. Mokhov, and V.P.
Nechaev, 2004: Impact of climate change on permafrost in the past, present
and future. Physics of Atmosphere and Oceans, 38(1), 25-39.
Anyamba, A., J-P. Chretien, J. Small, C.J. Tucker, and K.J. Linthicum, 2006: Developing
global climate anomalies suggest potential disease risks for
2006-2007. International Journal of Health Geographics, 5(60),
doi:10.1186/1476-072X-5-60.
Arianoutsou, M., S. Koukoulas, and D. Kazanis, 2011: Evaluating post-fire forest
resilience using GIS and multi-criteria analysis: an example from Cape Sounion
National Park, Greece. Journal of Environmental Management, 47, 384-397.
Arkell, B.P. and G.J.C. Darch, 2006: Impact of climate change on London’s transport
network. Proceedings of the Institute of Civil Engineers – Municipal Engineer,
159(4), 231-237.
Arnone, III, J.A., P.S.J. Verburg, D.W. Johnson, J.D. Larsen, R.L. Jasoni, L.J. Annmarie,
C.M. Batts, C. von Nagy, W.G. Coulombe, D.E. Schorran, P.E. Buck, B.H. Braswell,
J.S. Coleman, R.A. Sherry, L.L. Wallace, Y. Luo, and D.S. Schimel,
2008: Prolonged suppression of ecosystem carbon dioxide uptake after an
anomalously warm year. Nature, 455, 383-386.
AusAID, 2005: Economic Impact of Natural Disasters on Development in the Pacific.
Volume 1: Research Report, Australian Agency for International Development,
Canberra, Australia, 92 pp., ausaid.gov.au/publications/pdf/impact_pacific_
report.pdf.
Baccini, M., T. Kosatsky, A. Analitis, H.R. Anderson, M. D’Ovidio, B. Menne, P.
Michelozzi, A. Biggeri, and the PHEWE Collaborative Group, 2011: Impact of
heat on mortality in 15 European cities: attributable deaths under different
weather scenarios. Journal of Epidemiology and Community Health,
65(1), 64-
70, doi:10.1136/jech.2008.085639.
Banks, S., M. Blyton, D. Blair, L. McBurney, and D. Lindenmayer, 2011: Adaptive
responses and disruptive effects: how major wildfire influences kinship-based
social interactions in a forest marsupial. Molecular Ecology, doi:10.1111/
j.1365-294X.2011.05282.x.
Barnett, T.P. and D.W. Pierce, 2008: When will Lake Mead run dry? Journal of Water
Resources Research, 44, W03201, doi:10.1029/2007WR006704.
Barnett, T.P. and D.W. Pierce, 2009: Sustainable water deliveries from the Colorado
River in a changing climate. Proceedings of the National Academy of Sciences,
106(18), 7334-7338, doi:10.1073/pnas.0812762106.
Barredo, J.I., 2009: Normalised flood losses in Europe: 1970-2006. Natural Hazards
and Earth System Sciences, 9, 97-104.
Barriopedro, D., E.M. Fischer, J. Luterbacher, R.M. Trigo, and R. Garcia-Herrara, 2011:
The hot summer of 2010: redrawing the temperature record map of Europe.
Science, 332, 220-224.
Basara, J.B., H.G. Basara, B.G. Illston, and K.C. Crawford, 2010: The impact of the
urban heat island during an intense heat wave in Oklahoma City. Advances in
Meteorology, (2010), 230365, doi:10.1155/2010/230365.
Battisti, D.S. and R.L. Naylor, 2009: Historical warnings of future food insecurity with
unprecedented seasonal heat. Science, 323, 240-244.
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
276
Bebi, P., D. Kulakowski, and C. Rixen, 2009: Snow avalanche disturbances in forest
ecosystems – State of research and implications for management. Forest
Ecology and Management, 257(9), 1883-1892.
Becken, S. and J.E. Hay, 2007: Tourism and climate change: risks and opportunities.
Climate Change, Economies and Society Series, Channel View Publications,
Clevedon, UK, 352 pp.
Bender, M.A., T.R. Knutson, R.E. Tuleya, J.J. Sirutis, G.A. Vecchi, S.T. Garner, and I.M.
Held, 2010: Modeled impact of anthropogenic warming on the frequency of
intense Atlantic hurricanes. Science, 327, 454-458.
Beniston, M., 2004: The 2003 heatwave in Europe: A shape of things to come? An
analysis based on Swiss climatological data and model simulations.
Geophysical Research Letters, 31, L02202.
Beniston, M. and H. Diaz, 2004: The 2003 heatwave as an example of summers in a
greenhouse climate? Observations and climate model simulations for Basel,
Switzerland. Global and Planetary Change, 44(1-4), 73-81.
Beniston, M., F. Keller, and S. Goyette, 2003: Snow pack in the Swiss Alps under
changing climatic conditions: An empirical approach for climate impact studies.
Theoretical and Applied Climatology, 74, 19-31.
Benito, G., T. Grodek, and Y. Enzel, 1998: The geomorphic and hydrologic impacts of
the catastrophic failure of flood-control-dams during the 1996-Biescas flood
(Central Pyrennes, Spain). Zeitschrift für Geomorphologie, 42(4), 417-437.
Benito, G., R.F. Rohde, M. Seely, C. Kulls, O. Dahan, Y. Enzel, S. Todd, B. Botero, and T.
Grodek, 2010: Management of alluvial aquifers in two southern African
ephemeral rivers: implications for IWRM. Water Resources Management, 24,
641-667.
Benson, C. and E. Clay, 1998: The Impact of Drought on Sub-Saharan African
Economies: A Preliminary Examination. World Bank Technical Paper Series No.
401, World Bank, Washington, DC, 80 pp.
Benson, C. and E.J. Clay, 2000: Developing countries and the economic impacts of
natural disasters. In: Managing disaster risk in emerging economies [Kreimer, A.
and A. Margaret (eds.)]. Disaster Risk Management Series No. 2, World Bank,
Washington, DC, pp. 11-21.
Benson, C. and E. Clay, 2003: Disasters, vulnerability, and global economy. In:
Building Safer Cities: The Future of Disaster Risk [Kreimer, A., M. Arnold, and A.
Carlin (eds.)]. Disaster Risk Management Series No. 3, World Bank, Washington,
DC, pp. 3-31.
Benson, C. and E. Clay, 2004: Understanding the Economic and Financial Impacts of
Natural Disasters. Disaster Risk Management Series, No. 4, World Bank,
Washington, DC, 119 pp..
Benson, C. and J. Twigg, 2004. Measuring Mitigation: Methodologies for assessing
natural hazard risks and the net benefits of mitigation — A Scoping Study. The
International Federation of Red Cross and Red Crescent Societies/The
ProVention Consortium, Geneva, Switzerland, 149 pp.
Berrittella M., A. Bigano, and R.S.J. Tol, 2006: A general equilibrium analysis of
climate change impacts on tourism. Tourism Management, 27(5), 913-924.
Berry, M., 2003: Why is it important to boost the supply of affordable housing in
Australia - and how can we do it? Urban Policy and Research, 21, 413-435.
Berry, H., K. Bowen, and T. Kjellstrom, 2010: Climate change and mental health: a
causal pathways framework. International Journal of Public Health
, 55, 123-
132.
Bertness, M.D. and P.J. Ewanchuk, 2002: Latitudinal and climate-driven variation in
the strength and nature of biological interactions in New England salt marshes.
Oecologia, 132, 392-401.
Beven, J.L., L.A. Avila, E.S. Blake, D.P. Brown, J.L. Franklin, R.D. Knabb, R.J. Pasch, J.R.
Rhome, and S.R. Stewart, 2008: Atlantic hurricane season of 2005. Monthly
Weather Review, 136, 1109-1173.
Bhuiyan, C., R.P. Singh, and F.N. Kogan, 2006: Monitoring drought dynamics in the
Aravalli region (India) using different indices based on ground and remote sens-
ing data. International Journal of Applied Earth Observation and
Geoinformation, 8(4), 289-302.
Bigano, A., F. Bosello, R. Roson, R.S.J. Tol, 2008: Economy-wide impacts of climate
change: a joint analysis for sea level rise and tourism. Mitigation and Adaption
Strategies for Global Change, 13, 765-791.
Bindoff, N., J. Willebrand, V. Artale, A. Cazenave, J. Gregory, S. Gulev, K. Hanawa, C.
Le Quéré, S. Levitus, Y. Nojiri, C.K. Shum, L.D. Talley and A. Unnikrishnan, 2007:
Observations: Oceanic climate change and sea level. In: Climate Change 2007:
The Physical Science Basis, Contribution of Working Group I to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change
[Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor,
and H.L. Miller (eds.)]. Cambridge University Press, Cambridge, UK, pp. 385-432.
Birch, E.L. and S.M. Wachter (eds.), 2006: Rebuilding Urban Places after Disaster:
Lessons from Hurricane Katrina. University of Pennsylvania Press, Philadelphia,
PA, 375 pp.
Bongaerts, P., T. Ridgeway, E.M. Sampayo, and O. Hoegh-Guldberg, 2010: Assessing
the ‘deep reef refugia’ hypothesis: focus on Caribbean reefs. Coral Reefs, 29,
309-327.
Borrell, C, M. Marí-Dell’Olmo, M. Rodríguez-Sanz, P.Garcia-Olalla, J.A. Caylà, J.
Benach, and C. Muntaner, 2006: Socioeconomic position and excess mortality
during the heat wave of 2003 in Barcelona. European Journal of Epidemiology,
21(9), 633-640.
Boruff, B.J., J.A. Easoz, S.D. Jones, H.R. Landry, J.D. Mitchem, and S.L. Cutter, 2003:
Tornado hazards in the United States. Climate Research, 24, 103-117.
Botzen, W.J.W., L.M. Bouwer, and J.C.J.M. van den Bergh, 2010: Climate change and
hailstorm damage: Empirical evidence and implications for agriculture and
insurance. Resource and Energy Economics, 32, 341-362.
Bouwer, L.M., 2010: Disasters and climate change: analysis and methods for
projecting future losses from extreme weather. Ph.D. thesis, Vrije Universiteit,
Amsterdam, The Netherlands, 141 pp., dare.ubvu.vu.nl/bitstream/1871/16355/1/
dissertation.pdf.
Bouwer, L.M., 2011: Have disaster losses increased due to anthropogenic climate
change? Bulletin of the American Meteorological Society, 92, 39-46.
Bouwer, L.M. and W.J.W. Botzen, 2011: How sensitive are US hurricane damages to
climate? Comment on a paper by W.D. Nordhaus. Climate Change Economics,
2, 1-7.
Bouwer, L.M., R.P. Crompton, E. Faust, P. Höppe, and R.A. Pielke Jr., 2007:
Confronting disaster losses. Science,
318, 753.
Bouwer, L.M., P. Bubeck, and J.C.J.H. Aerts, 2010: Changes in future flood risk due to
climate and development in a Dutch polder area. Global Environmental
Change, 20, 463-471.
Breshears, D.D. and C.D. Allen, 2002: The importance of rapid, disturbance-induced
losses in carbon management and sequestration. Global Ecology and
Biogeography, 11, 1-5.
Breshears, D.D., N.S. Cobb, P.M. Rich, K.P. Price, C.D. Allen, R.G. Balice, W.H. Romme,
J.H. Kastens, M.L. Floyd, J. Belnap, J.J. Anderson, O.B. Myers, and C.W. Meyer,
2005: Regional vegetation die-off in response to global-change-type drought.
Proceedings of the National Academy of Sciences, 102(42), 15144-15148.
Bronstert, A., A. Bárdossy, C. Bismuth, H. Buiteveld, M. Disse, H. Engel, U. Fritsch, Y.
Hundecha, R. Lammersen, D. Niehoff, and N. Ritter, 2007: Multi-scale modelling
of land-use change and river training effects on floods in the Rhine basin. River
Research and Applications, 23(10), 1102-1125.
Brooks, N., 2004: Drought in the Africa Sahel: long-term perspectives and future
prospects. Working paper 61, Tyndall Centre for Climate Research, University of
East Anglia, Norwich, UK, 31 pp.
Brooks, N., W.N. Adger, and P.M. Kelly, 2005: The determinants of vulnerability and
adaptive capacity at the national level and the implications for adaptation.
Global Environmental Change, 15, 151-163.
Brookshire, D.S., S.E. Chang, H. Cochrane, R.A. Olson, A. Rose, and J. Steenson, 1997:
Direct and indirect economic losses from earthquake damage. Earthquake
Spectra, 13, 683-701.
Brouwer, R., S. Akter, L. Brander, and E. Haque, 2007: Socioeconomic vulnerability
and adaptation to environmental risk: acase study of climate change and
flooding in Bangladesh. Risk Analysis, 27, 313-326.
BTE, 2001: Economic costs of natural disasters in Australia. Report 103, Bureau of
Transport Economics, Canberra, Australia, 170 pp.
Burnley, I. and P. Murphy, 2004: Sea Change: Movement from Metropolitan to
Arcadian Australia. UNSW Press, Sydney, Australia, 272 pp.
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
277
Burton, I. and E. May, 2004: The adaptation deficit in water resources management.
IDS Bulletin, 35(3), 31-37.
Burton, K., R. Kates, and G. White, 1993: The Environment as Hazard. 2nd ed.
Guilford Press, New York, NY, 290 pp.
Bussotti, F., E. Cenni, M. Ferretti, A. Cozzi, L. Brogi, and A. Mecci, 1995: Forest condition
in Tuscany (Central Italy) - Field surveys 1987-1991. Forestry, 68(1), 11-24.
Butcher, G.V. and S. Ford, 2009: Modeling the regional economic impacts of the
2007/08 drought: results and lessons. Paper presented at the 2009 NZARES
Conference, Nelson, New Zealand, 27-28 August 2009, 11 pp.,
ageconsearch.umn.edu/bitstream/97167/2/2009_22_Economic%20Impact%2
0of%202007-08%20Drought_ButcherG.pdf.
Butry, D., E. Mercer, J. Prestemon, J. Pye, and T. Holmes, 2001: What is the price of
catastrophic wildfire? Journal of Forestry, 99(11), 9-17.
Buzinde, C.N., D. Manuel-Navarrete, E.E. Yoo, and D. Morais, 2010: Tourists’
perceptions in a climate of change: Eroding destinations. Annals of Tourism
Research, 37(2), 333-354.
Buzna, L., K. Peters, and D. Helbing, 2006: Modelling the dynamics of disaster
spreading in networks. Physica A: Statistical Mechanics and its Applications,
363(1), 132-140.
Cadier, E., M. Villacis, A. Graces, P. Lhuissier, L. Maisincho, R. Laval, D. Paredes, B.
Caceres, and B. Francou, 2007: Variation of a low latitude Andean glacier
according to global and local climate variations: first results. In: Glacier Mass
Balances and Meltwater Discharge [Ginot, P. and J.E. Sicart (eds.)]. IAHS Pub.
No. 318, International Association of Hydrological Sciences, Wallingford, UK, pp.
66-74.
Campbell, J.R., 2006: Traditional Disaster Reduction in Pacific Island
Communities.GNS Science Report 2006/38, Institute of Geological and Nuclear
Sciences, Lower Hutt, New Zealand, 46 pp.
Cardemil, L., J.C. Di Tata, and F. Frantischek, 2000: Central America: adjustment and
reforms in the 1990s. Finance and Development, 37(1), 34-37,
www.imf.org/external/pubs/ft/fandd/2000/03/cardemil.htm.
Carson, R., R. Mitchell, M. Hanemann, R. Kopp, S. Presser, and P. Ruud, 2003:
Contingent valuation and lost passive use: damages from the Exxon Valdez oil
spill. Environmental and Resource Economics, 25, 257-286.
Carter, L.M., S.J. Cohen, N.B. Grimm, and J.L. Hatfield, 2008: Weather and Climate
Extremes in a Changing Climate, Regions of Focus: North America, Hawaii,
Caribbean and U.S. Pacific Islands. Synthesis and Assessment Product 3.3,
Report by the U.S. Climate Change Science Program and the Subcommittee on
Global Change, Washington, DC, 162 pp.
Carter, M.R., P.D. Little, T. Mogues, and W. Negatu, 2007: Poverty traps and natural
disasters in Ethiopia and Honduras. World Development, 35(5), 835-856,
doi:10.1016/j.worlddev.2006.09.010.
Caselli, F. and P. Malhotra, 2004: Natural Disasters and Growth: from Thought
Experiment to Natural Experiment, International Monetary Fund, Washington,
DC, 29 pp.
Casimiro, E. and J.M. Calheiros, 2002: Human health. In: Climate Change in
Portugal. Scenarios, Impacts and Adaptation Measures—SIAM Project [Santos
F.D., K. Forbes, and R. Moita (eds.)]. Gradiva Publishers, Lisbon, Portugal, pp.
241-300.
Cavallo, E. and I. Noy, 2009: The Economics of Natural Disasters: A Survey. IDB Working
Paper 09-19, Inter-American Development Bank, Washington, DC, 48 pp.
Cavallo, E. and I. Noy, 2010: The Economics of Natural Disasters: A Survey. IDB
Working Paper Series, No. IDB-WP-124, Inter-American Development Bank,
Washington, DC, 50 pp., www.iadb.org/res/publications/pubfiles/pubIDB-WP-
124.pdf.
CCSP, 2008: Impacts of Climate Change and Variability on Transportation Systems
and Infrastructure: Gulf Coast Study, Phase I. A Report by the U.S. Climate
Change Science Program and the Subcommittee on Global Change Research
[Savonis, M.J., V.R. Burkett, and J.R. Potter (eds.)]. Department of Transportation,
Washington, DC, 445 pp.
CDHS, 2007: Review of July 2006 Heatwave Related Fatalities in California. CDHS
Report, Epidemiology and Prevention for Injury Control Branch, California
Department of Health Services, Sacramento, CA, 10 pp.
Chang, H., J. Franczyk, and C. Kim, 2009: What is responsible for increasing flood
risks? The case of Gangwon Province, Korea. Natural Hazards, 48, 399-354.
Changnon, S. (ed.), 1996: The Great Flood of 1993. Westview Press, Boulder, CO, 321
pp.
Changnon, S.A., 2001: Damaging thunderstorm activity in the United States. Bulletin
of the American Meteorological Society, 82, 597-608.
Changnon, S.A., 2006: Railroads and Weather: From Fogs to Floods and Heat to
Hurricanes, the Impacts of Weather and Climate on American Railroading.
American Meteorological Society, Boston, MA, 136 pp.
Changnon, S.A., 2009: Increasing major hail losses in the U.S. Climatic Change, 96,
161-166.
Chapin, F.S., T.V. Callaghan, Y. Bergeron, M. Fukuda, J.F. Johnstone, G. Juday, and S.A.
Zimov, 2004: Global change and the boreal forest: thresholds, shifting states or
gradual change? Ambio, 33, 361-365.
Charveriat, C., 2000: Natural Disasters in Latin America and the Caribbean: An
Overview of Risk. IDB Working Paper 434, Inter-American Development Bank,
Washington, DC, 104 pp.
Chen, K. and J. McAneney, 2006: High-resolution estimates of Australia’s coastal
population with validations of global population, shoreline and elevation
datasets. Geophysical Research Letters, 33, L16601, doi:10.1029/
2006GL026981.
Cheng, S. and R. Wang, 2002: An approach for evaluating the hydrological effects of
urbanization and its application. Hydrological Processes, 16(7), 1403-1418.
Choi, O. and A. Fisher, 2003: The impacts of socioeconomic development and climate
change on severe weather catastrophe losses: Mid-Atlantic Region (MAR) and
the US. Climatic Change, 58(1-2), 149-170.
Christensen, J.H., B. Hewitson, A. Busuioc, A. Chen, X. Gao, I. Held, R. Jones, R.K.
Kolli, W. Kwon, R. Laprise, V.M. Rueda, L. Mearns, C.G. Menéndez, J. Räisänen,
A. Rinke, A. Sarr, and P. Whetton, 2007: Regional climate projections. In: Climate
Change 2007: The Physical Science Basis. Contribution of Working Group I to
the Fourth Assessment Report of the Intergovernmental Panel on Climate
Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M.
Tignor, and H.L. Miller (eds.)]. Cambridge University Press, Cambridge, UK, pp.
847-933.
Christensen, N.S., A.W. Wood, N. Voisin, D.P. Lettenmaier, and R.N. Palmer, 2004: The
effects of climate change on the hydrology and water resources of the Colorado
River basin. Climatic Change, 62(1), 337-363.
Ciais, P., M. Reichstein, N. Viovy, A. Granier, J. Ogée, V. Allard, M. Aubinet, N.
Buchmann, Chr. Bernhofer, A. Carrara, F. Chevallier, N. De Noblet, A.D. Friend, P.
Friedlingstein, T. Grünwald, B. Heinesch, P. Keronen, A. Knohl, G. Krinner, D.
Loustau, G. Manca, G. Matteucci, F. Miglietta, J.M. Ourcival, D. Papale, K.
Pilegaard, S. Rambal, G. Seufert, J.F. Soussana, M.J. Sanz, E.D. Schulze, T. Vesala,
and R. Valentini, 2005: Europe-wide reduction in primary productivity caused by
the heat and drought in 2003. Nature, 437(7058), 529-533.
Ciscar, J.C. (ed.), 2009: Climate Change Impacts in Europe: Final Report of the PESETA
research project. EUR 24093 EN-2009, European Commission, Joint Research
Centre, Institute for Prospective Technological Studies, EU Publications,
Luxembourg, Luxembourg, 113 pp.
Clarke, M.L. and H.M. Rendell, 2006: Hindcasting extreme events: the occurrence
and expression of damaging floods and landslides in Southern Italy. Land
Degradation and Development, 17(4), 365-380.
Collier, P., V. Elliot, H. Hegre, A. Hoeffler, M. Reynal-Querol, and N. Sambanis, 2003:
Breaking the Conflict Trap: Civil War and Development Policy. World Bank,
Washington, DC.
Corominas, J., 2005: Impacto sobre los riesgos naturales de origen climático:
inestabilidad de laderas. In: Proyecto ECCE. Evaluación Preliminar de los
impactos en España por efecto del Cambio Climático [Moreno, J.M. (ed.)].
Ministerio de Medio Ambiente, Madrid, Spain, pp. 549-579.
Costa, M.H. and G.F. Pires, 2010: Effects of Amazon and Central Brazil deforestation
scenarios on the duration of the dry season in the arc of deforestation.
International Journal of Climatology, 30(13), 1970-1979.
Costanza, R. and J. Farley. 2007. Ecological economics of coastal disasters:
Introduction to the Special Issue. Ecological Economics, 63, 249-253.
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
278
Costello, L., 2007: Going bush: the implications of urban-rural migration.
Geographical Research, 45, 85-94.
Costello, L., 2009: Urban-rural migration: housing availability and affordability.
Australian Geographer, 40, 219-233.
CRED, 2010: EM-DAT: The OFDA/CRED International Disaster Data Base. Centre for
Research on the Epidemiology of Disasters, Université Catholique de Louvain,
Louvain, Belgium, www.emdat.be/Database/.
Crichton, D., 2007: What can cities do to increase resilience? Philosophical
Transactions of the Royal Society A, 365(1860), 2731-2739.
Crompton, R.P. and K.J. McAneney, 2008: Normalised Australian insured losses from
meteorological hazards: 1967-2006. Environmental Science and Policy, 11,
371-378.
Crompton, R.P., K.J. McAneney, and K. Chen, 2010: Influence of location, population,
and climate on building damage and fatalities due to Australian bushfire:
1925–2009. Weather, Climate, and Society, 2, 300-310.
Crompton, R., R.A. Pielke Jr., and J. McAneney, 2011: Emergence timescales for
detection of anthropogenic climate change in US tropical cyclone loss data.
Environmental Research Letters, 6(1), doi:10.1088/1748-9326/6/1/014003.
Crosta, G.B., H. Chen, and C.F. Lee, 2004: Replay of the 1987 Val Pola Landslide,
Italian Alps. Geomorphology, 60(2, 3), 127-146.
Crowards, T., 2000: Comparative Vulnerability to Natural Disasters in the Caribbean.
Staff Working Paper No. 1/00, Caribbean Development Bank, Charleston, SC, 21
pp.
Crozier, M.J., 2010: Deciphering the effect of climate change on landslide activity: A
review. Geomorphology, 124(3-4), 260-267, doi:10.1016/j.geomorph.2010.04.009.
Cruz, A., 2005: NaTech disasters: a review of practices, lessons learned and future
research needs. Paper presented at the 5th Annual IIASA-DPRI Forum,
Integrated Disaster Risk Management: Innovations in Science and Policy,
Beijing, China, 14-18 September 2005, 10 pp.
Cruz, R.V., H. Harasawa, M. Lal, S. Wu, Y. Anokhin, B. Punsalmaa, Y. Honda, M. Jafari,
C. Li and N. Huu Ninh, 2007: Asia. In: Climate Change 2007: Impacts, Adaptation
and Vulnerability. Contribution of Working Group II to the Fourth Assessment
Report of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F.
Canziani, J.P. Palutikof, P.J. van der Linden and C.E. Hanson (eds.)]. Cambridge
University Press, Cambridge, UK, pp. 469-506.
Cummins, J., and O. Mahul, 2009: Catastrophe Risk Financing in Developing
Countries: Principles for Public Intervention. World Bank, Washington, DC.
Cuny, F.C., 1983: Disasters and Development. Oxford University Press, Oxford, UK,
278 pp.
Curriero, F.C., J.A. Patz, J.B. Rose, and S. Lele, 2001: The association between extreme
precipitation and waterborne disease outbreaks in the United States, 1948-
1994. American Journal of Public Health, 91(8), 1194-1199, doi:10.2105/
AJPH.91.8.1194.
Cutter, S. and C. Emrich, 2005: Are natural hazards and disaster losses in the U.S.
increasing? EOS,
86(41), 381-396.
Cutter, S.L. and C.T. Emrich, 2006: Moral hazard, social catastrophe: the changing
face of vulnerability along the hurricane coasts. Annals of the American
Academy of Political and Social Science, 604, 102-112.
Cutter, S.L. and C. Finch, 2008: Temporal and spatial changes in social vulnerability
to natural hazards. Proceedings of the National Academy of Sciences, 105(7),
2301-2306.
Cutter, S.L., L. Barnes, M. Berry, C. Burton, E. Evans, E. Tate, and J. Webb, 2008: A
place-based model for understanding community resilience to natural disasters.
Global Environmental Change, 18, 598-606.
D’Almeida, C., C.J. Vorosmarty, G.C. Hurtt, J.A. Marengo, S.L. Dingman, and B.D.
Keim, 2007: The effects of deforestation on the hydrological cycle in Amazonia:
a review on scale and resolution. International Journal of Climatology, 27,
633-648.
Dao, H. and P. Peduzzi, 2004: Global evaluation of human risk and vulnerability to
natural hazards In: Proceedings of ENVIROINFO 2004, 18th International
Conference Informatics for Environmental Protection, Vol. 1, CERN, Geneva,
Switzerland, 21-23 October 2004, Sh@ring, Editions du Tricorne, Geneva,
Switzerland, pp. 435-446.
Dasgupta, S., B. Laplante, S. Murray, and D. Wheeler, 2009: Sea-Level Rise and Storm
Surges: A Comparative Analysis of Impacts in Developing Countries. Policy
Research Working Paper 4901, World Bank Development Research Group,
Environment and Energy Team, Washington, DC, 41 pp.
Dash, S., R. Jenamani, S. Kalsi, and S. Panda, 2007: Some evidence of climate change
in twentieth-century India, Climatic Change, 85, 299-321.
Davis, R.E, P.C. Knappenberger, P.J. Michaels, and W.M. Novicoff, 2003: Changing
heat-related mortality in the United States. Environmental Health Perspectives,
111, 1712-1718.
Dawson, R. J., J.W. Hall, P.D. Bates, and R.J. Nicholls, 2005: Quantified analysis of the
probability of flooding in the Thames Estuary under imaginable worst-case sea
level rise scenarios. International Journal of Water Resources Development,
21(4), 577-591, doi:10.1080/07900620500258380.
Dawson, R.J., M.E. Dickson, R.J. Nicholls, J.W. Hall, M.J.A. Walkden, P.K. Stansby, M.
Mokrech, J. Richards, J. Zhou, J. Milligan, A. Jordan, S. Pearson, J. Rees, P.D. Bates,
S. Koukoulas, and A.R. Watkinson, 2009: Integrated analysis of risks of coastal
flooding and cliff erosion under scenarios of long term change. Climatic
Change, 95, 249-288.
de Bruin, K., R.B.Dellink, A.Ruijs, L. Bolwidt, A. vanBuuren, J. Graveland, R.S.
deGroot, P.J.Kuikman, S. Reinhard, R.P. Roetter, V.C. Tassone, A. Verhagen, and
E.C. van Ierland, 2009: Adapting to climate change in The Netherlands: an
inventory of climate adaptation options and ranking of alternatives. Climatic
Change, 95(1-2), 23-45.
de Wit, M. and J. Stankiewicz, 2006: Changes in water supply across Africa with
predicted climate change. Science, 311, 1917-1921.
DEFRA, 2004: Scientific and Technical Aspects of Climate Change, including Impacts
and Adaptation and Associated Costs. Department for Environment, Food and
Rural Affairs, London, UK, 19 pp., www.newforestnpa.gov.uk/__data/assets/
pdf_file/0016/20473/152-07_annex_3_climate_change.pdf.
DEFRA, 2006: The UK’s Fourth National Communication under the United Nations
Framework Convention on Climate Change
. United Kingdom Department for
Environment, Food and Rural Affairs, London, UK, 135 pp., unfccc.int/resource/
docs/natc/uknc4.pdf.
Delpla, I., A.-V. Jung, E. Baures, M. Clement, and O. Thomas, 2009: Impacts of climate
change on surface water quality in relation to drinking water production.
Environment International, 35(8), 1225-1233.
Deschênes, O.and M. Greenstone, 2007: The economic impacts of climate change:
Evidence from agricultural output and random fluctuations in weather. The
American Economic Review, 97(1), 354-385.
Devoy, R.J.N., 2008: Coastal vulnerability and the implications of sea-level rise for
Ireland. Journal of Coastal Research, 24(2), 325-341.
DHS, 2009: January 2009 Heatwave in Victoria: an Assessment of Health Impacts.
Victorian Govt. Department of Human Services, Melbourne, Australia, 16 pp.
Dodman, D. and D. Satterthwaite, 2008: Institutional capacity, climate change
adaptation and the urban poor. Institute of Development Studies Bulletin,
39(4), 67-74.
Donat, M.G., G.C. Leckebusch, J.G. Pinto, and U. Ulbrich, 2010: European storminess
and associated circulation weather types: future changes deduced from a multi-
model ensemble of GCM simulations. Climate Research, 42, 27-43.
Donat, M.G., G.C. Leckebusch, S. Wild, and U. Ulbrich, 2011: Future changes in
European winter storm losses and extreme wind speeds inferred from GCM and
RCM multi-model simulations. Natural Hazards and Earth System Sciences, 11,
1351-1370.
Dorland, C., R.S.J. Tol, and J.P. Palutikof, 1999: Vulnerability of the Netherlands and
Northwest Europe to storm damage under climate change. Climatic Change,
43, 513-535.
Douglas, I., 2009: Climate change, flooding and food security in south Asia, Food
Security, 1(2), 127-136.
Douglas, I., K. Alam, M. Maghenda, Y. Mcdonnell, L. Mclean, and J. Campbell, 2008:
Unjust waters: climate change, flooding and the urban poor in Africa,
Environment and Urbanization, 20 (1), 187-205.
Downton, M. and R.A. Pielke, Jr., 2005: How accurate are disaster loss data? The
case of U.S. flood damage. Natural Hazards, 352, 211-228.
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
279
Downton, M.W., J.Z.B. Miller, and R.A. Pielke, 2005: Reanalysis of US National
Weather Service flood loss database. Natural Hazards Review, 6, 13-22.
Dregne, H.E., 1986: Desertification of arid lands. In: Physics of Desertification [El-Baz,
F. and M.H.A. Hassan (eds.)]. Martinus Nijhoff, Dordrecht, The Netherlands, pp.
4-34.
DSD,1994: Barbados Declaration. A/CONF.167/9,I, Annex I. Report of the Global
Conference on the Sustainable Development of Small Island Developing States,
Bridgetown, Barbados, 25 April-6 May 1994, United Nations Division for
Sustainable Development, Department of Economic and Social Affairs, New
York, NY, USA, 77 pp., islands.unep.ch/dbardecl.htm.
Eakin, C.M., J. A. Morgan, S. F. Heron, T. B. Smith, G. Liu, et al., 2010: Caribbean corals
in crisis: record thermal stress, bleaching, and mortality in 2005. Public Library
of Science ONE, 5(11), e13969, doi:10.1371/journal.pone.0013969.
Easterling, W. and M. Apps, 2005: Assessing the consequences of climate change for
food and forest resources: A view from the IPCC. Climate Change, 70, 165-
189.
Easterling, W.E., P.K. Aggarwal, P. Batima, K.M. Brander, L. Erda, S.M. Howden, A.
Kirilenko, J. Morton, J.-F. Soussana, J. Schmidhuber, and F.N. Tubiello, 2007: Food,
fibre and forest products. In: Climate Change 2007: Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani,
J.P. Palutikof, P.J. van der Linden and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 273-313.
Ebersole, B.A., J.J. Westerink, S. Bunya, J.C. Dietrich, and M.A. Cialone, 2010:
Development of storm surge which led to flooding in St. Bernard Polder during
Hurricane Katrina. Ocean Engineering, 37, 91-103.
Ebi, K.L. and G.A. Meehl, 2007: The Heat is On: Climate Change & Heatwaves in the
Midwest. Pew Center on Global Climate Change, Arlington, VA, 14 pp.,
www.pewclimate.org/docUploads/Regional-Impacts-Midwest.pdf.
Ebi, K.L., J. Balbus, P.L. Kinney, E. Lipp, D. Mills, M.S. O’Neill, and M. Wilson, 2008:
Effects of global change on human health. In: Analyses of the Effects of Global
Change on Human Health and Welfare and Human Systems [Gamble, J.L.
(ed.)]. A Report by the U.S. Climate Change Science Program and the
Subcommittee on Global Change Research. U.S. Environmental Protection
Agency, Washington, DC, pp. 2-1 to 2-78.
ECA, 2009: Shaping Climate-Resilient Development: A Framework for Decision-
Making Study. Report of the Economics of Climate Adaptation Working Group,
New York, NY, 159 pp.
ECLAC, 2003: Handbook for Estimating the Socio-economic and Environmental
Effects of Disasters. LC/MEX/G.5, LC/L 1874, Economic Commission for Latin
America and the Caribbean, Santiago, Chile, 354 pp., www.undp.org/cpr/
disred/documents/publications/eclac_handbook.pdf.
EEA, 2004: Impacts of Europe’s changing climate. An indicator-based assessment.
EEA Report No. 2/2004, European Environmental Agency, Copenhagen,
Denmark, 100 pp.
EEA, 2008: Impacts of Europe’s changing climate- 2008 indicator-based assessment.
EEA Report No.4/2008, European Environmental Agency, Copenhagen,
Denmark, 19 pp., doi:10.2800/48117.
EEA, 2009: Regional Climate Change and Adaptation: The Alps Facing the Challenge
of Changing Water Resources. EEA Report No. 8, European Environmental
Agency, Copenhagen, Denmark, 143 pp.
Ehmer, P. and E. Heymann, 2008:
Climate change and tourism: Where will the journey
lead? Deutsche Bank Research, Frankfurt am Main, Germany, 28 pp.
Ellson, R.W., J.W. Milliman, and R.B. Roberts, 1984: Measuring the regional economic
effects of earthquakes and earthquake predictions. Journal of Regional Science,
24(4), 559-579.
Elsasser, H. and R. Bürki, 2002: Climate change as a threat to tourism in the Alps.
Climate Research, 20, 253-257.
Emanuel, K., 2005: Increasing destructiveness of tropical cyclones over the past 30
years. Nature, 436(7051), 686-688.
Endfield, G.H. and I.F. Tejedo, 2006: Decades of drought, years of hunger: archival
investigations of multiple year droughts in late colonial Chihuahua. Climatic
Change, 75(4), 391-419, doi:10.1007/s10584-006-3492-7.
Endlicher, W., G. Jendritzky, J. Fischer, and J.-P. Redlich, 2008: Heat waves, urban
climate and human health. In Urban Ecology [Marzluff, J.M., E. Shulenberger, W.
Endlicher, M. Alberti, G. Bradley, C. Ryan, U. Simon, and C. ZumBrunnen (eds.)].
Springer, New York, NY, pp. 269-278.
Engelthaler, D.M., D.G. Mosley, J.E. Cheek, C.E. Levy, K.K. Komatsu, P. Ettestad, T.
Davis, D.T. Tanda, L. Miller, J.W. Frampton, R. Porter, and R.T. Bryan, 1999:
Climatic and environmental patterns associated with hantavirus pulmonary
syndrome, Four Corners region, United States. Emerging Infectious Diseases,
5(1), 87-94.
Eriksen, S.H., K. Brown, and P.M. Kelly, 2005: The dynamics of vulnerability: locating
coping strategies in Kenya and Tanzania. Geographical Journal, 171, 287-305.
Esteban-Talaya, A., F. López Palomeque, and E. Aguiló Pérez, 2005: Impacts in the
touristic sector. In: A Preliminary Assessment of the Impacts in Spain due to the
Effect of Climate Change [Moreno, J.M. (ed.)]. Ministry of Environment, Madrid,
Spain, pp. 653-690.
Eurostat, 2010: Ageing in the European Union: where exactly? Rural areas are losing
the young generation quicker than urban areas. Eurostat Statistics in Focus,
26/2010, Eurostat, European Commission, Luxembourg, 16 pp.
Evans, B. and M. Webster, 2008: Adapting to Climate Change in Europe and Central
Asia. Background Paper on Water Supply and Sanitation. World Bank
Sustainable Development Department, Europe and Central Asia Region,
Washington, DC, 34 pp.
Evans, E.P., R. Ashley, J.W. Hall, E.C. Penning-Rowsell, A. Saul, P.B. Sayers, C.R. Thorne,
and A. Watkinson, 2004. Foresight Flood and Coastal Defence Project: Scientific
Summary: Volume I, Future risks and their drivers. Office of Science and
Technology, London, UK, 366 pp., www.bis.gov.uk/foresight/our-work/projects/
published-projects/flood-and-coastal-defence/project-outputs/volume-1.
Faccio, S.D., 2003: Effects of ice storm-created gaps on forest breeding bird communities
in central Vermont. Forest Ecology and Management, 186(1-3), 133-145.
Fankhauser, S., J.B. Smith, and R.S.J. Tol, 1999. Weathering climate change: some
simple rules to guide adaptation decisions. Ecological Economics, 30(1), 67-
78.
FAO, 2008: The State of Food Insecurity in the World 2008, High food prices and food
security – threats and opportunities. Food and Agriculture Organization of the
United Nations, Rome, Italy, 56 pp.
FAO
, 2009: The State of Agricultural Commodity Markets 2009. Food and Agriculture
Organization of the United Nations, Rome, Italy, 63 pp.
Fallon, P. and R. Betts. 2010: Climate impacts on European agriculture and water
management in the context of adaptation and mitigation – the importance of
an integrated approach. Science of Total Environment, 408, 5667-5687.
Fengqing, J., Z. Cheng, M. Guijin, H. Ruji, and M. Qingxia, 2005: Magnification of
flood disasters and its relation to regional precipitation and local human
activities since the 1980s in Xinxiang, Northwestern China. Natural Hazards,
36, 307-330.
Few, R., 2003: Flooding, vulnerability and coping strategies: local responses to a
global threat. Progress in Development Studies, 3(43), 43-58.
Few, R., M. Ahern, F. Matthies, and S. Kovats, 2004: Floods, Health and Climate
Change: A Strategic Review. Working Paper 63, Tyndall Centre for Climate
Change Research, University of East Anglia, Norwich, UK, 138 pp.
Feyen, L., J.I. Barredo, and R. Dankers, 2009: Implications of global warming and
urban land use change on flooding in Europe. Water and urban development
paradigms – towards an integration of engineering. In: Design and
Management Approaches [Feyen, J., K. Shannon, and M. Neville (eds.)]. CRC
Press, Boca Raton, FL, pp. 217-225.
Field, C.B., L.D. Mortsch, M. Brklacich, D.L. Forbes, P. Kovacs, J.A. Patz, S.W. Running,
and M.J. Scott, 2007: North America. In: Climate Change 2007: Impacts,
Adaptation and Vulnerability. Contribution of Working Group II to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change [Parry,
M.L., O.F. Canziani, J.P. Palutikof, P.J. van der Linden and C.E. Hanson (eds.)].
Cambridge University Press, Cambridge, UK, pp. 617-652.
Field, R.D., G.R. Van Der Werf, and S.S.P. Shen, 2009: Human amplification of
drought-induced biomass burning in Indonesia since 1960. Nature Geoscience,
2, 185-188.
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
280
Fink, A., T. Brücher, A. Krüger, G. Leckebusch, J. Pinto, and U. Ulbrich, 2004: The 2003
European summer heatwaves and drought – synoptic diagnosis and impacts.
Weather, 59, 209-216.
Fischlin, A., G.F. Midgley, J.T. Price, R. Leemans, B. Gopal, C. Turley, M.D.A. Rounsevell,
O.P. Dube, J. Tarazona, and A.A. Velichko, 2007: Ecosystems, their properties,
goods, and services. In: Climate Change 2007: Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani,
J.P. Palutikof, P.J. van der Linden and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 211-272.
FitzGerald, D.M., M.S. Fenster, B.A. Argow, and I.V. Buynevich, 2008: Coastal impacts
due to sea-level rise, Annual Review of Earth and Planetary Sciences, 36(1),
601-647.
Folke, C., S.R. Carpenter, T. Elmquist, L.H. Gunderson, C.S. Holling, and B.H. Walker,
2002: Resilience and sustainable development: Building adaptive capacity in a
world of transformations. Ambio, 31, 437-440.
Folke, C., S. Carpenter, B. Walker, M. Scheffer, T. Elmqvist, L. Gunderson, and C.S. Holling,
2004: Regime shifts, resilience, and biodiversity in ecosystem management.
Annual Review of Ecology, Evolution, and Systematics, 35, 557-581.
Fouillet, A, G. Rey, V. Wagner, K. Laaidi, P. Empereur-Bissonnet, A. Le Tertre, P.
Frayssinet, P. Bessemoulin, F. Laurent, P. De Crouy-Chanel, E. Jougla, D. Hémon,
2008: Has the impact of heat waves on mortality changed in France since the
European heat wave of summer 2003? A study of the 2006 heat wave.
International Journal of Epidemiology, 37(2), 309-317.
Fowler, H.J., C.G. Kilsby, and P.E. O’Connell, 2003: Modeling the impacts of climatic
change and variability on the reliability, resilience, and vulnerability of a water
resource system. Water Resources Research, 39(8), 1222-1232.
Fowler, H.J., C.G. Kilsby, and J. Stunell, 2007: Modelling the impacts of projected
future climate change on water resources in north-west England. Hydrology
and Earth System Sciences, 11(3), 1115-1124.
Frazier, T.G., N. Wood, B. Yarnal, and D.H. Bauer, 2010: Influence of potential sea level
rise on societal vulnerability to hurricane storm-surge hazards, Sarasota
County, Florida. Applied Geography, 30(4), 490-505, doi:10.1016/
j.apgeog.2010.05.005.
Freeman, P.K., 2000: Estimating chronic risk from natural disasters in developing
countries: A case study in Honduras. Presented at the Annual Bank Conference
on Development Economics – Europe Development Thinking at the Millennium,
Paris, France, 26-28June, 2000, 10 pp.
Freeman, P.K., L. Martin, R. Mechler, K. Warner, with P. Hausman, 2002: Catastrophes
and Development, Integrating Natural Catastrophes into Development
Planning. World Bank Disaster Risk Management Working Paper Series No.4,
World Bank, Washington, DC, 64 pp.
Frey, A.E., F. Olivera, J.L. Irish, L.M. Dunkin, J.M. Kaihatu, C.M. Ferreira, and B.L. Edge,
2010: The impact of climate change on hurricane flooding inundation, population
affected, and property damages. Journal of the American Water Resources
Association, 46(5), 1049-1059.
Fritze, J., G. Blashki, S. Burke, and J. Wiseman, 2008: Hope, despair and transformation:
climate change and the promotion of mental health and wellbeing.
International Journal of Mental Health Systems, 2,13.
Fuhrer, J., M. Beniston, A. Fischlin, C. Frei, S. Goyette, K. Jasper, and C. Pfister, 2006:
Climate risks and their impact on agriculture and forests in Switzerland.
Climatic Change, 79(1-2)
, 79-102.
Gaddis, E., B. Miles, S. Morse, and D. Lewis, 2007: Full-cost accounting of coastal
disasters in the United States: Implications for planning and preparedness.
Ecological Economics, 63, 307-318.
Ghesquiere, F., O. Mahul, M. Forni, and R. Gartley, 2006: Caribbean Catastrophe Risk
Insurance Facility: A solution to the short-term liquidity needs of small island
states in the aftermath of natural disasters. IAT03-13/3, World Bank,
Washington, DC, 4 pp., siteresources.worldbank.org/PROJECTS/Resources/
Catastrophicriskinsurancefacility.pdf.
Gillett, N.P., A.J. Weaver, F.W. Zwiers, and M.D. Flannigan, 2004: Detecting the effect
of climate change on Canadian forest fires. Geophysical Research Letters,
31(18), L18211, doi:10.1029/2004GL020876.
Githeko, A.K., S.W. Lindsay, U.E. Confalonieri, and J.A. Patz, 2000: Climate change
and vector-borne diseases: a regional analysis. Bulletin of the World Health
Organization, 78, 1136-1147.
Gitlin, A.R., C.M. Sthultz, M.A. Bowker, S. Stumpof, K.L. Paxton, K.A. Kennedy, A.
Munoz, J.K. Bailey, and T.G. Whitham, 2006: Mortality gradients within and
among dominant plant populations as barometers of ecosystem change
during extreme drought. Conservation Biology, 20(5), 1477-1486.
Goldstein, R., 2003: A Survey of Water Use and Sustainability in the United States
with a Focus on Power Generation. Electric Power Research Institute, Palo Alto,
CA, 64 pp.
Gössling, S. and C.M. Hall, 2006a: Tourism and Global Environmental Change.
Ecological, Social, Economic and Political Interrelationships. Routledge, London,
UK, 330 pp.
Gössling, S. and C.M. Hall, 2006b: Uncertainties in predicting tourist flows under
scenarios of climate change. Climatic Change, 79(3-4), 163-173.
Gowdy, J., 2007: Toward an experimental foundation for benefit-cost analysis.
Ecological Economics, 63, 649-655.
Granier, A., M. Reichstein, N. Bréda, I.A. Janssens, E. Falge, P. Ciais, T. Grünwald, M.
Aubinet, P. Berbigier, C. Bernhofer, N. Buchmann, O. Facini, G. Grassi, B.
Heinesch, H. Ilvesniemi, P. Keronen, A. Knohlc, B. Köstner, F. Lagergren, A.
Lindroth, B. Longdoz, D. Loustau, J. Mateusq, L. Montagnanir, C. Nyst, E. Moors,
D. Papale, M. Peiffer, K. Pilegaard, G. Pita, J. Pumpanen, S. Rambal, C. Rebmann,
A. Rodrigues, G. Seufert, J. Tenhunen, T. Vesala, and Q. Wang, 2007: Evidence for
soil water control on carbon and water dynamics in European forests during
the extremely dry year: 2003. Agricultural and Forest Meteorology, 143(1-2),
123-145.
Griffith, G.P., E.A. Fulton, and A.J. Richardson, 2011: Effects of fishing and
acidification-related benthic mortality on the southeast Australian marine
ecosystem. Global Change Biology, 17(10), 3058-3074, doi:10.1111/j.1365-
2486.2011.02453.x.
Grothmann, T. and A. Patt, 2005: Adaptive capacity and human cognition: The
process of individual adaptation to climate change. Global Environmental
Change, Part A: Human and Policy Dimensions, 3(15), 199-213.
Guha-Sapir, D., F. Vos, R. Below, with S. Ponserre, 2011: Annual Disaster Statistical
Review 2010: The numbers and trends. Centre for Research on the
Epidemiology of Disasters, Université Catholique de Louvain, Brussels, Belgium,
42 pp.,cred.be/sites/default/files/ADSR_2010.pdf.
Guimaraes, P., F.L. Hefner, and D.P. Woodward, 1993: Wealth and income effects of
natural disasters: An econometric analysis of Hurricane Hugo. Review of
Regional Studies, 23, 97-114.
Guo, Y., J. Zhang, L. Zhang, and Y. Shen, 2009: Computational investigation of
typhoon-induced storm surge in Hangzhou Bay, China. Estuarine, Coastal and
Shelf Science, 85, 530-553.
Haeberli, W., A. Kääb, D. Vonder Mühll, and P. Teysseire, 2001: Prevention of
outburst floods from periglacial lakes at Grubengletscher, Valais, Swiss Alps.
Journal of Glaciology, 47, 111-122.
Haines, A., R.S. Kovats, D. Campbell-Lendrum, and C. Corvalan, 2006: Climate change
and human health: impacts, vulnerability, and mitigation. The Lancet,367,
2101-2109.
Hall, J.W., E.P. Evans, E.C. Penning-Rowsell, P.B. Sayers, C.R. Thorne, and A.J. Saul,
2003: Quantified scenarios analysis of drivers and impacts of changing flood
risk in England and Wales: 2030-2100. Environmental Hazards, 5, 51-65.
Hall, J.W., P.B. Sayers, and R.J. Dawson, 2005: National-scale assessment of current
and future flood risk in England and Wales. Natural Hazards, 36(1-2), 147-164.
Hallegatte, S., 2007: The use of synthetic hurricane tracks in risk analysis and
climate change damage assessment. Journal of Applied Meteorology and
Climatology, 46, 1956-1966.
Hallegatte, S., 2008: A roadmap to assess the economic cost of climate change with
an application to hurricanes in the United States. In: Hurricanes and Climate
Change [Elsner, J.B. and T.H. Jagger (eds.)]. Springer-Verlag, Berlin and
Heidelberg, Germany, pp. 361-386.
Hallegatte, S., 2009: Strategies to adapt to an uncertain climate change. Global
Environmental Change, 19(2), 240-247.
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
281
Hallegatte, S. and P. Dumas, 2009: Can natural disasters have positive
consequences? Investigating the role of embodied technical change. Ecological
Economics, 68, 777-786.
Hallegatte, S. and M. Ghil, 2007: Endogenous Business Cycles and the Economic
Response to Exogenous Shocks. Working Papers 2007.20, Fondazione Eni
Enrico Mattei, Milan, Italy, 20 pp.
Hallegatte, S., J.C. Hourcade, and P. Dumas, 2007: Why economic dynamics matter
in assessing climate change damages: illustration on extreme events. Ecological
Economics, 62, 330-340.
Hallegatte, S., F. Henriet, A. Patwardhan, K. Narayanan, S. Ghosh, S. Karmakar, U.
Patnaik, A. Abhayankar, S. Pohit, J. Corfee-Morlot, C. Herweijer, N. Ranger, S.
Bhattacharya, M. Bachu, S. Priya, K. Dhore, F. Rafique, P. Mathur, and N. Naville,
2010: Flood Risks, Climate Change Impacts and Adaptation Benefits in
Mumbai: An Initial Assessment of Socio-Economic Consequences of Present and
Climate Change Induced Flood Risks and of Possible Adaptation Options. OECD
Environment Working Papers No. 27, Organisation for Economic Co-operation
and Development Publishing, Paris, France, 61 pp. dx.doi.org/10.1787/
5km4hv6wb434-en.
Hamilton, J.M., D.J. Maddison, and R.S.J. Tol, 2005: Climate change and international
tourism: a simulation study. Global Environmental Change, 15, 253-266.
Handmer, J., 2003: We are all vulnerable. Australian Journal of Emergency
Management, 18(3), 55-60.
Handmer, J. and S. Dovers, 2007: The Handbook of Disaster and Emergency Policy
and Institutions. Earthscan, London, UK, 192 pp.
Handmer, J., C. Reed, and O. Percovich, 2002: Disaster Loss Assessment Guidelines.
Department of Emergency Services, State of Queensland, Emergency
Management Australia, Commonwealth Australia, 90 pp.
Handmer, J., S. O’Neil, and D. Killalea, 2010: Review of Fatalities in the February 7,
2009, Bushfires. Victorian Bushfires Royal Commission, Melbourne, Australia.
Hanson, S., R. Nicholls, N. Ranger, S. Hallegatte, J. Corfee-Morlot, C. Herweijer, and J.
Chateau, 2011: A global ranking of port cities with high exposure to climate
extremes. Climatic Change, 104, 89-111.
Hardoy, J.E., D. Mitlin, and D. Satterthwaite, 2001: Environmental Problems in an
Urbanizing World: Finding Solutions for Cities in Africa, Asia and Latin America.
Earthscan, London, UK, 448 pp.
Harvey, N. and B. Caton, 2003: Coastal Management in Australia. Oxford University
Press, Melbourne, Australia, 342 pp.
Hashizume, M., Y. Wagatsuma, A.S. Faruque, T. Hayashi, P.R. Hunter, B.G. Armstrong,
D.A. Sack, 2008: Factors determining vulnerability to diarrhoea during and after
severe floods in Bangladesh. Journal of Water and Health, 6(3), 323-332.
Hatfield, J.L., 2010: Climate impacts on agriculture in the United States: The value of
past observations. In: Handbook of Climate Change and Agroecosystems:
Impact, Adaptation and Mitigation [Hillel, D. and C. Rosenzweig(eds.)]. Imperial
College Press, London, UK, pp. 239-253.
Hatfield, J.L., K.J. Boote, P. Fay, L. Hahn, C. Izaurralde, B.A. Kimball, T. Mader, J.
Morgan, D. Ort, W. Polley, A. Thomson, and D. Wolfe, 2008: Agriculture. In: The
Effects of Climate Change on Agriculture, Land Resources, Water Resources, and
Biodiversity in the United States. A report by the U.S. Climate Change Science
Program and the Subcommittee on Global Change Research, Washington, DC,
pp. 21-74.
Hatfield, J.L., K.J. Boote, B.A. Kimball, L. Ziska, R.C. Izaurralde, D. Ort, A.M. Thomson,
and D.A. Wolfe, 2011: Climate impacts on agriculture: Implications for crop
production. Agronomy Journal, 103(2), 351-370.
Hay, J. and N. Mimura, 2010: The changing nature of extreme weather and climate
events: risks to sustainable development. Geomatics, Natural Hazards and Risk,
1(1), 3-18.
Heger, M., A. Julca, and O. Paddison, 2008: Analysing the Impact of Natural Hazards
in Small Economies: The Caribbean Case. UNU/WIDER Research Paper 2008/25,
The World Institute for Development Economics Research, Helsinki, Finland, 27
pp.
Hein, L., M.J. Metzger, and A. Moren, 2009: Potential impacts of climate change on
tourism; a case study for Spain. Current Opinion in Environmental
Sustainability, 1, 170-178.
Hennessy, K., C. Lucas, N. Nicholls, J. Bathols, R. Suppiah, and J. Ricketts, 2006:
Climate change impacts on fire-weather in south-east Australia. CSIRO Marine
and Atmospheric Research, Bushfire CRC and Australia Bureau of Meteorology,
Clayton South Victoria, Australia, 88 pp.
Hennessy, K., B. Fitzharris, B.C. Bates, N. Harvey, S.M. Howden, L. Hughes, J. Salinger,
and R. Warrick, 2007: Australia and New Zealand. In: Climate Change 2007.
Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the
Fourth Assessment Report of the Intergovernmental Panel on Climate Change
[Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der Linde, and C.E. Hanson
(eds.)]. Cambridge University Press, Cambridge, UK, pp. 507-540.
Hess, J.J., J.N. Malilay, and A.J. Parkinson, 2008: Climate change: The importance of
place. American Journal of Preventive Medicine, 35(5), 468-478.
Hilker, N., A. Badoux, and C. Hegg, 2009: The Swiss flood and landslide damage
database 1972–2007. Natural Hazards and Earth System Sciences, 9, 913-925.
Hinkel, J., R.J. Nicholls, A.T. Vafeidis, R.S.J. Tol, and T. Avagianou, 2010: Assessing risk
of and adaptation to sea-level rise in the European Union: an application of
DIVA. Mitigation and Adaptation Strategies for Global Change, 15, 703-719.
Hirabayashi, Y. and S. Kanae, 2009: First estimate of the future global population at
risk of flooding. Hydrological Research Letters, 3, 6-9.
Hjelle, B. and G.E. Glass, 2000: Outbreak of Hantavirus infection in the Four Corners
region of the United States in the wake of the 1997-1998 El Niño-Southern
Oscillation. Journal of Infectious Diseases, 181(5), 1569-1573.
Hlavinka, P., M. Trnka, D. Semeradova, M. Dubrovsky, Z. Zalud, and M. Mozny, 2009:
Effect of drought on yield variability of key crops in Czech Republic. Agricultural
and Forest Meteorology, 149, 431-442.
Hochrainer, S., and R. Mechler, 2011: Natural disaster risk in Asian megacities. A
case for risk pooling? Cities, 28(1), 53-61.
Hoes, O.A.C. 2007: Aanpak wateroverlast in polders op basis van risicobeheer. Ph.D.
thesis, Delft University of Technology, Delft, The Netherlands, 188 pp.,
repository.tudelft.nl/assets/uuid:69de94f9-385c-4c50-beff-8cc7b955beba/
ceg_hoes_20070319.pdf.
Hoes, O., and W. Schuurmans, 2006: Flood standards or risk analyses for polder
management in The Netherlands. Irrigation and Drainage, 55
, S113-S119.
Hoes, O.A.C., W. Schuurmans, and J. Strijker, 2005: Water systems and risk analysis.
Water Science and Technology, 51, 105-112.
Hoffman, R.N., P. Dailey, S. Hopsch, R.M. Ponte, K. Quinn, E.M. Hill, and B. Zachry,
2010: An estimate of increases in storm surge risk to property from sea level
rise in the first half of the twenty-first century. Weather, Climate and Society, 2,
271-293.
Holling, C.S. (ed.), 1978: Adaptive environmental assessment and management.
John Wiley, New York, NY, 377 pp.
Holper, P.N., S. Lucy, M. Nolan, C. Senese, and K. Hennessy, 2007: Infrastructure and
Climate Change Risk Assessment for Victoria. Consultancy Report to the
Victorian Government prepared by CSIRO, Maunsell Australia, and Phillips Fox,
Aspendale, Victoria, Australia,84 pp.
Höppe, P. and T. Grimm, 2009: Rising natural catastrophe losses: What is the role of
climate change? In: Economics and Management of Climate Change: Risks,
Mitigation and Adaptation [Hansjürgens, B. and R. Antes (eds.)]. Springer-
Verlag, Heidelberg, Germany, pp. 13-22.
Huey, C.A.D., J.J. Tewksbury, L.J. Vitt, P.E. Hertz, H.J.A. Pérez, and T. Garland, Jr., 2009:
Why tropical forest lizards are vulnerable to climate warming. Proceedings of
the Royal Society, Biology, 276, 1939-1948.
Huggel, C., W. Haeberli, A. Kaab, D. Bieri, and S. Richardson, 2004: An assessment
procedure for glacial hazards in the Swiss Alps. Canadian Geotechnical Journal,
41, 1068-1083.
Hughes, T.P., A.H. Baird, D.R. Bellwood, M. Card, S.R. Connolly, C. Folke, R. Grosberg,
O. Hoegh-Guldberg, J.B.C. Jackson, J. Kleypas, J.M. Lough, P. Marshall, M.
Nystrom, S.R. Palumbi, J.M. Pandolfi, B. Rosen, and J. Roughgarden, 2003:
Climate change, human impacts, and the resilience of coral reefs. Science, 301,
929-933.
Huigen, M.G.A. and I.C. Jens, 2006: Socio-economic impact of Super Typhoon
Harurot in San Mariano, Isabela, the Philippines. World Development, 34(12),
2116-2136.
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
282
Huq, S., S. Kovats, H. Reid, and D. Satterthwaite, 2007: Editorial: Reducing risks to cities
from disasters and climate change. Environment and Urbanization, 19(1), 3-15.
Hurkman, W.J., W.H. Vensel, C.K. Tanaka, L. Whitehead, and S.B. Altenbach, 2009:
Effect of high temperature on albumin and globulin accumulation in the
endosperm proteome of the developing wheat grain. Journal of Cereal Science,
49, 12-23.
Iglesias, A., L. Garrote, and F. Martín-Carrasco, 2009: Drought risk management in
Mediterranean river basins. Integrated Environmental Assessment and
Management, 5, 11-16, doi:10.1897/IEAM_2008-044.1.
Innes, J.L., 1992: Observations on the condition of beech (Fagus sylvatica L.) in
Britain in 1990. Forestry, 65(1), 35-60.
IPCC, 2001: Climate Change 2001: Impacts, Adaptation and Vulnerability.
Contribution of Working Group II to the Third Assessment Report of the
Intergovernmental Panel on Climate Change [McCarthy, J.J., O.F. Canziani, N.A.
Leary, D.J. Dokken, and K. S. White (eds.)]. Cambridge University Press,
Cambridge, UK, 1032 pp.
IPCC, 2007: Climate Change 2007: Impacts, Adaptation and Vulnerability.
Contribution of Working Group II to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P.
Palutikof, P.J. van der Linden, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, 976 pp.
IPCC, 2010: Meeting Report of the Intergovernmental Panel on Climate Change
Expert Meeting on Detection and Attribution Related to Anthropogenic Climate
Change [Stocker, T.F., C.B. Field, D. Qin, V. Barros, G.-K. Plattner, M. Tignor, P.M.
Midgley, and K.L. Ebi (eds.)]. IPCC Working Group I Technical Support Unit,
University of Bern, Bern, Switzerland, 55 pp.
Irish, J.L., A.E. Frey, J.D. Rosati, F. Olivera, L.M. Dunkin, J.M. Kaihatu, C.M. Ferreira, and
B.L. Edge, 2010: Potential implications of global warming and barrier island
degradation on future hurricane inundation, property damages, and population
impacted. Ocean & Coastal Management, 53(10), 645-657, doi:10.1016/
j.ocecoaman.2010.08.001.
Jones, B.M., C.D. Arp, M.T. Jorgenson, K.M. Hinkel, J.A. Schmutz, and P.L. Flint, 2009:
Increase in the rate and uniformity of coastline erosion in Arctic Alaska.
Geophysical Research Letters, 36, L03503, doi:10.1029/2008GL036205.
Jones, R.N., 2004: Managing climate change risks. In: The Benefits of Climate
Change Policies: Analytical and Framework Issues [Agrawal, S. and J. Corfee-
Morlot (eds.)]. Organisation for Economic Co-operation and Development,
Paris, France, pp. 249-297.
Jonkman, S.N., 2007: Loss of Life Estimation in Flood Risk Assessment: Theory And
Applications. Ph.D. Thesis, Delft University of Technology, Delft, The
Netherlands, 354 pp.
Jonkman, S.N., M. Bočkarjova, M. Kok, and P. Bernardini, 2008: Integrated
hydrodynamic and economic modelling of flood damage in The Netherlands.
Ecological Economics, 66, 77-90.
Kahn, M.E., 2005: The death toll from natural disasters: The role of income, geography
and institutions. The Review of Economics and Statistics, 87(2), 271-284.
Kalkstein, L., S. Greene, D. Mills, and J. Samenow, 2011: An evaluation of the
progress in reducing heat-related human mortality in major U.S. cities. Natural
Hazards, 56(1), 113-129, doi:10.1007/s11069-010-9552-3.
Kaplan
, I.C., P.S. Levin, M. Burden, and E.A. Fulton, 2010: Fishing catch shares in the
face of global change: a framework for integrating cumulative impacts and
single species management. Canadian Journal of Fisheries and Aquatic
Sciences, 67, 1968-1982.
Karl, T.R., J.M. Melillo, and T.C. Peterson (eds.), 2009: Global Climate Change Impacts
in the United States. U.S. Global Change Research Program, Cambridge
University Press, New York, NY, 188 pp., downloads.globalchange.gov/
usimpacts/pdfs/climate-impacts-report.pdf.
Karoly, D., 2010: The recent bushfires and extreme heat wave in southeast
Australia. Bulletin of Australian Meteorological and Oceanographical Societies,
22, 10-13.
Kazama, S., T. Kono, K. Kakiuchi, and M. Sawamoto, 2009: Evaluation of flood control
and inundation conservation in Cambodia using flood and economic growth
models. Hydrological Processes, 23, 623-632.
Kellenberg, D.K. and A.M. Mobarak, 2008: Does rising income increase or decrease
damage risk from natural disasters? Journal of Urban Economics, 63(3), 788-
802.
Kettlewell, P.S., R.B. Sothern, and W.L. Koukkari, 1999: U.K. wheat quality and
economic value are dependent on the North Atlantic Oscillation. Journal of
Cereal Science, 29, 205-209.
Kim, H.Y., T. Horie, H. Nakagawa, and K. Wada, 1996: Effects of elevated CO
2
concentration and high temperature on growth and yield of rice. II. The effect
of yield and its component of Akihikari rice. Japanese Journal of Crop Science,
65, 644-651.
Kim, S., Y. Tachikawa, E. Nakakita, and K. Takara, 2009: Reconsideration of reservoir
operations under climate change: case study with Yagisawa Dam, Japan.
Annual Journal of Hydraulic Engineering, JSCE, 53, 597-611.
Kim, U. and J.J. Kaluarachchi, 2009: Climate change impacts on water resources in the
upper blue Nile river basin, Ethiopia. Journal of the American Water Resources
Association, 45(6), 1361-1378.
Kleinen, T. and G. Petschel-Held, 2007: Integrated assessment of changes in flooding
probabilities due to climate change. Climatic Change, 81(3), 283-312.
Knowlton, K., M. Rotkin-Ellman, G. King, H.G. Margolis, D. Smith, G. Solomon, R. Trent,
and P. English, 2008: The 2006 California heatwave: Impacts on hospitalizations
and emergency department visits. Environmental Health Perspectives, 117(1),
61-67.
Koppe, C., G. Jendritzky, R.S. Kovats, and B. Menne, 2004: Heat Waves: Risks and
Responses. Health and Global Environmental Change, Series No. 2, World
Health Organization, WHO Regional Office for Europe, Copenhagen, Denmark,
123 pp., www.euro.who.int/__data/assets/pdf_file/0008/96965/E82629.pdf.
Korf, B., 2004: War, livelihoods and vulnerability in Sri Lanka. Development and
Change, 35(2), 275-295.
Kosatsky, T., 2005: The 2003 European heatwaves. Euro Surveillance, 10(7), 148-149.
Kovats, R.S. and R. Akhtar, 2008: Climate, climate change and human health in Asian
cities. Environment and Urbanization, 20(1), 165-175.
Kovats, R.S. and K. Ebi, 2006: Heatwaves and public health in Europe. European
Journal of Public Health, 16(6), 592-599.
Kramer, R.A. 1995. Advantages and limitations of benefit-cost analysis for evaluating
investments in natural disaster mitigation. In: Disaster Prevention for
Sustainable Development: Economic and Policy Issues [Munasinghe, M. and C.
Clarke (eds.)]. World Bank, Washington, DC, pp. 61-76.
Kshirsagar, N., R. Shinde, and S. Mehta, 2006: Floods in Mumbai: Impact of public
health service by hospital staff and medical students. Journal of Postgraduate
Medicine, 52(4), 312.
Kumar, R., R. Venuprasad, and G.N. Atlin, 2007: Genetic analysis of rainfed lowland
rice drought tolerance under naturally-occurring stress in eastern India:
Heritability and QTL effects. Field Crops Research, 103, 42-52.
Kundzewicz, Z., 2003: Water and climate – The IPCC TAR perspective. Nordic
Hydrology, 34(5), 387-398.
Kundzewicz, Z.W. and H.-J. Schellnhuber, 2004: Floods in the IPCC TAR perspective.
Natural Hazards,
31, 111-128.
Kundzewicz, Z.W., L.J. Mata, N. Arnell, P. Döll, P. Kabat, B. Jiménez, K. Miller, T. Oki,
Z. Sen, and I. Shiklomanov, 2007: Freshwater resources and their management.
Climate Change 2007. Impacts, Adaptation and Vulnerability. Contribution of
Working Group II to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. van der
Linden, and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK,
pp.173-210.
Kunst, A.E., C.W.N. Looman, and J.P. Mackenbath, 1993: Outdoor air temperature
and mortality in The Netherlands: a time-series analysis. American Journal of
Epidemiology, 137, 331-341.
Kurukulasuriya, P. and R. Mendelsohn, 2008a: Crop switching as a strategy for
adapting to climate change. African Journal of Agricultural and Resource
Economics, 2(1), 105-126.
Kurukulasuriya, P. and R. Mendelsohn, 2008b: How Will Climate Change Shift Agro-
Ecological Zones and Impact African Agriculture? Policy Research Working
Paper Series 4717, World Bank, Washington, DC, 31 pp.
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
283
L’Hôte, Y., G. Mahé, B. Some, and J.P. Triboulet, 2002: Analysis of a Sahelian annual
rainfall index from 1896 to 2000: the drought continues. Hydrological Science
Journal, 47, 563-572.
Lal, P.N., 2010. Vulnerability to natural disasters: an economic analysis of the impact
of the 2009 floods on the Fijian sugar belt. Pacific Economic Bulletin, 25(2),
62-77.
Langner, J., R. Bergström, and V. Foltescu, 2005: Impact of climate change on surface
ozone and deposition of sulphur and nitrogen in Europe. Atmospheric
Environment, 39(6), 1129-1141.
Laschewski, G. and G. Jendritzky, 2002: Effects of the thermal environment on human
health: an investigation of 30 years of daily mortality data from southwest
Germany. Climate Research, 21, 91-103.
Lavorel, S. and W. Steffen, 2004: Cascading impacts of land use through time: the
Canberra bushfire disaster. In: Global Change and the Earth System: A Planet
Under Pressure [Steffen, W., A. Sanderson, P.D. Tyson, J. Jäger, P.A. Matson, B.
Moore III, F. Oldfield, K. Richardson, H.J. Schellnhuber, B.L. Turner II, and R.J.
Wasson (eds.)]. Springer-Verlag, Berlin, Germany, pp. 186-188.
Leckebusch, G.C., U. Ulbrich, L. Fröhlich, and J.G. Pinto, 2007: Property loss potentials
for European midlatitude storms in a changing climate. Geophysical Research
Letters, 34, L05703, doi:10.1029/2006GL027663.
Legesse, D., C. Vallet-Coulomb, and F. Gasse, 2003: Hydrological response of a
catchment to climate and land use changes in Tropical Africa: case study South
Central Ethiopia. Journal of Hydrology, 275, 67-85.
Lehner, B., P. Doell, J. Alcamo, T. Henrichs, and F. Kaspar, 2006: Estimating the impact
of global change on flood and drought risks in Europe: A continental, integrated
analysis. Climatic Change, 75(3), 273-299.
Lemke, P., J. Ren, R.B. Alley, I. Allison, J. Carrasco, G. Flato, Y. Fujii, G. Kaser, P. Mote,
R.H. Thomas, and T. Zhang, 2007: Observations: Changes in snow, ice and frozen
ground. In: Climate Change 2007: The Physical Science Basis. Contribution of
Working Group I to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M.
Marquis, K.B. Averyt, M. Tignor and H.L. Miller (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 337-383
Lemmen, D., F. Warren, J. Laeroix, and E. Bush (eds.), 2008: From Impacts to
Adaptation: Canada in a Changing Climate 2007. Climate Change Impacts and
Adaptation Division, Earth Sciences Sector, Natural Resources Canada, Ottawa,
Canada, 448 pp.
Lenton, T., A. Footitt, and A. Dlugolecki, 2009: Major Tipping Points in the Earth’s
Climate System and Consequences for the Insurance Sector. World Wide
Fund for Nature, Gland, Switzerland and Allianz SE, Munich, Germany, 89
pp.
Létard, V., H. Flandre, and S. Lepeltier, 2004: France and the French response to the
heatwave: Lessons from a crisis. In: Information report of the Senate number
195(2003-2004), Paris, France, 391 pp. (in French).
Levy, B.S. and V.W. Sidel (eds.), 2000: War and Public Health. American Public Health
Association, Washington DC, 436 pp.
Li, L.H., H.G. Xu, X. Chen, and S.P. Simonovic, 2010: Streamflow forecast and reservoir
operation performance assessment under climate change. Water Resources
Management, 24(1), 83-104.
Lim, J.H., J.H. Chun, and J.H. Shin, 2010: Global warming related dieback of evergreen
coniferous forests in Korea due to high temperature and drought stress in winter
season. Proceedings of the XXIII IUFRO 2010 World Congress, 23-28 August,
Seoul. The International Forestry Review, 12(5), 44-44.
Lindenmayer, D., D. Blair, L. McBurney, and S. Banks, 2010: Forest Phoenix: How A
Great Forest Recovers After Wildfire. CSIRO, Melbourne, Australia.
Lise, W. and R.S.J. Tol, 2002: Impact of climate on tourism demand. Climatic Change,
55(4), 429-449.
Liu, X.P., J.Q. Zhang, D.W. Zhou, Z.S. Song, and X.T. Wuet, 2006: Study on grassland
fire risk dynamic distribution characteristic and management policy. Chinese
Journal of Grassland, 28(6), 77-83 (in Chinese).
Loayza, N., O. Eduardo, R. Jamele, and L. Christiaensen, 2009: Natural Disasters and
Growth - Going Beyond the Averages. Policy Research Working Paper Series
4980, World Bank, Washington, DC, 40 pp.
Lobell, D.B., M.B. Burke, C. Tebaldi, M.D. Mastrandrea, W.P. Falcon, and R.L. Naylor,
2008: Prioritizing climate change adaptation needs forfood security in 2030.
Science, 319, 607.
Lobell, D.B., W. Schlenker, and J. Costa-Roberts, 2011: Climate trends and global crop
production since 1980. Science,333(6042), 616-620, doi:10.1126/science.1204531.
Love, G., A. Soares, and H. Püempel, 2010: Climate change, climate variability and
transportation. Procedia Environmental Sciences, 1, 130-145.
Loveridge, F.A., T.W. Spink, A.S. O’Brien, K.M. Briggs, and D. Butcher, 2010: The
impact of climate and climate change on infrastructure slopes, with particular
reference to southern England. Quarterly Journal of Engineering Geology and
Hydrogeology, 43(4), 461-472.
Lugeri, N., Z.W. Kundzewicz, E. Genovese, S. Hochrainer, and M. Radziejewski, 2010:
River flood risk and adaptation in Europe – assessment of the present status.
Mitigation and Adaptation Strategies for Global Change, 15(7), 621-639,
doi:10.1007/s11027-009-9211-8.
Luo, Q., M.A.J. Williams, W. Bellotti, and B. Bryan, 2003: Quantitative and visual
assessments of climate change impacts on South Australian wheat production.
Agricultural Systems, 77, 173-186.
Lyon, B., 2003: Enhanced seasonal rainfall in northern Venezuela and the extreme
events of December 1999. Journal of Climate, 16, 2302-2306.
Maaskant, B., S.N. Jonkman, and L.M. Bouwer, 2009: Future risk of flooding: an
analysis of changes in potential loss of life in South Holland (The Netherlands).
Environmental Science and Policy, 12, 157-169.
MacDonald, A.M., R.C. Calow, D.M.J. MacDonald, W.G. Darling, and B.E.O.
Dochortaigh, 2009: What impact will climate change have on rural groundwater
supplies in Africa? Hydrological Sciences Journal, 54(4), 690-703.
MacDonald, G.M., 2010: Climate change and water in Southwestern North America
special feature: water, climate change, and sustainability in the southwest.
Proceedings of the National Academy of Sciences, 107, 21256-21262.
Magrin, G., C. Gay García, D. Cruz Choque, J.C. Giménez, A.R. Moreno, G.J. Nagy, C.
Nobre, and A. Villamizar, 2007: Latin America. In: Climate Change 2007.
Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the
Fourth Assessment Report of the Intergovernmental Panel on Climate Change
[Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der Linde, and C.E. Hanson
(eds.)]. Cambridge University Press, Cambridge, UK, pp. 581-615.
Maloney, S.K. and C.F. Forbes, 2011: What effect will a few degrees of climate
change have on human heat balance? Implications for human activity.
International Journal of Biometeorology,55(2), 147-160.
Mark, B.G. and G.O. Seltzer, 2003: Tropical glacier meltwater contribution to stream
discharge: a case study in the Cordillera Blanca, Peru. Journal of Glaciology, 49,
271-281.
Markandya, A. and P. Watkiss, 2009: Potential Costs and Benefits of Adaptation
Options: A Review of Existing Literature. Technical paper prepared for the United
Nations Framework Convention on Climate Change, UNFCCC, Bonn, Germany.
Marques, S.C., U.M. Azeiteiro, F. Martinhoa, and M.A. Pardala, 2007: Climate
variability and planktonic communities: The effect of an extreme event (severe
drought) in a southern European estuary. Estuarine Coastal and Shelf Science,
73(3-4), 725-734.
McClung, D.M. and Schaerer, P., 2006: The Avalanche Handbook. The Mountaineers
Books, Seattle, WA, 342 pp.
McGranahan, G., D. Balk, and B. Anderson, 2007: The rising tide: assessing the risks
of climate change and human settlements in low elevation coastal zones.
Environment and Urbanization, 19(1), 17-37.
McGregor, G., M. Pelling, T. Wolf, and S. Gosling, 2007: The Social Impacts of Heat
Waves. Science Report SC20061/SR6, Environment Agency, Bristol, UK.
McInnes, K.L., K.J.E. Walsh, G.D. Hubbert, and T. Beer, 2003: Impact of sea-level rise
and storm surges on a coastal community. Natural Hazards, 30, 187-207.
McKechnie, A.E. and B.O. Wolf, 2010: Climate change increases the likelihood of
catastrophic avian mortality events during extreme heatwaves. Biology Letters,
6(2), 253-256.
McKenzie, E., B. Prasad, and A. Kaloumaira, 2005: Economic Impact of Natural
Disasters on Development in the Pacific. Volume I, South Pacific Applied
Geoscience Commission and University of the South Pacific, Suva, Fiji.
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
284
McMaster, H.J., 1999: The potential impact of global warming on hail losses to
winter cereal crops in New South Wales. Climatic Change, 43, 455-476.
McMichael, A.J., D.H. Campbell-Lendrum, C.F. Corvalán, K.L. Ebi, A.K. Githeko, J.D.
Scheraga, and A. Woodward (eds.), 2003a: Climate Change and Human
Health: Risks and Responses. World Health Organization, Geneva, Switzerland,
322 pp.
McMichael, A., R. Woodruff, P. Whetton, K. Hennessy, N. Nicholls, S. Hales, A.
Woodward, and T. Kjellstrom, 2003b: Human Health and Climate Change in
Oceania: A Risk Assessment 2002. Commonwealth Department of Health and
Ageing, Canberra, Australia, 128 pp.
McMichael, A.J., R.E. Woodruff, and S. Hales, 2006: Climate change and human
health: present and future risks. The Lancet, 367, 859-869.
McMichael, A.J., P. Wilkinson, R.S. Kovats, S. Pattenden, B. Armstrong, N.
Vajanapoom, E.M. Niciu, H. Mahomed, C. Kingkeow, M. Kosnik, M.S. O’Neill, I.
Romieu, M. Ramirez-Aguilar, M.L. Barreto, N. Gouveia, and B. Nikiforov, 2008:
International study of temperature, heat and urban mortality: the ‘ISOTHURM’
project. International Journal of Epidemiology, 37, 1121-1131.
Mechler, R., 2004: Natural Disaster Risk Management and Financing Disaster Losses
in Developing Countries. Verlag Versicherungswirtsch, Karlsruhe, Germany, 235
pp.
Mechler, R., 2008: From Risk to Resilience. The Cost-Benefit Analysis Methodology.
Working Paper 1, Provention Consortium, Geneva, Switzerland.
Mechler, R. and Z.W. Kundzewicz, 2010: Assessing adaptation to extreme weather
events in Europe. Mitigation and Adaptation Strategies for Global Change,
15(7), 611-620.
Meehl, G.A. and C. Tebaldi, 2004: More intense, more frequent, and longer lasting
heat waves in the 21st century. Science, 305(5686), 994-997.
Meizhu, C., X. Guangji, W. Shaopeng, and Z. Shaoping, 2010: High-temperature
hazards and prevention measurements for asphalt pavement. In: Proceedings
of the International Conference on Mechanic Automation and Control
Engineering (MACE), Wuhan, China, 26-28 June 2010.Institute of Electrical and
Electronics Engineers, New York, NY, pp. 1341-1344, doi:10.1109/
MACE.2010.5536275.
Mendelsohn, R., 2007: What causes crop failure? Climatic Change, 81(1), 61-70.
Mendelsohn, R., A. Dinar, A. Dafelt, 2000: Climate Change Impacts on African
Agriculture. World Bank, Washington, DC, 25 pp., www.ceepa.co.za/
Climate_Change/pdf/%285-22-01%29afrbckgrnd-impact.pdf.
Meyer, M.D., 2008: Design Standards for U.S. Transportation Infrastructure. The
Implications of Climate Change. Accession number 01104982, Transportation
Research Board of the National Academies, Washington, DC, 30 pp.,
onlinepubs.trb.org/onlinepubs/sr/sr290Meyer.pdf.
Micklin, P., 2007: The Aral Sea disaster. Annual Review of Earth and Planetary
Sciences, 35, 47-72.
Miller, S., R. Muir-Wood, and A. Boissonnade, 2008: An exploration of trends in
normalised weather-related catastrophe losses. In: Climate Extremes and
Society [Diaz, H.F. and R.J. Murnane (eds.)]. Cambridge University Press,
Cambridge, UK, pp. 225-247.
Mills, E., 2005: Insurance in a climate of change. Science, 309, 1040-1044.
Mimura, N., L. Nurse, R.F. McLean, J. Agard, L. Briguglio, P. Lefale, R. Payet, and G.
Sem, 2007: Small islands. In: Climate Change 2007. Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani,
J.P. Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 687-716.
Ministry for the Environment, 2008: Coastal Hazards and Climate Change. A
Guidance Manual for Local Government in New Zealand. 2nd ed. Revised by
Ramsay, D. and R. Bell, Ministry of Agriculture and Forestry, Government of New
Zealand, Wellington, New Zealand, 129 pp., mfe.govt.nz/publications/
climate/coastal-hazards-climate-change-guidance-manual/coastal-hazards-
climate-change-guidance-manual.pdf .
Ministry of Agriculture and Forestry, 2005: Finland’s National Strategy for
Adaptation to Climate Change. Ministry of Agriculture and Forestry of Finland,
Helsinki, Finland, 281 pp.
Ministry of Environment and Forest, 2005: National Adaptation Programme of
Action (NAPA), Final Report. Government of the People’s Republic of
Bangladesh, Dhaka, Bangladesh.
Mirvis, V.M., 1999: Regularities of a change in the air temperature regime in Russia.
In: Investigations in the Main Geophysical Observatory. Gidrometeoizdat, St.
Petersburg, Russia.
Mitchell, J.K., 1998: Introduction: Hazards in changing cities. Applied Geography,
18, 1-6.
Moench, M., R. Mechler, and S. Stapleton, 2007. Guidance note on the costs and
benefits of disaster risk reduction. In: UNISDR High Level Platform on Disaster
Risk Reduction, Geneva, Switzerland, 4-7 June 2007. International Strategy for
Disaster Reduction, Geneva, Switzerland, 29 pp.
Mohammed, A.R. and L. Tarpley, 2009: High nighttime temperatures affect rice
productivity through altered pollen germination and spikelet fertility.
Agricultural and Forest Meteorology, 149, 999-1008.
Mokrech, M., R.J. Nicholls, J.A. Richards, C. Henriques, I.P. Holman, and S. Shackley,
2008: Regional impact assessment of flooding under future climate and
socioeconomic scenarios for East Anglia and North West England. Climatic
Change, 90, 31-55.
Molesworth, A.M., L.E. Cuevas, S.J. Connor, A.P. Morse, and M.C. Thomson, 2003:
Environmental risk and meningitis epidemics in Africa. Emerging Infectious
Diseases, 9, 1287-1293.
Moline, M.A., N.J. Karnovsky, Z. Brown, G.J. Divoky, T.K. Frazer, C.A. Jacoby, J.J.
Torres, and W.R. Fraser, 2008: High latitude changes in ice dynamics and their
impact on polar marine ecosystems. Annals of the New York Academy of
Sciences, 1134, 267-319, doi:10.1196/annals.1439.010.
Molua, E.L. and C.M. Lambi, 2006: The Economic Impact of Climate Change on
Agriculture in Cameroon. CEEPA Discussion Paper No. 17, Centre for
Environmental Economics and Policy in Africa, Pretoria, South Africa, 33 pp.
Morin, E., T. Grodek, O. Dahan, G. Benito, C. Kulls, Y. Jacoby, G. Van Langenhove, M.
Seely, and Y. Enzel, 2009: Flood routing and alluvial aquifer recharge along the
ephemeral arid Kuiseb River, Namibia. Journal of Hydrology, 368, 262-275.
Morrissey, S.A. and J.P. Reser, 2007: Natural disasters, climate change and mental
health considerations for rural Australia. Australian Journal of Rural Health, 15
,
120-125.
Morton, D.C., M.E. Roessing, A.E. Camp, and M.L. Tyrrell, 2003: Assessing the
Environmental, Social, and Economic Impacts of Wildfire. Research Paper 001,
Yale University Global Institute of Sustainable Forestry, New Haven, CT, 54 pp.
Mouillot, F., S. Rambal, and R. Joffre, 2002: Simulating climate change impacts on
fire frequency and vegetation dynamics in a Mediterranean-type ecosystem.
Global Change Biology, 8, 423-437.
Moulin, C., C.E. Lambert, F. Dulac, and U. Dayan, 1997: Control of atmospheric export
of dust from North Africa by the North Atlantic Oscillation. Nature, 387, 691-694.
Mozumder, P., E. Flugman, and T. Randhir, 2011: Adaptation behavior in the face of
global climate change: Survey responses from experts and decision makers
serving the Florida Keys. Ocean & Coastal Management, 54(1), 37-44,
doi:10.1016/j.ocecoaman.2010.10.008.
Muller, C., W. Cramer, W.L. Hare, and H. Lotze-Campen, 2011: Climate change risks
for African agriculture. Proceedings of the National Academy of Sciences, 108,
4313-4315.
Munich Re, 2011: TOPICS GEO, Natural catastrophes 2010, Analyses, assessments,
positions. Munich Reinsurance Company, Munich, Germany,
www.munichre.com/publications/302-06735_en.pdf
Narita, D., R.S.J. Tol, and D. Anthoff, 2009: Damage costs of climate change through
intensification of tropical cyclone activities: an application of FUND. Climate
Research, 39, 87-97.
Narita, D., R.S.J. Tol, and D. Anthoff, 2010: Economic costs of extratropical storms
under climate change: an application of FUND. Journal of Environmental
Planning and Management, 53, 371-384.
NCDC, 2005: State of the Climate: Hurricanes & Tropical Storms for Annual 2005.
National Oceanic and Atmospheric Administration, National Climatic Data
Center, Washington, DC, 7 pp., www.ncdc.noaa.gov/sotc/tropical-cyclones/
2005/13.
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
285
NCDC, 2011: State of the Climate: Global Hazards for December 2010. National
Oceanic and Atmospheric Administration, National Climatic Data Center,
Washington, DC, lwf.ncdc.noaa.gov/sotc/hazards/2010/12.
Neria, Y., A. Nandi, and S. Galea, 2008. Post-traumatic stress disorder following
disasters: a systematic review. Psychological Medicine, 38, 467-480.
Neumayer, E. and F. Barthel, 2011: Normalizing economic loss from natural disasters:
a global analysis. Global Environmental Change, 22(1), 13-24.
Neumayer, E. and T. Plümper, 2007: The gendered nature of natural disasters: The
impact of catastrophic events on the gender gap in life expectancy, 1981-2002.
Annals of the Association of American Geographers, 97(3), 551-566.
New, M., 2002: Climate change and water resources in the southwestern Cape,
South Africa. South African Journal of Science, 98, 369-373.
Niall, S. and K. Walsh, 2005: The impact of climate change on hailstorms in
southeastern Australia. International Journal of Climatology, 25, 1933-1952.
Nicholls, R.J., P.P. Wong, V.R. Burkett, J.O. Codignotto, J.E. Hay, R.F. McLean, S.
Ragoonaden, and C.D. Woodroffe, 2007: Coastal systems and low-lying areas.
In: Climate Change 2007. Impacts, Adaptation and Vulnerability. Contribution of
Working Group II to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. van der
Linden, and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK, pp.
315-356.
Nicholls, R.J., S. Hanson, C. Herweijer, N. Patmore, S. Hallegatte, J. Corfee-Morlot, J.
Château, and R. Muir-Wood, 2008: Ranking Port Cities With High Exposure And
Vulnerability to Climate Extremes: Exposure Estimates. OECD ENV/WKP 2007-
1, Organisation for Economic Co-operation and Development, Paris, France, 62
pp.
Nicholson, S.E., X. Yin, and M.B. Ba, 2000: On the feasibility of using a lake water
balance model to infer rainfall: An example from Lake Victoria. Hydrological
Sciences Journal, 45, 75-95.
Nilsson, C., I. Stjerquist, L. Bärring, P. Schlyter, A.M. Jönsson, and H. Samuelson, 2004:
Recorded storm damage in Swedish forests 1901-2000. Forest Ecology and
Management, 199, 165-173.
Nkomo, J.C. and G. Bernard, 2006: Estimating and Comparing Costs and Benefits of
Adaptation Projects: Case Studies in South Africa and The Gambia. AIACC
Project No. AF 47, Assessments of Impacts and Adaptations to Climate Change,
International START Secretariat, Washington, DC, 137 pp.
Nordhaus, W.D., 2010: The economics of hurricanes and implications of global
warming. Climate Change Economics, 1, 1-20.
Nordhaus, W.D. and J. Boyer, 2000: Warming the World: Economic Models of Global
Warming. MIT Press, Cambridge, MA, 232 pp.
Nöthiger, C. and H. Elsasser, 2004: Natural hazards and tourism: new findings on
the European Alps. Mountain Research and Development, 24(1), 24-27.
Noy, I., 2009: The macroeconomic consequences of disasters. Journal of Development
Economics, 88(2), 221-231.
NRC, 2008: Potential Impacts of Climate Change on U.S. Transportation. National
Research Council, Transportation Research Board Special Report 290, National
Academy of Sciences, Washington, DC.
NWS, 2011: National Weather Service Natural Hazard Statistics. National Oceanic
and Atmospheric Administration, National Weather Service, Office of Climate,
Water, and Weather Services, Silver Spring, MD, www.nws.noaa.gov/om/
hazstats.shtml.
OCDESC, 2007: National Hazardscape Report. Officials Committee for Domestic and
External Security Coordination, Department of the Prime Minister and Cabinet,
Wellington, New Zealand, 139 pp.
OECD-FAO, 2008: OECD-FAO Agricultural Outlook 2008-2017, Highlights.
Organisation for Economic
Co-operation and Development and Food and Agriculture Organization, Paris,
France, 72 pp., www.fao.org/es/ESC/common/ecg/550/en/AgOut2017E.pdf.
Oh, C.H. and R. Reuveny, 2010: Climatic natural disasters, political risk, and
international trade. Global Environmental Change, 20, 243-254.
Ojeda, J. and L. Donnelly, 2006: Landslides in Colombia and their Impact on Towns
and Cities. IAEG2006 Paper 112, Geological Society of London, London, UK, 13
pp., iaeg.info/iaeg2006/papers/iaeg_112.pdf.
Okuyama, Y. and S. Sahin, 2009: Impact Estimation of Disasters: A Global Aggregate
for 1960 to 2007. World Bank Policy Research Working Paper 4963, World Bank,
Washington, DC, 40 pp.
Opdam, P. and D. Wascher, 2004: Climate change meets habitat fragmentation: linking
landscape and biogeographical scale levels in research and conservation.
Biological Conservation, 117, 285-297.
Osman-Elasha, B., M. Medany, I. Niang-Diop, T. Nyong, R. Tabo, and C. Vogel, 2006:
Impacts, Vulnerability and Adaptation to Climate Change in Africa. Background
paper for the African Workshop on Adaptation, Implementation of Decision
1/CP.10 of the UNFCCC Convention, Accra, Ghana, 21-23 September 2006,
United Nations Framework Convention on Climate Change, Bonn, Germany, 54 pp.
Otero, R.C. and R.Z. Marti, 1995. The impacts of natural disasters on developing
economies: Implications for the international development and disaster community.
In: Disaster Prevention for Sustainable Development: Economic and Policy Issues
[Munasinghe, M. and C. Clarke (eds.)]. World Bank, Washington, DC, pp.11-40.
Oxfam, 2007: Adapting to Climate Change. What is Needed in Poor Countries and
Who Should Pay? Oxfam Briefing Paper 104, Oxfam International, Oxford, UK,
47 pp.
PACJA, 2009: The Economic Cost of Climate Change in Africa. The Pan African
Climate Justice Alliance, Nairobi, Kenya, 49 pp.
Pagiola, S., K. von Ritter, and J. Bishop, 2004: Assessing the Economic Value of
Ecosystem Conservation. World Bank Environment Department Paper No.101,
World Bank, Washington, DC, 58 pp.
Palecki, M.A., S.A. Changnon, and K.E. Kunkel, 2001: The nature and impacts of the
July 1999 heatwave in the midwestern United States: Learning from the lessons
of 1995. Bulletin of the American Meteorological Society, 82(7), 1353-1368.
Pandey, S., D.D. Behura, R. Villano, and D. Naik, 2000: Economic Costs of Drought
and Farmers’ Coping Mechanisms: A Study of Rainfed Rice Systems in Eastern
India. IRRI Discussion Paper Series 39, International Rice Research Institute, Los
Banos, Philippines, 35 pp.
Park, J.H., L. Duan, B. Kim, M.J. Mitchell, and H. Shibata, 2010: Potential effects
of climate change and variability on watershed biogeochemical processes and
water quality in Northeast Asia. Environment International, 36(2), 212-225.
Parmenter, R., E. Yadav, C. Parmenter, P. Ettestad, and K. Gage,1999: Incidence of
plague associated with increased winter-spring precipitation in New Mexico.
American Journal of Tropical Medicine and Hygiene, 61(5), 814-821.
Parmesan, C., T.L. Root, and M.R. Willig, 2000: Impacts of extreme weather and
climate on terrestrial biota. Bulletin of the American Meteorological Society,
81, 443-450.
Parry, M.L., O.F. Canziani, J.P. Palutikof and co-authors, 2007. Technical summary. Climate
Change 2007. Impacts, Adaptation and Vulnerability. Contribution of Working
Group II to the Fourth Assessment Report of the Intergovernmental Panel on
Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der Linde, and C.E.
Hanson (eds.)]. Cambridge University Press, Cambridge, UK, pp. 23-78.
Parry, M., N. Arnell, P. Berry, D. Dodman, S. Fankhauser, C. Hope, S. Kovats, R.
Nichollas, D. Satterthwaite, R. Tiffin, and R. Wheeler, 2009: Assessing the Costs
of Adaptation to Climate Change: A review of the UNFCCC and other recent
estimates. International Institute for Environment and Development and
Grantham Institute for Climate Change, London, UK, 111 pp.
Pascal, M., K. Laaidi, M. Ledrans, E. Baffert, C. Caserio-Schönemann, A. Le Tertre, J.
Manach, S. Medina, J. Rudant, and P. Empereur-Bissonnet, 2006: France’s heat
health watch warning system. International Journal of Biometeorology, 50,
144-153.
Patt, A.G., M. Tadross, P. Nussbaumer, K. Asante, M. Metzger, J. Rafael, A. Goujon, and
G. Brundrit, 2010: Estimating least-developed countries’ vulnerability to climate-
related extreme events over the next 50 years. Proceedings of the National
Academy of Sciences, 107(4), 1333-1337.
Pearce, D.W., W.R. Cline, A.N. Achanta, S.R. Fankhauser, R.K. Pachauri, R.S.J. Tol, and
P. Vellinga, 1996: The social costs of climate change: Greenhouse damage and
the benefits of control. In: Intergovernmental Panel on Climate Change. Working
Group III. Climate Change 1995: Economic and Social Dimensions of Climate
Change [Bruce, J.P., H. Yi, and E.F. Haites (eds.)]. Cambridge University Press,
Cambridge, UK, pp. 179-224.
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
286
Pearce, G., A.B. Mullan, M.J. Salinger, T.W. Opperman, D. Woods, and J.R. Moore, 2005:
Impact of Climate Change on Long-Term Fire Danger. Research Report Number
50, New Zealand Fire Service Commission, Wellington, New Zealand, 75 pp.
Peduzzi, P., H. Dao, C. Herold, and F. Mouton, 2009: Assessing global exposure and
vulnerability towards natural hazards: the Disaster Risk Index. Natural Hazards
and Earth System Sciences, 9, 1149-1159.
Peduzzi, P., B. Chatenoux, H. Dao, C. Herold, and G. Giuliani, 2011: Preview Global
Risk Data Platform. UNEP/GRID and UNISDR, Geneva, Switzerland,
preview.grid.unep.ch/index.php?preview=tools&cat=1&lang=eng.
Pelling, M. and J.I. Uitto, 2001: Small island developing states: natural disaster
vulnerability and global change. Global Environmental Change Part B:
Environmental Hazards, 3(2), 49-62.
Pelling, M., A. Özerdem, and S. Barakat, 2002: The macro-economic impact of
disasters. Progress in Development Studies, 2(4), 283-305.
Perry, A., 2003: Impacts of climate change on tourism in the Mediterranean. In:
Climate Change in the Mediterranean: socio-economic perspectives of impacts,
vulnerability and adaptation [Gioponi, C., M. Schechter, and E. Elgar (eds.)].
Edward Elgar Publishing, Cheltenham, UK, pp. 279-289.
Petley, D.N., G.J. Hearn, A. Hart, N.J. Rosser, S.A. Dunning, K. Oven, and W.A. Mitchell,
2007: Trends in landslide occurrence in Nepal. Natural Hazards, 43(1), 23-44,
doi:10.1007/s11069-006-9100-3.
Phillips, O.L., L. Aragao, S.L. Lewis, J.B. Fisher, J. Lloyd, G. Lopez-Gonzalez, Y. Malhi, A.
Monteagudo, J. Peacock, and C.A. Quesada, 2009: Drought sensitivity of the
Amazon rainforest. Science, 323, 1344-1347.
Pielke Jr., R.A. 2007: Future economic damage from tropical cyclones: sensitivities
to societal and climate changes. Philosophical Transactions of the Royal Society
A, 365, 2717-2729.
Pielke Jr., R.A. and Downton, M.W., 2000: Precipitation and damaging floods: trends
in the United States, 1932-1997. Journal of Climate, 13(20), 3625-3637.
Pielke Jr., R.A., J. Rubiera, C. Landsea, M.L. Fernandez, and R. Klein, 2003: Hurricane
vulnerability in Latin America and the Caribbean: normalized
damage and loss potentials. Natural Hazards Review, 4, 101-114.
Pielke Jr., R.A., S. Agrawala, L.M. Bouwer, I. Burton, S. Changnon, M.H. Glantz, W.H.
Hooke, R.J.T. Klein, K. Kunkel, D. Mileti, D. Sarewitz, E.L. Thompkins, N. Stehr, and
H. von Storch, 2005: Clarifying the attribution of recent disaster losses: a
response to Epstein and McCarthy. Bulletin of the American Meteorological
Society, 86, 1481-1483.
Pielke Jr., R.A., J. Gratz, C.W. Landsea, D. Collins, M. Saunders, and R. Musulin, 2008:
Normalized hurricane damages in the United States: 1900-2005. Natural
Hazards Review, 9, 29-42.
Pomeranets, K.S., 2005: Three Centuries of Floods in St-Petersburg. Iskusstvo, St.
Petersburg, Russia, 214 pp.
Pompe, J.J. and J.R. Rinehart, 2008: Mitigating damage costs from hurricane strikes
along the southeastern U.S. Coast: A role for insurance markets. Ocean &
Coastal Management, 51(12), 782-788.
Power, S.A., 1994: Temporal trends in twig growth of Fagus sylvatica L and their
relationships with environmental factors. Forestry,67(1), 13-30.
Prasad, P.V.V., K.J. Boote, L.H. Allen, Jr., J.E. Sheehy, and J.M.G. Thomas, 2006: Species,
ecotype and cultivar differences in spikelet fertility and harvest index of rice in
response to high temperature stress. Field Crops Research, 95, 398-411.
Pulhin, J., M. Tapia, and R. Perez, 2010: Integrating disaster risk reduction and climate
change adaptation: Initiatives and challenges in the Philippines. Climate
Change Adaptation and Disaster Risk Reduction: An Asian Perspective.
Community, Environment and Disaster Risk Management, 5, 217-235.
Quarantelli, E.L. (ed.), 1998: What is a Disaster? Perspectives on the Question.
Routledge, London, UK, 312 pp.
Quinn, W.H., D.O. Zopf, K.S. Short, R.T.W. Kuo Yang, 1978: Historical trends and
statistics of the Southern Oscillation, El Niño, and Indonesian droughts. Fishery
Bulletin, 76, 663-678.
Rachold, V., H. Lantuit, N. Couture, and W. Pollard (eds.), 2005: Arctic Coastal Dynamics.
Report of the 5th International Workshop, McGill University, Montreal, Canada,
13-16 October 2004.Reports on Polar and Marine Research 506, Alfred Wegener
Institute for Polar and Marine Research, Bremerhaven, Germany, 131 pp.
Raddatz, C., 2009: The Wrath of God: Macroeconomic Costs of Natural Disasters.
Policy Research Working Paper 5039, Development Research Group,
Macroeconomics and Growth Team, World Bank, Washington, DC, 35 pp.
Radeloff, V.C., R.B. Hammer, S.I. Stewart, J.S. Fried, S.S. Holcomb, and J.F. Mckeefry,
2005: The wildland–urban interface in the United States. Ecological
Applications, 15(3), 799-805.
Raghavan, S. and S. Rajesh, 2003: Trends in tropical cyclone impact: a study in
Andhra Pradesh, India. Bulletin of the American Meteorological Society, 84,
635-644.
Ranger, N., S. Hallegatte, S. Bhattacharya, M. Bachu, S. Priya, K. Dhore, F. Rafique, P.
Mathur, N. Naville, F. Henriet, C. Herweijer, S. Pohit, and J. Corfee-Morlot, 2011:
An assessment of the potential impact of climate change on flood risk in
Mumbai. Climatic Change, 104, 139-167.
Raschky, P.A. 2008: Institutions and the losses from natural disasters. Natural
Hazards Earth Systems Science, 8, 627-634.
Rasmussen, T.N., 2004: Macroeconomic Implications of Natural Disasters in the
Caribbean. IMF Working Papers WP/04/224, International Monetary Fund,
Washington, DC.
Ready, R. and S. Navrud, 2006: International benefit transfer methods and validity
tests. Ecological Economics, 60, 429-434.
Reason, C.J.C. and A. Keibel, 2004: Tropical cyclone Eline and its unusual penetration
and impacts over the Southern African mainland. Weather Forecast, 19
, 789-805.
Reason, C.J.C., S. Hachigonta, and R.F. Phaladi, 2005: Interannual variability in rainy
season characteristics over the Limpopo region of southern Africa. International
Journal of Climatology, 25, 1835-1853.
Rebetez, M. and M. Dobbertin, 2004: Climate change may already threaten Scots
pine stands in the Swiss Alps. Theoretical and Applied Climatology, 79, 1-9.
Rehfeldt, G.E., N.L. Crookston, M.V. Warwell, and J.S. Evans, 2006: Empirical analyses
of plant-climate relationships for the western United States. International
Journal of Plant Sciences, 167(6), 1123-1150.
Reid, H., L. Sahlén, J. MacGregor, and J. Stage, 2007: The economic impact of climate
change in Namibia: How climate change will affect the contribution of
Namibia’s natural resources to its economy. Environmental Economics
Programme Discussion Paper 07-02, International Institute for Environment and
Development, London, UK, 46 pp.
Reid, W.V., H.A. Mooney, A. Cropper, D. Capistrano, S.R. Carpenter, K. Chopra, P.
Dasgupta, T. Dietz, A.K. Duraiappah, R. Hassan, R. Kasperson, R. Leemans, R.M.
May, A.J. McMichael, P. Pingali, C. Samper, R. Scholes, R.T. Watson, A.H. Zakri, Z.
Shidong, N.J. Ash, E. Bennett, P. Kumar, M.J. Lee, C. Raudsepp-Hearne, H. Simons,
J. Thonell, and M.B. Zurek (eds.), 2005: Ecosystems and Human Well-being:
Synthesis. Island Press, Washington, DC, 155 pp.
Reilly, J. and D. Schimmelpfennig, 2000: Irreversibility, uncertainty, and learning: por-
traits of adaptation to long term climate change. Climatic Change, 45(1), 253-
278.
Revi, A., 2008: Climate change risk: an adaptation and mitigation agenda for Indian
cities. Environment and Urbanization, 20(1), 207-229, doi:10.1177/
0956247808089157.
Revich, B., 2008: Climate Change Impact on Public Health in the Russian Arctic.
United Nations in the Russian Federation, Moscow, Russia, 24 pp.
Rey, G., A. Fouillet, É. Jougla, and D. Hémon, 2007: Heat waves, ordinary temperature
fluctuations and mortality in France since 1971. Population (English Edition),
62, 457-485, doi:10.3917/pope.703.0457,cairn.info/revue-population-english-
2007-3-page-457.htm.
Rich, P.M., D.D. Breshears, and A.B. White, 2008: Phenology of mixed woody-
herbaceous ecosystems following extreme events: net and differential
responses. Ecology, 89(2), 342-352.
Richard, Y., N. Fauchereau, I. Poccard, M. Rouault, and S. Trzaska, 2001: 20th century
droughts in Southern Africa: spatial and temporal variability, teleconnections
with oceanic and atmospheric conditions. International Journal of Climatology,
21, 873-885.
Rico, M., G. Benito, A.R. Salgueiro, A. Díez-Herrero, and H. Pereira, 2008: Reported
tailings dam failures. A review of the European incidents in the worldwide
context. Journal of Hazardous Materials, 152, 846-852.
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
287
Rignot, E., M. Koppes, and I. Velicogna, 2010: Rapid submarine melting of the
calving faces of West Greenland glaciers. Nature Geoscience, 3, 187-191,
doi:10.1038/ngeo765.
Rittmaster, R., W. Adamowicz, B. Amiro, and R.T. Pelletier, 2006: Economic analysis
of health effects from forest fires. Canadian Journal of Forest Research, 36(4),
868-877.
Rizwan, A.M., L. Dennis, and C. Liu, 2008: A review on the generation, determination,
and mitigation of urban heat island. Journal of Environmental Sciences, 20(1),
120-128.
Rodell, M., I. Velicogna, and J. Famiglietti, 2009: Satellite-based estimates of
groundwater depletion in India. Nature, 460, 999-1002.
Rodriguez-Oreggia, E., A. de la Fuente, R. de la Torre, H. Moreno, and C. Rodriguez,
2010: The Impact of Natural Disasters on Human Development and Poverty at
the Municipal Level in Mexico. CID Working Paper No. 43, Center for
International Development at Harvard University, Cambridge, MA, USA,
34 pp.
Rogers, K. and T. Ralph (eds.), 2010: Floodplain Wetland Biota in the Murry-Darling
Basin: Water and Habitat Requirements. CSIRO Publishing, Collingwood,
Victoria, Australia, 348 pp.
Rose, A., 2007: Economic resilience to natural and man-made disasters: Multidisciplinary
origins and contextual dimensions. Environmental Hazards, 7(4), 383-398.
Rosenzweig, C. and M.L. Parry, 1994: Potential impact of climate change on world
food supply. Nature, 367, 133-138.
Rosenzweig, C., F.N. Tubiello, R. Goldberg, E. Mills, and J. Bloomfield, 2002:
Increased crop damage in the US from excess precipitation under climate
change. Global Environmental Change, 12, 197-202.
Rosenzweig, C., G. Casassa, D.J. Karoly, A. Imeson, C. Liu, A. Menzel, S. Rawlins, T.L.
Root, B. Seguin, and P. Tryjanowski, 2007: Assessment of observed changes and
responses in natural and managed systems. In: Climate Change 2007. Impacts,
Adaptation and Vulnerability. Contribution of Working Group II to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change [Parry,
M.L., O.F. Canziani, J.P.Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)].
Cambridge University Press, Cambridge, UK, pp. 79-131.
Roshydromet, 2005: Strategic Prediction for the Period of up to 2010—2015 of
Climate Change Expected in Russia and Its Impact on Sectors of the Russian
National Economy. Federal Service for Hydrometeorology and Environmental
Monitoring (Roshydromet), Moscow, Russia, 24pp.
Roshydromet, 2008: Assessment Report on Climate Change and Its Impacts on the
Territory of the Russian Federation. Federal Service for Hydrometeorology and
Environmental Monitoring (Roshydromet),Vol.II, Climate Change Impacts,
RIHMI-WDC, Obninsk, Kaluga Region, Russia, 288pp.
Rübbelke, D. and S. Vogele, 2011: Impacts of climate change on European critical
infrastructures: The case of the power sector. Environmental Science & Policy,
14, 53-63.
Saenz, R., R.A. Bissell, and F. Paniagua, 1995: Post-disaster malaria in Costa Rica.
Prehospital Disaster Medicine, 10(3), 154-160.
Sanderson, M.G., C.D. Jones, W.J. Collins, C.E. Johnson, and R.G. Derwent, 2003:
Effect of climate change on isoprene emissions and surface ozone levels.
Geophysical Research Letters
, 30(18), 1936, doi:10.1029/2003GL017642.
Satterthwaite, D., 2008: Climate Change and Urbanization: Effects and
Implications for Urban Governance. Paper presented at the United Nations
Expert Meeting on Population Distribution, Urbanization, Internal Migration
and Development, New York, NY, 21-23 January 2008, United Nations
Population Division, New York, NY, 29 pp., un.org/esa/population/meetings/
EGM_PopDist/P16_Satterthwaite.pdf.
Satterthwaite, D., S. Huq, M. Pelling, H. Reid, and P. Romero-Lankao, 2007:
Adapting to Climate Change in Urban Areas: The Possibilities and Constraints in
Low- and Middle-income Nations. Human Settlements Working Paper Series,
Climate Change and Cities No.1, International Institute for Environment and
Development, London, UK, 107 pp., pubs.iied.org/pdfs/10549IIED.pdf.
Savelieva, N.I., I.P. Semiletov, G.E. Weller, and L.N. Vasilevskaya, 2004: Climate
change in the northern Asia in the second half of the 20th century. Pacific
Oceanography, 2(1-2), 74-84.
Schär, C. and G. Jendritzky, 2004: Climate change: Hot news from summer 2003.
Nature, 432, 559-560.
Schär, C., P.L. Vidale, D. Luthi, C. Frei, C. Haberli, M.A. Liniger, and C. Appenzeller,
2004: The role of increasing temperature variability in European summer
heatwaves. Nature, 427, 332-336.
Scheffer, M., S. Carpenter, J.A. Foley, C. Folke, and B. Walker, 2001: Catastrophic
shifts in ecosystems. Nature, 413, 591-596.
Schelhaas, M.J., G.J. Nabuurs, and A. Schuck, 2003: Natural disturbances in the
European forests in the 19th and 20th centuries. Global Change Biology, 9(11),
1620-1633.
Schlenker, W. and D. Lobell, 2010: Robust negative impacts of climate change on
African agriculture. Environmental Research Letters, 5(1), doi:10.1088/1748-
9326/5/1/014010.
Schlenker, W. and M.J. Roberts, 2009: Nonlinear temperature effects indicate severe
damages to U.S. crop yields under climate change. Proceedings of the National
Academy of Sciences, 106, 15594-15598.
Schmid, D., I. Lederer, P. Much, A. Pichler, and F. Allerberger, 2005: Outbreak of
norovirus infection associated with contaminated flood water, Salzburg, 2005.
Eurosurveillance, 10(24), E050616.
Schmidt, S., C. Kemfert, and E. Faust, 2009: Simulation of Economic Losses from
Tropical Cyclones in the Years 2015 and 2050: The Effects of Anthropogenic
Climate Change and Growing Wealth. Discussion paper 914, German Institute
for Economic Research, Berlin, Germany, 29 pp.
Schneebeli, M., M. Laternser and W. Ammann, 1997: Destructive snow avalanches and
climate change in the Swiss Alps. Eclogae Geologicae Helvetiae, 90, 457-461.
Schneider, F., A. Buehn, and C.E. Montenegro, 2010. Shadow Economies All over the
World: New Estimates for 162 Countries from 1999 to 2007. Policy Research
Working Paper 5356, World Bank Development Research Group, Poverty and
Inequality Team and Europe and Central Asia Region, Human Development
Economics Unit, Washington, DC, 52 pp. www-wds.worldbank.org/
external/default/WDSContentServer/IW3P/IB/2010/10/14/000158349_2010101
4160704/Rendered/PDF/WPS5356.pdf.
Schneider, S.H., W.E. Easterling, and L.O. Mearns, 2000: Adaptation: Sensitivity to
natural variability, agent assumptions and dynamic climate changes. Climatic
Change,
45, 203-221.
Scholes, R.J. and R. Biggs (eds.), 2004: Ecosystem Services in Southern Africa: A
Regional Assessment. A contribution to the Millennium Ecosystem Assessment,
CSIR, Pretoria, South Africa, 76 pp., www.icp-confluence-sadc.org/sites/
default/files/SAfMA_Regional_Report_-_final.pdf.
Schreider, S.Y., D.I. Smith, and A.J. Jakeman, 2000: Climate change impacts on urban
flooding. Climatic Change, 47, 91-115.
Schulze, R., J. Meigh, and M. Horan, 2001: Present and potential future vulnerability
of eastern and southern Africa’s hydrology and water resources. South African
Journal of Science, 97, 150-160.
Schwierz, C., P. Köllner-Heck, E. Zenklusen Mutter, D.N. Bresch, P.L. Vidale, M. Wild,
and C. Schär, 2010: Modelling European winter wind storm losses in current
and future climate. Climatic Change, 101, 485-514.
Scott, D., G. McBoyle, B. Mills, and A. Minogue, 2006: Climate change and the
sustainability of ski-based tourism in eastern North America. Journal of
Sustainable Tourism, 14(4), 367-375
Scott, D., B. Amelung, S. Becken, J.P. Ceron, G. Dubois, S. Gossling, P. Peeters, and M.C.
Simpson, 2008: Climate Change and Tourism: Responding to Global Challenges.
United Nations World Tourism Organization, Madrid, Spain, 256 pp.
Semyonov, V.A. and A.A. Korshunov, 2006: Floods on the Russian rivers in the late 20th
century and the early 21st century. Issues of Geography and Geo-ecology, 5, 6-12.
Sénat. 2004. France and the French face the canicule: The lessons of a crisis.
Information Report no. 195, Sénat, Paris, France, pp. 59-62 (in French),
www.senat.fr/rap/r03-195/r03-19510.html.
Seo, N. and R. Mendelsohn, 2008: An analysis of crop choice: Adapting to climate
change in Latin American farms. Ecological Economics, 67, 109-116.
Shankman, D., B. D. Keim, and J. Song, 2006: Flood frequency in China’s Poyang Lake
Region: trends and teleconnections. International Journal of Climatology, 26,
1255-1266.
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
288
Shen, C., W.-C. Wang, Z. Hao, and W. Gong, 2008: Characteristics of anomalous
precipitation events over eastern China during the past five centuries. Climate
Dynamics, 31(4), 463-476, doi:10.1007/s00382-007-0323-0.
Sherstyukov, A.B., 2009: Climate Change and Its Impact in the Russian Permafrost
Zone. RIHMI-WDC, Obninsk, Russia, 127 pp.
Skidmore, M. and H. Toya, 2002: Do natural disasters promote long-run growth?
Economic Inquiry, 40(4), 664-687.
Skylas, D.J., S.J. Cordwell, P.G. Hains, M.R. Larsen, D.J. Basseal, B.J. Walsh, C.
Blumenthal, W. Rathmell, L. Copeland, and C.W. Wrigley, 2002: Heat shock of
wheat during grain filling: proteins associated with heat-intolerance. Journal of
Cereal Science, 35, 175-188.
Smit, B., O. Pilifosova, I. Burton, B. Challenger, S. Huq, R. Klein, and G. Yohe, 2001:
Adaptation to climate change in the context of sustainable development and
equity. In: Climate Change 2001: Impacts, Adaptation, and Vulnerability.
Contribution of Working Group II to the Third Assessment Report of the
Intergovernmental Panel on Climate Change [McCarthy, J.J., O.F. Canziani, N.A.
Leary, D.J. Dokken, and K.S. White (eds.)]. Cambridge University Press,
Cambridge, UK, pp. 879-906.
Smith, D.E, S.B. Raper, S. Zerbini, and A. Sánchez-Arcilla (eds.), 2000: Sea Level
Change and Coastal Processes: Implications for Europe. Proceedings of the
Climate change and Sea level research in Europe: State of the art and future
research needs Conference, Mataro, Spain, April 1997. Office for Official
Publications of the European Communities, Luxembourg, 247 pp.
Smith, O., 2006: Trends in Transportation Maintenance Related to Climate Change.
Paper presented to the Climate Change Workshop, Center for Transportation
and Environment, Washington, DC, 29 March, 2006.
Smyth, A.W., G. Altay, G. Deodatis, M. Erdik, G. Franco, P. Gulkan, H. Kunreuther, H.
Lus, E. Mete, N. Seeber, and O. Yuzugullu, 2004: Probabilistic benefit-cost analysis
for earthquake damage mitigation: Evaluating measures for apartment houses
in Turkey. Earthquake Spectra, 20(1), 171-203.
Spranger, M., 2008: Special nature of disaster risk in megacities. In: Natural
Catastrophe Risk Insurance Mechanisms for Asia and the Pacific, Tokyo, Japan,
4-5 November 2008. Munich Reinsurance Company, Munich, Germany, 11 pp.
Stedman, J.R., 2004: The predicted number of air pollution related deaths in the UK
during the August 2003 heatwave. Atmospheric Environment, 38, 1083-1085.
Steffen, W., G. Love, and P.H. Whetton, 2006: Approaches to defining dangerous
climate change: an Australian perspective. In: Avoiding Dangerous Climate
Change [Schellnhuber, H.J., W. Cramer, N. Nakicenovic, T. Wigley, and G. Yohe
(eds.)]. Cambridge University Press, Cambridge, UK, 392 pp.
Steiger, R., 2010: The impact of climate change on ski season length and
snowmaking requirements in Tyrol, Austria. Climate Research, 43, 251-262.
Stern, N., 2007: The Economics of Climate Change: The Stern Review. Cambridge
University Press, Cambridge, UK.
Stewart, I.D., 2011. A systematic review and scientific critique of methodology in modern
urban heat island literature. International Journal of Climatology, 31(2), 200-217.
Stirling, I. and C.L. Parkinson, 2006: Possible effects of climate warming on selected
populations of polar bears in the Canadian Arctic. Arctic,
59(3), 261-275.
Stone, D.A. and M.R. Allen, 2005: The end-to-end attribution problem: from emissions
to impacts. Climatic Change, 71, 303-318.
Stott, P.A., D.A. Stone, and M.R. Allen, 2004: Human contribution to the European
heatwave of 2003. Nature, 432, 610-614.
Straneo, F., G.S. Hamilton, D.A. Sutherland, L.A. Stearns, F. Davidson, M.O. Hammill,
G.B. Stenson and A. Rosing-Asvid, 2010: Rapid circulation of warm subtropical
waters in a major glacial fjord in East Greenland. Nature Geoscience, 3, 182-
186, doi:10.1038/ngeo764.
Stribley, G.H. and M.R. Ashmore, 2002: Quantitative changes in twig growth pattern
of young woodland beech (Fagus sylvatica L.) in relation to climate and ozone
pollution over 10 years. Forest Ecology and Management, 157(1-3), 191-204.
Sturm, M., C. Racine, and K. Tape, 2001: Climate change: Increasing shrub abundance
in the Arctic. Nature, 411, 546-547.
Su, H. and G.X. Liu, 2004: Elementary analyses on the progress of grassland fire
disaster information management technique. Grassland of China, 26(3), 69-71
(in Chinese).
Suzuki, T., 2009: Estimation of inundation damage caused by global warming in
three major bays and western parts of Japan. Global Environmental Research,
14(2), 237-246 (in Japanese).
Swenson, S. and J. Wahr, 2009: Monitoring the water balance of Lake Victoria, East
Africa, from space. Journal of Hydrology, 370, 163-176.
SWIPA, 2011: SWIPA 2011 Assessment Executive Summary: Snow, Water, Ice and
Permafrost in the Arctic. Snow, Water, Ice and Permafrost in the Arctic (SWIPA),
Arctic Monitoring and Assessment Programme Secretariat, Oslo, Norway, 15pp,
amap.no/swipa/SWIPA2011ExecutiveSummaryV2.pdf.
Swiss Re, 2009: The Effects of Climate Change: An Increase in Coastal Flood Damage
in Northern Europe. Focus Report, Swiss Reinsurance Company, Zurich,
Switzerland, 4 pp., media.swissre.com/documents/the_effects_of_climate_
change_an_increase_in_coastal_flood_damage_in_northern_europe.pdf.
Swiss Re, 2010: Natural Catastrophes and Man-Made Disasters in 2009:
Catastrophe Claims Few Victims, Insured Losses Fall. Sigma, No 1/2010, Swiss
Reinsurance Company, Zurich, Switzerland, 36 pp.
Swiss Re, 2011: Natural Catastrophes and Man-Made Disasters in 2010: A Year of
Devastating and Costly Events. Sigma, No 1/2011, Swiss Reinsurance Company,
Zurich, Switzerland, 36 pp.
Takahashi, T., H. Nakagawa, Y. Satofuka, and K. Kawaike, 2001: Flood and sediment
disasters triggered by 1999 rainfall in Venezuela; a river restoration plan for an
alluvial fan. Journal of Natural Disaster Science, 23(2), 65-82.
Takara, K., S. Kim, Y. Tachikawa, and E. Nakita, 2009: Assessing climate change
impact on water resources in the Tone River Basin, Japan, using super-high-
resolution atmospheric model output. Journal of Disaster Research, 4(1),
12-23.
Tallis, H.M. and P. Kareiva, 2006: Shaping global environmental decisions using
socio-ecological models. Trends in Ecology & Evolution, 21(10), 562-568.
Tan, J., Y. Zheng, X. Tang, C. Gup, L. Li, G. Song, X. Zhen, D. Yuan, A. Kalkstein, and F.
Li, 2010: The urban heat island and its impact on heat waves and human health
in Shanghai. Journal of Biometeorology, 54(1),
Tape, K., M. Sturm, and C. Racine, 2006: The evidence for shrub expansion in
Northern Alaska and the Pan-Arctic. Global Change Biology, 12, 686-702.
TEEB, 2009: The Economics of Ecosystems and Biodiversity for National and
International Policy Makers – Summary: responding to the value of nature. The
Economics of Ecosystems and Biodiversity, UNEP, Geneva, Switzerland, 39 pp.
Thibault, K.M. and J.H. Brown, 2008: Impact of an extreme climatic event on
community assembly. Proceedings of the National Academy of Sciences,
105(9), 3410-3415.
Thomas, M.F. and M.B. Thorp, 2003: Palaeohydrological reconstructions for tropical
Africa – evidence and problems. In: Palaeohydrology: Understanding Global
Change [Benito, G. and K.J. Gregory (eds.)]. John Wiley and Sons, Chichester, UK,
pp. 167-192.
Thonicke, K. and W. Cramer, 2006: Long-term trends in vegetation dynamics and
forest fires in Brandenburg (Germany) under a changing climate. Natural
Hazards, 38, 283-300.
Thornton, P.K., P.G. Jones, T. Owiyo, R.L. Kruska, M. Herrero, P. Kristjanson, A.
Notenbaert, N. Bekele, and A. Omolo, with contributions from V. Orindi, B.
Otiende, A. Ochieng, S. Bhadwal, K. Anantram, S. Nair, V. Kumar, and U. Kulkar,
2006: Mapping Climate Vulnerability and Poverty in Africa. ILRI Report to the
Department for International Development, International Livestock Research
Institute, Nairobi, Kenya, 171 pp.
Tian, X, T. Matsui, S. Li, M. Yoshimoto, K. Kobaysai, and T. Hasegawa. 2010: Heat-
induced floret sterility of hybrid rice (Oryza sativa L.) cultivars under humid and
low wind conditions in the field of Jainghan basin, China. Plant Production
Science, 13(3), 243-251.
Tol, R.S.J., 1995: The damage costs of climate change: Toward more comprehensive
calculations. Environmental and Resource Economics, 5, 353-374.
Tol, R.S.J., 2001: Estimates of the damage costs of climate change. Environmental
and Resource Economics, 21, 47-73.
Tol, R.S.J. and F.P.M. Leek, 1999: Economic analysis of natural disasters. In: Climate,
Change and Risk. [Downing, T.E., A.A. Olsthoorn, and R.S.J. Tol (eds.)].
Routledge, London, UK, pp. 308-327.
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
289
Toole, M.J., 1997: Communicable diseases and disease control. In: The Public Health
Consequences of Disasters. [Noji, E.K. (ed.)]. Oxford University Press, New York,
NY, pp. 79-100.
Toya, H. and M. Skidmore, 2007: Economic development and the impacts of natural
disasters. Economics Letters, 94, 20-25.
Trenberth, K.E., P.D. Jones, P. Ambenje, R. Bojariu, D. Easterling, A. Klein Tank, D.
Parker, F. Rahimzadeh, J.A. Renwick, M. Rusticucci, B. Soden and P. Zhai, 2007:
Observations: Surface and atmospheric climate change. In: Climate Change
2007: The Physical Science Basis. Contribution of Working Group I to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change
[Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor, and
H.L. Miller (eds.)]. Cambridge University Press, Cambridge, UK, pp. 237-336.
Trewin, B. and H. Vermont, 2010. Changes in the frequency of record temperatures
in Australia, 1957-2009. Australian Meteorological and Oceanographic Journal,
60, 113-119.
Truong, G., A.E. Palm, and F. Felber, 2006: Recent invasion of the mountain birch
Betula pubescens ssp. tortuosa above the treeline due to climate change:
genetic and ecological study in northern Sweden. Journal of Evolutionary
Biology, 20(1), 369-380.
Tsubo, M., S. Fukai, J. Basnayake, M. Ouk, 2009: Frequency of occurrence of various
drought types and its impact on performance of photoperiod-sensitive and
insensitive rice genotypes in rainfed lowland conditions in Cambodia. Field
Crops Research, 113, 287-296.
U.S. Census Bureau, 2010: Coastline Population Trends in the United States: 1960
to 2008. U.S. Department of Commerce, Economics and Statistics
Administration, U.S. Census Bureau, Washington, DC, 27 pp., www.census.gov/
prod/2010pubs/p25-1139.pdf.
UN-HABITAT, 2008: State of the World’s Cities 2008/2009: Harmonious Cities. United
Nations Human Settlements Programme, Earthscan, London, UK, 259 pp.
UNCTAD, 2009: Multi-year Expert Meeting on Transport and Trade Facilitation:
Maritime Transport and the Climate Change Challenge, 16-18 February 2009.
Summary of Proceedings, UNCTAD/SDTE/TLB/2009/1, UN Conference on Trade
and Development, Geneva, Switzerland, 47 pp.
UNDP, 2004: Reducing Disaster Risk: A Challenge for Development. United Nations
Development Programme, Bureau for Crisis Prevention and Recovery, New
York, NY, 146 pp.
UNDP, 2007: Fighting Climate Change: Human Solidarity in a Divided World [Watkins,
K. (ed.)]. Human Development Report 2007/2008 of the United Nations
Development Programme, Palgrave Macmillan, Basingstoke, UK, 384 pp.
UNECE, 2009: Self-Made Cities: In search of sustainable solutions for informal
settlements in the United Nations Economic Commission for Europe region.
United Nations Economic Commission for Europe, United Nations, Geneva,
Switzerland, 113 pp.
UNFCCC, 1992: United Nations Framework Convention on Climate Change. United
Nations, FCCC/INFORMAL/84 GE.05-62220 (E) 200705, Secretariat of the
United Nations Framework Convention on Climate Change, Bonn, Germany, 24
pp., unfccc.int/resource/docs/convkp/conveng.pdf.
UNFCCC, 2007: Investment and Financial Flows to Address Climate Change.
Secretariat of the United Nations Framework Convention on Climate Change,
Bonn, Germany, 272 pp.
UNHCR, 2010: 2009 Global Trends: Refugees, Asylum-seekers, Returnees, Internally
Displaced and Stateless Persons. United Nations High Commissioner for
Refugees, Division of Programme Support and Management, Geneva,
Switzerland, 29 pp.
UNISDR, 2005: Framework for Action 2005-2015. Building the Resilience of Nations
and Communities to Disasters. United Nations International Strategy for
Disaster Reduction, Geneva, Switzerland, 22 pp., http://www.unisdr.org/
2005/wcdr/intergover/official-doc/L-docs/Hyogo-declaration-english.pdf.
UNISDR, 2009, Global Assessment Report on Disaster Risk Reduction: Risk and
Poverty in a Changing Climate – Invest Today for a Safer Tomorrow. United Nations
International Strategy for Disaster Reduction Secretariat, Geneva, Switzerland,
Oriental Press, Manama, Kingdom of Bahrain, 207 pp., www.preventionweb.net/
english/hyogo/gar/report/index.php?id=1130&pid:34&pih:2.
UNISDR, 2011: Global Assessment Report on Disaster Risk Reduction: Revealing
Risk, Redefining Development. United Nations International Strategy for
Disaster Reduction Secretariat, Geneva, Switzerland, Information Press, Oxford,
UK, 178 pp.
UNWTO, 2003. Tourism, Peace, and Sustainable Development for Africa: Luanda,
Angola, 29-30 May 2003. United Nations World Tourism Organization, Madrid,
Spain, 276 pp.
Usbeck, T., T. Wohlgemuth, M. Dobbertin, C. Pfister, A. Bürgi, and M. Rebetez, 2010:
Increasing storm damage to forests in Switzerland from 1858 to 2007.
Agricultural and Forest Meteorology, 150, 47-55.
Uyarra, M.C., I.M. Côté, J.A. Gill, R.T.T. Tinch, D. Viner, and A.R. Watkinson, 2005:
Island-specific preferences of tourists for environmental features: implications
of climate change for tourism-dependent states. Environmental Conservation,
32(1), 11-19.
van der Veen, A., 2004: Disasters and economic damage: macro, meso and micro
approaches. Disaster Prevention and Management, 13(4), 274-279.
van der Werf, G.R., J. Dempewolf, S.N. Trigg, J.T. Randerson, P.S. Kasibhatla, L. Giglio,
D. Murdiyarso, W. Peters, D.C. Morton, G.J. Collatz, A.J. Dolman, and R.S. DeFries,
2008: Climate regulation of fire emissions and deforestation in equatorial Asia.
Proceeding of the National Academy of Sciences, 105(51), 20350-20355.
Vanacker, V., M. Vanderschaeghe, G. Govers, E. Willems, J. Poesen, J. Deckers, and B.
De Bievre, 2003: Linking hydrological, infinite slope stability and land use change
models through GIS for assessing the impact of deforestation on landslide
susceptibility in High Andean watersheds. Geomorphology, 52, 299-315.
Vandentorren, S., P. Bretin, A. Zeghnoun, L. Mandereau-Bruno, A. Croisier, C.
Cochet, J. Ribéron, I. Siberan, B. Declercq, and M. Ledrans, 2003: August 2003
heat wave in France: risk factors for death of elderly people living at home.
European Journal of Public Health, 16(6), 583-591.
Vanham, D., E. Fleischhacker, and W. Rauch, 2009: Impact of an extreme dry and hot
summer on water supply security in an alpine region. Water Science and
Technology, 59(3), 469-477.
Veijalainen, N., E. Lotsari, P. Alho, B. Vehviläinen, and J. Käyhkö, 2010: National scale
assessment of climate change impacts on flooding in Finland. Journal of
Hydrology, 391(3-4), 333-350.
Venton, C., and P. Venton, 2004: Disaster Preparedness Programmes in India: A Cost
Benefit Analysis. Network Paper 49, Humanitarian Practice Network at
Overseas Development Institute, London, UK, 21 pp., odihpn.org/documents/
networkpaper049.pdf.
Verschuren, D., K.R. Laird, and B.F. Cumming, 2000: Rainfall and drought in equatorial
east Africa during the past 1,100 years. Nature
, 403, 410-414.
Victorian Bushfires Royal Commission, 2010: The 2009 Victorian Bushfires
Royal Commission Final Report. State Government of Victoria, Melbourne,
Australia, 42 pp., royalcommission.vic.gov.au/Commission-Reports/Final-Report.
Villalba, R. and T.T. Veblen, 1997: Regional patterns of tree population age
structures in northern Patagonia: climatic and disturbance influences. Journal of
Ecology, 85, 113-124.
Villers, L. and I. Trejo, 2004: Evaluación de la vulnerabilidad en los sistemas forestales.
Cambio Climático. In: Una Visión desde México [Martínez, J. and A. Fernández
Bremauntz (eds.)]. Secretaría de Medio Ambiente y Recursos Naturalesand
Instituto Nacional de Ecologia, Tlalpan, Mexico, pp. 239-254.
Voigt, T., H.-M. Füssel, I. Gärtner-Roer, C. Huggel, M. Zemp, and C. Marty, 2010:
Impacts of Climate Change On Snow, Ice, and Permafrost in Europe: Observed
Trends, Future Projections, and Socio-Economic Relevance. ETC/ACC Technical
Paper 2010/13, The European Topic Centre on Air and Climate Change,
Bilthoven, The Netherlands, 117 pp.
Vörösmarty, C.J., 2002: Global change, the water cycle, and our search for Mauna
Loa. Hydrological Processes, 16, 135-139, doi:10.1002/hyp.527.
Vörösmarty, C.J., E.M. Douglas, P.A. Green, and C. Revenga, 2005: Geospatial indicators
of emerging water stress: an application to Africa. Ambio, 34, 230-236.
Vos, F., J. Rodriguez, R. Below, and D. Guha-Sapir, 2010: Annual Disaster Statistical
Review 2009: The Numbers and Trends. Centre for Research on the
Epidemiology of Disasters, Université Catholique de Louvain, Brussels, Belgium,
38 pp.
Chapter 4 Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
290
Wamsley, T.V., M.A. Cialone, J.M. Smith, J.H. Atkinson, and J.D. Rosati, 2010: The
potential of wetlands in reducing storm surge. Ocean Engineering, 37, 59-68.
Wang, J.X.L. and J.K. Angell, 1999: Air Stagnation Climatology for the United States
(1948-1998). NOAA/Air Resources Laboratory ATLAS, No.1, Silver Spring, MD,
73 pp.
Wardekker, J.A., A. de Jong, J.M. Knoop, and J.P. van der Sluijs, 2010: Operationalising
a resilience approach to adapting an urban delta to uncertain climate changes.
Technological Forecasting & Social Change, 77, 987-998.
Washington, R. and A. Preston, 2006: Extreme wet years over southern Africa: role
of Indian Ocean Sea surface temperatures. Journal of Geophysical Research,
111, D15104, doi:10./029/2005 JD006724.
Washington, R., M. Harrison, and D. Conway, 2004: African Climate Report. Report
commissioned by the UK Government to review African climate science, policy
and options for action, UK Department for Environment Food and Rural Affairs
and the Department for International Development, London, UK, 45 pp.
West, C.T. and D.G. Lenze, 1994: Modeling the regional impact of natural disaster
and recovery: A general framework and an application to Hurricane Andrew.
International Regional Science Review, 17(2), 121-150.
West, J.J., M.J. Small, and H. Dowlatabadi, 2001: Storms, investor decisions, and the
economic impacts of sea level rise. Climatic Change, 48, 317-342.
Westerling, A.L. and B.P. Bryant, 2008: Climate change and wildfire in California.
Climatic Change, 87, 231-249.
Westerling, A.L., H.G. Hidalgo, D.R. Cayan, and T.W. Swetnam, 2006: Warming and
earlier spring increases western U.S. forest wildfire activity. Science, 313, 940-
943.
Wheeler T.R., P.Q. Craufurd, R.H. Ellis, J.R. Porter, and P.V. Vara Prasad, 2000:
Temperature variability and the yield of annual crops. Agriculture, Ecosystems
and Environment, 82, 159-167.
Whitehead, P.G., R.L. Wilby, R.W. Battarbee, M. Kernan, and A.J. Wade, 2009: A
review of the potential impacts of climate change on surface water quality.
Hydrological Sciences Journal, 54(1), 101-123.
WHO, 2003: The Health Impacts of 2003 Summer Heat-Waves. Briefing note for the
delegations of the fifty-third session of the World Health Organization Regional
Committee for Europe, Vienna, Austria, 8-11 September 2003, WHO Regional
Committee for Europe, Copenhagen, Denmark, 12 pp.
WHO/UNICEF, 2000: Global Water Supply and Sanitation Assessment: 2000 Report.
World Health Organization, Geneva, 87 pp., who.int/entity/water_
sanitation_health/monitoring/jmp2000.pdf.
Widawsky, D.A., and J.C. O’Toole, 1990: Prioritizing the rice biotechnology research
agenda for Eastern India. Rockefeller Foundation Press, New York, NY, 86 pp.
Wilbanks, T.J., P. Romero Lankao, M. Bao, F. Berkhout, S. Cairncross, J.-P. Ceron, M.
Kapshe, R. Muir-Wood and R. Zapata-Marti, 2007: Industry, settlement and
society. In: Climate Change 2007: Impacts, Adaptation and Vulnerability.
Contribution of Working Group II to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P.
Palutikof, P.J. van der Linden, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 357-390.
Wilbanks, T., V. Bhatt, D. Bilello, S. Bull, J. Ekmann, W. Horak, J. Huang, M. Levine, M.
Sale, D. Schmalzer, and M. Scott, 2008. Effects of Climate Change on Energy
Production and Use in the United States. U.S. Climate Change Science Program
Synthesis and Assessment Product 4.5, Report by the U.S. Climate Change
Science Program and the Subcommittee on Global Change Research,
Washington, DC, 84 pp.
Wilby, R.L., 2003: Past and projected trends in London’s urban heat island. Weather,
58, 251-260.
Wilby, R.L., 2007: A review of climate change impacts on the built environment. Built
Environment, 33(1), 31-45.
Wildavsky, A. 1988: Searching for Safety. Transaction Books, New Brunswick, NJ,
253 pp.
Wiley, J.W. and J.M. Wunderle, Jr., 1994: The effects of hurricanes on birds, with special
reference to Caribbean islands. Bird Conservation International, 3, 319-349.
Wisner, B., P. Blaikie, T. Cannon, and I. Davis, 2004: At Risk: Natural Hazards, People’s
Vulnerability and Disasters. Routledge, London, UK, 471 pp.
Woodruff, R.E., S. Hales, C. Butler, and A.J. McMichael, 2005: Climate Change Health
Impacts in Australia: Effects of Dramatic CO
2
Emission Reductions. Australian
Conservation Foundation and the Australian Medical Association, Canberra,
Australia, 44 pp.
Woodworth, P.L., J.M. Gregory, and R.J. Nicholls, 2005: Long term sea level changes
and their impacts. In: The Global Coastal Ocean: Multiscale Interdisciplinary
Processes [Robinson, A.R. and K.H. Brink (eds.)]. Harvard University Press,
Cambridge, MA, pp. 715-753.
World Bank, 2000: Cities, Seas and Storms: Managing Change in Pacific Island
Economies. Volume IV, Adapting to Climate Change, Papua New Guinea and
Pacific Islands Country Unit, International Bank for Reconstruction and
Development/World Bank, Washington, DC, 48 pp., web.worldbank.org/
WBSITE/EXTERNAL/COUNTRIES/EASTASIAPACIFICEXT/PACIFICISLAND-
SEXTN/0,,contentMDK:20218394~pagePK:141137~piPK:217854~theSitePK:4
41883,00.html.
World Bank, 2006: Clean Energy and Development: Towards an Investment
Framework. World Bank, Washington, DC, 146 pp.
World Bank, 2007: Disasters, Climate Change, and Economic Development in Sub-
Saharan Africa: Lessons and Future Directions. Evaluation Brief 3, Independent
Evaluation Group, World Bank, Washington, DC, 31 pp.
World Bank, 2010: Economics of Adaptation to Climate Change: Synthesis Report.
World Bank, Washington, DC, 101 pp.
World Bank and UN, 2010: Natural Hazards, UnNatural Disasters: The Economics
of Effective Prevention. International Bank for Reconstruction and Development/
World Bank, Washington, DC, 254 pp., gfdrr.org/gfdrr/nhud-home.
Xiao, F. and Z. Xiao, 2010: Characteristics of tropical cyclones in China and their
impacts analysis. Natural Hazards, 54, 827-837
Yee, E.L., H. Palacio, R.L. Atmar, U. Shah, C. Kilborn, M. Faul, T.E. Gavagan, R.D.
Feigin, J. Versalovic, and F.H. Neill, 2007: Widespread outbreak of norovirus
gastroenteritis among evacuees of Hurricane Katrina residing in a large
“megashelter” in Houston, Texas: lessons learned for prevention. Clinical
Infectious Diseases, 44, 1032-1039.
Yezer, A.M. and C.B. Rubin, 1987: The Local Economic Effects of Natural Disasters.
National Hazard Research, George Washington University, Washington, DC.
Yohe, G., J. Neumann, and H. Ameden, 1995: Assessing the economic cost of
greenhouse induced sea level rise: Methods and applications in support of a
national survey. Journal of Environmental Economics and Management, 29,
S78-S97.
Yohe, G., J. Neumann, P. Marshall, and H. Ameden, 1996: The economic cost of
greenhouse induced sea level rise in the United States. Climatic Change, 32,
387-410.
Yu, G., Z. Schwartz, and J.E. Walsh, 2009: A weather-resolving index for assessing the
impact of climate change on tourism related climate resources. Climatic
Change, 95, 551-573.
Zebisch, M., T. Grothmann, D. Schröter, C. Hasse, U. Fritsch, and W. Cramer, 2005:
Climate change in Germany–Vulnerability and adaptation of climate sensitive
sectors. Report commissioned by the Federal Environmental Agency, Germany
(UFOPLAN 201 41 253), Potsdam Institute of Climate Impact Research,
Potsdam, Germany, 205 pp.
Zhang, J.Q., D.W. Zhou, Z.S. Song, and Z.J. Tong, 2006: A new perception on risk
assessment and risk assessment of grassland fire disaster. Journal of Basic
Science and Engineering, 14, 56-62 (in Chinese).
Zhang, Q., L. Wu, and Q. Liu, 2009: Tropical cyclone damages in China: 1983-2006.
Bulletin of the American Meteorological Society, 90, 489-495.
Zimmerman, R. and C. Faris, 2010: Infrastructure impacts and adaptation challenges.
In:
Annals of the New York Academy of Sciences. 2010. Blackwell Publishing
Inc., New York, NY, pp. 63-86.
Chapter 4Changes in Impacts of Climate Extremes: Human Systems and Ecosystems
291
Coordinating Lead Authors:
Susan Cutter (USA), Balgis Osman-Elasha (Sudan)
Lead Authors:
John Campbell (New Zealand), So-Min Cheong (Republic of Korea), Sabrina McCormick (USA),
Roger Pulwarty (USA), Seree Supratid (Thailand), Gina Ziervogel (South Africa)
Review Editors:
Eduardo Calvo (Peru), Khamaldin Daud Mutabazi (Tanzania)
Contributing Authors:
Alex Arnall (UK), Margaret Arnold (World Bank), Joanne Linnerooth Bayer (USA), Hans-Georg Bohle
(Germany), Christopher Emrich (USA), Stephane Hallegatte (France), Bettina Koelle (South Africa),
Noel Oettle (South Africa), Emily Polack (UK), Nicola Ranger (UK), Stephan Rist (Switzerland),
Pablo Suarez (Argentina),Gustavo Wilches-Chaux (Colombia)
This chapter should be cited as:
Cutter, S., B. Osman-Elasha, J. Campbell, S.-M. Cheong, S. McCormick, R. Pulwarty, S. Supratid, and G. Ziervogel, 2012:
Managing the risks from climate extremes at the local level. In: Managing the Risks of Extreme Events and Disasters to
Advance Climate Change Adaptation [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea,
K.J. Mach, G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups I and II of the
Intergovernmental Panel on Climate Change (IPCC). Cambridge University Press, Cambridge, UK, and New York, NY, USA,
pp. 291-338.
5
Managing the Risks
from Climate Extremes
at the Local Level
Managing the Risks from Climate Extremes at the Local Level
292
Executive Summary .................................................................................................................................293
5.1. Introduction: Why the Local is Important ................................................................................296
5.2. How Local Places Currently Cope with Disaster Risk ..............................................................298
5.2.1. Emergency Assistance and Disaster Relief ......................................................................................................................299
5.2.2. Population Movements ....................................................................................................................................................300
5.2.3. Recovery and Reconstruction...........................................................................................................................................301
5.3. Anticipating and Responding to Future Disaster Risk.............................................................302
5.3.1. Communicating Risk.........................................................................................................................................................302
5.3.1.1. Message Design..................................................................................................................................................
..............................302
5.3.1.2. Modes and Timing of Risk Communication ......................................................................................................................................302
5.3.1.3. Warnings and Warning Systems........................................................................................................................................................303
5.3.2. Structural Measures ...
....
..................................................................................................................................................304
5.3.3. Land Use and Ecosystem Protection ................................................................................................................................306
5.3.4. Storage and Rationing of Resources ...............................................................................................................................307
5.4. Building Capacity at the Local Level for Risk Management in a Changing Climate...............308
5.4.1. Proactive Behaviors and Protective Actions ....................................................................................................................308
5.4.2. Empowerment for Local Decisionmaking ........................................................................................................................310
5.4.3. Social Drivers ...................................................................................................................................................................310
5.4.4. Integrating Local Knowledge...........................................................................................................................................311
5.4.5. Local Government and Nongovernment Initiatives and Practices ..................................................................................312
5.5. Challenges and Opportunities..................................................................................................313
5.5.1. Differences in Coping and Risk Management..................................................................................................................313
5.5.1.1. Gender, Age, and Wealth ................................................................................................................................................
..................313
5.5.1.2. Livelihoods and Entitlements............................................................................................................................................................314
5.5.1.3. Health and Disability ........................................................................................................................................................................316
5.5.1.4. Human Settlements ..........................................................................................................................................................................316
5.5.2. Costs of Managing Disaster Risk and Risk from Climate Extremes ...
.............................................................................317
5.5.2.1. Costs of Impacts, Costs of Post-Event Responses.............................................................................................................................317
5.5.2.2.
Adaptation and Risk Management – Present and Future.................................................................................................................318
5.5.2.3. Consistency and Reliability of Cost and Loss Estimations at the Local Level ...................................................................................318
5.5.3. Limits to Local Adaptation...
....
........................................................................................................................................319
5.5.4. Advancing Social and Environmental Justice...................................................................................................................320
5.6. Management Strategies...........................................................................................................320
5.6.1. Basics of Planning in a Changing Climate .......................................................................................................................320
5.6.2. Community-Based Adaptation .........................................................................................................................................321
5.6.3. Risk Sharing and Transfer at the Local Level...................................................................................................................321
5.6.4. A Transformative Framework for Management Strategies .............................................................................................323
5.7. Information, Data, and Research Gaps at the Local Level ......................................................323
5.8. Summary...................................................................................................................................325
References ...............................................................................................................................................326
Chapter 5
Table of Contents
293
Disasters are most acutely experienced at the local level (high agreement, robust evidence). The reality of
disasters in terms of loss of life and property occurs in local places and to local people. These localized impacts can
then cascade to have national and international consequences. In this chapter, local refers to a range of places, social
groupings, experience, management, institutions, conditions, and sets of knowledge that exist at a sub-national scale.
[5.1]
Developing strategies for disaster risk management in the context of climate change requires a range of
approaches, informed by and customized to specific local circumstances (high agreement, robust evidence).
These differences and the context (national to global, urban to rural) in which they are situated shape local
vulnerability and local impacts. [5.1]
The impacts of climate extremes and weather events may threaten human security at the local level (high
agreement, medium evidence). Vulnerability at the local level is attributed to social, political, and economic
conditions and drivers including localized environmental degradation and climate change. Addressing disaster risk and
climate extremes at the local level requires attention to much wider issues relating to sustainable development. [5.1]
While structural measures provide some protection from disasters, they may also create a false sense of
safety (high agreement, robust evidence). Such measures result in increased property development, heightened
population density, and more disaster exposure. Current regulations and design levels for structural measures may be
inadequate under conditions of climate change. [5.3.2]
Sustainable land management is an effective disaster risk reduction tool (high agreement, robust evidence).
Land management includes land use, planning, zoning, conservation zones, buffer zones, or land acquisition. Often it
is difficult for local jurisdictions to implement such measures as a result of political and economic pressures for
development. However, such measures are often less disruptive to the environment and more sustainable at the local
level than structural measures. [5.3.3]
Humanitarian relief is often required when disaster risk reduction measures are absent or inadequate
(high agreement, robust evidence). Such assistance is more effective when it takes local social, cultural, and
economic conditions into account, acknowledges local agency in disaster response, and recognizes that the initial
assistance during and immediately after disasters is nearly always locally generated. [5.2.1]
Post-disaster recovery and reconstruction provide an opportunity for reducing weather- and climate-related
disaster risk and for improving adaptive capacity (high agreement, robust evidence). An emphasis on rapidly
rebuilding houses, reconstructing infrastructure, and rehabilitating livelihoods often leads to recovering in ways that
recreate or even increase existing vulnerabilities, and that preclude longer-term planning and policy changes for
enhancing resilience and sustainable development. Including local actors benefits the recovery process. [5.2.3]
Disasters associated with climate extremes influence population mobility and relocation affecting host and
origin communities (medium agreement, medium evidence). Most people return and participate in the post-
disaster recovery in their local areas. If disasters occur more frequently and/or with greater magnitude, some local
areas will become increasingly marginal as places to live or in which to maintain livelihoods. In such cases, migration
and displacement could become permanent and could introduce new pressures in areas of relocation. For locations
such as atolls, in some cases it is possible that many residents will have to relocate. In other cases, migration is an
adaptation to climate change, with remittances supporting community members who remain at home. [5.2.2]
Integration of local knowledge with additional scientific and technical knowledge can improve disaster
risk reduction and climate change adaptation (high agreement, robust evidence). Local populations document
their experiences with the changing climate, particularly extreme weather events, in many different ways, and this type
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
Executive Summary
294
of self-generated knowledge induces discussions of proactive adaptation strategies and can uncover existing capacity
within the community and important current shortcomings. [5.4.4]
Effectively communicating risk involves multiple pathway exchanges between decisionmakers and local
citizens (high agreement, medium evidence). Viewing risk communication as a social process allows for effective
participatory approaches, relationship building, and the production of visual, compelling, and engaging information for
use by local stakeholders. [5.3.1]
Inequalities influence local coping and adaptive capacity and pose disaster risk management and adaptation
challenges (high agreement, robust evidence). These inequalities reflect differences in gender, age, wealth, class,
ethnicity, health, and disability. They may also be reflected in differences in access to livelihoods and entitlements.
Understanding and increasing the awareness of coping mechanisms in the context of local-level livelihood is important
to climate change adaptation planning and risk management. This signifies the need for the identification and
accommodation of these differences to enhance opportunities arising from their incorporation into adaptation planning
and disaster response. [5.5.1]
Ecosystem management and restoration activities that focus on addressing deteriorating environmental
conditions are essential to protecting and sustaining people’s livelihoods in the face of climate extremes
(high agreement, robust evidence). Such activities include, among others, watershed rehabilitation, agro-ecology,
and forest landscape restoration. Moreover, provision of better access to and control of resources will improve people’s
livelihoods, and build long-term adaptive capacity. Such approaches have been recommended in the past, but have
not been incorporated into capacity building to date. [5.3.3]
Local-level institutions and self-organization are critical for social learning, innovations, and action; all are
essential elements for local risk management and adaptation (high agreement, medium evidence). Adaptive
capacities are not created in a vacuum – local institutions provide the enabling environment for community-based
adaptation planning and implementation. Local participation (community-based organizations, development committees)
contributes to empowering the most vulnerable and strengthening innovations. Addressing political and cultural issues
at the local levels are fundamental to the development of any strategy aiming at sustained disaster risk management
and adaptation. [5.4]
The rapid urbanization of the sub-national populations and the growth of megacities, especially in
developing countries, have led to the emergence of highly vulnerable urban communities, particularly
through informal settlements and inadequate land management, presenting challenges to disaster
management (high agreement, robust evidence). Addressing these critical vulnerabilities means consideration of
the social, political, and economic driving forces, including rural-to-urban migration, changing livelihoods, and wealth
inequalities as key inputs into decisionmaking. [5.5.1]
Effective local adaptation strategy requires addressing a number of factors that limit the ability of local
people to undertake necessary measures to protect themselves against climate extremes and disasters
(high agreement, robust evidence). Closing the information gap is critical to reducing vulnerability of natural-
resource dependent communities. Maintaining the ability of a community to ensure equitable access and entitlement
to key resources and assets is essential to building local adaptive capacity in a changing climate. Moreover, capacity
building and development of new skills for diversifying local livelihoods are key to flexibility in disaster reduction,
improving local adaptation, and managing disasters. [5.5]
Comprehensive assessments of local disaster risk are lacking in many places (high agreement, medium
evidence). As a foundation for management options, the methodology for locally based vulnerability assessments
(exposure and sensitivity) and potential costs needs more development and testing for applications to the local
context. [5.6]
Insurance is a risk transfer mechanism used at the local level (medium agreement, medium evidence). Risk
sharing (formal insurance, micro-insurance, crop insurance) can be a tool for risk reduction and for recovering livelihoods
Chapter 5Managing the Risks from Climate Extremes at the Local Level
295
after a disaster. Under certain conditions such tools can provide disincentives for reducing disaster risk at the local
level through the transfer of the risk spatially (to other places) or temporally (to the future). [5.6.3]
Local participation supports community-based adaptation to benefit management of disaster risk and
climate extremes (medium agreement, medium evidence). However, improvements in the availability of human
and financial capital and of disaster risk and climate information customized for local stakeholders can enhance
community-based adaptation. [5.6]
Data on natural disasters anddisaster risk reduction are lacking at the local level, which can constrain
improvements in local vulnerability reduction (high agreement, medium evidence). This is the case in all areas
but especially so in developing countries. Local knowledge systems are often neglected in disaster risk management.
There is considerable potential for adapting geographic information systems to include local-level knowledge to support
disaster management activities. [5.7]
Disaster loss estimates are inconsistent and highly dependent on the scale of the analysis, and result in
wide variations among community, state, province, and sub-national regions (high agreement, robust
evidence). Indirect losses are increasingly taken into account as significant factors in precipitating negative economic
impacts. Adaptation costs, though hard to estimate, can be reduced if climate change adaptation is integrated into
existing disaster risk management and disaster risk management is in turn embedded in development strategies and
decisionmaking. [5.5]
Mainstreaming disaster risk management into policies and practices provides key lessons that apply to
climate change adaptation at the local level (high agreement, medium evidence). Addressing social welfare,
quality of life, infrastructure, and livelihoods, and incorporating a multi-hazards approach into planning and action for
disasters in the short term, facilitates adaptation to climate extremes in the longer term. [5.4, 5.5, 5.6]
The main challenge for local adaptation to climate extremes is to apply a balanced portfolio of approaches
as a one-size-fits-all strategy may prove limiting for some places and stakeholders (high confidence,
medium evidence). Successful measures simultaneously address fundamental issues related to the enhancement of
local collective actions, and the creation of approaches at national and international scales that complement, support,
and legitimize such local actions. [5.4, 5.6]
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
296
5.1. Introduction: Why the Local is Important
Disasters occur first at the local level and affect local people. These
localized impacts can then cascade to have national and international
ramifications. As a result, the responsibility for managing such risks
requires the linkage of local, national, and global scales (Figure 5-1).
Some disaster risk management options are bottom-up strategies,
designed by and for local places, while other management options are
products of global negotiations (Chapter 7) that are then implemented
through national institutions (Chapter 6) to local levels. Institutions,
actors, governance, and geographic units of analysis are not uniform
across these scales. Even within each scale there are differences. While
some communities are able to cope with disaster risks, others have
limited disaster resilience and capacity to cope with present disaster risk
let alone adapt to climate variability and extremes. This is the topic of
this chapter: to present evidence on where disasters are experienced,
how disaster risks are managed at present, and the variability in coping
mechanisms and capacity in the face of climate variability and change,
all from the perspective of local places and local actors. The chapter
explores three themes: how disaster risks are managed at present; how the
impact of climate extremes threatens human security at the local level; and
the role of scale and context in shaping variability in vulnerability,
coping, adaptive capacity, and the management of disaster risks and
climate extremes at the local level.
The idea of local has many connotations. For the purposes of this report,
local refers to a range of places, management structures, institutions,
social groupings, conditions, and sets of experiences and knowledge
that exist at a scale below the national level. As administrative units,
local can range from villages, districts, suburbs, cities, and metropolitan
areas, through to regions, states, and provinces. The conception of local
includes the set of institutions (public and private) that maintain and
protect local people as well as those that have some administrative
control over space and resources. In these places, choices and actions
for disaster risk management and adaptation to climate extremes can
be initially independent of national interventions. At the local level there
is traditional knowledge about disaster risk and grassroots actions to
manage it. Functional or physical units such as watersheds, ecological
zones, or economic regions operate at the local level, including the
private and public institutions that govern their use and management.
Each of the differing connotations of local means that there are differing
approaches to and contents of disaster risk management practice,
differing stakeholders and interest groups, and more significantly,
differing relations with the national and international levels (Adger
et al., 2005). We recognize that states and provinces in many countries
are large complex entities with similar powers as smaller nations.
Where we discuss states and provinces and similar administrative
structures in this chapter, we refer to them as sub-national for clarification
purposes.
Chapter 5Managing the Risks from Climate Extremes at the Local Level
LOCAL
NATIONAL
GLOBAL
Actors
Households
Businesses
NGOs, Faith-based
Military
Local leaders
Actors
Businesses
NGOs
Governments
Actors
Elected officials
Businesses
NGOs
Faith-based
Military
Governance
Laws & regulations
Federal agencies
International treaties
Administrative Units
Villages
Community
City, Town, County,
Parish
State, Province
Geographic Units
Watershed
Coastal zone
Ecosystem
Megacity
Disaster Risk Management
Climate Change Adaptation
Information and
Knowledge Flows
Financial
Flow
Figure 5-1 | Linking local to global actors and responsibilities.
297
Local places vary in their disaster experience, who and what is at risk, the
potential geographical extent of the potential impact and responses,
and in stakeholders and decisionmakers. Local places have considerable
experience with short-term coping responses and adjustments to disaster
risk (UNISDR, 2004), as well as with longer-term adjustments such as
the establishment of local flood defenses, the selection of drought
resistant crops, or seasonal or longer migration by one or more family
members. For example, the use of remittances is a substantial source of
post-disaster income and is regularly used as a means for diversifying
livelihoods to enhance resilience and to proactively cope with extremes
(Adger et al., 2002).
Climate-sensitive hazards such as flooding, tropical cyclones, drought,
heat, and wildfires regularly affect many localities with frequent, yet
low-level, losses (UNISDR, 2009). Because of their frequent occurrence,
localities have developed extensive reactive disaster risk management
practices. However, disaster risk management also entails the day-to-day
struggle to improve livelihoods, social services, and environmental
services. Local response and long-term adaptation to climate extremes
will require disaster risk management that acknowledges the role of
climate variability. This can mean a modification and expansion of local
disaster risk management principles and experience through innovative
organizational, institutional, and governmental measures at all
jurisdictional levels (local, national, international). Institutionally-driven
arrangements may constrain or impede local actions and ultimately
limit the coping capacity and adaptation of local places.
Local communities routinely experience hazard impacts, with many
resulting from extreme weather and climate events (see Chapter 3). The
significance of discussing these from the local perspective is that extreme
weather and climate events will vary from place to place and not all
places have the same experience with that particular initiating event.
Research demonstrates that disaster experience influences proactive
behaviors in preparing for and responding to subsequent events (see
Section 5.4.1). In the context of climate change, some localities could be
experiencing certain types of hazards for the first time and not have the
existing capacities for preparedness and response. For example,
Hurricane Catarina, the first South Atlantic hurricane that made landfall
as a category 1 storm just north of Porto Alegre, Brazil, in March 2004
(McTaggart-Cowan et al., 2006), was the region’s first local experience
with a hurricane. However, there is low confidence in attribution of any
long-term increases in hurricane formation in this ocean region where
tropical cyclones had not been previously recorded (Table 3-2). Finally,
not all of the extreme events become severe enough to cause a disaster
of national or international magnitude, yet they will create ongoing
problems for local disaster risk management.
The second theme of the chapter examines how climate extremes could
threaten the human security of local populations. Because these risks
often affect the basic functioning of society, it is increasingly recognized
that climate change adaptation and disaster risk management should
be integral components of development planning and implementation
to increase sustainability (Thomalla et al., 2006; see Box 5-1). In other
words, both have to be mainstreamed into national development plans,
poverty reduction strategies, sectoral policies, and other development
tools and techniques (UNDP, 2007). For example, rural communities in
many world regions face greater risks of livelihood loss resulting from
flooding of low-lying coastal areas, water scarcity and drought, decline
in agricultural yields and fisheries resources, and loss of biological
resources (Osman-Elasha and Downing, 2007). In some African countries
where recurrent floods are closely linked with El Niño-Southern
Oscillation (ENSO) events (Ward et al., 2010), the result is major
economic and human loss seen in places such as Mozambique (Mirza,
2003; Obasi, 2005) and Somalia. For such communities, with less
developed infrastructure and health services, the impacts of floods are
often further exacerbated by health problems associated with water
scarcity and quality, such as malnutrition, diarrhea, cholera, and malaria
(Kabat et al., 2002).
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
Box 5-1 | Climate Change and Violent Conflict
Linking climate change and violent conflict is controversial. The conceptual debate links climate change to resource scarcity (or those
essential resources to support livelihoods), which in turn leads to human insecurity. At the local scale, there are two distinct outcomes:
armed conflict or migration, with the latter potentially leading to increased conflict in the receiving locality (Barnett and Adger, 2007;
Nordås and Gleditsch, 2007). For example, some research suggests that environmental stresses feed the tensions between localities as
they compete for land to support their livelihoods (Kates, 2000; Barnett, 2003; Osman-Elasha and El Sanjak, 2009). Extreme events such
as droughts and heat waves could increase these tensions in areas already facing situations of water scarcity and environmental
degradation, giving rise to conflicts resulting in the dislocation of large numbers of refugees and people within and across borders.
However, there is limited agreement and evidence to support the link between climate change and violent conflict, especially in Africa
(Burke et al., 2009; Buhaug, 2010). While the causal chain suggested in the literature (climate change increases the risk of violent conflict)
has found currency within the policy community, it has not been adequately substantiated in the scientific literature (low agreement and
limited evidence). Where empirical studies exist, they are methodologically flawed in a number of ways: not controlling for population
size; focusing only on conflict cases and not all migration instances; using aggregated, not disaggregated climate data at sub-national
scales; and having inherent inconsistencies in the time frames used (short-term variability in violent conflict; longer-term variability in
climate). More research on the local climate-conflict nexus is warranted in order to determine if a causal linkage exists.
298
In order to develop preparedness measures for disaster risk management
and climate adaptation, the vast contextual differences of localities
will have to be considered. They include differences in population
characteristics that influence vulnerability, differences in settlement
patterns ranging from urban to rural, differences in administrative units
from municipalities to provincial governments, and differences within
developing and developed country contexts. Given the wide disparities,
it is clear that single solutions for disaster risk management are not
possible. For example, there are differences between urban and rural
communities in terms of disaster and climate change vulnerability and
disaster risk and adaptation options. Given the rapid pace of urbanization
and diffusion of communication and transportation networks into distant
areas, the sharp distinction between urban and rural is less visible in
many areas. In its place is a continuum with local places exhibiting
both rural and urban characteristics with a mix of vulnerabilities and
jurisdictional issues that are neither totally urban nor rural (McGregor
et al., 2006; Aragon-Durand, 2007).
Scalar considerations must also be emphasized in planning. Efforts to
forge greater and more equitable capacity at the local scale have to be
supported by policies at the national level to increase the ability of local
institutions and communities to cope with present and future risks from
climate-sensitive hazards. To effectively reduce vulnerabilities to hazards
associated with climate change, coordination across different levels and
sectors is required, in addition to the involvement of a broad range of
stakeholders beginning at the local level (UNISDR, 2004; DFID, 2006;
Tearfund., 2006; Devereux and Coll-Black, 2007; Davies, 2009). The larger
global context within which a locality is situated affects outcomes. It is
possible that the history of resource exploitation, globalization, and the
processes of development as currently practiced may be increasing, rather
than reducing disaster vulnerability at the local level (see Chapter 2).
Those choosing strategies for reducing disaster risk and adapting to
climate change, especially in developing countries, need to take these
processes into account (UNISDR, 2009).
These contextual factors are critical to planning for climate extremes.
They suggest the need for strengthening coordination between climate
change adaptation and disaster risk management locally that will in turn
improve the implementation of plans (Mitchell and van Aalst, 2008).
Such coordination is also needed in order to avoid any negative impacts
across different sectors or scales that could potentially result from
fragmented adaptation and development plans. This is evident in the
implementation of some of the adaptation strategies, such as large-
scale agriculture, irrigation, and hydroelectric development, that may
benefit large groups or the national interests but may also harm local,
indigenous, and poor populations (Kates, 2000; Rojas Blanco, 2006).
Some sources believe that it is essential that any new disaster risk
reduction or climate change adaptation strategies must be built on
strengthening local actors and enhancing their livelihoods (Osman-Elasha,
2006a). Moreover, a key aspect of planning for adaptation at the local
level is the identification of the differentiated social impacts of climate
change based on gender, age, disability, ethnicity, geographical location,
livelihood, and migrant status (Tanner and Mitchell, 2008). Emphasis
needs to be given to identifying the adaptation measures that serve the
most vulnerable groups, address their urgent needs, and increase their
resilience. This often means using a more coordinated and integrated
management approach with the involvement of diverse stakeholder
groups (Sperling and Szekely, 2005), which may assist in avoiding
maladaptation across sectors or scales and provide for win-win solutions.
5.2. How Local Places Currently Cope
with Disaster Risk
Local people everywhere have developed skills, knowledge, and
management systems that enable them to interact with their environment.
Often these interactions are beneficial and provide the livelihoods that
people living in local places depend on. At the same time communities
have developed ways of responding to disruptive environmental events.
These coping mechanisms include measures that seek to modify
the impacts of disruptive events, modify some of the attributes or
environmental aspects of the events themselves, and/or actions to share
or reduce the disaster risk burdens (Burton et al., 1993). By the same
token, some actions taken at local levels (e.g., deforestation and coral
mining) may also increase disaster risks. It is important to acknowledge
that while climate change may alter the magnitude and/or frequency of
some climatic extremes (see Chapter 3), other environmental, social,
political, or economic processes (many of them also global in scale) are
affecting the abilities of communities to cope with disaster risks and
climate-sensitive hazards (Wisner et al., 2004; Adger and Brown, 2009).
Accordingly, disaster losses have increased significantly in recent decades
(UNDP, 2004; UNISDR, 2004). These social, economic, and political
processes are complex and deep seated and present major obstacles to
reducing disaster risk, and may constrain efforts to reduce community
Chapter 5Managing the Risks from Climate Extremes at the Local Level
FAQ 5.1 | Why is the local context important
in climate change adaptation and
disaster risk management?
In the context of this report, the local refers to a range of
places (community, city, province, region, state), management
structures, institutions, social groupings, conditions, and sets
of experiences and knowledge that exist at a scale below the
national level. It also includes the set of institutions (public
and private) that maintain and protect social relations as well
as those that have some administrative control over space
and resources. The definition of the local influences the context
for disaster risk management, the experience of disasters, and
conditions, actions, and adaptation to climate changes. Local
is important because locals respond and experience disasters
at first hand, they retain local and traditional knowledge
valuable for disaster reduction and adaptation plans, and
lastly they implement adaptation plans.
299
vulnerabilities to extreme events under conditions of climate change. In
Section 5.2 we outline three common local-level coping strategies:
emergency assistance and disaster relief, population movements, and
recovery and reconstruction.
5.2.1. Emergency Assistance and Disaster Relief
Humanitarian assistance is often required when other measures to
reduce disasters have been unsuccessful, and plays a critical role in
helping local people cope with the effects of disasters. Such relief often
helps to offset distress and suffering at the local level and to assist in
recovery and rehabilitation. Sometimes external relief is unnecessary or
inappropriate because the local people affected by disasters often are not
completely helpless or passive and are capable of helping themselves
(Cuny, 1983; De Ville de Groyet, 2000). This view is sustained by
commonplace definitions of disasters as situations where communities
or even countries cannot cope without external assistance (Cuny, 1983;
Quarantelli, 1998).
It is important to realize that the first actors providing assistance during
and after disasters are members of the affected community (De Ville de
Groyet, 2000) who provide relief through local charities, kinship networks,
or local governments. In isolated communities such as those in the outer
islands of small-island developing states, external assistance may be
subject to considerable delay and self-help is an essential element of
response, especially in the period before assistance arrives. Typically,
emergency assistance and disaster relief in developed countries comes
in the form of assistance from national and state/provincial level
governments to local communities. The provision of international relief
is usually from members of the Organisation for Economic Co-operation
and Development to developing countries (Development Initiatives,
2009). The international provision of disaster relief to local places has
become highly sophisticated and much broader in scope over the past two
decades, involving both development and humanitarian organizations,
with the increasing recognition that external relief providers make use
of local knowledge in planning their relief efforts (Morgan, 1994; Darcy
and Hofmann, 2003; Méheux et al., 2010). The relief itself includes such
things as assistance in post-disaster assessment, food provision, water
and sanitation, medical assistance and health services, household goods,
temporary shelter, transport, tools and equipment, security, logistics,
communications, and community services (Bynander et al., 2005; Cahill,
2007). Many of these activities are organized into clusters of specialists
from multilateral organizations and nongovernmental organizations
(NGOs), among others, coordinated by the United Nations.
While much of the relief tends to be organized at more of a national and
international scale than local scale, the distribution and use of relief
occur at the local level. From this perspective it is vital to understand
what is locally appropriate in terms of the type of relief provided, and
how it is distributed (Darcy and Hofmann, 2003; Kovác and Spens,
2007). Similarly, local resources and capacities should be utilized as
much as possible (Beamon and Balcik, 2008). There has also been a
trend towards international humanitarian organizations working with
local partners, although on occasion this can result in the imposition of
external cultural values resulting in resentment or resistance (Hillhorst,
2002).
Relief, nevertheless, is often a critically important strategy for coping.
Relief organizations have built capacity based on experience in recent
years, have become increasingly accountable, and are obliged to follow
humanitarian principles. Despite these improvements, some problems
remain. Relief cannot cover all losses, most of which are borne locally.
Relief can undermine local coping capacities and reduce resilience and
sustainability (Susman et al., 1983; Waddell, 1989), reinforce the status
quo that was characterized by vulnerability (O’Keefe et al., 1976), and in
some cases, serve to remove independence or autonomy from disaster
‘victims’ so that ownership of the event and control over the recovery
phase is lost at the local level (Hillhorst, 2002). Relief is often
inequitably distributed and in some disasters there is insufficient relief.
Corruption is also a factor in some disaster relief operations with local
elites often benefiting more than others (Pelling and Dill, 2010).
Humanitarian organizations are increasingly aware of these concerns
and many are addressing them through coordination of activities,
addressing gendered inequalities, and working in partnership with local
organizations in disaster relief. There is also a growing recognition of
the need for accountability in humanitarian work (Humanitarian
Accountability Partnership International, 2011).
Not all disasters engender the same response, as local communities
receive different levels of assistance. For example, those people most
affected by a small event can suffer just as much as a globally publicized
big event but are sometimes overlooked by relief agencies. Fast onset
and unusual disasters such as tsunamis generate considerably more
public interest and contributions from governments, NGOs, and the
public, sometimes referred to as the CNN factor (Olsen et al., 2003).
Disasters that are overshadowed by other newsworthy or media events,
such as coverage of major sporting events, are often characterized by
lower levels of relief support (Eisensee and Stromberg, 2007). Where
there is widespread media coverage, NGOs and governments are often
pressured to respond quickly with the possibility of an oversupply of
relief and personnel. This has worsened in recent times when reporters
are ‘parachuted’ into disaster sites often in advance of relief teams but
they have little understanding of the contextual factors that often
underlie vulnerability to disasters (Silk, 2000). Such media coverage
often perpetrates disaster myths such as the prevalence of looting,
helplessness, and social collapse, putting pressure on interveners to
select military options for relief when humanitarian assistance would be
more helpful (Tierney et al., 2006).
Relief is politically more appealing than disaster risk management (Seck,
2007) and it often gains much greater political support and funding than
measures that would help offset the need for it in the first place.
Providing relief reflects well on politicians (both in donor and recipient
countries) who are seen to be caring, taking action, and responding to
public demand (Eisensee and Stromberg, 2007).
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
300
Major shares of the costs of disaster relief and recovery still fall on the
governments of disaster-affected countries. Bilateral relief is limited to
materials from donor countries and most relief is subject to relatively
strict criteria to reduce perceived levels of corruption. In both cases,
flexibility is heavily restricted. Relief can also produce local economic
distortions such as causing shops to lose business as the market
becomes flooded with relief supplies. These problems can be overcome
by directly transferring cash to local people to buy building materials,
seed, and the like. Such programs have performed well where local
supplies are available (Farrington and Slater, 2009). At the same time,
there is the view that disaster relief can create a culture of dependency
and expectation at the local level (Burby, 2006), where disaster relief
becomes viewed as an entitlement program as local communities are
not forced to bear the responsibility for their own locational choices,
land use, and lack of mitigation practices.
5.2.2. Population Movements
A second coping strategy is population movements. Natural disasters
are linked with population movements in a number of ways (Hunter,
2005; Perch-Nielson et al., 2008; Warner et al., 2010). Evacuations occur
before, during, and after some disaster events. Longer-term relocation of
affected communities sometimes occurs. Relocations can be temporary or
permanent. These different forms of population movements have variable
social, psychological, health, and financial implications for the communities
concerned. Population movements may also be differentiated on the
basis of whether the mobility is voluntary or forced (displacement) and
whether or not international borders are crossed. Most contemporary
research views population mobility as a continuum from completely
voluntary movements to completely forced migrations (Laczko and
Aghazarm, 2009). The United Nations Office for the Coordination of
Humanitarian Affairs and the Internal Displacement Monitoring Centre
estimated that at least 36 million people were displaced by natural
disasters in 2008. While these displaced people would come from and
arrive at local origins and destinations there is little information on the
local implications and time frames of the displacement (UNOCHA and
IDMC, 2009).
Where climate change increases the marginality of livelihoods and
settlements beyond a sustainable level, communities may be forced to
migrate or be displaced (McLeman and Smit, 2006). While migration
typically has many causes, of which the environment (including climate)
is just one factor, extremes often serve as precipitating events (Hugo,
1996). Furthermore, a number of researchers consider that climate-
related migration, other than forced displacement, may not necessarily
be a problem and indeed may be a positive adaptive response, with
people who remain at the place of origin benefitting from remittances
(Barnett and Webber, 2009; Tacoli, 2009). Nomadic pastoralists migrate
as part of their livelihoods but often respond to disruptive events by
modifying their patterns of mobility (Anderson et al., 2010). Migration
is highly gendered in terms of both drivers and impacts, which differ
between men and women, although it is not clear how these differences
might be played out in the context of climate change (Hugo, 2010).
Global estimations provide little insight into the local implications of such
large-scale migratory patterns. Migration will have local effects, not only
for the communities generating the migrants, but those communities
where they may settle. Barnett and Webber (2009) also note that the less
voluntary the migration choice is, the more disruptive it will become. In
the context of dam construction, for example Hwang et al. (2007) found
that communities anticipating forced migration experienced stress.
Hwang et al. (2010) also found that forced migration directly led to
increased levels of depression and the weakening of social safeguards
in the relocation process. Much post-disaster relocation is temporary,
which is also associated with psychological and social effects such as
disruption of social networks and trauma (Neria et al., 2009).
One outcome of climate change is that entire communities could be
required to relocate and in some cases, such as those living in atoll
Chapter 5Managing the Risks from Climate Extremes at the Local Level
FAQ 5.2 | What lessons have been learned about effective disaster management and climate
change adaptation at the local scales?
In fostering sustainable and disaster-resilient areas, local response to climate extremes will require disaster risk management that
acknowledges the role of climate variability and change and the associated uncertainties and that will contribute to long-term adaptation.
In order to anticipate the risks and uncertainties associated with climate change there are a number of emerging approaches and
responses at the local level. One set of responses focus on integrating information about changing climate risks into disaster planning
and scenario assessments of the future. Setting up plans in advance, for example, enabled communication systems to be strengthened
before the extreme event struck. Another is community-based adaptation (CBA), which helps to define solutions for managing risks
while considering climate change. CBA responses provide increased participation by locals and recognition of the local context and the
access to adaptation resources and promote adaptive capacity within communities. A critical factor in community-based actions is that
community members are empowered to take control of the processes involved. Scaling up community-based approaches poses a
challenge as well as integrating climate information and other interventions such as ecosystem management and restoration, watershed
rehabilitation, agroecology, and forest landscape restoration. These types of interventions protect and enhance natural resources at the
local scale, improve local capacities to adapt to future climate, and may also address immediate development needs.
301
countries, the relocation will be international. Such relocation can have
significant social, cultural, and psychological impacts (Campbell, 2010b).
Community relocation schemes are those in which whole communities
are relocated to a new non-exposed site. Perry and Lindell (1997)
examined one such instance in Allenville, Arizona. They developed a set
of five principles for achieving positive outcomes in relocation projects:
1) the community to be relocated should be organized; 2) all potential
relocatees should be involved in the relocation decisionmaking process;
3) citizens must understand the multi-organizational context in which
the relocation is to be conducted; 4) special attention should be
given to the social and personal needs of the relocatees; and 5) social
networks need to be preserved. For many communities relocation is
difficult, especially in those communities with communal land ownership.
In the Pacific Islands, for example, relocation within one’s own lands is
least disruptive but leaving it completely is much more difficult, as is
making land available for people who have been relocated (Campbell,
2010b).
5.2.3. Recovery and Reconstruction
Recovery and reconstruction include actions that seek to establish or
re-establish the everyday life of the locality affected by disaster (Hewitt,
1997). Often reconstruction enables communities and businesses to
return to the same conditions that existed prior to the disaster, and in
so doing create the potential for further similar losses, thus reproducing
the same exposure that resulted in disaster in the first place (Jha et al.,
2010). Recovery and reconstruction (especially housing rehabilitation
and rebuilding) are among the more contentious elements of disaster
response. One of the major issues surrounding recovery is the lack
of clarity between recovery as a process and recovery as an outcome.
The former emphasizes betterment processes where pre-existing
vulnerability issues are addressed. The latter focuses on the material
manifestation of recovery such as building houses or infrastructure.
Often following large disasters, top-down programs result in rebuilding
houses but fail to provide homes (Petal et al., 2008). Moreover, haste
in reconstruction, while achieving short-term objectives, often results
in unsustainable outcomes and increasing vulnerability (Ingram et
al., 2006). As seen in the aftermath of Hurricane Katrina, there are
measureable local disparities in recovery, leading to questions of
recovery for whom and recovery to what (Curtis et al., 2010; Finch et
al., 2010; Stevenson et al., 2010). There are a number of obstacles to
effective and timely reconstruction including lack of labor, lack of
capacity among local construction companies, material shortages,
resolution of land tenure considerations, and insufficiency of funds
(Keraminiyage et al., 2008). While there is urgency to have people
re-housed and livelihoods re-established, long-term benefits may be
gained through carefully implemented reconstruction (Hallegatte, 2008;
Hallegatte and Dumas, 2009) in order to achieve greater disaster
resilience.
Most research on recovery and reconstruction has tended to focus on
housing and the so-called lifelines of infrastructure: electricity, water
supply, and transport links. Less is published on the equally important,
if indeed not more so, rehabilitation of livelihoods, and addressing the
problems of power inequities that often include land and resource
grabbing by the economic and politically powerful after disaster in both
developed and developing countries. Agricultural rehabilitation (e.g.,
the provision of seeds, planting material, fertilizers, and stock, and the
remediation of land) is particularly important where local livelihoods
are directly affected such as in subsistence or semi-subsistence societies
(Dorosh et al., 2010). In addition, some climate-related disaster events,
such as droughts, do not always directly destroy the built environment
infrastructure (like flooding or tropical cyclones) so the rehabilitation of
livelihoods, in particular sustainable livelihoods, becomes an important
aspect of disaster risk reduction and development (Nakagawa and
Shaw, 2004).
As with relief, major problems can occur where planning and
implementation of recovery and reconstruction is taken out of the
hands of the local communities concerned. In addition, the use of
inappropriate (culturally, socially, or environmentally) materials and
techniques may render rebuilt houses unsuitable for their occupants
(Jha et al., 2010). As Davidson et al. (2007) found, this is often the
case and results in local community members having little involvement
in decisionmaking for the recovery process; instead they are used to
provide labor. It is also important to acknowledge that post-disaster
recovery often does not reach all community members and in many
recovery programs, the most vulnerable, those who have suffered the
greatest losses, often do not recover from disasters and endure long-
term hardship (Wisner et al., 2004). In this context, it is important to
take into account the diversity of livelihoods in many local areas and to
work with local residents and stakeholders to develop strategies that
are potentially more resilient in the face of future events (Pomeroy et
al., 2006).
During the post-recovery phase, reconstruction requires weighing,
prioritizing, and sequencing of policy programming, given the multiple,
and sometimes competing agendas for most decisionmakers and
operational actors. Often there are opportunities for change in policy
directions and agenda setting at local to national levels at this time
(Birkland, 1997). The post-event lobbying for action and resources
requires a balance between short-term needs and long-term goals.
The most significant is the pressure to quickly return to conditions prior
to the event rather than incorporate longer-term and more sustainable
development policies (Christoplos, 2006; Kates et al., 2006). How long
such a window will stay open or precisely what factors will make it
close under a given set of conditions is not well known even though
three to six months has been recognized in specific cases (Kates et al.,
2006).
The most often used strategies for coping with present disaster risk at
the local level are emergency assistance (including disaster relief),
population movements, and recovery and reconstruction. As illustrated
above, there is considerable variability among and between local places
in how these actions are implemented and the impacts of their use.
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
302
5.3. Anticipating and Responding
to Future Disaster Risk
This section examines how local places anticipate future risks and how
they respond to them. In addition to enhanced communication, other
approaches to anticipating and responding to future risks include
structural interventions such as dikes or dams, natural resources planning
and ecosystem protection, and storage and rationing of resources.
5.3.1. Communicating Risk
Effective communication is necessary across the full cycle of disaster
management: reduction, preparedness, response, and recovery, especially
at the local level where communications face particular constraints and
possibilities. A burgeoning field of research explores the barriers to
communicating the impacts of climate change to motivate constructive
behaviors and policy choices (Frumkin and McMichael, 2008).
Communicating the likelihood of extreme impacts of climate change
also presents an important and difficult challenge (Moser and Dilling,
2007). Research on climate communications addresses how information
can be designed, and the mechanisms and timing of its distribution.
5.3.1.1. Message Design
As used here, the term risk communication refers to intentional efforts
on the part of one or more sources (e.g., international agencies, national
governments, local government) to provide information about hazards
and hazard adjustments through a variety of channels to different audience
segments (e.g., the general public, specific at-risk communities). The
characteristics of messages that have a significant impact on local adoption
of adjustments involve information quality (specificity, consistency, and
source certainty), information reinforcement (number of warnings and
repetition) (Mileti and O’Brien, 1992; O’Brien and Mileti, 1992; Mileti
and Fitzpatrick, 1993), and the ways in which information is designed.
Messages targeted to specific audiences are more readily received
(Maibach et al., 2008) than those which are not. Targeting threats to
future generations may generate more concern than overt actions to
reduce contemporary climate change impacts (Maibach et al., 2008). In
addition, communication is more effective when the information
regarding risk does not exceed the capacity for coping and therefore
galvanizes resilience (Fritze et al., 2008). Some research suggests that a
focus on personal risk of specific damages of climate change can be a
central element in motivating interest and behavior change (Leiserowitz,
2007). Risk messages vary in threat specificity, guidance specificity,
repetition, consistency, certainty, clarity, accuracy, and sufficiency (Mileti
and Sorensen, 1990; Mileti and Peek, 2002).
Communications that include social,
interpersonal, physical, environmental,
and policy factors can foster civic engagement and social change
fundamental to reducing risk (Brulle, 2010). A participatory approach
highlights the need for multiple pathways of communication that
engenders credibility, trust, and cooperation (NRC, 1989; Frumkin
and McMichael, 2008), which are especially important in high-stress
situations such as those associated with climate extremes. For example,
participatory video production is effective in communicating the
extreme impacts of climate change (Suarez et al., 2008; Baumhardt et
al., 2009). Participatory video involves a community or group creating
their own videos through storyboarding and production (Lunch and
Lunch, 2006). Such projects are traditionally used in contexts, such as
poor communities, where there are constraints to accessing accurate
climate information (Patt and Gwata, 2002; Patt and Schröter, 2008).
Engaging with community leaders or opinion leaders in accessing social
networks through which to distribute information is another approach,
traditionally used by health educators but also applicable to the
translation of climate risks in a community context (Maibach et al.,
2008). Another approach used in health communications that is relevant
to climate education is the ‘community drama’ in which community
members engage in plays to communicate health risks (Middlekoop et al.,
2006). These types of communication projects can motivate community
action necessary to promote preparedness (Semenza, 2005; Jacobs et
al., 2009).
Visualizing methods such as mapping, cartographic animations, and
graphic representations are also used to engage with stakeholders who
may be impacted by extreme events (McCall, 2008; A. Shaw et al.,
2009). Many programs are developing ways to use visualizations to
help decisionmakers adapt to a changing environment, suggesting that
such tools can increase climate literacy (Niepold et al., 2008).
Visualizations can be powerful tools, but issues of validity, subjectivity,
and interpretation must be seriously considered in such work
(Nicholson-Cole, 2004). These communications are most effective when
they take local experiences or points of view and locally relevant places
into account (O’Neill and Ebi, 2009). Little evaluation has been done of
visualization projects, therefore leaving a gap in understanding of how
to most effectively communicate future risks of extreme events.
5.3.1.2. Modes and Timing of Risk Communication
The generation and receipt of risk information occurs through a diverse
array of channels. They include: interpersonal contact with particular
researchers; planning and conceptual foresight; outside consultation on
the planning process; user-oriented transformation of information; and
individual and organizational leadership (NRC, 2006). Researchers have
long recognized a variety of informal information source vehicles including
peers (friends, relatives, neighbors, and coworkers) and news media
(Drabek, 1986). These sources systematically differ in terms of such
characteristics as perceived expertise, trustworthiness, and protection
responsibility (Lindell and Perry, 1992; Lindell and Whitney, 2000; Pulwarty,
2007). Risk-area residents use information channels for different purposes:
the internet, radio, and television are useful for immediate updates;
meetings are useful for clarifying questions; and newspapers and
brochures are useful for retaining information that might be needed
later. In addition, within-community discussions on risks to livelihoods,
Chapter 5Managing the Risks from Climate Extremes at the Local Level
303
such as during droughts, act as mechanisms for risk communication and
response actions (Dekens, 2007).
Policies and actions affecting communications and advanced warning
have a major impact on the adaptive capacity and resilience of
livelihoods. The collection and transmittal of weather- (and climate-)
related information is often a governmental function and timely
issuance remains a key weakness in climate information systems,
especially for communication passed on to communities from the
national early warning units (UNISDR, 2006). There are other localized
forms of communication that can be used rapidly, such as neighborhood
watch systems (Lichterman, 2000). Some private communication methods,
such as text messaging, Facebook, and Twitter, may reach affected
populations before government directives (Palen et al., 2007). However,
some research shows that there has been too much reliance on one-way
devices for communication (such as the radio), which were felt to be
inadequate for agricultural applications (for example, farmers are not
able to ask further questions regarding the information provided)
(Ziervogel, 2004). Within many rural communities, low bandwidth and
poor computing infrastructure pose serious constraints to risk-message
receipt. Such gaps are evident in developed as well as lesser-developed
regions.
The degree of acceptability of information and trust in the providers
dictates the context of communicating disaster and climate information
(see Box 5-2). Pre-decisional processes (reception, attention, and
comprehension) influence the disaster message’s effectiveness (Lindell
and Perry, 2004). Several studies have identified the characteristics of
pre-decisional practices that lead to effective communication over the
long-term (Fischhoff, 1992; Cutter, 2001; Pulwarty, 2007). These include:
1) understanding the goals, objectives, and constraints of communities
in the target system; 2) mapping practical pathways to different outcomes
carried out as joint problem definition and fact-finding strategies
among research, extension, and farmer communities; 3) bringing the
delivery persons (e.g., extension personnel, research community, etc.) to
an understanding of what has to be done to translate current information
into usable information; 4) interacting with actual and potential users
to better understand informational needs, desired formats of information,
and timeliness of delivery; 5) assessing impediments and opportunities
to the flow of information including issues of credibility, legitimacy,
compatibility (appropriate scale, content, match with existing practice),
and acceptability; and 6) relying on existing stakeholders’ networks and
organizations to disseminate and assess climate information and
forecasts.
Much research has yet to be done regarding risk communication on
climate change. There has been little systematic investigation, for
example, on message effectiveness in prompting local action based on
differing characteristics such as the precision of message dissemination,
penetration into normal activities, message specificity, message
distortion, rate of dissemination over time, receiver characteristics,
sender requirements, and feedback (Lindell and Perry, 1992; NRC, 2006).
Little research attention has been devoted to how information can be
distributed within a family, although the existing research does show
there are emotional, social, and structural barriers to such distribution
(Norgaard, 2009).
5.3.1.3. Warnings and Warning Systems
The disaster research community has shown that warnings of impending
hazards need to be complemented by information on the risks actually
posed by the hazards as well as the potential strategies and pathways
to mitigate the damage in the particular context in which they arise
(Drabek, 1999; UNISDR, 2006). Local-level early warnings based on
traditional knowledge (e.g., water turning a different color, winds
shifting) are frequently used. The use of radios, megaphones, and cell
phones are also used at the local level to warn.
Effective early warning implies information interventions into an
environment where vulnerability is assumed (Olson, 2000). This backdrop
is reinforced through significant lessons that have been identified from
the use of seasonal climate forecasts over the past 15 years (Podestá et
al., 2002; Pulwarty, 2007). It is now widely accepted that the existence
of predictable climate variability and impacts are necessary but not
sufficient to achieve effective use of climate information, including
seasonal forecasts. The practical obstacles to using information about
future conditions at the local scale are diverse. They include: limitations
in modeling the climate system’s complexities (e.g., projections having
coarse spatial and temporal resolution; limited predictability of some
relevant variables; and forecast skill characterization; see Chapter 3);
procedural, institutional, and cognitive barriers in receiving or
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
Box 5-2 | Successful Communication of Local
Risk-Based Climate Information
The following questions have been identified as shaping the
successful communication of risk-based climate information
(Ascher, 1978; Fischhoff, 1992; Pulwarty, 2003):
What do people know and believe about the risks being
posed?
What is the past experience/outcomes of information
use?
Is the new information relevant for decisions in the
particular community?
Are the sources/providers of information credible to the
intended user?
Are practitioners (e.g., farmers) receptive to the
information and to research?
Is the information accessible to the decisionmaker?
Is the information compatible with existing decision
models (e.g., for farming practice)?
Does the community (or individuals in the community)
have the capacity to use information?
304
understanding climatic information; and the capacity and willingness of
decisionmakers to modify actions (Kasperson et al., 1988; Stern and
Easterling, 1999; Roncoli et al., 2001; Patt and Gwata, 2002; Marx et al.,
2007). In addition, functional, structural, and social factors inhibit joint
problem identification and collaborative knowledge production
between providers and users. These include divergent objectives, needs,
scope, and priorities; different institutional settings and standards; as
well as differing cultural values, understanding, and mistrust (Pulwarty
et al., 2004; Rayner et al., 2005; Weichselgartner and Kasperson, 2010).
Significant advances in warning systems in terms of improved monitoring,
instrumentation, and data collection have occurred (Case Study 9.2.11),
but the management of the information and its dissemination to at-risk
populations is still problematic (Sorensen, 2000). Researchers have
identified several aspects of information communication, such as
stakeholder awareness, key relationships, and language and terminology,
that are socially contingent in addition to the nature of the predictions
themselves. More is known about the effects of these message
characteristics on warning recipients than is known about the degree to
which generators and providers of information including hazards
researchers address them in their risk communication messages. For
example, warnings may be activated (such as the tsunami early warning
system), yet fail to reach potentially affected communities (Oloruntoba,
2005). Similarly, many communities do not have access to climate-
sensitive hazard warning systems such as tone alert radio, emergency
alert system, emergency phone numbers (reverse 911 in the United States,
but in other parts of the world you call 110, 112, or other numbers), and
thus never hear the warning message, let alone act upon the information
(Sorensen, 2000). On the other hand, Valdes (1997) demonstrated that
flood warning systems based on community operation and participation
in Costa Rica make a difference as to whether early warnings are acted
upon to save lives and property. Implementing community early warning
systems (such as the correlations of rain data and water levels among
monitoring stations along the river) serves to encourage communities to
become more proactive in their hazard mitigation approaches.
Part of the research gap regarding risk communication stems from the
lack of projects that can be tested and shown to affect preparedness.
On the most basic level, there is considerable understanding of the
information needed for preparing for disasters, but less-specific
understanding of what information and trusted communication
processes are necessary to generate local confidence and preparedness
for climate change (Fischhoff, 2007). The very discussion of climate
forecasts and projections within potentially impacted communities has
served as a vehicle for democratizing the drought discourse in Ceará in
Northeast Brazil (Finan and Nelson, 2001). Developing a seamless
continuum across emergency responses, preparedness, and coping and
adaptation requires insight into the demands that different types of
disasters will place upon the local area and the need to perform
basic emergency functions – pre-event assessments, proactive hazards
mitigation, and incident management (Lindell and Perry, 1996). As noted
in previous IPCC reports (IPCC, 2007), preparing for short-term disasters
enhances the capacity to adapt to longer-term climate change.
5.3.2. Structural Measures
Structural measures may be used to reduce the effects of climate-related
events such as floods, droughts, coastal erosion, and heat waves.
Structural interventions to reduce the effects of extreme events often
Chapter 5Managing the Risks from Climate Extremes at the Local Level
Box 5-3 | Large Dams in Brazil: Scalar Challenges to Climate Adaptation
Effective climate adaptation requires consideration of cross-scale management concerns. Any project or impact that crosses jurisdictions
from local to regional to national to transnational is best planned using a perspective that takes into account all levels of management
(Adger et al., 2005). The planned or built large dams in Amazonia, Brazil (McCormick, 2011), exemplify these issues. These dams are
related to water management and would cross local, regional, and national boundaries. At the national level, these dams would provide
large-scale energy needs and serve major urban centers and industrial sectors across the country. At the regional level, the large
Amazonian dams could both generate energy and assist in drought management through storage of hydrological resources (Postel et al.,
1996). Because of the expansive range and impacts of large dams, their planning and management raise a variety of scalar concerns
about climate adaptation. While on one level a dam may present benefits regionally and nationally, it may also cause serious
environmental and social problems locally (McCormick, 2009). For example, dams upstream may lead to erosion and inundation of
deltas (Yang et al., 2011).
While there are many environmental benefits of hydroelectric power and large-scale water management, the uncertainty of climate
change could alter such benefits at local to global scales and influence the social and environmental ramifications of these projects. For
example, the flooding caused by the construction of reservoirs could result in migration of locally affected communities, thereby increasing
community fragmentation, poverty, and ill health of humans and biota (Kingsford, 2000). This becomes a local and regional impact of
dam construction that may increase vulnerability to climate change in many localities. Changing rainfall patterns that affect reservoir
levels could impact the availability of energy generation at the national level (DeLucena et al., 2009). Degradation of flora and fauna
also results in additional greenhouse gas emissions at local to global scales (Fearnside, 1995).
305
employ engineering works to provide protection from flooding such as
dikes, embankments, seawalls, river channel modification, flood gates,
and reservoirs. However, structural measures also include those that
strengthen buildings (during construction and retrofitting), that
enhance water collection in drought-prone areas (e.g., roof catchments,
water tanks, wells), and that reduce the effects of heat waves (e.g.,
insulation and cooling systems). Although many of these structural
interventions can achieve success in reducing disaster impacts, they can
also fail due to lack of maintenance, age, or due to extreme events that
exceed the engineering design level (Galloway, 2007; Doyle et al., 2008;
Galloway et al., 2009). In the event that the frequency and magnitude
of extreme events increase as a result of climate change, new design
levels may be necessary. Technical considerations should also include
local social, cultural, and environmental considerations (WMO, 2003;
Opperman et al., 2009).
Implementing structural measures from planning through implementation
that involve participatory approaches with local residents who are
proactively involved often leads to increased local ownership and more
sustainable outcomes. Such an example is a program of building,
managing, and maintaining cyclone shelters in Bangladesh (Zimmermann
and Stössel, 2011). One of the key reasons why sub-national structural
projects are often ineffective is that they are approved on the basis
of technical information alone, rather than based on both technical
information and local knowledge (ActionAid, 2005; Prabhakar et al.,
2009; see also Section 5.4.4). In addition, national legislation has
important influences on the choice of disaster risk reduction strategies
at the local level as local and national institutional arrangements that
often favor structural responses over other non-structural approaches
(Burby, 2006; Galloway, 2009). Technological responses alone may also
have unintended geomorphologic and social consequences, including
increasing flood hazard in downstream locations, increasing costs of
long-term flood protection works, or increasing coastal erosion in areas
deprived of sediments by coastal protection works (Adger et al., 2005;
Hudson et al., 2008; Box 5-3).
The method of protecting an entire area by building a dike has been in
use for thousands of years and is still being applied by communities in
flood-prone countries. Embankments, dikes, levees, and floodwalls are
all designed to protect areas from flooding by confining the water to a
river channel, thus protecting the areas immediately behind them.
Building dikes is one of the most economical means of flood control
(Asian Disaster Preparedness Centre, 2005). Dikes built by communities
normally involve low technology and traditional knowledge (such as
earth embankments). Sand bagging is also a very common form of
flood-proofing. Generally, structures that are built of earth are highly
susceptible to erosion, leading to channel siltation and reduced water
conveyance on the wet side and slope instability and failure on the dry
side. Slopes can be stabilized by various methods, including turfing by
planting vegetation such as Catkin grass and Vetiver grass in
Bangladesh and Thailand, respectively. However there is a continuing
debate in the region as to whether the grass strips prevent erosion,
whether erosion is in fact the main problem instead of soil fertility, and
whether farmers still need slope stabilization (Forsyth and Walker,
2008).
Decisionmaking for large-scale structural measures is often based on
cost-benefit analyses and technical approaches. In many cases,
particularly in developed countries, structural measures are subsidized
by national governments and local governments and communities are
required to cover only partial costs. In New Zealand, this led to a
preponderance of structural measures despite planning legislation that
enabled non-structural measures. As a result, the potential for major
disasters was increased and development intensified in areas with
structural measures only to be seriously devastated by events that
exceeded the engineering design level (Ericksen, 1986). Reduction of
centralized subsidies in the mid-1980s and changes in legislation saw
greater responsibility for the costs of disaster risk management falling
on the communities affected and a move toward more integrated
disaster risk reduction processes within New Zealand (Ericksen et al.,
2000). Similar trends have been observed in relation to coastal protection,
where structural measures are often favored over non-structural
options (Titus et al., 2009; Titus, 2011).
Building codes closely align with engineering and architectural structural
approaches to disaster risk reduction (Petal et al., 2008; Kang et al.,
2009). This is accompanied by the elevation of buildings and ground
floor standards in the case of flooding (Aerts et al., 2009; Kang et al.,
2009). Though building code regulations exist, non-adoption, especially
in developing countries, is problematic (Spence, 2004). Damages to the
structure occur not only because of noncompliance with the codes, but
also through a lack of inspections, the ownership status of the structure,
and the political context and mechanisms of local governance (May and
Burby, 1998). Insurance arrangements can provide incentives to local
governments and households to implement building codes (Botzen et
al., 2009).
Short-term risk reduction strategies can actually produce greater
vulnerability to future events as shown in diverse contexts such as
ENSO-related impacts in parts of Latin America, induced development
below dams or levees in the United States, and flooding in the United
Kingdom (Bowden et al., 1981; Pulwarty et al., 2004; Berube and Katz,
2005; Penning-Rowsell et al., 2006). While locally based protection
works often enable areas to be productively used and will continue to
be needed for areas that are already densely settled, they are commonly
misperceived as providing complete protection, and actually increase
development – and thus vulnerability – in hazard-prone areas, resulting
in the so-called ‘levee effect’ (Tobin, 1995; Montz and Tobin, 2008). A
more general statement of this proposition is found in the safe development
paradox in which increased safety measures such as dams or levees
induce increased development, which in turn leads to increased exposure
and ultimately losses (Burby, 2006). The conflicting policy goals of rapid
recovery, safety, betterment, and equity and their relative strengths and
weaknesses largely reflect experience with large disasters in other
places and times. The actual decisions and rebuilding undertaken to
date clearly demonstrate the rush by government at all levels and the
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
306
residents themselves to rebuild the familiar or increase risks in new
locations through displacement (Kates et al., 2006). Similarly, in drought-
prone areas, provision of assured water supplies encourages the
development of intensive agricultural systems – and for that matter,
domestic water use habits – that are poorly suited to the inherent
variability of supply and will be even more so in areas projected to
become increasingly arid in a changing climate (Chapter 3).
5.3.3. Land Use and Ecosystem Protection
Changes in land use not only contribute to global climate change but
they are equally reflective of adaptation to the varying signals of
economic, policy, and environmental change (Lambin et al., 2001). Local
land use planning embedded in zoning, local comprehensive plans, and
retreat and relocation policies is a useful approach to disaster risk
management to keep people and property away from locations exposed
to risk (Burby, 1998). However, some countries and rural areas may not
have formal land use regulations that restrict development or settlement.
As land use management regulates the movement of people and
industries in hazard-prone zones, such policies face strong opposition
from development pressures, real estate interests accompanied by
property rights, and local resistance against land acquisition (Burby,
2000; Thomson, 2007). Buffer zones, setback lines in coastal zones, and
inundation zones based on flood and sea level rise projections can
result in controversies and lack of enforcement that bring temporary
resettlement, land speculation, and creation of new vulnerabilities
(Ingram et al., 2006; Jha et al., 2010). The government of Sri Lanka, for
example, created buffer zones after the Indian Ocean tsunami of 2004,
and relocated people to safer locations. Distance from people’s coastal
livelihoods and social disruptions led to the revision of buffers and new
resettlement (Ingram et al., 2006). In the United States, coastal retreat
measures are difficult to implement as coastal property carries high value
and wealthy property owners can exert political pressure to build along
the coast (Ruppert, 2008). Shorefront property owners and realtors
especially oppose setback regulations because they consider the
regulation to deter growth (NOAA, 2007).
Land use planning and the application of spatial hazard information is
another avenue for disaster risk management, especially hazard mitigation.
Berke and Beatley (1992) examined a range of hazard mitigation measures
and ranked them according to effectiveness and ease of enforcement.
The most effective measures include land acquisition, density reduction,
clustering of development, building codes for new construction, and
mandatory retrofit of existing structures. The high cost of land acquisition
programs can make them unattractive to small communities. There has
been limited systematic scientific characterization of the ways in which
different hazard agents vary in their threats and characteristics, thus
require different pre-impact interventions and post-impact responses by
households, businesses, and community hazard management organizations.
However, Burby et al. (1997) have found evidence for some communities
that previous occurrence of a disaster did not have a strong effect on
the number of hazard mitigation techniques subsequently employed.
Formal approaches to land use planning as a means of disaster risk
management are often less appropriate for many rural areas in developing
countries where traditional practices and land tenure systems operate.
Systems of land tenure often are very complex and flexible and can
contribute to vulnerability reduction. In the case of pastoralists in dryland
environments, sharing of land for grazing and of access to water are
important drought responses (Anderson et al., 2010). This is not always
the case and some land tenure systems marginalize certain groups and
increase their vulnerability (Clot and Carter, 2009; Robledo et al., 2011).
There are also restrictions on land use planning in regards to slums and
squatter settlements. Poverty and the lack of infrastructure and services
increase the vulnerability of urban poor to adverse impacts from disasters,
and national governments and international agencies have had little
success in reversing such trends. As a result, successful efforts to reduce
hazard exposure have been locally led, built upon successful initiatives,
and composed of informal measures rather than those imposed by
governments at the local level (Satterthwaite et al., 2007; Zimmermann
and Stössel, 2011).
Land acquisition is another means of protecting property and people by
relocating them away from hazardous areas (Olshansky and Kartez,
1998). Many jurisdictions have the power of eminent domain to purchase
property but this is rarely used as a form of disaster risk management
(Godschalk et al., 2000) or climate change adaptation. Voluntary
acquisition of land, for example, requires local authorities to purchase
exposed properties, which in turn enables households to obtain less
risky real estate elsewhere without suffering large economic losses in
the process (Handmer, 1987), but this is rarely used in developing
countries because of lack of resources and political support. Given the
rapid population growth in coastal areas and in flood plains in many
parts of the world, and the large number and high value of exposed
properties in coastal zones in developed countries such as the United
States and Australia, this buy out strategy is cost-prohibitive and thus
rarely used (Anning and Dominey-Howes, 2009). Similarly, voluntary
acquisition schemes for developing countries are equally fraught with
problems as people have strong ties to the land, and land is held
communally in places like the Pacific Islands where community identity
cannot be separated from the land to which its members belong
(Campbell, 2010b). Land use planning alone, therefore, may not be
successful as a singular strategy but when coupled with related policies
such as tax incentives or disincentives, insurance, and drainage and
sewage systems it could be effective (Yohe et al., 1995; Cheong, 2011b).
However, if sea level rise adversely affects local coastal areas some form
of relocation may become necessary in all exposed jurisdictions. In the
United States, some state and local governments have adopted rolling
easement policies, which allow construction in vulnerable areas subject
to the requirement that the structures will be removed if and when the
landward edge of a wetland or beach encroaches (Titus, 2011).
Ecosystem conservation offers long-term protection from climate
extremes. The mitigation of soil erosion, landslides, waves, and storm
surges are some of the ecosystem services that protect people and
infrastructure from extreme events and disasters (Sudmeier-Rieux et al.,
Chapter 5Managing the Risks from Climate Extremes at the Local Level
307
2006). The 2004 Indian Ocean tsunami attests to the utility of mangroves,
coral reefs, and sand dunes in alleviating the influx of large waves to the
shore (Das and Vincent, 2009). The use of dune management districts to
protect property along developed shorelines has achieved success in
many places along the US eastern shore and elsewhere (Nordstrom,
2000, 2008). Carbon sequestration is another benefit of ecosystem-
based adaptation based on sustainable watershed and community forest
management (McCall, 2010). While the extent of their protective
ecosystem functions is still debated (Gedan et al., 2011), the merits of
ecosystem services in general are proven, and development of quantified
models of the services is well under way (Barbier et al., 2008; Nelson et
al., 2009). These nonstructural measures are considered to be less
intrusive and more sustainable, and when integrated with engineering
responses provide mechanisms for adapting to disasters and climate
extremes (Galloway, 2007; Opperman et al., 2009; Cheong, 2011a).
5.3.4. Storage and Rationing of Resources
Communities may take a range of approaches to cope with disaster-
induced shortages of resources, including producing surpluses and
storing them. If the surpluses are not available, rationing of food may
occur. Many localities produce food surpluses that enable them to
manage during periods of seasonal or disaster-initiated disruptions to
their food supplies, although such practices were more prevalent in pre-
capitalist societies. In Pacific Island communities, for example, food
crops such as taro and breadfruit were often stored for periods up to
and exceeding a year by fermentation in leaf-lined pits. Yams could be
stored for several years in dry locations, and most communities maintained
famine foods such as wild yams, swamp taro, and sago, which were only
harvested during times of food shortage (Campbell, 2006). The provision
of disaster relief, among other factors, has seen these practices decline
(Campbell, 2010a). In Mali, women store part of their harvest as a hedge
against drought (Intercooperation, 2008). Stockpiling and prepositioning
of emergency response equipment, materials, foods, and pharmaceuticals
and medical equipment is also an important form of disaster preparedness
at the local level, especially for many indigenous communities.
Rationing may be seen as the initial response to food shortages at or
near the onset of a food crisis. However, in many cases rationing is
needed on a seasonal basis. Rationing at the local level is often self-
rationing instituted at the level of households – particularly poor ones
without the ability to accumulate wealth or surpluses. Often, rationing
initially occurs among women and children (Hyder et al., 2005;
Ramachandran, 2006). Most rationing takes place in response to
food shortages and is, for most poor communities, the first response
to the disruption of livelihoods (Walker, 1989; Barrett, 2002; Devereux
and Sabates-Wheeler, 2004; Baro and Deubel, 2006). In many cases
increases in food prices force those with insufficient incomes to ration
as well.
When the food shortage becomes too severe, households may reduce
future security by eating seeds or selling livestock, followed by severe
illness, migration, starvation, and death if the shortages persist. While
climate change may alter the frequency and severity of droughts (see
Section 3.5.1), the causes of food crises are multi-faceted and often
lie in social, economic, and political processes in addition to climatic
variability (Sen, 1981; Corbett, 1988; Bohle et al., 1994; Wisner et al., 2004).
Food rationing is unusual in developed countries where most communities
are not based on subsistence production. Welfare systems and NGO
agencies respond to needs of those with livelihood deficits in these
countries. However, other forms of rationing do exist particularly in
response to drought events. Reductions in water use are achieved
through a number of measures including: metering, rationing (fixed
amounts, proportional reductions, or voluntary reductions), pressure
reduction, leakage reduction, conservation devices, education, plumbing
codes, market mechanisms (e.g., transferable quotas, tariffs, pricing),
and water use restrictions (Lund and Reed, 1995; Froukh, 2001).
Electricity supplies may also be disrupted by disaster events resulting in
partial or total blackouts. While a number of countries have national
electrical grids, decisions on responses to shortages are often made at
local levels causing considerable disruption to other services, domestic
customers, and to businesses. Rose et al. (2007) show that many
American businesses can be quite resilient in such circumstances,
adapting a variety of strategies including conserving energy, using
alternative forms of energy, using alternative forms of generation,
rescheduling activities to a future date, or focusing on the low- or no-
energy elements of the business operation. Rose and Liao (2005) had
similar findings for water supply disruption. Electricity storage (in
advance) and rationing may also be required when low precipitation
reduces hydroelectricity production, a possible scenario in some places
under climate projections (Vörösmarty et al., 2000; Boyd and Ibarrarán,
2009). In some cases there may be competition among a range of sectors
including industry, agriculture, electricity production, and domestic
water supply (Vörösmarty et al., 2000) that may have to be addressed
through rationing and other measures such as those listed above.
Clear rules outlining which consumers have priority in using water or
electricity is important.
Other elements that may be rationed as a result of natural hazards or
disasters include prioritization of medical and health services where
disasters may simultaneously cause a large spike in numbers requiring
medical assistance and a reduction in medical facilities, equipment,
pharmaceuticals, and personnel. This may require classifying patients
and giving precedence to those with the greatest need and the highest
likelihood of a positive outcome. This approach seeks to achieve the
best results for the largest number of people (Alexander, 2002; Iserson
and Moskop, 2007).
Responding to future disaster risk will entail multiple approaches at the
local level. Starting with risk communication and warning information,
the following dominate the range of adjustments local areas presently
undertake in responding to future risks: structural measures, land use
planning, ecosystem protection, and storage and rationing of resources.
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
308
5.4. Building Capacity at the Local Level for
Risk Management in a Changing Climate
Local risk management has traditionally dealt with extreme events
without considering the climate change context. This section provides
examples of adaptations to disaster risk and how such proactive behaviors
at the community level by local government and NGOs provide guidance
for reducing the longer-term impacts of climate change. Although
reacting to extreme events and their impacts is important, it is crucial to
focus on building the resilience of communities, cities, and sectors in
order to ameliorate the impacts of extreme events now and into the
future.
5.4.1. Proactive Behaviors and Protective Actions
Researchers have identified some of the physical and social characteristics
that allow for the adoption of effective partnerships and implementation
practices during events (Birkland, 1997; Pulwarty and Melis, 2001). These
include the occurrence of previous strong focusing events (such as
catastrophic extreme events) that generate significant public interest
and the personal attention of key leaders, a social basis for cooperation
including close inter-jurisdictional partnerships, and the existence of a
supported collaborative framework between research and management.
Factors conditioning this outcome have been summed up by White et al.
(2001) as “knowing better and losing even more.” In this context,
knowing better indicates the accumulation of readily available knowledge
on drivers of impacts and effective risk management practices. For
instance, researchers had understood the consequences of a major
hurricane hitting New Orleans prior to Hurricane Katrina with fairly
detailed understanding of planning and response needs. This knowledge
appears to have been ignored at all levels of government including the
local level after Hurricane Katrina (Kates et al., 2006). White et al. (2001)
offer four explanations for why such conditions exist from an information
standpoint: 1) knowledge continues to be flawed by areas of ignorance;
2) knowledge is available but not used effectively; 3) knowledge is used
effectively but takes a long time to have an impact; and 4) knowledge
is used effectively in some respects but is overwhelmed by increases in
vulnerability and in population, wealth, and poverty. Another possibility
is that some individuals or communities choose to take the risk. For
example, there is some evidence that the value of living near the coast,
especially in developed countries, pays back the cost of the structure in
a few years due to increases in housing values (Kunreuther et al., 2009),
so the risk is worth taking by the individual and the community. Finally,
knowledge is often discounted.
Individuals can make choices to reduce their risk but social relations,
context, and certain structural features of the society in which they live
and work mediate these choices and their effects. The recognition that
dealing with risk and insecurity is a central part of how poor people
develop their livelihood strategies is giving rise to prioritizing disaster
mitigation and preparedness as important components of many poverty
alleviation agendas (Cuny, 1983; Olshansky and Kartez, 1998; UNISDR,
2009). A number of long-standing challenges remain as the larger and
looser coalitions of interests that sometimes emerge after disasters rarely
last long enough to sustain the kind of efforts needed to reduce present
disaster risk, let alone climate extremes in a climate change context.
At the household and community level, individuals often engage in
protective actions to minimize the impact of extreme events on
themselves, their families, and their friends and neighbors. In some
cases individuals ignore the warning messages and choose to stay in
places of risk. The range and choice of actions are often event specific
and time dependent, but they are also constrained by location, adequate
infrastructure, socioeconomic characteristics, access to disaster risk
information, and risk perception (Tierney et al., 2001). For example,
evacuation is used when there is sufficient warning to temporarily
relocate out of harm’s way such as for tropical storms, flooding, and
wildfires. Collective evacuations are not always possible given the
location, population size, transportation networks, and the rapid onset of
the event. At the same time, individual evacuation may be constrained
by a host of factors ranging from access to transportation, monetary
resources, health impairment, job responsibilities, gender, and the
reluctance to leave home. There is a consistent body of literature on
hurricane evacuations in the United States, for example, that finds:
1) individuals tend to evacuate as family units, but they often use more
than one private vehicle to do so; 2) social influences (neighbors, family,
friends) are key to individual and households evacuation decisionmaking;
if neighbors are leaving then the individual is more inclined to evacuate
and vice versa; 3) risk perception, especially the personalization of risk
by individuals, is a more significant factor in prompting evacuation than
prior adverse experience with hurricanes; 4) pets and concerns about
property safety reduce household willingness to evacuate; and 5) social
and demographic factors (age, presence of children, elderly, or pets in
households, gender, income, disability, and race or ethnicity) either
constrain or motivate evacuation depending on the particular context
(Perry and Lindell, 1991; Dow and Cutter, 1998, 2000, 2002; Whitehead
et al., 2000; Bateman and Edwards, 2002; Van Willigen et al., 2002;
Sorensen et al., 2004; Lindell et al., 2005; Dash and Gladwin, 2007;
McGuire et al., 2007; Sorensen and Sorensen, 2007; Edmonds and
Cutter, 2008; Adeola, 2009). Culture also plays an important role in
evacuation decisionmaking (Clot and Carter, 2009). For example, recent
studies in Bangladesh have shown that there are high rates of non-
evacuation despite improvements in warning systems and the construction
of shelters. While there are a variety of reasons for this, gender issues
(e.g., shelters were dominated by males, shelters didn’t have separate
spaces for males and females) have a major influence upon females not
evacuating (Paul and Dutt, 2010; Paul et al., 2010).
A different protective action, shelter-in-place, occurs when there is little
time to act in response to an extreme event or when leaving the
community would place individuals more at risk (Sorensen et al., 2004).
Seeking higher ground or moving to higher floors in residential structures
to get out of rising waters is one example. Another is the movement into
interior spaces within buildings to seek refuge from strong winds. In the
Chapter 5Managing the Risks from Climate Extremes at the Local Level
309
case of wildfires, shelter-in-place becomes a back-up strategy when
evacuation routes are restricted because of the fire and include protecting
the structure with garden hoses or finding a safe area such as a water
body (lake or backyard swimming pool) as temporary shelter (Cova et
al., 2009). In Australia, the shelter-in-place action is slightly different.
Here the local community engagement with wildfire risks has two options:
stay and defend or leave early. In this context, the decisions to remain
are based on social networks, prior experience with wildfires, gender
(males will remain to protect and guard property), and involvement
with the local fire brigade (McGee and Russell, 2003). The study also
found that rural residents were more self-reliant and prepared to defend
then suburban residents (McGee and Russell, 2003).
The social organization of societies dictates the flexibility in the choice
of protective actions – some are engaged in voluntarily (such as in the
United States, Australia, and Europe), while other protective actions for
individuals or households are coordinated by centralized authorities
such as Cuba and China. Planning for disasters is a way of life for Cuba,
where everyone is taught at an early age to mobilize quickly in the case
of a natural disaster (Sims and Vogelmann, 2002; Bermejo, 2006). The
organization of civil defense committees at block, neighborhood, and
community levels working in conjunction with centralized governmental
authority makes the Cuban experience unique (Sims and Vogelmann, 2002;
Bermejo, 2006). Recent experience with hurricanes affecting Cuba suggests
that such efforts are successful because there has been little loss of life.
In many traditional or pre-capitalist societies it appears that mechanisms
existed that protected community members from periodic shocks such as
natural hazards. These mechanisms, sometimes referred to as the moral
economy, were underpinned by reciprocity and limiting exploitation so
that everyone had basic security. The mechanisms are often linked to
kinship networks, and serve to redistribute resources to reduce the impacts
on those who had sustained severe losses, and have been identified in
Southeast Asia (Scott, 1976), Western Africa (Watts, 1983), and the
Pacific Islands (Paulson, 1993). The moral economy incorporated social,
cultural, political, and religious arrangements, which ensured that all
community members had a minimal level of subsistence (see Box 5-4).
For example, traditional political systems in the semiarid Limpopo Basin
in northern South Africa enabled chiefs to reallocate surpluses during
bad years, but this practice has declined under contemporary systems
where surpluses are sold (Dube and Sekhwela, 2008). In Northern
Kenya, social security networks existed among some groups of nomadic
pastoralists that enabled food and livestock stock to be redistributed
following drought events, but these are also breaking down with the
monetization of the local economy, among other factors (Oba, 2001).
Although the concept of moral economy is generally associated with
pre-capitalist societies and those in transition to capitalism (in the past),
significant features of moral economy, such as reciprocity, barter, crop
sharing, and other forms of cooperation among families and communities
or community-based management of agricultural lands, waters, or woods
are still part of the social reality of developing countries that cannot be
considered anymore as pre-capitalist. Many studies show that moral
economy-based social relationships are still present such as traditional
institutions regulating access, use, and on-going redistribution of
community-owned land (Sundar and Jeffery, 1999; Rist, 2000; Hughes,
2001; Trawick, 2001; Rist et al., 2003). The revitalization, enhancement,
and innovation of such moral economy-based knowledge, technologies,
and forms of cooperation and interfamily organization represent an
important and still existing source of fostering collective action that
serves as an enabling condition for preventing and coping with hazards
related to natural resource management. While aspects of the traditional
moral economy have declined in many societies, informal networks
remain important in disaster risk reduction (see Section 5.4.3).
The notion of the moral economy does not recognize the inequalities in
some of the social systems that enabled such practices to be sustained
(e.g., gender-based power relationships) and tended to perhaps provide
an unrealistic notion of a less risky past. In addition, kinship-based
sharing networks may foster freeloading among some members (diFalco
and Bulte, 2009). Nevertheless, a reduction in traditional coping
mechanisms including the moral economy is reflected in growing
disaster losses and increasing dependency on relief (Campbell, 2006).
Collective action to prepare for or respond to disaster risk and extreme
climate impacts can also be driven by localized organizations and social
movements. Many such groups represent networks or first responders
for climate-sensitive disasters. However, there are many constraints that
these movements face in building effective coalitions including the need
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
Box 5-4 | Collective Behavior and the
Moral Economy at Work
A variety of socio-political networks that were used to offset
disaster losses existed throughout the Pacific region prior to
colonization (Sahlins, 1962; Paulson, 1993). One example of
such a system is the Suqe, or graded society, which existed in
northern Vanuatu, a small island nation in the South West
Pacific Ocean. In the Suqe ‘big men’ achieved the highest
status by accumulating surpluses of valued goods such as shell
money, specially woven mats, and pigs. Men increased their
grade within the system by making payments of these goods
to men of higher rank. In accumulating the items men would
also accumulate obligations to those they had borrowed from.
Accordingly, networks and alliances emerged among the
islands of northern Vanuatu. When tropical cyclones destroyed
crops, the obligations could be called in and assistance given
from members of the networks who lived in islands that
escaped damage (Campbell, 1990). A number of processes
associated with colonialism (changes to the socio-political
order), the introduction of the cash economy (the replacement
of shell money), and religious conversion that resulted in the
banning of the Suqe), as well as the provision of post-disaster
relief, have caused a number of elements of the moral economy
to fall into disuse (Campbell, 2006).
310
to connect with other movement organizations and frame the problem in
an accessible way (McCormick, 2010). One means of mobilizing collective
responses at the local level is through participatory approaches to
disaster risk reduction such as Community-Based Disaster Reduction or
Community-Based Disaster Preparedness (see Section 5.6.2). Such
approaches build on local needs and priorities, knowledge, and social
structures and are increasingly being used in relation to climate change
adaptation (Reid et al., 2009).
5.4.2. Empowerment for Local Decisionmaking
A critical factor in community-based disaster risk reduction is that
community members are empowered to take control of the processes
involved. Marginalization (Mustafa, 1998; Adger and Kelly, 1999; Polack,
2008) and disempowerment (Hewitt, 1997) are critical factors in creating
vulnerability and efforts to reduce these characteristics play an important
role in building resilient communities. In this chapter, empowerment
refers to giving community members control over their lives with support
from outside (Sagala et al., 2009). This requires external facilitators to
respect community structures and traditional and local knowledge
systems, to assist but not take a dominating role, to share knowledge,
and to learn from community members (Petal et al., 2008). A key element
in empowering communities is building trust between the community
and the external facilitators (Sagala et al., 2009). In the Philippines, for
example, Allen (2006) found that many aspects of community disaster
preparedness such as building on local institutions and structures, building
local capacity to act independently, and building confidence through
achieving project outcomes were already present. She also found that
where agencies focused on the physical hazard as the cause of disasters
and neglected the underlying causes of the social vulnerability within
these small specific projects, the result was disempowerment. It is also
important to note that communities have choices from a range of
disaster management options (Mercer et al., 2008). Empowerment in
community-based disaster risk management may also be applied to
groups within communities whose voice may otherwise not be heard or
who are in greater positions of vulnerability (Wisner et al., 2004). These
include women (Wiest et al., 1994; Bari, 1998; Clifton and Gell, 2001;
Polack, 2008) and disabled people (Wisner, 2002).
Another key element of empowerment is ownership of or responsibility
for disaster management (Buvinić et al., 1999). This applies to all
aspects of the disaster, from the ownership of a disaster itself so that
the community has control of relief and reconstruction, to a local
project to improve preparedness. Empowerment and ownership ensure
that local needs are met, that community cohesion is sustained, and
a greater chance of success of the disaster management process.
Empowerment and ownership of the disaster impacts may be
particularly important in achieving useful (for the locality) post-disaster
assessments (Pelling, 2007). It is important for external actors to
identify those voices who speak for the local constituencies. Also,
accountability and governance of disaster and climate management
issues is growing in importance.
5.4.3. Social Drivers
Similar to empowerment is the role of localized social norms, social
capital, and social networks as these also shape behaviors and actions
before, during, and after extreme events. Each of these factors operates
on their own and in some cases also intersects with the others. As
vulnerability to disasters and climate change is socially constructed
(Sections 2.4 and 2.5.2), the breakdown of collective action often leads
to increased vulnerability. Norms regarding gender also play a role in
determining outcomes. For example, women were more prone to
drowning than men during the Asian tsunami because they were less
able to swim and because they were attempting to save their children
(Rofi et al., 2006; Section 2.5.8).
Social norms are rules and patterns of behavior that reflect expectations
of a particular social group (Horne, 2001). Norms structure many different
kinds of action regarding climate change (Pettenger, 2007). Norms are
embedded in formal institutional responses, as well as informal groups
that encounter disasters (Raschky, 2008). Norms of reciprocity, trust,
and associations that bridge social divisions are a central part of social
cohesion that fosters community capacity (Kawachi and Berkman, 2000).
A number of types of groups drive norms and, consequently, vulnerability,
including religious, neighborhood, cultural, and familial groups (Brenkert
and Malone, 2005). In the occurrence of extreme events, affected
groups interact with one another in an attempt to develop a set of
norms appropriate to the situation, otherwise known as the emergent
norm theory of collective behavior (NRC, 2006). This is true of those first
affected at the local level whose norms and related social capital affect
capacity for response (Dolan and Walker, 2004).
Social capital is a multifaceted concept that captures a variety of social
engagements within the community that bonds people and generates a
positive collective value. It is also an important element in disaster
preparedness, coping, and response. It is an important element in the
face of climate extremes because community social resources such as
networks, social obligations, trust, and shared expectations create social
capital to prevent, prepare, and cope with disasters (Dynes, 2006). In
climate change adaptation, scholars and policymakers increasingly
promote social capital as a long-term adaptation strategy (Adger, 2003;
Pelling and High, 2005). Although often positive, social capital can have
some negative outcomes. Internal social networks are oftentimes self-
referential and insular (Portes and Landolt, 1996; Dale and Newman, 2010).
This results in a closed society that lacks innovation and diversity, which
are essential elements for climate change adaptation. Disaster itself is
overwhelming, and can lead to the erosion of social capital and the
demise of the community (Ritchie and Gill, 2007). This invites external
engagement beyond local-level treatment of the disaster and extreme
events (Brondizio et al., 2009; Cheong, 2010). The inflow of external aids,
expertise, and the emergence of new groups to cope with disaster are
indicative of bridging and linking social capital beyond local boundaries.
Social capital is embedded in social networks (Lin, 2001), or the social
structure composed of individuals and organizations, through multiple
Chapter 5Managing the Risks from Climate Extremes at the Local Level
311
types of dependency, such as kinship, financial exchange, or prestige
(Wellman and Berkowitz, 1988). Social networks provide a diversity of
functions, such as facilitating sharing of expertise and resources across
stakeholders (Crabbé and Robin, 2006). Networks can function to promote
messages within communities through preventive advocacy, or the
engagement of advocates in promoting preventive behavior (Weibel,
1988). Information about health risks has often been effectively distributed
through a social network structure using opinion leaders as a guide
(Valente and Davis, 1999; Valente et al., 2003), and has promising
application for changing behavior regarding climate adaptation
(Maibach et al., 2008). Such opinion leaders may span a range of types,
from formally elected officials, celebrities, and well-known leaders, to
local community members who are well-embedded in local social
networks. It is important to note that more potential has been shown in
influencing behavior through community-level interventions than
through individual-level directives at the population level (Kawachi and
Berkman, 2000). Local and international networks can support the
development of policies and practices that result in greater preparedness
(Tompkins, 2005). Local resilience in the face of climate change can be
fostered by strong social networks that support effective responses
(Ford et al., 2006). For example, networks facilitate the transmission of
information about risks (Berkes and Jolly, 2001). Therefore, communities
with stronger social networks appear to be better prepared for extreme
climate impacts because of access to information and social support
(Buckland and Rahman, 1999).
At the same time, it is important to note that social networks may not
always be sufficient to foster effective adaptation to extreme events.
Some social networks actually discourage people from moving away
from high risk zones, such as has been the case in storms and floods
when residents have not wanted to leave (Eisenman et al., 2007). The
impacts of climate change itself may also change the structure and utility
of social networks. As people migrate away from climate and other risks
or are pulled toward alternative locations for social or ecological
resources, those left behind can experience fragmented or weakened
social networks. The utilization of social networks can also be prevented
by the status of particular social groups, such as illegal and legal settlers
or immigrants (Wisner et al., 2004). Other social and environmental
contextual factors must be considered when conceptualizing the role of
social networks in managing extreme events. For example, strong social
networks have facilitated adaptability in Inuit communities, but are being
undermined by the dissolution of traditional ways of life (Ford et al., 2006).
5.4.4. Integrating Local Knowledge
Local and traditional knowledge is increasingly valued as important
information to include when preparing for disasters (McAdoo et al.,
2009; R. Shaw et al., 2009). It is embedded in local culture and social
interactions and transmitted orally over generations (Berkes, 2008). Place-
based memory of vulnerable areas, know-how for responding to recurrent
extreme events, and detection of abnormal environmental conditions
manifest the power of local knowledge. Because local knowledge is
often tacit and invisible to outsiders, community participation in disaster
management is essential to tap this information as it can offer alternative
perspectives and approaches to problem-solving (Battista and Baas,
2004; Turner and Clifton, 2009).
Within a climate change context, indigenous people as well as long-term
residents often conserved their resources in situ, providing important
information about changing environmental conditions as well as actively
adapting to the changes (Salick and Byg, 2007; Macchi et al., 2008;
Salick and Ross, 2009; Turner and Clifton, 2009). Research is emerging
that helps to document changes that local people are experiencing
(Ensor and Berger, 2009; Salick and Ross, 2009). Although this evidence
might be similar to scientific observations from external researchers, the
fact that local communities are observing it is initiating discussions
about existing and potential adaptation to climate changes from within
the community.
The following example is illustrative. In six villages in eastern Tibet, near
Mt. Khawa Karpo, local documentation of warmer temperatures, less
snow, and glacial retreat across areas were consistent, whereas other
observations were more varied, including those for river levels and
landslide incidences (Byg and Salick, 2009). In Gitga’at (Coast Tsimshian)
Nation of Hartley Bay, British Columbia, indigenous people observe the
decline of some species but also new appearances of others, anomalies
in weather patterns, and declining health of forests and grasslands that
have affected their ability to harvest food (Turner and Clifton, 2009). The
Alaska Native Tribal Health Consortium generated climate change and
health impact assessment reports from observations, data, and traditional
ecological knowledge (ANTHC, 2011). Other than knowledge from
indigenous groups, local knowledge associated with contemporary
societies and cities exists though more research is needed in this area
(Hordijk and Baud, 2011).
Integration of local knowledge with external scientific, global, and
technical knowledge is an important dimension of climate change
adaptation and disaster management. Experiences in environmental
management and integrated assessment suggest mechanisms for such
knowledge transfers from the bottom up and from the top down (Burton
et al., 2007; Prabhakar et al., 2009). For example, communities set up
trusted intermediaries to transfer and communicate external knowledge
such as technology-based early warning systems and innovative and
sustainable farming techniques that incorporate the local knowledge
system (Bamdad, 2005; Kristjanson et al., 2009). Another example is the
re-engineering of local practices to adapt to climate change as shown
in the conversion of traditional dry-climate adobe construction to more
stabilized earth construction built to withstand regular rainfall. The
utilization of participatory methods to draw in the perspectives of local
stakeholders for subsequent input into hazards vulnerability assessments
or climate change modeling or scenario development is well documented.
Stakeholder interactions and related workshops using participatory or
mediated modeling elicit discussions of model assumptions, local impacts,
consistencies of observed and modeled patterns, and adaptation
strategies (Cabrera et al., 2008; Langsdale et al., 2009).
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
312
Obstacles to utilizing local knowledge as part of adaptation strategies
exist. Climate-induced biodiversity change threatens historical coping
strategies of indigenous people as they depend on the variety of wild
plants, crops, and their environments particularly in times of disaster
(Turner and Clifton, 2009). In dryland areas such as in Namibia and
Botswana, one of the indigenous strategies best adapted to frequent
droughts is livestock herding, including nomadic pastoralism (Ericksen
et al., 2008). Decreased access to water sources through fencing and
privatization has inhibited this robust strategy. Also in Botswana, it has
been suggested that government policies have weakened traditional
institutions and practices, as they have not adequately engaged with local
community institutions and therefore the mechanisms for redistributing
resources have not been strengthened sufficiently (Dube and Sekhwela,
2008).
5.4.5. Local Government and Nongovernment
Initiatives and Practices
Governance structures are pivotal to addressing disaster risk and
informing responses as they help shape efficiency, effectiveness, equity,
and legitimacy (Adger et al., 2003; UNISDR, 2009). Some places centralize
climate change management practices at the national level (see
Chapter 6). This may be due to the ways in which many climate
extremes affect environmental systems that cross political boundaries
resulting in discordance if solely locally managed (Cash and Moser,
2000) but could also be based on old practices of operations. In other
places, actions are more decentralized, emerging at the local level and
tailored to local contexts (Bizikova et al., 2008). If multiple levels of
planning are to be implemented, mechanisms for facilitation and guidance
at the local level are needed in order that fairness is guaranteed during
the implementation of national policies at the local scale (Thomas and
Twyman, 2005). Local governments play an important role as they are
responsible for providing infrastructure, preparing and responding to
disasters, developing and enforcing planning, and connecting national
government programs with local communities (Huq et al., 2007; UNISDR,
2009). The quality and provision of these services have an impact on
disaster and climate risk (Tanner et al., 2009). Effective localized
planning, for example, can minimize both the causes and consequences
of climate change (Bulkeley, 2006).
Though local government-led climate adaptation policies and initiatives
are less pronounced than climate change mitigation measures, a growing
number of cities and sub-national entities are developing adaptation
plans, but few have implemented their strategies (Heinrichs et al., 2009;
Birkmann et al., 2010). The Greater London Authority (2010), for example,
has prepared a Public Consultation Draft of their climate change
adaptation strategy for London. The focus of this is on the changing risk
of flood, drought, and heat waves through the century and actions for
managing them. Some of the actions include improvement in managing
surface water flood risk, an urban greening program to buffer the
impacts from floods and hot weather, and retrofitting homes to improve
water and energy efficiency. ICLEI, a non-profit network of more than
1,200 local government members across the globe, provides web-based
information (www.iclei.org) in support of local sustainability efforts
using customized tools and case studies on assessing climate resilience
and climate change adaptation.
Some assessments of urban adaptations exist. For example, adaptation
efforts in eight cities (Bogotá, Cape Town, Delhi, Pearl River Delta, Pune,
Santiago, Sao Paulo, and Singapore) tend to support existing disaster
management strategies (Heinrichs et al., 2009). Another study comparing
both formal adaptation plans and less formal adaptation studies in nine
cities including Boston, Cape Town, Halifax, Ho Chi Minh City, London,
New York, Rotterdam, Singapore, and Toronto demonstrates that the
focus is mostly on risk reduction and the protection of citizens and
infrastructure, with Rotterdam seeing adaptation as opportunity for
transformation (Birkmann et al., 2010). These nine cities have focused
more on expected biophysical impacts than on socioeconomic impacts
and have not had a strong focus on vulnerability and the associated
susceptibility or coping capacity. Despite the intention that city adaptation
responses aim at an integrated approach, they tend to have sectoral
responses, with limited integration of local voices. There is a good
understanding of the impacts, but the implementation of policy and
outcomes on the ground are harder to see (Bulkeley, 2006; Burch and
Robinson, 2007).
In these adaptation strategies, the size of the local government is
important, and it varies depending on the population and location.
Primate and large cities exert more independence, whereas smaller
municipalities depend more on higher levels of the government units,
and often form associations to pool their resources (Lundqvist and
Borgstede, 2008). In the latter case, state-mandated programs and state-
generated grants are the main incentives to formulate mitigation policies
(Aall et al., 2007) and can be applicable to adaptation policies. Lack of
resources and capabilities has led to outsourcing of local adaptation
plans, and can generate insensitive and unrefined local solutions and
more reliance on technological fixes (Crabbé and Robin, 2006).
The history and process of decentralization are significant in the capacity
of the local government to formulate and implement adaptation policies.
Aligning local climate adaptation policies with the state/provincial and
national/federal units is a significant challenge for local governments
(Roberts, 2008; van Aalst et al., 2008). The case of decentralization in
climate change adaptation is relatively new, and we can draw some
lessons from decentralized natural resource management and crisis
management. One of the problems of decentralization has been the
complexity and uniqueness of each locality that policy planners often
failed to take into account because of the lack of understanding and
consultation with the local community, and this could result in
recentralizing the entire process in some instances (Ribot et al., 2006;
Geiser and Rist, 2009). Some remedies include working with local
institutions, ensuring appropriate transfer of various rights and access,
and providing sufficient time for the process (Ribot, 2003). The crisis
management literature also points out that there has been a lack of
coordination and integration between central and local governments
Chapter 5Managing the Risks from Climate Extremes at the Local Level
313
(Waugh and Streib, 2006; Schneider, 2008). Moynihan (2009) suggests a
networked collaboration as a solution and posits that even a hierarchical
disaster management structure such as the incident command system in
the United States operates on the network principles of negotiation, trust,
and reciprocity.
Although government actors play a key role, it is evident that partnerships
between public, civic, and private actors are crucial in addressing climate
hazards-related adaptation (Agrawal, 2010). While international agencies,
the private sector, and NGOs play a norm-setting agenda at provincial,
state, and national levels, community-based organizations (CBOs) often
have greater capacity to mobilize at the local scale (Milbert, 2006). NGO
and CBO networks play a critical role in capturing the realities of local
livelihoods, facilitating sharing information, and identifying the role of
local institutions that lead to strengthened local capacity (Bull-
Kamanga et al., 2003). Strong city-wide initiatives are often based on
strategic alliances, and local community organizations are essential to
making city planning operational (Hasan, 2007). This can be seen in the
case of the New York City Panel on Climate Change that acted as a
scientific advisory group to both Mayor Bloomberg’s Office of Long-
term Planning and Sustainability and the New York City Climate Change
Adaptation Task Force, a stakeholder group of approximately 40 public
agencies and private sector organizations that manage the critical
infrastructure of the region (Rosenzweig et al., 2011). The Panel and
stakeholders separated functions between scientists (knowledge
provision) and stakeholders (planning and action) and communicated
climate change uncertainties with coordination by the Mayor’s office
(Rosenzweig et al., 2011).
Many nongovernment actors charged with managing climate risks use
community risk assessment tools to engage communities in risk reduction
efforts and influence planning at district and sub-national levels (van
Aalst, 2006; Twigg, 2007). NGO engagement in risk management
activities ranges from demonstration projects, training and awareness-
raising, legal assistance, alliance building, small-scale infrastructure,
and socioeconomic projects, to mainstreaming and advocacy work (Luna,
2001; Shaw, 2006). Bridging citizen-government gaps is a recognized
role of civil society organizations and NGOs often act as social catalysts
or social capital, an essential for risk management in cities (Wisner,
2003). Conversely, the potential benefits of social capital are not always
maximized due to mistrust, poor communications, or lack of functioning
either within municipalities or nongovernment agencies. This has major
implications for risk reduction (Wisner, 2003) and participation of the
most vulnerable in nongovernment initiatives at municipal or sub-
national level is not guaranteed (Tanner et al., 2009).
This section highlighted mechanisms for building capacity for local
adaptation to climate extremes ranging from empowerment for
decisionmaking to utilization of social networks. A balanced portfolio of
approaches that capture local knowledge, proactive behaviors, and
governmental and nongovernmental initiatives and practices will prove
most successful in managing the risk of climate extremes at the local
level.
5.5. Challenges and Opportunities
As illustrated earlier in the chapter, disaster risk management actions
increase the coping capacity of local places to disasters in the short
term and benefit a community’s resilience in the long term. Differences
in coping, risk management, and adaptation along with the costs of
managing disaster risk at the local level present challenges and
opportunities for adaptation to climate extremes. They not only
influence human security, but the scale and context of the differences
highlight opportunities for proactive actions for risk reduction and
climate change adaptation, and also identify constraints to such
actions.
5.5.1. Differences in Coping and Risk Management
There are significant differences among localities and population
groups in the ability to prepare for, respond to, recover from, and adapt
to disasters and climate extremes. During the last century, social science
researchers have examined those factors that influence coping responses
by households and local entities through post-disaster field investigations
as well as pre-disaster assessments (Mileti, 1999; NRC, 2006). Among
the most significant individual characteristics are gender, age, wealth,
ethnicity, livelihoods, entitlements, health, and settlements. However, it
is not only these characteristics operating individually, but also their
synergistic effects that give rise to variability in coping and managing
risks at the local level.
5.5.1.1. Gender, Age, and Wealth
The literature suggests that at the local level gender makes a difference
in vulnerability (Section 2.4) and in the differential mortality from
disasters (Neumayer and Plümper, 2007). The evidence is robust with
high agreement. In disasters, women tend to have different coping
strategies and constraints on actions than men (Fothergill, 1996;
Morrow and Enarson, 1996; Peacock et al., 1997). These are due to
socialized gender factors such as social position (class), marital status,
education, wealth, and caregiver roles, as well as physical differences in
stature and endurance. At the local level for example, women’s lack of
mobility, access to resources, lack of power and legal protection, and
social isolation found in many places across the globe tend to augment
disaster risk, and vulnerability (Schroeder, 1987; IFRC, 1991; Mutton and
Haque, 2004; UNIFEM, 2011). Relief and recovery operations are often
insensitive to gender issues (Hamilton and Halvorson, 2007), and so the
provision of such supplies and services also influences the differential
capacities to cope (Enarson, 2000; Ariyabandu, 2006; Wachtendorf et
al., 2006; Fulu, 2007), especially at the local level. However, the active
participation of women has been shown to increase the effectiveness
of prevention, disaster relief, recovery, and reconstruction, thereby
improving disaster management (Enarson and Morrow, 1997, 1998;
Fothergill, 1999, 2004; Hamilton and Halvorson, 2007; Enarson, 2010;
see Box 5-5).
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
314
Age acts as an important factor in coping with disaster risk (Cherry, 2009).
Older people are more prone to ill health, isolation, disabilities, and
immobility (Dershem and Gzirishvili, 1999; Ngo, 2001), which negatively
influence their coping capacities in response to extreme events (see
Case Study 9.2.1). In North America, for example, retired people often
choose to live in hazardous locations such as Florida or Baja California
because of warmer weather and lifestyles, which in turn increases their
potential exposure to climate-sensitive hazards. Often because of hearing
loss, mental capabilities, or mobility, older persons are less able to receive
warning messages or take protective actions, and are more reluctant to
evacuate (O’Brien and Mileti, 1992; Hewitt, 1997). However, older people
have more experience and wisdom with accumulated know-how on
specific disasters/extreme events as well as the enhanced ability to
transfer their coping strategies arising from life experiences.
Children have their own knowledge of hazards, hazardous places, and
vulnerability that is often different than adults (Plush, 2009; Gaillard
and Pangilinan, 2010). Research has shown significant diminishment of
coping skills (and increases in post-traumatic stress disorder and other
psychosocial effects) among younger children following Hurricane
Katrina (Barrett et al., 2008; Weems and Overstreet, 2008). In addition
to physical impacts and safety (Lauten and Lietz, 2008; Weissbecker et
al., 2008), research also suggests that emotional distress caused by
fear of separation from the family, and increased workloads following
disasters affects coping responses of children (Babugura, 2008; Ensor,
2008). However, the research also suggests that children are quite
resilient and can adapt to environmental changes thereby enhancing
the adaptive capacity of households and communities (Bartlett, 2008;
Manyena et al., 2008; Mitchell et al., 2008; Pfefferbaum et al., 2008;
Ronan et al., 2008; Williams et al., 2008).
Wealth, especially at the local level affects the ability of a households
or localities to prepare for, respond to, and rebound from disaster events
(Cutter et al., 2003; Masozera et al., 2007). Wealthier places have a
greater potential for large monetary losses, but at the same time, they
have the resources (insurance, income, political cache) to cope with the
impacts and recover from extreme events, and they are less socially
vulnerable. In Asia, for example, wealth shifted construction practices
from wood to masonry, which made many of the cities more vulnerable
and less able to cope with disaster risk (Bankoff, 2007), especially in
seismic regions. Poorer localities and populations often live in cheaper
hazard-prone locations, and face challenges not only in responding to
the event, but also recovering from it. Poverty also enhances disaster
risk (Carter et al., 2007). In some instances, it is neither the poor nor
the rich that face recovery challenges, but rather localities that are
in between, such as those not wealthy enough to cope with the
disaster risk on their own, but not poor enough to receive full federal or
international assistance.
In some localities, it is not just wealth or poverty that influence coping
strategies and disaster risk management, but rather the interaction
between wealth, power, and status, that through time and across space
has led to a complicated system of social stratification (Heinz Center,
2002). One of the best examples of this is the human experience with
Hurricane Katrina (see Box 5-6).
5.5.1.2. Livelihoods and Entitlements
Adaptive capacity is influenced to a large extent by the institutional
rules and behavioral norms that govern individual responses to hazards
(Dulal et al., 2010). It is also socially differentiated along the lines of age,
ethnicity, class, religion, and gender (Adger et al., 2007). Local institutions
regulate the access to adaptation resources: those that ensure equitable
opportunities for access to resources promote adaptive capacity within
communities and other local entities (Jones et al., 2010). Institutions, as
purveyors of the rules of the game (North, 1990), mediate the socially
differential command over livelihood assets, thus determining protection
or loss of entitlements.
Livelihood is the generic term for all the capabilities, assets, and activities
required for a means of living. Livelihood influences how families and
communities cope with and recover from stresses and shocks (Carney,
1998). Another definition of livelihoods gives more emphasis to access
to assets and activities that is influenced by social relations (gender,
class, kin, and belief systems) and institutions (Ellis, 2000). Understanding
Chapter 5Managing the Risks from Climate Extremes at the Local Level
Box 5-5 | The Role of Women in Proactive
Behavior
Women’s involvement in running shelters and processing food
was crucial to the recovery of families and communities after
Hurricane Mitch hit Honduras. One-third of the shelters were
run by women, and this figure rose to 42% in the capital. The
municipality of La Masica in Honduras, with a mostly rural
population of 24,336 people, stands out in the aftermath of
Mitch because, unlike other municipalities in the northern
Atlanta Department, it reported no mortality. Some attributed
this outcome to a process of community emergency
preparedness that began about six months prior to the
disaster. Gender lectures were given and, consequently, the
community decided that men and women should participate
equally in all hazard management activities. When Mitch
struck, the municipality was prepared and vacated the area
promptly, thus avoiding deaths. Women participated actively in
all relief operations. They went on rescue missions, rehabilitated
local infrastructure (such as schools), and along with men,
distributed food. They also took over from men who had
abandoned the task of continuous monitoring of the early
warning system. This case study illustrates the more general
finding that the active incorporation of women into disaster
preparedness and response activities helps to ensure success
in reducing the impacts of disasters (Buvinić et al., 1999;
Cupples, 2007; Enarson, 2009).
315
how natural resource-dependent people cope with climate change in
the context of wider livelihood influences is critical to formulating valid
adaptation frameworks.
Local people’s livelihoods and their access to and control of resources
can be affected by events largely beyond their control such as climatic
extremes (e.g., floods, droughts), conflict, or agricultural problems such
as pests and disease and economic shocks that can largely impact their
livelihoods (Chambers and Conway, 1992; Jones et al., 2010). For poor
communities living on fragile and degraded lands such as steep hillsides,
dry lands, and floodplains, climate extremes present additional threats
to their livelihoods that could be lost completely if exposed to repeated
disastrous events within short intervals leaving insufficient time for
recovery. Actions aiming at improving local adaptive capacity focus
more on addressing the deteriorating environmental conditions. A
central element in their adaptation strategies involves ecosystem
management and restoration activities such as watershed rehabilitation,
agroecology, and forest landscape restoration (Ellis, 2000; Ellis and
Allison, 2004; Osman-Elasha, 2006b). As some suggest (Spanger-Siegfried
et al., 2005) these types of interventions often protect and enhance
natural resources at the local scale and address immediate development
priorities, but can also improve local capacities to adapt to future
climate change. The buffering capacities of local people’s livelihoods
and their institutions are critical for their adaptation to extreme climate
stress. They often rest on the ability of communities to generate
potentials for self-organization and for social learning and innovations
(Adger et al., 2006).
A number of studies indicated that sustainable strategies for disaster
reduction help improve livelihoods (UNISDR, 2004), while social capital
and community networks support adaptation and disaster risk reduction
by diminishing the need for emergency relief in times of drought
and/or crop failure (Devereux and Coll-Black, 2007; see Section 5.2.1).
A research study in South Asia suggests that adaptive capacity and
livelihood resilience depend on social capital at the household level (i.e.,
education and other factors that enable individuals to function within a
wider economy), the presence or absence of local enabling institutions
(local cooperatives, banks, self-help groups), and the larger physical and
social infrastructure that enables goods, information, services, and people
to flow. Interventions to catalyze effective adaptation are important at
all these multiple levels (Moench and Dixit, 2004). Diversification within
and beyond agriculture is a widely recognized strategy for reducing risk
and increasing well-being in many developing countries (Ellis, 2000;
Ellis and Allison, 2004).
Entitlements are assets of the individuals and household. Assets are
broadly defined and include not only physical assets such as land, but
also human capital such as education and training. At the local scale
assets include institutional assets such as technical assistance or credit;
social capital such as mutual assistance networks; public assets such as
basic infrastructure like water and sanitation; and environmental assets
such as access to resources and ownership of them (Leach et al., 1999).
The link between disaster risk, access to resources, and adaptation has
been widely documented in the literature (Sen, 1981; Adger, 2000;
Brooks, 2003). Extreme climate events generally lead to entitlement
decline in terms of the rights and opportunities that local people have
to access and command the livelihood resources that enable them to
deal with and adapt to climate stress.
Assessment of livelihoods provides the explanation as to the differences
in responses based on the understanding of endowments, entitlements,
and capabilities, within the organizational structure and power relations
of individuals, households, communities, and other local entities (Scoones,
1998). Access to assets and entitlements is an important element in
improving the ability of localities to lessen their vulnerability and to
cope with and respond to disasters and environmental change. In some
instances, this may not be true. For example, if a disaster affects a
household asset, but the household is still paying off its debt regarding
the initial cost of the asset and assuming that the asset is not protected
or insured against hazards, the asset loss coupled with the need to pay off
the loan renders the household more vulnerable, not less (Twigg, 2001).
Entitlement protection thus requires adaptive types of institutions and
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
Box 5-6 | Race, Class, Age, and Gender: Hurricane Katrina Recovery and Reconstruction
The intersection of race, class, age, and gender influenced differential decisionmaking; the uneven distribution of vulnerability and
exposure; and variable access to post-event aid, recovery, and reconstruction in New Orleans before, during, and after Hurricane Katrina
(Elliott and Pais, 2006; Hartman and Squires, 2006; Tierney, 2006). Evacuation can protect people from injury and death, but there are
inequalities in who can evacuate and when, with the elderly, poor, and minority residents least able to leave without assistance (Cutter
and Smith, 2009). Extended evacuations (or temporary displacements lasting weeks to months) produce negative effects. Prolonged
periods of evacuation can result in a number of physical and mental health problems (Curtis et al., 2007; Mills et al., 2007). Furthermore,
separation from family and community members and not knowing when a return home will be possible also adds to stress among evacuees
(Curtis et al., 2007). DeSalvo et al. (2007) found that long periods of displacement were among the key causes of post-traumatic stress
disorder in a study of New Orleans workers. These temporary displacements can also lead to permanent outmigration by specific social
groups as shown by the depopulation of New Orleans five years after Hurricane Katrina (Myers et al., 2008). In terms of longer-term
recovery, New Orleans is progressing, however large losses in population, housing, and employment suggest a pattern of only partial
recovery for the city with significant differences in the location and the timing at the neighborhood or community level (Finch et al., 2010).
316
patterns of behavior (Meehl et al., 2007), with a focus on local people’s
agency within specific configurations of power relations. The challenge is,
therefore, to empower the most vulnerable to pursue livelihood options
that strengthen their entitlements and protect what they themselves
consider the social sources of adaptation and resilience in the face of
extreme climate stress. Better management of disaster risk also
maximizes use of available resources for adapting to climate change
(Kryspin-Watson et al., 2006).
5.5.1.3. Health and Disability
The changes in extreme events and impacts of climate change influence
the morbidity and mortality of many populations now, and even more
so in the future (Campbell-Lendrum et al., 2003). The extreme impacts
of climate change (see Sections 3.1.1 and 4.4.6) directly or indirectly
affect the health of many populations and these will be felt first at the
local level. Heat waves lead to heatstroke and cardiovascular disease,
while shifts in air pollution concentrations such as ozone that often
increase with higher temperatures cause morbidity from other diseases
(Bernard et al., 2001). Heat waves differentially affect populations
based on their ethnicity, gender, age (Díaz et al., 2002), and medical and
socioeconomic status (O’Neill and Ebi, 2009), consequently raising
concerns about health inequalities (see Case Study 9.2.1), especially at
the local scale. Health inequalities are of concern in extreme impacts of
climate change more generally, as those with the least resources often
have the least ability to adapt, making the poor and disenfranchised
most vulnerable to climate-related illnesses (McMichael et al., 2008).
For extreme events, pre-existing health conditions that characterize
vulnerable populations can exacerbate the impact of disaster events
since these populations are more susceptible to additional injuries from
disaster impacts (Brauer, 1999; Brown, 1999; Parati et al., 2001). Chronic
health conditions/disabilities can also lead to subsequent communicable
diseases and illnesses in the short term, to lasting chronic illnesses, and
to longer-term mental health conditions (Shoaf and Rottmann, 2000;
Bourque et al., 2006; Few and Matthies, 2006).
A range of vector-borne illnesses has been linked to climate, including
malaria, dengue, Hantavirus, Bluetongue, Ross River Virus, and cholera
(Patz et al., 2005). Cholera, for example, has seasonal variability that
may be directly affected by climate change (Koelle et al., 2004). Vector-
borne illnesses have been projected to increase in geographic reach and
severity as temperatures increase (McMichael et al., 2006), but these
changes depend on a variety of human interventions like deforestation
and land use. The areas of habitation by mosquitoes and other vectors
are moving to areas previously free from such vectors of transmission
(Lafferty, 2009). Pools of standing water that are breeding grounds for
mosquitoes promise to expand, therefore increasing illness exposure
(Depradine and Lovell, 2004; Meehl et al., 2007). At the same time, some
literature shows that illnesses like malaria are less prone to increase
than originally thought (Gething et al., 2010). Much of the nuance of
this literature is due to the location-specific nature of these outcomes.
Therefore, vector control programs will be best suited to the local
characteristics of changing risks. Some programs, like those geared
toward surveillance, need common characteristics to support national
programs and also need to be coordinated across scales from local to
national and between local places. In addition, there are a variety of
social factors that have the potential to influence disease rates that are
most suitably managed at the sub-national level or urban scale. For
instance, certain types of population growth or change may increase
risk and affect disease rates (Patz et al., 2005). Increased population and
related land use changes can also increase disease rates. Vector control
programs generally implemented at the local level also have the potential
to influence health outcomes (Tanser et al., 2003). Infectious disease
patterns also have the potential to change dramatically, necessitating
improved prevention on the part of local providers who have knowledge
of local environmental change (Parkinson and Butler, 2005).
There is concern regarding the mental health impacts of storms and floods
that lead to destruction of livelihoods and displacement, especially for
vulnerable populations (Balaban, 2006). In some hurricanes, the mental
health of residents in affected communities is extremely negatively
impacted over an extended period of time (Weisler et al., 2006). Policy
responses to the event were insufficient to manage these impacts, and
provide a lesson for future events where greater mental health services
may be necessary (Lambrew and Shalala, 2006). Managing public
health and disability is important in the response to disasters (Shoaf
and Rottmann, 2000).
Human health is at risk from many extreme events linked to climate
change. While resources from scales above the local are often necessary,
the direction and application of those resources by local actors who
know how to best apply them could make significant differences in
human morbidity and mortality linked to climate extremes.
5.5.1.4. Human Settlements
Settlement patterns are another factor that influences disaster risk
management and coping with extremes. Human settlements differ in
their physical and governance structures, population growth patterns,
as well as in the types, drivers, impacts, and responses to disasters. As
noted earlier (see Section 5.5.1.2), rural livelihoods and poverty are
drivers of disaster risk, but not the only ones. Poverty, resource scarcity,
access to resources, as well as inaccessibility constrain disaster risk
management. When these are coupled with climate variability, conflict,
and health issues they reduce the coping capacity of rural places
(UNISDR, 2009). At the other extreme are the concentrated settlements
of towns and cities where the disaster risks are magnified because of
population densities, poor living conditions including overcrowded and
substandard housing, lack of sanitation and clean water, and health
impairments from pollution and lack of adequate medical care (Bull-
Kamanga et al., 2003; De Sherbinin et al., 2007). Strengthening local
capacity in terms of housing, infrastructure, and disaster preparedness
is one mechanism shown to improve urban resilience and the adaptive
capacity of cities to climate-sensitive hazards (Pelling, 2003). It is also
Chapter 5Managing the Risks from Climate Extremes at the Local Level
317
instructive to note how communities with differing capacities address
similar problems (Walker and Sydneysmith, 2008).
One important locality receiving considerable research and policy attention
is megacities (see Case Study 9.2.8) due to the density of infrastructure,
the population at risk, the growing number and location of informal
settlements, complex governance, and disaster risk management
(Mitchell, 1999). Given the rapid rate of growth in the largest of these
world’s cities and the increasing urbanization, the disaster risks will
increase in the next decade, placing more people in harm’s way with
billions of dollars in infrastructure located in highly exposed areas
(Munich Re Group, 2004; Kraas et al., 2005; Wenzel et al., 2007).
For many regions, the ability to limit urban exposure has already been
achieved through building codes, land management, and disaster risk
mitigation, yet losses keep increasing. For disaster reduction to become
more effective, megacities will need to address their societal vulnerability
and the driving forces that produce it (rural to urban migration,
livelihood pattern changes, wealth inequities, informal settlements)
(Wisner and Uitto, 2009). Many megacities are seriously compromised
in their ability to prepare for and respond to present disasters, let alone
adapt to future ones influenced by climate change (Fuchs, 2009;
Heinrichs et al., 2009; Prasad et al., 2009).
However, it is not only the megacities that pose challenges, but the
overall growth in urban populations. Currently more than half of the
global population lives in urban areas with an increasing population
exposed to multiple risk factors (UNFPA, 2009). Risk is increasing in urban
agglomerations of different size due to unplanned urbanization and
accelerated migration from rural areas or smaller cities (UN-HABITAT,
2007). The 2009 Global Assessment Report on Disaster Risk Reduction
(UNISDR, 2009) lists unplanned urbanization and poor urban governance
as two main underlying factors accelerating disaster risk. It highlighted
that the increase in global urban growth of informal settlements in hazard-
prone areas reached 900 million in informal settlements, increasing by
25 million per year (UNISDR, 2009). Urban hazards exacerbate disaster risk
by the lack of investment in infrastructure as well as poor environmental
management, thus limiting the adaptive capacity of these areas.
5.5.2. Costs of Managing Disaster Risk
and Risk from Climate Extremes
5.5.2.1. Costs of Impacts, Costs of Post-Event Responses
It is extremely difficult to assess the total cost of a large disaster, such
as Hurricane Katrina, especially at the local scale since most economic
data are only available at the national scale. Direct losses consist of
direct market losses and direct non-market losses (intangible losses).
The latter include health impacts, loss of lives, natural asset damages
and ecosystem losses, and damages to historical and cultural assets.
Indirect losses (also labeled higher-order losses (Rose, 2004) or hidden
costs (Heinz Center, 1999)) include all losses that are not provoked by
the disaster itself, but by its consequences. Measuring indirect losses is
important as it evaluates the overall economic impact of the disaster on
society. Another difficulty with the measurement of economic losses
at the local level has to do with the boundary delineation for local
analyses. For example, local losses can be compensated from various
inflows of goods, workers, capital, and governmental or foreign aid from
outside the affected area (Eisensee and Stromberg, 2007). Local
disasters also provide ripple effects and influence world markets, for
instance, oil prices in the case of Hurricane Katrina due to the
temporary shutdown of oil rigs. It is important to consider tradeoffs at
different spatial scales especially when estimating indirect losses at the
local level. Disaster loss estimates are, therefore, highly dependent on the
scale of the analysis, and result in wide variations among community,
state, province, and sub-national regions.
Despite the difficulties in assessing local economic impact, several studies
exist. For example, Strobl (2008) provided an econometric analysis of
the impact of the hurricane landfall on county-level economic growth in
the United States. This analysis showed that a county struck by at least
one hurricane over a year led to a decline in economic growth on average
by 0.79% and an increase by 0.22% the following year. The economic
impact of the 1993 Mississippi flooding in the United States showed
significant spatial variability within the affected regions. In particular,
states with a strong dependence on the agricultural sector had a
disproportionate loss of wealth compared to states that had a more
diversified economy (Hewings and Mahidhara, 1996). Noy and Vu (2010)
investigated the impact of disasters on economic growth in Vietnam at
the provincial level, and found that fatal disasters decreased economic
production while costly disasters increased short-term growth.
Rodriguez-Oreggia et al. (2010) focused on poverty and the World
Bank’s Human Development Index at the municipality level in Mexico,
and demonstrated that municipalities affected by disasters saw an
increase in poverty of 1.5 to 3.6%. Studies also found that regional
indirect losses increase nonlinearly with direct losses (Hallegatte, 2008),
and can be compensated by importing the means for reconstruction
(workers, equipment, finance) from outside the affected area.
The U.S. Bureau of Labor Statistics (2006) also provided a detailed
analysis of the labor market consequences of Hurricane Katrina within
Louisiana and found a marked economic and employment loss for
Louisiana businesses, high unemployment rates immediately after the
disaster, and continued unemployment into 2006 for returning evacuees.
At the household level, Smith and McCarty (2006) show that households
are more often forced to move outside the affected area due to
infrastructure problems than due to structural damages to their home.
Modeling approaches are also used to assess disaster indirect losses at
sub-national levels. These approaches include input-output models
(Okuyama, 2004; Haimes et al., 2005; Hallegatte, 2008) and Computable
General Equilibrium models (Rose et al., 1997; Rose and Liao, 2005;
Tsuchiya et al., 2007). Most of the published analyses were carried out
in developed countries. In the United States, West and Lenze (1994), for
example, discuss the merits of combining different impact models to
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
318
triangulate, obtaining better primary data to reduce uncertainty, and
developing tools for estimating the impact of Hurricane Andrew on
Florida using reconstruction scenarios. The lack of research on disaster
loss estimates in developing countries creates problems of underreported
economic losses or overestimation of disaster losses depending on
political or other interests. This is a big research gap.
5.5.2.2. Adaptation and Risk Management – Present and Future
Studies on the costs of local disaster risk management are scarce,
fragmented, and conducted mostly in rural areas. One study estimated the
benefit/cost ratio of disaster management and preparedness programs in
the villages of Bihar and Andra Pradesh, India to be 3.76 and 13.38,
respectively (Venton and Venton, 2004), suggesting higher benefits than
costs. Research undertaken by the Institute for Social and Environmental
Transition on a number of cases in India, Nepal, and Pakistan
demonstrated that benefits exceed the costs for local interventions
(Dixit et al., 2008; Moench and Risk to Resilience Study Team, 2008).
For example, they note that return rates are particularly robust for
lower-cost interventions (e.g., raising house plinths and fodder storage
units, community-based early warning, establishing community grain or
seed banks, and local maintenance of key drainage points), when
compared to embankment infrastructure strategies that require capital
investment (Moench and Risk to Resilience Study Team, 2008). The
studies demonstrated a sharp difference in the effectiveness of the two
approaches, concluding that the embankments historically have not had
an economically satisfactory performance in that study area. In contrast,
the benefit/cost ratio for the local-level strategies indicated economic
efficiency over time and for all climate change scenarios (Dixit et al.,
2008). In developed countries, there are cost differences in adaptation
strategies between urban and rural areas. For example, in Japan disaster
damage is several hundred times more costly in urban than in rural
areas, necessitating different disaster risk management strategies
depending on the benefit to cost analysis (Kazama et al., 2009).
Though disaster risk management and adaptation policies are closely
linked, few integrated cost analyses of risk management and adaptation
are available at the local level. One example draws from recent studies
of the cost of city-scale adaptation. Rosenzweig et al. (2007, 2011)
developed a sophisticated analytical response to a projected fall in water
availability in New York. This frames adaptation assessment within a
step-wise decision analysis by identifying and quantifying impact risks
before identifying adaptation options that are then screened, evaluated,
and finally implemented. Another series of studies used simplified
catastrophe risk assessment to calculate the direct costs of storm surges
under scenarios of sea level rise coupled with an economic input-output
model for Copenhagen and Mumbai (Hallegatte et al., 2008a,b, 2011;
Ranger et al., 2011). The output is an assessment of the direct and indirect
economic impacts of storm surge under climate change including
production, job losses, reconstruction time, and the benefits of investment
in upgraded coastal defenses. Results show that the consideration of
adaptation is an important element in the economic assessment of
extreme disaster risks related to climate change (Hallegatte et al.,
2011). Ranger et al. (2011) show that by improving the drainage system
in Mumbai, losses associated with a 1-in-100 year flood event could be
reduced by as much as 70%. This means that the annual losses could be
reduced in absolute terms compared with the current level, even with
climate change. Full insurance coverage of flooding could also cut the
indirect cost by half. These analyses highlight the fact that adaptation to
extreme events and climate change can focus on reducing the direct
losses (e.g., through the upgrade of coastal defenses) or indirect losses
by making the economy more robust, utilizing insurance schemes, or
enacting public policies to support small businesses after the disaster.
5.5.2.3. Consistency and Reliability of Cost and
Loss Estimations at the Local Level
There are inconsistencies in disaster-related economic loss data at all
levels – local, national, global – that ultimately influence the accuracy
of such estimates (Guha-Sapir and Below, 2002; Downton and Pielke Jr.,
2005; Pielke Jr. et al., 2008). The reliability of disaster economic loss
estimates is especially problematic at the local level. First, the spatial
coverage and resolution of databases are global in coverage, but only
have data that represent the entire country, not sub-units within it such
as provinces, states, or counties. Second, thresholds for inclusion, where
only large economically significant disasters are included, bias the data
toward singular events with large losses, rather than multiple, smaller
events with fewer losses. Third, what gets counted varies between
databases (e.g., insured versus uninsured losses; direct versus indirect; Gall
et al., 2009). Moreover, disaster loss estimates have various purposes
Chapter 5Managing the Risks from Climate Extremes at the Local Level
FAQ 5.3 | Is it possible to estimate the cost
of risk management and adaptation
at the local scale?
Studies on the costs of local disaster risk management are
scarce, fragmented, and conducted mostly in rural areas. Most
economic data (e.g., input-output table, income data) are
available at the national scale. Moreover, there is a clear lack of
research on disaster estimates in developing countries, which
presents a big gap in need of further research. In developed
countries, there are cost differences in adaptation strategies
between urban and rural areas. The reliability of disaster
economic loss estimates is especially problematic at the local
level due to factors associated with the global nature of
spatial coverage and resolution. In addition there is some
ambiguity on impact and adaptation costs that affect local-
level economic analyses, such as the lack of consensus on
physical impacts of climate change and adaptive capacity and
on the evaluation of non-market costs (e.g., biodiversity or
cultural heritage), which creates some uncertainty about local
impact and adaptation costs.
319
(e.g., assessment of foreign aid needs; cost-benefit analysis of protection
investments; World Bank, 2010). Depending on the purpose, spatial and
conceptual gaps exist depending on the inclusion of loss-only data or a
combination of loss and gain estimates as well as the calculation of
non-market losses.
Similarly, there is some ambiguity about impact and adaptation costs
that affect local-level economic analyses. The lack of consensus on
physical impacts of climate change and adaptive capacity (see Section
4.5) is one issue. Another is the discount rate (Heal, 1997; Tol, 2003;
Nordhaus, 2007; Stern, 2007; Weitzman, 2007) and the evaluation of
non-market costs, especially the value of biodiversity or cultural heritage
(Pearce and Moran, 1994), the latter contributing some uncertainty
about local impact and adaptation costs. Finally, the possibility of low-
probability high-consequence climate change is not fully included in
most analyses (Stern, 2007; Weitzman, 2007; Lonsdale et al., 2008;
Nicholls et al., 2008).
5.5.3. Limits to Local Adaptation
Local adaptation is set within larger spatial and temporal scales (Adger
et al., 2005), which influence the range of actors involved and the types
of potential barriers to the adaptation process (Moser and Ekstrom,
2010; see Sections 6.3 and 7.6). At the local scale, limits and barriers to
local adaptation generally fall into three interconnected categories:
ecological and physical; human informational related to knowledge,
technology, economics, and finances; and psychological, behavioral,
and socio-cultural barriers (ICIMOD, 2009; Adger et al., 2010). The social
and cultural limits to adaptation are not well researched, with little
attention within the climate change literature devoted to this thus far.
Lack of access to information by local people has restricted improvements
in knowledge, understanding, and skills – needed elements in helping
localities undertake improved measures to protect themselves against
disasters and climate change impacts (Agrawal et al., 2008). The
information gap is particularly evident in many developing countries with
limited capacity to collect, analyze, and use scientific data on mortality
and demographic trends as well as evolving environmental conditions
(IDRC, 2002; Carraro et al., 2003; NRC, 2007). Based on Fischer et al.
(2002), closing the information gap is critical to reducing climate
change-related threats to rural livelihoods and food security in Africa.
Lack of capacities and skills, particularly for women, also has been
identified as a limiting factor for effective local adaptation actions
(Osman-Elasha et al., 2006). For example, localities in areas prone to
climate extremes such as frequent drought have developed certain
coping responses that assist them in surviving harsh conditions. Over
time, such coping responses proved inadequate due to the magnitude of
the problem (Ziervogel et al., 2006). For example, in Mali, one initiative
involves empowering women and giving them the skills to diversify
their livelihoods, thus linking environmental management, disaster
risk reduction, and the position of women as key resource managers
(UNISDR and UNOCHA, 2008).
In financial terms, microfinance services typically do not reach the
poorest and most vulnerable groups at local levels who have urgent and
immediate needs to be addressed (Amin et al., 2001; Helms, 2006). The
ability of a community to ensure equitable access and entitlement to key
resources and assets is a key factor in building local adaptive capacity.
In developed countries, household decisions regarding disaster risk
reduction and adaptation are often guided by factors other than cost.
For example, Kunreuther et al. (2009) found that most individuals
underestimate the risk and do not make cost-benefit tradeoffs in their
decisions to purchase hazard insurance and/or have adequate coverage.
They also found empirical evidence to suggest that the hazard insurance
purchase decision was driven not only by the need to protect assets, but
also to reduce anxiety, satisfy mortgage requirements, and social
norms. For other types of disaster mitigation activities, households do
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
FAQ 5.4 | What are the limits to adaptation at the local level?
Traditionally, local risk management strategies focused only on short-term climatic events without considering the long-term trajectories
presented by a changing climate. Although reacting to climate extreme events and their impacts is important, it is more crucial now to
focus on building the resilience of communities, cities, and sectors in order to ameliorate the impacts of future climatic changes. The
range and choice of actions that can be taken at the levels of individual or households are often event-specific and time-dependent.
They are also constrained by location, adequate infrastructure, socioeconomic characteristics, and access to disaster risk information. For
example, the increased urban vulnerability due to urbanization and rising population exacerbates disaster risk by the lack of investment
in infrastructure as well as poor environmental management, and can have spillover effects to rural areas.
The obstacles to information transfer and communications are diverse, ranging from limitations in modeling the climate system to
procedural, institutional, and cognitive barriers in receiving or understanding climatic information and advance warnings and the capacity
and willingness of decisionmakers to modify action. Within many rural communities, low bandwidth and poor computing infrastructure
pose serious constraints to risk message receipt. Such gaps are evident in developed as well as lesser-developed regions. Constraints
exist in locally-organized collective action because of the difficulties of building effective coalitions with other organizations.
320
not voluntarily invest in cost-effective mitigation because of
underestimating the risk, taking a short-term rather than long-term
view, and not learning from previous experience. However, they found
social norms significant: if homeowners in the neighborhood installed
hurricane shutters, most would follow suit; the same was true of
purchasing insurance (Kunreuther et al., 2009). For municipal governments,
adoption of building codes in hurricane-prone areas reduces damages
by US$ 108 per square meter for homes built from 1996 to 2004 in Florida
(Kunreuther et al., 2009). However, enforcement of building codes by
municipalities is highly variable and becomes a limiting factor in disaster
risk management and adaptation.
Local-level adaptation actions in many cases are portrayed as reactive
and short term, unlike the higher-level national or regional plans that are
considered anticipatory and involve formulation of policies and programs
(Bohle, 2001; Burton et al., 2003). Poverty, increased urbanization, and
extreme climate events limit the capacity to initiate planned livelihood
adaptations at the local scale. If extreme events happen more frequently
or with greater intensity or magnitude some locations may be
uninhabitable for lengthy and repeated periods, rendering sustainable
development impossible. In such a situation, not all places will be able
to adapt without considerable disruption and costs (economic, social,
cultural, and psychological). In some cases forced migration may be the
only alternative (Brown, 2008).
As the above shows, the main challenge for local adaptation to climate
extremes is to find a good balance of measures that simultaneously
address fundamental issues related to the enhancement of local
collective actions, and the creation of subsidiary structures at national
and international scales that complement such local actions. This
means that the localized expression of the type, frequency, and
extremeness of climate-sensitive hazards will be set within these
national and international contexts.
5.5.4. Advancing Social and Environmental Justice
One of the key issues in examining outcomes of local strategies for
disaster risk management and climate change adaptation is the principle
of fairness and equity. There is a burgeoning research literature on climate
justice looking at the differential impacts of adaptation policies
(Kasperson and Kasperson, 2001; Adger et al., 2006) at local, national,
and global scales. The primary considerations at the local level are the
differential impacts of policies on communities, subpopulations, and
regions from present management actions (or inactions) (Thomas and
Twyman, 2005). There is also concern regarding the impact of present
management (or inactions) in transferring the vulnerability of disaster risk
from one local place to another (spatial inequity) or from one generation
to another (intergenerational equity) (Cooper and McKenna, 2008).
There is less research on the mechanisms or practical actions needed for
advancing social and environmental justice at the local scale, independent
of the larger issues of accountability and governance at all scales. This
is an important gap in the literature.
5.6. Management Strategies
5.6.1. Basics of Planning in a Changing Climate
Prior to the development and implementation of management strategies
and adaptation alternatives, local entities need baseline assessments on
disaster risk and the potential impacts of climate extremes. The assessment
of local disaster risk includes three distinct elements: 1) exposure
hazard assessment, or the identification of hazards and their potential
magnitudes/severities as they relate to specific local places (see
below); 2) vulnerability assessments that identify the sensitivity of the
population to such exposures and the capacity of the population to cope
with and recover from them (see below and Sections 2.6.2 and 4.4); and
3) damage assessments that determine direct and indirect losses
from particular events (either ex-post in real events or ex-ante through
modeling of hypothetical events) (described in Section 5.5.2; see
Section 4.5.1). Each of these plays a part in understanding the hazard
vulnerability of a particular locale or characterizing not only who is at risk
but also the driving forces behind the differences in disaster vulnerabilities
in local places.
There are numerous examples of exposure and vulnerability assessment
methodologies and metrics (Birkmann, 2006; see Chapter 2). Of particular
note are those studies focused on assessing the sub-national exposure
to coastal hazards (Gornitz et al., 1994; Hammar-Klose and Thieler, 2001),
drought (Wilhelmi and Wiilhite, 2002; Alcamo et al., 2008; Kallis, 2008),
or multiple hazards such as the Federal Emergency Management Agency’s
multi-hazard assessment for the United States (FEMA, 1997).
Vulnerability assessments highlight the interactive nature of disaster risk
exposure and societal vulnerability. While many of them are qualitatively
based (Bankoff et al., 2004; Birkmann, 2006), there is an emergent
literature on quantitative metrics in the form of vulnerability indices. The
most prevalent vulnerability indices, however, are national in scale
(SOPAC and UNEP, 2005; Cardona, 2007) and compare countries to one
another, not places at sub-national geographies. The exceptions are the
empirically-based Social Vulnerability Index (Cutter et al., 2003) and
extensions of it (Fekete, 2009).
Vulnerability assessments are normally hazard-specific and many have
focused on climate-sensitive threats such as extreme storms in Revere,
Massachusetts (Clark et al., 1998), sea level rise in Cape May, New
Jersey (Wu et al., 2002), or flooding in Germany (Fekete, 2009) and the
United States (Burton and Cutter, 2008; Zahran et al., 2008). Research
focused on multi-hazard impact assessments ranges from locally based
county-level assessments for all hazards (Cutter et al., 2000) to sub-
national studies such as those involving all hazards for Barbados and
St. Vincent (Boruff and Cutter, 2007) to those involving a smaller subset
of climate-related threats (O’Brien et al., 2004; Brenkert and Malone,
2005; Alcamo et al., 2008). The intersection of local exposure to climate-
sensitive hazards and social vulnerability was recently assessed for the
northeast (Cox et al., 2007) and southern region of the United States
(Oxfam, 2009).
Chapter 5Managing the Risks from Climate Extremes at the Local Level
321
However, the full integration of hazard exposure and social vulnerability
into a comprehensive vulnerability assessment for the local area or region
of concern is often lacking for many places. Part of this is a function of
the bifurcation of the science inputs (e.g., natural scientists provide
most of the relevant data and models for exposure assessments while
social scientists provide the inputs for the populations at risk). It also
relates to the difficulties of working across disciplinary or knowledge
boundaries.
The development of methodologies and metrics for climate adaptation
assessments is emerging and mostly derivative of the methodologies
employed in vulnerability assessments noted above. For example, some
are extensions or modifications of community vulnerability assessment
methodologies and employ community participatory approaches such
as those used by World Vision (Greene, n.d.), the Red Cross (van Aalst et
al., 2008), and others. Still others begin with livelihood or risk assessment
frameworks and use a wide range of techniques including multi-criteria
decision analyses (Eakin and Bojorquez-Tapia, 2008); index construction
(Vescovi et al., 2009); segmentation and regional to global comparisons
(Torresan et al., 2008); and scenarios (Wilby et al., 2009).
5.6.2. Community-Based Adaptation
Community-based adaptation (CBA) empowers communities to decide
how they want to prepare for climate risks and coordinate community
action to achieve adaptation to climate change (Ebi, 2008). Part of this
entails community risk assessment for climate change adaptation that
assesses the hazards, vulnerabilities, and capacities of the community
(van Aalst et al., 2008), which has also been called community-based
disaster preparedness among other names. The intention is to foster
active participation in collecting information that is rooted in the
communities and enables affected people to participate in their own
assessment of risk and identify responses than can enhance resilience
by strengthening social-institutional measures including social relations
(Allen, 2006; Patiño and Gauthier, 2009). In assessing short- and long-
term risks, the needs of vulnerable groups are often excluded (Douglas
et al., 2009). The tools for engaging vulnerable groups in the process
include transect walks and risk maps that capture the climate-related
hazards and risks and storylines about possible future climate change
impacts (Ebi, 2008; van Aalst et al., 2008; Patiño and Gauthier, 2009),
although these tools often require input from participants external to
the community who have long-term climate information.
The challenges in using community-based adaptation approaches
include the challenge of scaling up information (Burton et al., 2007), the
fact that it is resource-intensive (van Aalst et al., 2008), and recognizing
that disempowerment occurs when local stories are distorted or not
valued sufficiently (Allen, 2006). The integration of climate change
information increases this challenge as it introduces an additional layer
of uncertainty and may conflict with the principle of keeping CBA simple.
There is little evidence that secondary data on climate change has
been used in CBA, partly because of the challenge of limited access to
downscaled climate change scenarios relevant at the local level (Ziervogel
and Zermoglio, 2009) and because of the uncertainty of projections.
Examples of community-based approaches illustrate some of the
processes involved. In northern Bangladesh, a flooding adaptation project
helped to establish early warning committees within villages that linked
to organizations outside the community, with which they did not usually
interact and that had historically blocked collective action and resource
distribution (Ensor and Berger, 2009). Through this revised governance
structure, the building of small roads, digging culverts, and planting
trees to alleviate flood impacts was facilitated. In Portland, Oregon,
another project involved a range of actors to reduce the impact of urban
heat islands through engaging neighborhoods and linking them to
experts to install green roofs, urban vegetation, and fountains that led
to an increased sense of ownership in the improvements (Ebi, 2008). In
the Philippines, the community-based approach enabled a deeper
understanding of locally specific vulnerability than in previous disaster
management contexts (Allen, 2006). While individually important, these
community-based approaches should be viewed as part of a wider
system that recognizes the drivers at multiple scales, including the
municipalities and national levels.
CBA responses provide increased participation and recognition of the
local context, which is important when adapting to climate change (see
Box 5-7). The need for coordinated collective action was seen in
Kampala, where land cover change and changing climate are increasing
the frequency and severity of urban flooding and existing response
activities are uncoordinated and consist of clearing drainage channels
(Douglas et al., 2009). However, residents felt more could be done to
adapt to frequent flooding, including increasing awareness of roles and
responsibilities in averting floods, improving the drainage system,
improving garbage and solid waste disposal, strengthening the building
inspection unit, and enforcing bylaws on the construction of houses and
sanitation facilities. Similarly, in Accra, residents felt that municipal laws
on planning and urban design need to be enforced, suggesting that
strong links are needed between community responses and municipal
responses (Douglas et al., 2009).
5.6.3. Risk Sharing and Transfer at the Local Level
Risk transfer and risk sharing are pre-disaster financing arrangements
that shift economic risk from one party to another and are more fully
discussed in Chapter 2, Section 7.4.4, and Case Study 9.2.13. Informal
risk sharing practices are common and important for post-disaster relief
and reconstruction. In the absence of more formal mechanisms like
insurance, those incurring losses may employ diverse non-insurance
financial coping strategies, such as relying on the solidarity of international
aid, remittances, selling and pawning fungible assets, and borrowing from
moneylenders. Traditional livestock loans are one example (Oba, 2001).
At-risk individuals in low-income countries rely extensively on reciprocal
exchange, kinship ties, and community self-help. For example, women in
high-risk areas often engage in innovative ways to access post-disaster
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
322
capital by joining informal risk-hedging schemes, becoming clients of
multiple microfinance institutions, or maintaining reciprocal social
relationships. Combined analysis of multiple surveys suggests that
about 40% of households in low- and lower-middle income countries
are involved in private transfers in a given year as recipients or donors
(Davies and Leavy, 2007).
Households in disaster-prone slum areas in El Salvador spend an average
of 9.2% of their yearly income on risk management, including financing
emergency relief and recovery (Wamsler, 2007). A particularly important
informal risk-sharing mechanism is remittances, or transfers of money from
foreign workers to their home countries (discussed further in Section
7.4.4). Household savings can be accessed from a bank, but they can also
be in the form of stockpiles of food, grains, seeds, and fungible assets.
Small savings institutions can be directly impacted by catastrophes, which
can result in insufficient liquidity to handle a run on their accounts, as
occurred during the 1998 floods in Bangladesh (Kull, 2006). Lacking
sufficient savings, many disaster victims take out loans to cover their post-
disaster expenses. The interest rate (18-60%) charged on formal micro-
credit, while relatively high, generally is far below the rate (120-300%)
charged by local moneylenders (Linnerooth-Bayer and Mechler, 2009).
Insurance, including micro-insurance, is the most common formal risk
transfer mechanism at the local level. An insurance contract spreads
stochastic losses geographically and temporally, and can assure timely
liquidity for the recovery and reconstruction process. As such, it is an
effective disaster risk reduction tool especially when combined with
other risk management measures. For example, in most industrialized
countries, insurance is utilized in combination with early warning systems,
risk information, disaster preparation, and disaster mitigation. Where
insurance is applied without adequate risk reduction, it can be a
disincentive for adaptation, as individuals may rely on insurance to
manage their risks and are left overly exposed to impacts (Rao and
Hess, 2009; see Section 5.4.1). Furthermore, insurance can provide the
necessary financial security to take on productive but risky investments
(Höppe and Gurenko, 2006). Examples include a pilot project in Malawi
where micro-insurance is bundled with loans that enable farmers to access
agricultural inputs that increase their productivity (Hess and Syroka, 2005),
and a project in Mongolia that protects herders’ livestock from extreme
winter weather to reduce livestock losses (Skees et al., 2008).
Micro-insurance is a financial arrangement to protect low-income people
against specific perils in exchange for regular premium payments
(Churchill, 2006, 2007). Several pilot projects have yielded promising
outcomes, yet experience is too short to judge if micro-insurance schemes
are viable in the long run for local places. Many of the ongoing micro-
insurance initiatives are index-based – a relatively new approach whereby
the insurance contract is not against the loss itself, but against an event
Chapter 5Managing the Risks from Climate Extremes at the Local Level
Box 5-7 | Taking Collective Action to Improve Livelihoods Strategies: Small-Scale Farmers
Adapting to Climate Change in Northern Cape, South Africa
The Northern Cape Province, South Africa, is a harsh landscape, with frequent and severe droughts and extreme conditions for the people,
animals, and plants living there. This has long had a negative impact on small-scale rooibos farmers living in some of the more marginal
production areas. Rooibos is an indigenous crop that is well adapted to the prevailing hot, dry summer conditions, but is sensitive to
prolonged drought. Rooibos tea has become well-accepted on world markets, but this success has brought little improvement to
marginalized small-scale producers.
In 2001, a small group of farmers decided to take collaborative action to improve their livelihoods and founded the Heiveld Co-operative
Ltd. Initially established as a trading cooperative to help the farmers produce and market their tea jointly, it subsequently became
apparent that the local organization was also an important vehicle for social change in the wider community (Oettlé et al., 2004). The
Heiveld became a repository and source of local and scientific knowledge related to sustainable rooibos production. Following a severe
drought (2003-2005) and a perceived increase in weather variability, the Heiveld farmers decided to monitor the local climate and to
discuss seasonal forecasts and possible strategies in quarterly climate change preparedness workshops. These workshops are facilitated
in collaboration with two local NGOs (Indigo and Environmental Monitoring Group). They are also supported by scientists to address
farmers’ questions in a participatory action research approach – to ensure that local knowledge and scientific input can be combined to
increase the resilience of local livelihoods. The Heiveld Co-operative has been an important organizational vehicle for this learning
process, strongly supported by their long-term partners, with the focus on supporting the development of possible adaptation strategies
through a joint learning approach to respond to and prepare for climate variability and change.
The extension of social, participatory, and organizational learning to climate change adaptation illustrated in this case study emphasizes
the significance of identifiable climate change signals, informal networks, and boundary organizations to enhance the preparation of
people and organizations for the changing climate (Berkhout et al., 2006; Pelling et al., 2008). Participatory learning is especially
emphasized (Berkhout, 2002; A. Shaw et al., 2009; R. Shaw et al., 2009). Focusing on what can be learned from managing current
climate risk is a good starting point, particularly for poor and marginalized communities (Someshwar, 2008).
323
that causes loss, such as insufficient rainfall during critical stages of
plant growth (Turvey, 2001). Weather index insurance is largely at a
pilot stage, with several projects operating around the globe, including
in Mongolia, Kenya, Malawi, Rwanda, and Tanzania (Hellmuth et al.,
2009). Index insurance for agriculture is more developed in India, where
the Agricultural Insurance Company of India has extended coverage
against inadequate rainfall to 700,000 farmers (Hellmuth et al., 2009).
Index-based contracts as an alternative to traditional crop insurance
have the advantages of greatly limiting transaction costs (from reduced
claims handling) and in improving emergency response (Chantarat et al.,
2007). A disadvantage is the potential of a mismatch between yield and
payout, a critical issue given the current lack of density of meteorological
stations in vulnerable regions – a challenge remote sensing may help
address (Skees and Barnett, 2006). Participants’ understanding of how
insurance operates, as well as their trust in the product and the
stakeholders involved, may also be a problem for scaling up index
insurance pilots, although simulation games and other innovative
communication approaches are yielding promising results (Patt et al.,
2009). Affordability can also be a problem. Disasters can affect whole
communities or regions (co-variant risks), and because of this insurers
must be prepared for meeting large claims all at once, with the cost of
requisite backup capital potentially raising the premium far above the
client’s expected losses – or budget. While valuable in reducing the
long-term effects on poverty and development, insurance instruments,
particularly if left entirely to the market, are not appropriate in all
contexts (Linnerooth-Bayer et al., 2010).
The insurance industry itself is vulnerable to climate change. The
continuing exit of private insurers in some market areas is seen with the
increasingly catastrophic local losses in the United States (Lecomte and
Gahagan, 1998), United Kingdom (Priest et al., 2005), and Germany
(Thieken et al., 2006; Botzen and van den Bergh, 2008), which in turn
reduces disaster management options at the sub-national scale. Climate
change could be particularly problematic for this sector at the local
scale (Vellinga et al., 2001), including the probable maximum loss and
pressures from regulators responding to changing prices and coverage
(Kunreuther et al., 2009).
One response to rising levels and volatility of risk has been to increase
insurance and reinsurance capacity through new alternative risk transfer
instruments, such as index-linked securities (including catastrophe bonds
and weather derivatives) (Vellinga et al., 2001). These tools could play
an increasingly important role in a new era of elevated catastrophe risks
(Kunreuther et al., 2009). Another approach is to reduce risks through
societal adaptation (Herweijer et al., 2009), and risk communication and
financial incentives from insurers (Ward et al., 2008). For example,
Lloyds of London (2008) demonstrated that in exposed coastal regions,
increases in average annual losses and extreme losses due to sea level
rise in 2030 could be offset through investing in property-level
resilience to flooding or sea walls. Similarly, RMS (2009) shows that
wind-related losses in Florida could be significantly reduced through
strengthening buildings.
Risk transfer is broader than shifting the economic burden from one
party to another. It also entails the transfer of risks from one generation
(intergenerational equity) to the next. Risk transfer also has a spatial
element in shifting the risk burdens from one geographic location to
another. Both of these larger transfer mechanisms are significant for
disaster risk management and climate change adaptation at the local
scale, but more research is required to assess the localized effects.
Spatial and intergenerational equity are considered in Chapter 8.
5.6.4. A Transformative Framework
for Management Strategies
Management strategies need to consider adaptation as a process rather
than measures and actions for a particular event or time period. Experience
in planning and implementing adaptation to climate change as well as
disaster response reveals that socio-institutional processes are important
in bringing together a set of intertwined elements (Tschakert and Dietrich,
2010; see Chapter 8). O’Brien et al. (2011) suggest an adaptation
continuum (see Figure 5-2), where the goal is to move toward partnerships
that enable social transformations and increased resilience.
A key component of the disaster risk management and adaptation
process is the ability to learn (Pahl-Wostl et al., 2007; Armitage et al.,
2008; Lonsdale et al., 2008). This focus on learning partly derives from the
fields of social-ecological resilience and sustainability science (Berkes,
2009; Kristjanson et al., 2009). As scenarios combine quantitative
indicators of climate, demographic, biophysical, and economic change,
as well as qualitative storylines of socio-cultural changes at the local
level, the participation of local stakeholders is essential to generate
values and understandings of climate extremes.
Adaptation is a process rather than an endpoint and requires a focus on
the institutions and policies that enable or hinder this process
(Inderberg and Eikeland, 2009) as well as the acknowledgement that
there are often competing stakeholder goals (Ziervogel and Ericksen,
2010). Fostering better adaptive capacity for disaster and climate risk
will help to accelerate future adaptation (Inderberg and Eikeland, 2009;
Moser, 2009; Patt, 2009). However, there are barriers, including lack of
coordination between actors, the complexity of the policy field
(Mukheibir and Ziervogel, 2007; Winsvold et al., 2009), and limited
human capacity to implement policies (Ziervogel et al., 2010). Lastly,
individual, sector, and institutional perceptions of risk and adaptive
capacity can determine whether adaptation responses are initiated or
not (Grothmann and Patt, 2005).
5.7. Information, Data, and Research Gaps
at the Local Level
The causal processes by which disasters produce systemic effects over
time and across space is reasonably well-known (Kreps, 1985; Cutter,
1996; Lindell and Prater, 2003; NRC, 2006). Yet, local emergency
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
324
management communities have by and large paid little attention to the
links between climate change and natural hazards (Bullock et al., 2009).
As a result, state and local disaster mitigation plans, even when required
by law, usually fail to include climate change, sea level rise, or climate
extreme events in hazard assessments or do so in entirely deterministic
ways. Decisions about development, hazard mitigation, and emergency
preparedness in the context of climate change give rise to critical
questions about social and economic adaptation, and the information
and data to support it, especially at the local scale (Mileti, 1999; Cutter,
2001; Mileti and Peek, 2002). For example: How do cumulative impacts
of smaller events over time compare to single high-impact events for
localities? Do increased levels of hazard mitigation and disaster
preparedness increase local risk-taking by individuals and social systems?
How do short-term adjustments or coping strategies enable or constrain
long-term vulnerabilities in localities? What are the tradeoffs among
decision acceptability versus decision quality, especially within local
contexts (Comfort et al., 1999; Travis, 2010)?
For many of these questions, sufficient empirical information is lacking,
especially at the sub-national scale (see also Section 5.4.2.3). Two recent
all-hazards studies for the United States found that from 1970 to 2004,
climate-sensitive hazards accounted for the majority of recorded fatalities
from natural hazards (Borden and Cutter, 2008; Thacker et al., 2008). Yet,
these are the only databases for monitoring mortality from natural hazards
at the local level and suffer from lack of consistency and completeness.
The hurricane recovery process includes ample evidence of how efforts
to ensure that the rush to return to normal have also led to depletion of
natural resources and increased risk. How decisions regarding the right
to migrate (even temporarily), the right to organize, and the right of
access to information are made will, as a result, have major implications
for the ability of different groups to adapt successfully to floods, droughts,
and storms. The idea of linking place-based recovery, preparedness, and
resilience to adaptation is intuitively appealing. However, the constituency
that supports improved disaster risk management has historically proven
too small to bring about many of the changes that have been
recommended by researchers, especially those that focus on strengthening
the social fabric to decrease vulnerability. Behind the specific questions
of the transparency of risk, are broader questions about the public
sphere. What public goods will be provided by governments at all levels
(and how will they be funded), what public goods will be provided by
private or organizations in civil society, what will be provided by
Chapter 5Managing the Risks from Climate Extremes at the Local Level
Enabling
Participatory
Top-Down
Reliant
Where Many Are
Partnership
Individual
Learning
Reflexive
Social Learning
Institutional Change
Increasing Local Capacity
Paradigm Shift
Learning Processes
Social
Transformations
Where All Should Be
Impacts
Vulnerability to
Adaptation
Adaptation and
Development
Development to
Resilience
Resilience
Figure 5-2 | Learning and transformation. Throughout the adaptation process, learning is expected to increase along with institutional change leading to the potential for
paradigmatic transformation – the community moves away from an impact-focus perspective to a resilience-centric one where there is an expectation of risk and where good
governance and key partnerships are the norm. Source: adapted from O’Brien et al., 2011.
325
market actors, and what will not? How will these influence local-level
disaster risk management, especially of climate-sensitive hazards
(Mitchell, 1988, 1999; Thomalla et al., 2006; van Aalst et al., 2008)?
While there has been increasing focus on the processes by which
knowledge has been produced, less time has been spent examining the
capacity of local communities to critically assess knowledge claims
made by others for their reliability and relevance to those communities
(Fischhoff, 2007; Pulwarty, 2007). There is the need to move beyond the
integration of physical and societal impacts to focus on practice and
evaluation. How are impediments to the flow of information created? Is
a focus on communication adequate to ensure effective response? How
are these nodes defined among differentially vulnerable groups, for
example, based on economic class, race, or gender? However, there is
little research on the extent to which local jurisdictions have adopted
policy options and practice and the ways in which they are being
implemented. Most of the studies to date have addressed factors that
lead to policy adoption and not necessarily successful implementation.
Beyond infrastructure and retrofitting concerns, successful adaptation
strategies integrate urban planning, water management, early warning
systems, and preparedness. One widely acknowledged goal is to address,
directly, the problem of an inadequate fit between what the research
community knows about the physical and social dimensions of uncertain
environmental hazards and what society chooses to do with that
knowledge. An even larger challenge is to consider how different
systems of knowledge about the physical environment and competing
systems of action can be brought together in pursuit of diverse goals
that we wish to pursue (Mitchell, 2003). Several sources (Comfort et al.,
1999; Bullock et al., 2009; McKinsey Group, 2009) have identified key
research and data requirements for addressing these challenges:
1) Multi-way information exchange systems – effective adaptation will
always be locally driven. Communities need reliable measurements
and assessment tools, integrated information about risks that those
tools reveal, and best approaches to minimize those risks. The goal is
to improve the assessment and transparency of risk in a geographic
place-based approach for vulnerable regions. Improving the collection
and quality control of locally based data on economic losses, disaster
and adaptation costs, and human losses (fatalities) will ensure
improved empirically based baseline assessments.
2) Maps of the decision processes for disaster mitigation, preparedness,
response, and recovery and guidance for using such decision support
tools are needed. Hazard maps developed through collaboration
between researchers and affected communities are the simplest
and often most powerful form of risk information. They capture the
likelihood and impact of a peril and are important for informing
many aspects of disaster risk management including disaster risk
reduction, risk-based pooling of resources, and risk transfer. Such
devices would identify segments of threatened social systems that
could suffer disproportionate disaster impacts; critical actors at each
jurisdictional level; their risk assumptions; their different types of
information needs; and the design of an information infrastructure
that would support decisions at critical entry points (Comfort, 1993).
3) People who face hazards often need assistance to manage their
own environments over the long term and develop systematic
actions to improve resilience in vulnerable localities. Research is
needed on how local governments and institutions can support,
provide incentives, and legitimize successful approaches to
increasing capacity and action.
4) Methodologies, indicators, and measurement of progress in reducing
vulnerability and enhancing community capacity at the local level
are under-researched at present. Locally based risk management,
cost-effectiveness methodologies and analyses, quantification of
societal impacts of catastrophic events at local to national scales,
and research on implementation and evaluation of risk management
and mitigation programs are needed. Similarly, there is a critical
need for the assessment and coordination of multi-jurisdictional
and multi-sectoral efforts to help avoid the unintended consequences
of actions and interventions especially at the local scale.
5) Underserved people require access to the social and economic
security that comes from sharing risk, through financial risk transfer
mechanisms such as insurance. There is a paucity of studies at the
local level to assess the efficacy of alternative risk reduction, risk-
based resource pooling and transfer methods, analysis of benefits
and costs to various stakeholder groups, analysis of complementary
roles of mitigation and insurance, and analysis of safeguards
against insurance industry insolvency.
Interdisciplinary collaboration is clearly needed to prioritize and address
these research needs. Situating the scientific understanding of hazards,
disaster risk, and climate change adaptation within a broader discourse
about different forms of knowledge will increase the likelihood of public
actions that are better grounded in scientific knowledge and customized
for the local context.
5.8. Summary
This chapter presented evidence on how climate extremes affect local
places; how local places currently cope with disasters such as emergency
assistance and disaster relief; and how they anticipate and plan for future
disaster risk using improved communication, structures such as dams and
levees, land use management and ecosystem protection, and storage of
resources. The role of scale and context shapes variability in building
adaptive capacity at the local level. Differences in coping and risk
management also are scale-dependent and context-specific, and could
affect or limit adaptations to climate extremes at the local level. Lastly,
climate extremes threaten human security at the local level. Localized
vulnerability attributed to social, economic, environmental, and climate
change drivers at a variety of scales heightens the impacts of climate
extremes on local places. While some places have considerable experience
with disasters and some inherent capacity to cope with climate extremes,
others do not. These differences in coping and management necessitate
a range of approaches for disaster risk management and climate change
adaptation, thus attention to a broader set of national and international
contexts relating to social welfare, quality of life, and sustainable livelihoods.
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
326
References
A digital library of non-journal-based literature cited in this chapter that
may not be readily available to the public has been compiled as part of
the IPCC review and drafting process, and can be accessed via either the
IPCC Secretariat or IPCC Working Group II web sites.
Aall, C., K. Groven, and G. Lindseth, 2007: The scope of action for local climate policy:
The case of Norway. Global Environmental Politics, 7(2), 83-101.
ActionAid, 2005: People-Centered Governance: Reducing Disaster for Poor and
Excluded People. ActionAid International, Johannesburg, South Africa.
Adeola, F.O., 2009: Katrina cataclysm: Does duration of residency and prior
experience affect impacts, evacuation, and adaptation behavior among
survivors? Environment and Behavior, 41(4), 459-489.
Adger, W.N., 2000: Social and ecological resilience: Are they related? Progress in
Human Geography, 24(3), 347-364.
Adger, W.N., 2003: Social capital, collective action, and adaptation to climate
change. Economic Geography, 79(4), 387-404.
Adger, W.N. and K. Brown, 2009: Adaptation, vulnerability and resilience: Ecological
and social perspectives. In: Companion to Environmental Geography [Castree,
N., D. Demeritt, D. Liverman, and B. Rhoads (eds.)]. Wiley-Blackwell, Chichester,
UK, pp. 109-122.
Adger, W.N. and P.M. Kelly, 1999: Social vulnerability to climate change and the
architecture of entitlements. Mitigation and Adaptation Strategies for Global
Change, 4(3-4), 253-266.
Adger, W.N., P.M. Kelly, A. Winkels, L.Q. Huy, and C. Locke, 2002: Migration, remittances,
livelihood trajectories, and social resilience. Ambio, 31(4), 358-366.
Adger, W.N., K. Brown, J. Fairbrass, A. Jordan, J. Paavola, S. Rosendo, and G. Seyfang,
2003: Governance for sustainability: Towards a ‘thick’ analysis of environmental
decision-making. Environment and Planning A, 35, 1095-1110.
Adger, W.N., N.W. Arnell, and E.L. Tompkins, 2005: Successful adaptation to climate
change across scales. Global Environmental Change, 15(1), 77-86.
Adger, W.N., J. Paavola, S. Huq, and M.J. Mace (eds.), 2006: Fairness in Adaptation to
Climate Change. MIT Press, Cambridge, MA.
Adger, W.N., S. Agrawala, M.M.Q. Mirza, C. Conde, K. O’Brien, J. Pulhin, R. Pulwarty,
B. Smit, and K. Takahashi, 2007: Assessment of adaptation practices, options,
constraints and capacity. In: Climate Change 2007. Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani,
J.P. Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 717-743.
Adger, W.N., S. Dessai, M. Goulden, M. Hulme, I. Lorenzoni, D.R. Nelson, L.O. Naess,
J. Wolf, and A. Wreford, 2010: Are there social limits to adaptation to climate
change? Climatic Change, 93(3-4), 335-354.
Aerts, J., D. Major, M. Bowman, and P. Dircke, 2009: Connecting Delta Cities. Coastal
Cities, Flood Risk Management and Adaptation to Climate. VU University Press,
Amsterdam, The Netherlands, 90 pp.
Agrawal, A., 2010: Local institutions and adaptation to climate change. In: Social
Dimensions of Climate Change: Equity and Vulnerability in a Warming World
[Mearns, R. and A. Norton (eds.)]. World Bank, Washington, DC, pp. 173-198.
Agrawal, A., C. McSweeney, and N. Perrin, 2008: Social Development Notes. Local
Institutions and Climate Change Adaptation - the Social Dimensions of Climate
Change. World Bank, Washington, DC.
Alcamo, J., L. Acosta-Michlik, A. Carius, F. Eierdanz, R. Klein, D. Krömker, and D.
Tänzler, 2008: A new approach to quantifying and comparing vulnerability to
drought. Regional Environmental Change, 8(4), 137-149.
Alexander, D., 2002: Principles of Emergency Planning and Management. Oxford
University Press, Oxford, UK.
Allen, K., 2006: Community-based disaster preparedness and climate adaptation:
Local capacity-building in the Philippines. Disasters, 30(1), 81-101.
Amin, S., A.S. Rai, and G. Topa, 2001: Does Microcredit Reach and Poor and
Vulnerable? Evidence from Northern Bangladesh. Center for International
Development, Harvard University, Cambridge, MA.
Anderson, S., J. Morton, and C. Toulmin, 2010: Climate change for agrarian societies
in drylands: Implications and future pathways. In: Social Dimensions of Climate
Change: Equity and Vulnerability in a Warming World [Mearns, R. and A. Norton
(eds.)]. World Bank, Washington, DC, pp. 199-230.
Anning, D. and D. Dominey-Howes, 2009: Valuing climate change impacts on Sydney
beaches to inform coastal management decisions: A research outline.
Management of Environmental Quality, 20(4), 408-421.
ANTHC, 2011: Climate Change & Health Impact Assessment Reports. Alaska Native
Tribal Health Consortium, www.anthctoday.org/community/reports.html.
Aragon-Durand, F., 2007: Urbanization and flood vulnerability in the peri-urban
interface of Mexico City. Disasters, 31(4), 477-494.
Ariyabandu, M.M., 2006: Gender issues in recovery from the December 2004 Indian
ocean tsunami: The case of Sri Lanka. Earthquake Spectra, 22(3), 759-775.
Armitage, D., M. Marschke, and R. Plummer, 2008: Adaptive co-management and
the paradox of learning. Global Environmental Change, 18, 86-98.
Ascher, W., 1978: Forecasting: An appraisal for policy-makers and planners. Johns
Hopkins Press, Baltimore, MD, 256 pp.
Asian Disaster Preparedness Centre, 2005: A Primer. Integrated Flood Risk
Management in South East Asia. Asian Disaster Preparedness Centre, Bangkok,
Thailand.
Babugura, A.A., 2008: Vulnerability of children and youth in drought disasters: A
case study of Botswana. Children, Youth and Environments, 18(1), 126-157.
Balaban, V., 2006: Psychological assessment of children in disasters and emergencies.
Disasters, 30, 178-198.
Bamdad, N., 2005: The Role of Community Knowledge in Disaster Management: The
Bam Earthquake Lesson in Iran. Institute of Management and Planning Studies,
Tehran, Iran.
Bankoff, B., 2007: Living with risk: Coping with disasters. Education about Asia,
12(2), 26-29.
Bankoff, G., G. Frerks, and D. Hilhorst (eds.), 2004: Mapping vulnerability: Disasters,
Development & People. Earthscan, London, UK.
Barbier, E.B., E.W. Koch, B.R. Silliman, S.D. Hacker, E. Wolanski, J. Primavera, E.F.
Granek, S. Polasky, S. Aswani, L.A. Cramer, D.M. Stoms, C.J. Kennedy, D. Bael,
C.V. Kappel, G.M.E. Perillo, and D.J. Reed, 2008: Coastal ecosystem-based
management with nonlinear ecological functions and values. Science, 319,
321-323.
Bari, F., 1998: Gender, disaster and empowerment: A case study from Pakistan. In: The
Gendered Terrain of Disaster: Through Women’s Eyes [Enarson, E.P. and B.H.
Morrow (eds.)]. Praeger, Westport, CT, pp. 125-132.
Barnett, J., 2003: Security and climate change. Global Environmental Change, 13(1),
7-17.
Barnett, J. and W.N. Adger, 2007: Climate change, human security and violent
conflict. Political Geography, 26, 639-655.
Barnett, J. and M. Webber, 2009: Accommodating Migration to Promote Adaptation
to Climate Change. Commission on Climate Change and Development,
Stockholm, Sweden.
Baro, M. and T. Deubel, 2006: Persistent hunger: Perspectives on vulnerability,
famine, and food security in sub-Saharan Africa. Annual Review of
Anthropology, 35, 521-538.
Barrett, C.B., 2002: Food security and food assistance programs. In: Handbook of
Agricultural Economics [Gardner, B. and G. Rausser (eds.)]. Elsevier Science,
Amsterdam, The Netherlands.
Barrett, E.J., M. Martinez-Cosio, and C.Y.B. Ausbrooks, 2008: The school as a source
of support for Katrina-evacuated youth. Children, Youth and Environments,
18(1), 202-236.
Bartlett, S., 2008: The implications of climate change for children in lower-income
countries. Children, Youth and Environments, 18(1), 71-98.
Bateman, J.M. and B. Edwards, 2002: Gender and evacuation: A closer look at why
women are more likely to evacuate for hurricanes. Natural Hazards Review,
3(3), 107-117.
Chapter 5Managing the Risks from Climate Extremes at the Local Level
327
Battista, F. and S. Baas, 2004: Consolidated Report on Case Studies and Workshop
Findings. The Role of Local Institutions in Reducing Vulnerability to Recurrent
Natural Disasters and in Sustainable Livelihoods Development. Rural Institutions
and Participation Service, Food and Agriculture Organization, Rome, Italy.
Baumhardt, F., R. Lasage, P. Suarez, and C. Chadza, 2009: Farmers become filmmakers:
Climate change adaptation in Malawi. Participatory Learning and Action, 60,
129-138.
Beamon, B.M. and B. Balcik, 2008: Performance measurement in humanitarian supply
chains. International Journal of Public Sector Management, 21(1), 4-25.
Berke, P.R. and T. Beatley, 1992: A national assessment of earthquake mitigation:
Implications for land use planning and public policy. Earthquake Spectra, 8, 1-17.
Berkes, F., 2008: Sacred Ecology. 2nd ed. Routledge, New York, NY.
Berkes, F., 2009: Evolution of co-management: Role of knowledge generation, bridging
organizations and social learning. Journal of Environmental Management,
90(5), 1692-1702.
Berkes, F. and D. Jolly, 2001: Adapting to climate change: Social-ecological resilience
in a Canadian western Arctic community. Ecology and Society, 5(2), 18.
Berkhout, F., 2002: Technological regimes, path dependency and the environment.
Global Environmental Change, Part A: Human and Policy Dimensions, 12(1), 1-4.
Berkhout, F., J. Hertin, and D. Gann, 2006: Learning to adapt: Organisational
adaptation to climate change impacts. Climatic Change, 78(1), 135-156.
Bermejo, P.M., 2006: Preparation and response in case of natural disasters: Cuban
programs and experience. Journal of Public Health Policy, 27(1), 13-21.
Bernard, S.M., J.M. Samet, A. Grambsch, K.L. Ebi, and I. Romieu, 2001: The potential
impacts of climate variability and change on air pollution-related health effects
in the United States. Environmental Health Perspectives, 109, 199-209.
Berube, A. and B. Katz, 2005: Katrina’s Window: Confronting Concentrated Poverty
Across America. The Brookings Institution Metropolitan Policy Program,
Washington, DC.
Birkland, T.A., 1997: After disaster: Agenda Setting, Public Policy, and Focusing
Events. Georgetown University Press, Washington, DC.
Birkmann, J., 2006: Measuring Vulnerability to Natural Hazards: Towards Disaster
Resilient Societies. United Nations Publications, New York, NY.
Birkmann, J., M.A. Garschagen, F. Krass, and N. Quang, 2010: Adaptive urban
governance–new challenges for the second generation of urban adaptation
strategies to climate change. Sustainability Science, 5(2), 185-206.
Bizikova, L., T. Neale, and I. Burton, 2008:
Canadian Communities’ Guidebook for
Adaptation to Climate Change. Including an Approach to Generate Mitigation
Co-Benefits in the Context of Sustainable Development. 1st ed. Environment
Canada and University of British Columbia, Vancouver, Canada, 100 pp.
Bohle, H.-G., 2001: Vulnerability and criticality: Perspectives from social geography.
Newsletter of the International Human Dimensions Programme on Global
Environmental Change, 2/2001.
Bohle, H., T. Downing, and M. Watts, 1994: Climate change and social vulnerability:
Toward a sociology and geography of food insecurity. Global Environmental
Change, 4(1), 371-396.
Borden, K.A. and S.L. Cutter, 2008: Spatial patterns of natural hazards mortality in
the United States. International Journal of Health Geographics, 7(64).
Boruff, B.J. and S.L. Cutter, 2007: The environmental vulnerability of Caribbean island
nations. Geographical Review, 97(1), 24-45.
Botzen, W.J.W. and J.C.J.M. van den Bergh, 2008: The insurance against climate
change and flooding in The Netherlands: Present, future, and comparison with
other countries. Risk Analysis, 28(2), 413-426.
Botzen, W., J. Aerts, and J. van den Bergh, 2009: Willingness of homeowners to
mitigate climate risk through insurance. Ecological Economics, 68, 2265-2277.
Bourque, L., J. Siegel, M. Kano, and M. Wood, 2006: Weathering the storm: The
impact of hurricanes on physical and mental health. The Annals of the American
Academy of Political and Social Science, 604, 129-151.
Bowden, M., R.W. Kates, P.A. Kay, W.E. Riebsame, R.A. Warrick, D.L. Johnson, H.A. Gould,
and D. Weiner, 1981: The effect of climatic fluctuations on human populations:
Two hypotheses. In: Climate and History: Studies in Past Climates and their
Impact on Man [Wigley, T.M., M.J. Ingrams, and G. Farmer (eds.)]. Cambridge
University Press, Cambridge, UK, pp. 479-513.
Boyd, R. and M.E. Ibarrarán, 2009: Extreme climate events and adaptation: An
exploratory analysis of drought in Mexico. Environment and Development
Economics, 14(3), 371-396.
Brauer, M., 1999: Health impacts of biomass air pollution. In: Health Guidelines for
Vegetation Fire Events [Goh, K.-T., D.H. Schwela, J.G. Goldammer, and O.
Simpson (eds.)]. World Health Organization, Singapore.
Brenkert, A.L. and E.L. Malone, 2005: Modeling vulnerability and resilience to climate
change: A case study of India and Indian states. Climatic Change, 72(1-2), 57-102.
Brondizio, E., E. Ostrom, and O. Young, 2009: Connectivity and the governance of
multilevel social-ecological systems: The role of social capital. Annual Review of
Environment and Resources, 34, 253-278.
Brooks, N., 2003: Vulnerability, Risk and Adaptation: A Conceptual Framework. Tyndall
Centre for Climate Change Research, University of East Anglia, Norwich, UK.
Brown, D.L., 1999: Disparate effects of the 1989 Loma Prieta and 1994 Northridge
earthquakes on hospital admissions for acute myocardial infarction: Importance
of superimposition of triggers. American Heart Journal,
7(5), 830-836.
Brown, O., 2008: Migration and Climate Change. International Organization for
Migration, Geneva, Switzerland.
Brulle, R., 2010: From environmental campaigns to advancing the public dialogues:
Environmental communication for civic engagement. Environmental
Communication, 4(1), 82-98.
Buckland, J. and M. Rahman, 1999: Community-based disaster management during
the 1997 Red River flood in Canada. Disasters, 23(3), 174-191.
Buhaug, H., 2010: Climate not to blame for African civil wars. Proceedings of the
National Academy of Sciences, 107(38), 16477-16482.
Bulkeley, H., 2006: A changing climate for spatial planning? Planning Theory and
Practice, 7(2), 203-214.
Bull-Kamanga, L., K. Diagne, A. Lavell, E. Leon, F. Lerise, H. MacGregor, A. Maskrey,
M. Meshack, M. Pelling, H. Reid, D. Satterthwaite, J. Songsore, K. Westgate, and
A. Yitambe, 2003: From everyday hazards to disasters: The accumulation of risk
in urban areas. Environment and Urbanization, 15(1), 193-203.
Bullock, J., G. Haddow, and K. Haddow, 2009: Global Warming, Natural Hazards, and
Emergency Management. CRC Press, Boca Raton, FL, 282 pp.
Burby, R.J. (ed.), 1998: Cooperating with Nature: Confronting Natural Hazards with
Land-Use Planning for Sustainable Communities. Joseph Henry Press/National
Academy of Sciences, Washington, DC, 356 pp.
Burby, R.J., 2000: Creating hazard resilient communities through land-use planning.
Natural Hazards Review, 1(2), 99-106.
Burby, R.J., 2006: Hurricane Katrina and the paradoxes of government disaster
policy: Bringing about wise government decisions for hazardous areas. Annals
of the American Academy of Political and Social Science, 604(1), 171-191.
Burby, R.J., May, P.J. with Berke, P. R., L.C. Dalton, S.P. French, and E.J. Kaiser, 1997:
Making Government Plan: State Experiments in Managing Land Use. Johns
Hopkins University Press, Baltimore, MD.
Burch, S. and J. Robinson, 2007: A framework for explaining the links between
capacity and action in response to global climate change. Climate Policy, 7(4),
304-316.
Burke, M.B., E. Miguel, S. Satyanath, J.A. Dykema, and D.B. Lobell, 2009: Warming
increases the risk of civil war in Africa. Proceedings of the National Academy of
Sciences, 106(49), 20670-20674.
Burton, C. and S.L. Cutter, 2008: Levee failures and social vulnerability in the
Sacramento–San Joaquin delta area, California. Natural Hazards Review, 9(3),
136-149.
Burton, I., R.W. Kates, and G.F. White, 1993:
The Environment as Hazard. 2nd ed.
Guilford Press, New York, NY.
Burton, I., J. Soussan, and A. Hammill, 2003: Livelihoods and Climate Change,
Combining Disaster Risk Reduction, Natural Resource Management and
Climate Change Adaptation in a New Approach to the Reduction of
Vulnerability and Poverty. A Conceptual Framework Paper Prepared by the Task
Force on Climate Change, Vulnerable Communities and Adaptation,
International Institute for Sustainable Development, Winnipeg, Canada.
Burton, I., L. Bizikova, T. Dickinson, and Y. Howard, 2007: Integrating adaptation into
policy: Upscaling evidence from local to global. Climate Policy, 7(4), 371-376.
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
328
Buvinić, M., G. Vega, M. Bertrand, A. Urban, R. Grynspan, and G. Truitt, 1999:
Hurricane Mitch: Women’s Needs and Contributions. Sustainable Development
Department Technical Papers Series, Inter-American Development Bank,
Washington, DC.
Byg, A. and J. Salick, 2009: Local perspectives on a global phenomenon – climate
change in eastern Tibetian villages. Global Environmental Change, 19, 156-166.
Bynander, F., L. Newlove, and B. Ramberg, 2005: Sida and the Tsunami of 2004 – A
Study of Organizational Crisis Response. Swedish International Development
Cooperation Agency, Stockholm, Sweden.
Cabrera, V.E., N.E. Breuer, and P.E. Hildebrand, 2008: Participatory modeling in dairy
farm systems: A method for building consensual environmental sustainability
using seasonal climate forecasts. Climatic Change, 89, 395-409.
Cahill, K., 2007: The Pulse of Humanitarian Assistance. Fordham University Press,
New York, NY.
Campbell, J.R., 1990: Disasters and development in historical context: Tropical
cyclone response in the Banks Islands of northern Vanuatu. International
Journal of Mass Emergencies and Disasters, 8(3), 401-424.
Campbell, J.R., 2006: Traditional Disaster Reduction in Pacific Island Communities.
Institute of Geological and Nuclear Sciences Ltd., Avalon, New Zealand.
Campbell, J.R., 2010a: An overview of natural hazard planning in the Pacific island
region. The Australasian Journal of Disaster and Trauma Studies, 2010(1).
Campbell, J.R., 2010b: Climate-induced community relocation in the Pacific: The
meaning and importance of land. In: Climate Change and Displacement:
Multidisciplinary Perspectives [McAdam, J. (ed.)]. Hart Publishing, Oxford, UK,
pp. 57-79.
Campbell-Lendrum, D., A. Pruss-Ustun, and C. Corvalan, 2003: How much disease
could climate change cause? In: Climate Change and Health: Risks and Responses
[Campbell-Lendrum, D., C. Corvalan, K. Ebi, A. Githeko, and J. Scheraga (eds.)].
World Health Organization, Geneva, Switzerland, pp. 133-158.
Cardona, O.D., 2007: Indicators of Disaster Risk and Risk Management: Program for
Latin America and the Caribbean: Summary Report. Inter-American
Development Bank, Washington, DC.
Carney, E. (ed.), 1998: Sustainable Rural Livelihoods: What Contributions Can We
Make? Department for International Development, London, UK.
Carraro, L., S. Khan, S. Hunt, G. Rawle, M. Robinson, M. Antoninis, and L. Street,
2003: Monitoring the Millennium Development Goals: Current Weaknesses and
Possible Improvements. Oxford Policy Management, Department for
International Development, Glasgow, Scotland.
Carter, M.R., P.D. Little, T. Mogues, and W. Negatu, 2007: Poverty traps and natural
disasters in Ethiopia and Honduras. World Development, 35(5), 835-856.
Cash, D.W. and S.C. Moser, 2000: Linking global and local scales: Designing dynamic
assessment and management processes. Global Environmental Change, 10(2),
109-120.
Chambers, R. and G. Conway, 1992: Sustainable Rural Livelihoods: Practical
Concepts for the 21st Century. Institute of Development Studies, Brighton, UK.
Chantarat
, S., C.B. Barrett, A.G. Mude, and C.G. Turvey, 2007: Using weather index
insurance to improve drought response for famine prevention. American
Journal of Agricultural Economics, 89(5), 1262-1268.
Cheong, S., 2010: Initial responses to the MT Hebei-Spirit oil spill. Marine Technology
Society Journal, 44(1), 69-74.
Cheong, S., 2011a: Guest editorial: Coastal adaptation. Climatic Change, 106(1),
1-4.
Cheong, S., 2011b: Policy solutions in the U.S. Climatic Change, 106(1), 57-70.
Cherry, K.E. (ed.), 2009: Lifespan Perspectives on Natural Disasters: Coping with
Katrina, Rita, and Other Storms. Springer, New York, NY.
Christoplos, I., 2006: The Elusive “Window of Opportunity” For Risk Reduction in
Post-Disaster Recovery. In: ProVention Consortium Forum 2006, Bangkok,
Thailand, 2-3 Feb 2006.
Churchill, C. (ed.), 2006: Protecting the Poor: A Microinsurance Compendium.
International Labor Organization, Geneva, Switzerland.
Churchill, C., 2007: Insuring the low-income market: Challenges and solutions for
commercial insurers. Geneva Papers on Risk and Insurance–Issues and Practice,
32(3), 401-412.
Clark, G., S. Moser, S. Ratick, K. Dow, W. Meyer, S. Emani, W. Jin, J. Kasperson, R.
Kasperson, and H. Schwartz, 1998: Assessing the vulnerability of coastal
communities to extreme storms: The case of Revere, MA, USA. Mitigation and
Adaptation Strategies for Global Change, 3(1), 59-82.
Clifton, D. and F. Gell, 2001: Saving and protecting lives by empowering women.
Gender and Development, 9(1), 8-18.
Clot, N. and J. Carter, 2009: Disaster Risk Reduction: A Gender and Livelihood Perspective.
Swiss Agency for Cooperation and Development, Berne, Switzerland, 13 pp.
Comfort, L.K., 1993: Integrating information technology into international crisis
management and policy. Journal of Contingencies and Crisis Management, 1, 15-26.
Comfort, L., B. Wisner, S. Cutter, R. Pulwarty, K. Hewitt, A. Oliver-Smith, J. Wiener, M.
Fordham, W. Peacock, and F. Krimgold, 1999: Reframing disaster policy: The
global evolution of vulnerable communities. Global Environmental Change Part
B: Environmental Hazards, 1, 39-44.
Cooper, J.A.G. and J. McKenna, 2008: Social justice in coastal erosion management:
The temporal and spatial dimensions. Geoforum, 39, 294-306.
Corbett, J., 1988: Famine and household coping strategies. World Development,
16(9), 1099-1112.
Cova, T.J., F.A. Drews, L.K. Siebeneck, and A. Musters, 2009: Protective actions in wildfires:
Evacuate or shelter in place? Natural Hazards Review
, 10(4), 151-162.
Cox, J.R., C. Rosenzweig, W.D. Solecki, R. Goldberg, and P.L. Kinney, 2007: Social
Vulnerability to Climate Change: A Neighborhood Analysis of the Northeast U.S.
Megaregion. Union of Concerned Scientists, Cambridge, MA.
Crabbé, P. and M. Robin, 2006: Institutional adaptation of water resource infrastructures
to climate change in eastern Ontario. Climatic Change, 78(1), 103-133.
Cuny, F., 1983: Disasters and Development. Oxford University Press, New York, NY.
Cupples, J., 2007: Gender and hurricane Mitch: Reconstructing subjectivities after
disaster. Disasters, 21(2), 155-175.
Curtis, A., J.W. Mills, and M. Leitner, 2007: Katrina and vulnerability: The geography
of stress. Journal of Health Care for the Poor and Underserved, 18(2), 315-330.
Curtis, A., A. Duval-Diop, and J. Novak, 2010: Identifying spatial patterns of recovery
and abandonment in the post-Katrina holy cross neighborhood of New Orleans.
Cartography and Geographic Information Science, 37(1), 45-56.
Cutter, S.L., 1996: Vulnerability to environmental hazards. Progress in Human
Geography, 20, 529-539.
Cutter, S.L. (ed.), 2001: American Hazardscapes: The Regionalization of Hazards and
Disasters. Joseph Henry Press/National Academies of Science, Washington, DC.
Cutter, S.L. and M.M. Smith, 2009: Fleeing from the hurricane’s wrath: Evacuation
and the two Americas. Environment, 51(2), 26-36.
Cutter, S.L., J.T. Mitchell, and M.S. Scott, 2000: Revealing the vulnerability of people
and places: A case study of Georgetown County, South Carolina. Annals of the
Association of American Geographers, 90(4), 713-737.
Cutter, S.L., B.J. Boruff, and W.L. Shirley, 2003: Social vulnerability to environmental
hazards. Social Science Quarterly, 84(1), 242-261.
Dale, A. and L. Newman, 2010: Social capital: A necessary and sufficient condition
for sustainable community development? Community Development Journal,
45(1), 5-21.
Darcy, J. and C. Hofmann, 2003: Humanitarian Needs Assessment and Decision-
Making. HPG Report, Overseas Development Institute (ODI), London, UK, 1-74 pp.
Das, S. and J. Vincent, 2009: Mangroves protected villages and reduced death toll
during Indian super cyclone. Proceedings of the National Academy of Sciences,
106(18), 7357-7360.
Dash, N. and H. Gladwin, 2007: Evacuation decision making and behavioral responses:
Individual and household. Natural Hazards Review, 8(3), 69-77.
Davidson
, C.H., C. Johnson, L. Gonzalo, N. Dikmen, and A. Sliwinski, 2007: Truths and
myths about community participation in post-disaster housing projects. Habitat
International, 31(1), 100-115.
Davies, M., 2009: DFID Social Transfers Evaluation Summary Report. Department for
International Development, London, UK.
Davies, M. and J. Leavy, 2007: Connecting social protection and climate change
adaptation. IDS in Focus, 2.
De Sherbinin, A., A. Schiller, and A. Pulsipher, 2007: The vulnerability of global cities
to climate hazards. Environment and Urbanization, 19(1), 39-64.
Chapter 5Managing the Risks from Climate Extremes at the Local Level
329
De Ville de Groyet, C., 2000: Stop propagating disaster myths. Lancet, 356(9231),
762-764.
Dekens, J., 2007: Local Knowledge for Disaster Preparedness: A Literature Review.
Hillside Press, Kathmandu, Nepal, 85 pp.
DeLucena, A.F.P., A.S. Szklo, R. Schaeffer, R.R. de Souza, B. S. M. C. Borba, I. V. L. da
Costa, A.O.P. Júnior, and S. H. F. da Cunha, 2009: The vulnerability of renewable
energy to climate change in Brazil. Energy Policy, 37(3), 879-888.
Depradine, C.A. and E.H. Lovell, 2004: Climatological variables and the incidence of
dengue fever in Barbados. International Journal of Dengue Fever in Barbados,
14, 429-441.
Dershem, L. and D. Gzirishvili, 1999: Informal social support networks and household
vulnerability: Empirical findings from Georgia. World Development, 26(10),
1827-1838.
DeSalvo, K., D. Hyre, D. Ompad, A. Menke, L. Tynes, and P. Muntner, 2007: Symptoms
of posttraumatic stress disorder in a New Orleans workforce following
Hurricane Katrina. Journal of Urban Health: Bulletin of the New York Academy
of Medicine, 84(2), 142-152.
Development Initiatives, 2009: Global Humanitarian Assistance. Development
Initiatives, Wells, Somerset, UK.
Devereux, S. and S. Coll-Black, 2007: DFID Social Transfers Evaluation. Review of
Evidence and Evidence Gaps on the Effectiveness and Impacts of DFID-
Supported Pilot Social Transfer Schemes. Institute of Development Studies,
Brighton, UK.
Devereux, S. and R. Sabates-Wheeler, 2004: Transformative Social Protection.
Institute of Development Studies, Brighton, UK.
DFID, 2006: Eliminating World Poverty: Making Governance Work for the Poor: A
White Paper on International Development. Department for International
Development, London, UK.
Díaz, J., C. López, J.C. Alberdi, A. Jordán, R. García, E. Hernández, and A. Otero, 2002:
Heat waves in Madrid 1986-1997: Effects on the health of the elderly.
International Archives of Occupational & Environmental Health, 75(163-170).
diFalco, S. and E. Bulte, 2009: Social Capital and Weather Shocks in Ethiopia: Climate
Change and Culturally-Induced Poverty Traps. London School of Economics,
London, UK.
Dixit, A., A. Pokhrel, and M. Moench, 2008: From Risk to Resilience: Costs and
Benefits of Flood Mitigation in the Lower Bagmati Basin: Case of Nepal Terai
and North Bihar. ISET-Nepal and ProVention, Kathmandu, Nepal.
Dolan, A.H. and I.J. Walker, 2004: Understanding vulnerability of coastal communities
to climate change related risks. In: Proceedings of the 8th International Coastal
Symposium, Itajai, Brazil, 2004. Journal of Coastal Research, S39.
Dorosh, P., S. Malik, and M. Krausova, 2010: Rehabilitating Agriculture and
Promoting Food Security Following the 2010 Pakistan Floods Insights from
South Asian Experience. IFPRI Discussion Paper, World Bank/International Food
Policy Research Institute, Washington, DC, 26 pp.
Douglas, I., K. Alam, M. Maghenda, Y. Mcdonnell, L. Mclean, and J. Campbell, 2009:
Unjust waters: Climate change, flooding and the urban poor in Africa. In:
Adapting Cities to Climate Change [Bicknell, J., D. Dodman, and D. Satterthwaite
(eds.)]. Earthscan, London, UK, pp. 201-224.
Dow, K. and S.L. Cutter, 1998: Crying wolf: Repeat responses to hurricane evacuation
orders. Coastal Management, 26(4), 237-252.
Dow, K. and S.L. Cutter, 2000: Public orders and personal opinions: Household strategies
for hurricane risk assessment. Environmental Hazards, 2(4), 143-155.
Dow, K. and S.L. Cutter, 2002: Emerging hurricane evacuation issues: Hurricane Floyd
and South Carolina. Natural Hazards Review, 3(1), 12-18.
Downton, M.W. and R.A. Pielke Jr., 2005: How accurate are disaster loss data? The
case of U.S. flood damage. Natural Hazards, 35, 211-228.
Doyle, M.W., E.H. Stanley, D.G. Havlick, M.J. Kaiser, G. Steinbach, W.L. Graf, G.E.
Galloway, and J.A. Riggsbee, 2008: Aging infrastructure and ecosystem
restoration. Science, 319(5861), 286-287.
Drabek, T.E., 1986: Human System Responses to Disaster: An Inventory of
Sociological Findings. Springer Verlag, New York, NY.
Drabek, T.E., 1999: Understanding disaster warning responses. Social Science
Journal, 36(3), 515-523.
Dube, O.P. and S. Sekhwela, 2008: Indigenous knowledge, institutions and practices
for coping with variable climate in the Limpopo Basin of Botswana. In: Climate
change and adaptation [Leary, N., J. Adejuwon, and V. Barros (eds.)]. Earthscan,
Sterling, UK, pp. 71-89.
Dulal, H., G. Brodnig, G. Onoriose, and H. Thakur, 2010: Capitalising on Assets:
Vulnerability and Adaptation to Climate Change in Nepal. World Bank,
Washington, DC.
Dynes, R., 2006: Social capital: Dealing with community emergencies. Homeland
Security Affairs, 2(2), 1-26.
Eakin, H. and L.A. Bojorquez-Tapia, 2008: Insights into the composition of household
vulnerability from multicriteria decision analysis. Global Environmental Change,
18(1), 112-127.
Ebi, K.L., 2008: Community-based adaptation to the health impacts of climate
change. American Journal of Preventive Medicine, 35(5), 501-507.
Edmonds, A. and S.L. Cutter, 2008: Planning for pet evacuations during disasters.
Journal of Homeland Security and Emergency Management, 5(1), 33.
Eisenman, D.P., K.M. Cordasco, S. Asch, J.F. Golden, and D. Glik, 2007: Disaster planning
and risk communication with vulnerable communities: Lessons from hurricane
Katrina. American Journal of Public Health, 97(S1), S109-S115.
Eisensee, T. and D. Stromberg, 2007: News droughts, news floods, and U.S. disaster
relief. The Quarterly Journal of Economics,
122(2), 693-728.
Elliott, J.R. and J. Pais, 2006: Race, class, and Hurricane Katrina: Social differences in
human responses to disaster. Social Science Research, 35, 295-321.
Ellis, F., 2000: The determinants of rural livelihood diversification in developing
countries. Journal of Agricultural Economics, 51(2), 289-302.
Ellis, F. and E. Allison, 2004: Livelihood Diversification and Natural Resource Access.
Food and Agriculture Organization, Rome, Italy.
Enarson, E., 2000: We will make meaning out this: Women’s cultural responses to
the Red River valley flood. International Journal of Mass Emergencies and
Disasters, 18(1), 39-64.
Enarson, E., 2009: Gender Mainstreaming in Emergency Management. Prairie
Women’s Health Centre of Excellence, Toronto, Canada.
Enarson, E., 2010: Gender. In: Social Vulnerability to Disasters [Phillips, B.D., D.S.K.
Thomas, A. Fothergill, and L. Blinn-Pike (eds.)]. CRC Press, Boca Raton, FL, pp.
123-154.
Enarson, E. and B.H. Morrow, 1997: A gendered perspective: The voices of women. In:
Hurricane Andrew: Ethnology, Gender, and the Sociology of Disasters [Peacock,
W.G., B.H. Morrow, and H. Gladwin (eds.)]. Routledge, London, UK, pp. 116-140.
Enarson, E. and B.H. Morrow (eds.), 1998: The Gendered Terrain of Disaster: Through
Women’s Eyes. Praeger, Westport, CT.
Ensor, J. and R. Berger, 2009: Understanding Climate Change Adaptation: Lesson
from Community-Based Approaches. Practical Action Publishing, Bourton-on-
Dunsmore, UK.
Ensor, M.O., 2008: Displaced once again: Honduran migrant children in the path of
Katrina. Children, Youth and Environments, 18(1), 280-302.
Ericksen, N.J., 1986: Creating Flood Disasters? New Zealand’s Need for a New Approach
to Urban Flood Hazard. Ministry of Works and Development, Wellingon, NZ.
Ericksen, N.J., J. Dixon, and P. Berke, 2000: Managing natural hazards under the
resource management act. In: Environmental Planning and Management in
New Zealand [Memon, A. and Perkins, H. (ed.)]. Dunmore Press, Palmerston
North, UK, pp. 123-132.
Ericksen, N.J., K. O’Brien, and L. Rosentrater, 2008: Climate Change in Eastern and
Southern Africa: Impacts, Vulnerability and Adaptation. GECHS Report.
University of Oslo, Oslo, Norway.
Farrington, J. and R. Slater, 2009: ODI Social Protection Cash Transfers Series. Lump
Sum Cash Transfers in Developmental and Post-Emergency Contexts: How Well
have they Performed? Overseas Development Institute, London, UK, 1-26 pp.
Fearnside, P., 1995: Hydroelectric dams in the Brazilian Amazon as sources of
“greenhouse” gases. Environmental Conservation, 22, 7-19.
Fekete, A., 2009: Validation of a social vulnerability index in context to river-floods
in Germany. Natural Hazards and Earth Systems Science, 9, 393403.
FEMA, 1997:
Multi Hazard Identification and Risk Assessment. Government Printing
Office, Washington, DC.
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
330
Few, R. and F. Matthies, 2006: Flood hazards and health. Earthscan, London, UK.
Finan, T.J. and D.R. Nelson, 2001: Making rain, making roads, making do: Public and
private responses to drought in Ceará, Brazil. Climate Research, 19, 97-108.
Finch, C., C.T. Emrich, and S.L. Cutter, 2010: Disaster disparities and differential
recovery in New Orleans. Population and Environment, 31, 179-202.
Fischer, G., H. van Velthuizen, M. Shah, and F.O. Nachtergaele, 2002: Global Agro-
Ecological Assessment for Agriculture in the 21st Century: Methodology and
Results. International Institute for Applied Systems Analysis, Laxenburg, Austria.
Fischhoff, B., 1992: What forecasts (seem to) mean. International Journal of
Forecasting, 10, 387-403.
Fischhoff, B., 2007: Nonpersuasive communication about matters of greatest urgency:
Climate change. Environmental Science and Technology, 41(21), 7204-7208.
Ford, J., B. Smit, and J. Wandel, 2006: Vulnerability to climate change in the Arctic: A
case study from Arctic Bay, Canada. Global Environmental Change, 16(2), 145-160.
Forsyth, T. and A. Walker, 2008: Forest Guardians, Forest Destroyers: The Politics of
Environmental Knowledge in Northern Thailand. University of Washington
Press, Seattle, WA, pp. 302.
Fothergill, A., 1996: Gender, risk, and disaster. International Journal of Mass
Emergencies and Disasters, 14(1), 33-56.
Fothergill, A., 1999: Women’s role in disaster. American Behavioral Science Review,
7(2), 125-143.
Fothergill, A., 2004: Heads above Water: Gender, Class and Family in the Grand
Forks Flood. SUNY Press, Albany, NY.
Fritze, J., G. Blashki, S. Burke, and J. Wieseman, 2008: Hope, despair and transformation:
Climate change and the promotion of mental health and well-being.
International Journal of Mental Health Systems, 2, 1-12.
Froukh, M.L., 2001: Decision-support system for domestic water demand forecasting
and management. Water Resources Management, 15(6), 363-382.
Frumkin, H. and A.J. McMichael, 2008: Climate change and public health: Thinking,
communicating, acting. American Journal of Preventive Medicine, 35(5),
403-410.
Fuchs, R., 2009: Cities at Risk: Developing Adaptive Capacity for Climate Change in
Asia’s Coastal Megacities. International START Secretariat and the East West
Center, Washington, DC and Honolulu, HI.
Fulu, E., 2007: Gender, vulnerability, and the experts: Responding to the Maldives
tsunami.
Development and Change, 38(5), 843-864.
Gaillard, J.C. and M.L.C.J.D. Pangilinan, 2010: Participatory mapping for raising
disaster risk awareness among the youth. Journal of Contingencies and Crisis
Management, 18(3), 175-179.
Gall, M., K.A. Borden, and S.L. Cutter, 2009: When do losses count? Six fallacies of
natural hazards loss data. Bulletin of the American Meteorological Society,
90(6), 799-809.
Galloway, G.E.J., 2007: New directions in floodplain management. Journal of the
American Water Resources Association, 31(3), 351-357.
Galloway, G.E.J., 2009: Corps of Engineers responses to the changing national
approach to the floodplain management since the 1993 Midwest flood. Journal
of Contemporary Water Research & Education, 130(1), 5-12.
Galloway, G.E., D.F. Boesch, and R.R. Twilley, 2009: Restoring and protecting coastal
Louisiana. Issues in Science and Technology, 25(2).
Gedan, K.B., M.L. Kirwan, E. Wolanski, E.B. Barbier, and B.R. Silliman, 2011: The present
and future role of coastal wetland vegetation in protecting shorelines:
Answering recent challenges to the paradigm. Climatic Change, 106(1), 7-29.
Geiser, U. and S. Rist, 2009: Decentralization meets local complexity: Conceptual
entry points, field-level findings, and insights gained. In: Decentralization Meets
Local Complexity: Local Struggles, State Decentralization and Access to Natural
Resources In South Asia And Latin America [Geiser, U. and S. Rist (eds.)].
Geographica Bernesia/Swiss National Centre of Competence in Research
(NCCR) North-South, Berne, Switzerland, pp. 15-55.
Gething, P.W., D.L. Smith, A.P. Patil, A.J. Tatem, R.W. Snow, and S.I. Hay, 2010:
Climate change and the global malaria recession. Nature, 465, 342-345.
Godschalk, D.R., R. Norton, C. Richardson, and D. Salvesen, 2000: Avoiding coastal
hazard areas: Best state mitigation practices. Environmental Geosciences, 7(1),
13-22.
Gornitz, V.N., R.C. Daniels, T.W. White, and K.R. Birdwell, 1994: The development of
a coastal assessment database: Vulnerability to sea-level rise in the U.S.
southeast. Journal of Coastal Research, 12, 327-338.
Greater London Authority, 2010: The Draft Climate Change Adaptation Strategy
for London: Public Consultation Draft. Greater London Authority, London, UK.
Greene, S., n.d.: Community Owned Vulnerability and Capacity Assessment. World
Vision, Nairobi, Kenya, .
Grothmann, T. and A. Patt, 2005: Adaptive capacity and human cognition: The
process of individual adaptation to climate change. Global Environmental
Change, Part A: Human and Policy Dimensions, 15(3), 199-213.
Guha-Sapir, D. and R. Below, 2002: The Quality and Accuracy of Disaster Data: A
Comparative Analysis of Three Global Data Sets, ProVention Consortium,
Geneva, Switzerland.
Haimes
, Y., B. Horowitz, J. Lambert, J. Santos, C. Lian, and K. Crowther, 2005:
Inoperability input-output model for interdependent infrastructure sectors. I:
Theory and methodology. Journal of Infrastructure Systems, 11, 67-79.
Hallegatte, S., 2008: An adaptive regional input-output model and its application to
the assessment of the economic cost of Katrina. Risk Analysis, 28(3), 779-799.
Hallegatte, S. and P. Dumas, 2009: Can natural disasters have positive consequences?
Investigating the role of embodied technical change. Ecological Economics,
68(3), 777-786.
Hallegatte, S., F. Henriet, and J. Corfee-Morlot, 2008a: The Economics of Climate
Change Impacts and Policy Benefits at City Scale: A Conceptual Framework.
OECD, Paris, France.
Hallegatte, S., N. Patmore, O. Mestre, P. Dumas, J. Corfee-Morlot, C. Herweijer, and
R.M. Wood, 2008b: Assessing Climate Change Impacts, Sea Level Rise and
Storm Surge Risk in Port Cities: A Case Study on Copenhagen. OECD, Paris,
France.
Hallegatte, S., N. Ranger, O. Mestre, P. Dumas, J. Corfee-Morlot, C. Herweijer, and R.
Muir Wood, 2011: Assessing climate change impacts, sea level rise and storm
surge risk in port cities: A case study on Copenhagen. Climatic Change, 104(1),
113-137.
Hamilton, J.P. and S.J. Halvorson, 2007: The 2005 Kashmir earthquake. Mountain
Research and Development, 27(4), 296-301.
Hammar-Klose, E. and R. Thieler, 2001: Coastal Vulnerability to Sea-Level Rise: A
Preliminary Database for the US Atlantic, Pacific and Gulf Of Mexico Coasts.
USGS, Woods Hole, MA.
Handmer, J.W., 1987: Guidelines for floodplain acquisition. Applied Geography,
7(3), 203-221.
Hartman, C. and G.D. Squires (eds.), 2006: There is no such thing as a natural
disaster: Race, class, and Hurricane Katrina. Routledge, New York, NY.
Hasan, A., 2007: The urban resource center, Karachi. Environment and Urbanization,
19(1), 275-292.
Heal, G., 1997: Discounting and climate change; an editorial comment. Climatic
Change, 37(2), 335-343.
Heinrichs, D., R. Aggarwal, J. Barton, E. Bharucha, C. Butsch, M. Fragkias, P. Johnston,
F. Kraas, K. Krellenberg, A. Lampis, and O.G. Ling, 2009: Adapting Cities to
Climate Change: Opportunities and Constraints (Findings from Eight Cities).
Paper presented at Fifth Urban Research Symposium, Marseilles, France, 28-30
June 2009. World Bank, Washington DC.
Heinz Center, 1999: The Hidden Costs of Coastal Hazards: Implications for Risk
Assessment and Mitigation. Island Press, Washington, DC.
Heinz Center, 2002: Human Links to Coastal Disasters. The H. John Heinz III Center
for Science, Economics and the Environment, Washington, DC.
Hellmuth, M.E., D.E. Osgood, U. Hess, A. Moorhead, and H. Bhojwani (eds.), 2009:
Index Insurance And Climate Risk: Prospects for Development and Disaster
Management. Climate and Society No. 2, International Research Institute for
Climate and Society (IRI), Columbia University, New York, NY.
Helms, B., 2006: Access For All: Building Inclusive Financial Systems. World Bank,
Washington, DC.
Herweijer, C., N. Ranger, and R.E.T. Ward, 2009: Adaptation to climate change:
Threats and opportunities for the insurance industry. The Geneva Papers, 34(3),
360-380.
Chapter 5Managing the Risks from Climate Extremes at the Local Level
331
Hess, U. and J. Syroka, 2005: Weather-Based Insurance in Southern Africa: The Case
of Malawi. World Bank, Washington, DC.
Hewings, G.J.D. and R. Mahidhara, 1996: Economic impacts: Lost income, ripple
effects, and recovery. In: The Great Flood of 1993: Causes, Impacts, Responses
[Changnon, S.A. (ed.)]. Westview, Boulder, CO, pp. 205-217.
Hewitt, K., 1997: Regions of risk: A geographical introduction to disasters. Longman,
Harlow, UK, 389 pp.
Hillhorst, D., 2002: Being good or doing good? Quality and accountability of
humanitarian NGOs. Disasters, 26(3), 193-212.
Höppe, P. and E. Gurenko, 2006: Scientific and economic rationales for innovative
climate insurance solutions. Climate Policy, 6(6), 607-620.
Hordijk, M. and I. Baud, 2011: Inclusive adaptation: Linking participatory learning
and knowledge management to urban resilience. Local Sustainability, 1(3),
111-121.
Horne, C., 2001: Sociological perspectives on the emergence of norms. In: Social
Norms [Hechter, M. and K.-. Opp (eds.)]. Russell Sage Foundation, New York, NY,
pp. 3-34.
Hudson, P.F., H. Middelkoop, and E. Stouthamer, 2008: Flood management along the
lower Mississippi and Rhine rivers (the Netherlands) and the continuum of
geomorphic adjustment. Geomorphology, 101(2), 209-236.
Hughes, D.M., 2001: Cadastral politics: The making of community-based resource
management in Zimbabwe and Mozambique. Development and Change, 32(4),
741-768.
Hugo, G., 1996: Environmental concerns and international migration. The
International Migration Review, 30(1), 105-131.
Hugo, G., 2010: Climate change-induced mobility and the existing migration regime
in Asia and the Pacific. In: Climate Change and Displacement. Multidisciplinary
Perspectives [McAdam, J. (ed.)]. Hart, Oxford, UK, pp. 9-35.
Humanitarian Accountability Partnership International, 2011: 2010
Humanitarian Accountability Report. Humanitarian Accountability Partnership
International, Geneva, Switzerland, 1 pp.
Hunter, L.M., 2005: Migration and environmental hazards. Population and
Environment, 26(4), 273-302.
Huq, S., S. Kovats, H. Reid, and D. Satterthwaite, 2007: Reducing risks to cities from
disasters and climate change. Environment and Urbanization, 19(1), 39-64.
Hwang, S., J. Xi, Y. Cao, X. Feng, and X. Qia, 2007: Anticipation of migration and
psychological stress and the Three Gorges dam project, China. Social Science
and Medicine, 65(5), 1012-1024.
Hwang, S., Y. Cao, and J. Xi, 2010: Project-induced migration and depression: A panel
analysis.
Social Science and Medicine, 70(11), 1765-1772.
Hyder, A.A., S. Maman, J.E. Nyoni, S.A. Khasiani, N. Teoh, Z. Premji, and S. Sohani,
2005: The pervasive triad of food security, gender inequity and women’s health:
Exploratory research from Sub-Saharan Africa. African Health Sciences, 5(4),
328-334.
ICIMOD, 2009: Biodiversity and Climate Change. International Centre for Integrated
Mountain Development, Kathmandu, Nepal.
IDRC, 2002: Population and Health in Developing Countries–Volume 1: Population,
Health and Survival at INDEPTH Sites. International Development Research
Centre, Ottawa, Canada.
IFRC, 1991: Working with Women in Emergency Relief and Rehabilitation
Programmes. International Federation of Red Cross and Red Crescent Societies,
Geneva, Switzerland.
Inderberg, T.H. and P.O. Eikeland, 2009: Limits to adaptation: Analysing institutional
constraints. In: Adapting to Climate Change: Thresholds, Values and Governance
[Adger, W.N., I. Lorenzoni, and K. O’Brien (eds.)]. Cambridge University Press,
Cambridge, UK, pp. 433-447.
Ingram, J., G. Franco, C.R. Rio, and B. Khazai, 2006: Post-disaster recovery dilemmas:
Challenges in balancing short-term and long-term needs for vulnerability
reduction. Environmental Science and Policy, 9(7-8), 607-613.
Intercooperation, 2008: Mali: Highlighting local coping strategies for drought. In:
Gender Perspectives: Integrating Disaster Risk Reduction into Climate Change
Adaptation Good Practices and Lessons Learned. United Nations Secretariat of the
International Strategy for Disaster Reduction, Geneva, Switzerland, pp. 57-61.
IPCC, 2007: Climate Change 2007: The Physical Science Basis. Contribution of
Working Group I to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M.
Marquis, K.B. Averyt, M. Tignor and H.L. Miller (eds.)]. Cambridge University
Press, Cambridge, UK.
Iserson, K.V. and J.C. Moskop, 2007: Triage in medicine, part I: Concept, history, and
types. Annals of Emergency Medicine, 49(3), 275-281.
Jacobs, L., F. Cook, and M. Carpini, 2009: Talking Together: Public Deliberation and
Political Participation in America. University of Chicago Press, Chicago, IL.
Jha, A.K., J.D. Barenstein, P.M. Phelps, D. Pittet, and S. Sena, 2010: Safer Homes,
Stronger Communities: A Handbook for Reconstructing After Natural Disasters.
World Bank, Washington, DC, 370 pp.
Jones, L., S. Jaspars, S. Pavanello, E. Ludi, R. Slater, A. Arnall, N. Grist, and S. Mtisi,
2010: Responding to a Changing Climate: Exploring how Disaster Risk
Reduction, Social Protection and Livelihoods Approaches Promote Features of
Adaptive Capacity. ODI, London, UK.
Kabat, P., R.E. Schulze, M.E. Hellmuth, and J.A. Veraart (eds.), 2002: Coping with
Impacts of Climate Variability and Climate Change in Water Management: A
Scoping Paper. DWC-Report no. DWCSSO-01(2002), International Secretariat of
the Dialogue on Water and Climate, Wageningen, The Netherlands, 114 pp.
Kallis, G., 2008: Droughts. Annual Review of Environment and Resources, 33, 85-118.
Kang, S.J., S.J. Lee, and K.H. Lee, 2009: A study on the implementation of non-
structural measures to reduce urban flood damage. Journal of Asian
Architecture and Building Engineering, 8(2), 385-392.
Kasperson, R.E. and J.X. Kasperson, 2001: Climate Change, Vulnerability, and Social
Justice. Stockholm Environment Institute, Stockholm, Sweden.
Kasperson
, R.E., O. Renn, P. Slovic, H.S. Brown, J. Emel, R. Goble, J.X. Kasperson, and
S. Ratick, 1988: The social amplification of risk: A conceptual framework. Risk
Analysis, 2, 177-187.
Kates, R.W., 2000: Cautionary tales: Adaptation and the global poor. Climatic
Change, 45, 5-17.
Kates, R.W., C.E. Colten, S. Laska, and S.P. Leatherman, 2006: Reconstruction of New
Orleans after Hurricane Katrina: A research perspective. Proceedings of the
National Academy of Sciences, 103, 14653-14660.
Kawachi, I. and L. Berkman, 2000: Social cohesion, social capital, and health. In:
Social Epidemiology [Kawachi, I. and L. Berkman (eds.)]. Oxford University
Press, New York, NY, pp. 174-190.
Kazama, S., A. Sato, and S. Kawagoe, 2009: Evaluating the cost of flood damage
based on changes in extreme rainfall in Japan. Sustainability Science, 4(1), 61-69.
Keraminiyage, K., S. Jayasena, R. Haigh, and D. Amaratunga (eds.), 2008: Post
Disaster Recovery Challenges in Sri Lanka. School of the Built Environment,
University of Salford, UK.
Kingsford, R.T., 2000: Ecological impacts of dams, water diversions and river
management on floodplain wetlands in Australia. Austral Ecology, 25(2), 109-127.
Koelle, K., M. Pascual, and M.D. Yunus, 2004: Pathogen adaptation to seasonal forcing
and climate change. Proceedings of the Royal Society, 272(1566), 971-977.
Kovác, G. and K.M. Spens, 2007: Humanitarian logistics in disaster relief operations.
International Journal of Physical Distribution & Logistics Management, 37(2),
99-114.
Kraas, F., S. Aggarwal, M. Coy, G. Heiken, E. de Mulder, B. Marker, K. Nenonen, and
W. Yu, 2005: Megacities–our Global Urban Future. Earth Sciences for Society
Foundation, International Year for Planet Earth, Leiden, The Netherlands.
Kreps, G.A., 1985: Disaster and the social order. Sociological Theory, 3, 49-64.
Kristjanson, P., R. Reid, N. Dickson, W. Clark, D. Romney, R. Puskur, S. MacMillan,
and D. Grace, 2009: Linking international agricultural research knowledge with
action for sustainable development. Proceedings of the National Academy of
Sciences, 106(13), 5047-5052.
Kryspin-Watson, J., J. Arkedis, and W. Zakout, 2006: Mainstreaming Hazard Risk
Management into Rural Projects. Disaster Risk Management . World Bank,
Washington, DC.
Kull, D., 2006: Financial services for disaster risk management for the poor. In:
MicroFinance and Disaster Risk Reduction [Bhatt, M. and P.G.D. Chakrabarti
(eds.)]. Knowledge World, Delhi, India, pp. 39-64.
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
332
Kunreuther, H.C., E.O Michel-Kerjan, with N.A. Doherty, M.F. Grace, R.W. Klein, and
M.V. Pauly, 2009: At War With the Weather: Managing Large-Scale Risks in a
New Era of Catastrophes. MIT Press, Cambridge, MA.
Laczko, F. and C. Aghazarm (eds.), 2009: Migration, Environment and Climate
Change: Assessing the Evidence. International Organization for Migration,
Geneva, Switzerland.
Lafferty, K.D., 2009: The ecology of climate change and infectious diseases. Ecology,
90, 888-900.
Lambin, E.F., B.L. Turner, H.J. Geist, S.B. Agbola, A. Angelsen, J.W. Bruce, O.T. Coomes,
R. Dirzo, G. Fischer, C. Folke, P.S. George, K. Homewood, J. Imbernon, R. Leemans,
X. Li, E.F. Moran, M. Mortimore, P.S. Ramakrishnan, J.F. Richards, H. Skånes, W.
Steffen, G.D. Stone, U.S. Svedin, T.A. Veldkamp, C. Vogel, and J. Xu, 2001: The
causes of land-use and land-cover change: Moving beyond the myths. Global
Environmental Change, 11(4), 261-269.
Lambrew, J.M. and D.E. Shalala, 2006: Federal health policy response to Hurricane
Katrina: What it was and what it should have been. Journal of the American
Medical Association, 296, 1394-1397.
Langsdale, S.M., A. Beall, J. Carmichael, S.J. Cohen, C.B. Forster, and T. Neale, 2009:
Exploring the implications of climate change on water resources through
participatory modeling: Case study of the Okanagan Basin, British Columbia.
Journal of Water Resources Planning and Management, 135(5), 373-382.
Lauten, A.W. and K. Lietz, 2008: A look at standards gap: Comparing child protection
responses in the aftermath of Hurricane Katrina and the Indian Ocean tsunami.
Children, Youth and Environments, 18(1), 158-201.
Leach, M., R. Mearns, and I. Scoones, 1999: Environmental entitlements: Dynamics
and institutions in community-based natural resource management. World
Development, 27(2), 225-247.
Lecomte, E. and K. Gahagan, 1998: Hurricane insurance protection in Florida. In:
Paying the Price: The Status and Role of Insurance Against Natural Disasters in
the United Sates [Kunreuther, H. and R. Roth Sr. (eds.)]. Joseph Henry Press,
Washington, DC, pp. 97-124.
Leiserowitz, A., 2007: Communicating the risks of global warming: American risk
perceptions, affective images and interpretive communities. In: Creating a
Climate for Change [Moser, S. and L. Dilling, L. (eds.)]. Cambridge University
Press, New York, NY, pp. 44-63.
Lichterman, J.D., 2000: A “community as resource” strategy for disaster response.
Public Health Reports, 115(2-3), 262-265.
Lin, N., 2001: Building a network theory of social capital. In: Social capital: Theory and
research [Lin, N., K.S. Cook, and R.S. Burt (eds.)]. Transaction Press, New
Brunswick, NJ, pp. 3-30.
Lindell, M.K. and R.W. Perry, 1992: Behavioral Foundations of Community
Emergency Management. Hemisphere Publishing Corp., Washington, DC.
Lindell, M.K. and R.W. Perry, 1996: Identifying and managing conjoint threats.
Journal of Hazardous Materials, 50, 31-46.
Lindell, M.K. and R.W. Perry, 2004: Communicating Environmental Risk in
Multiethnic Communities. Sage Publications, Thousand Oaks, CA.
Lindell, M.K. and C.S. Prater, 2003: Assessing community impacts of natural disasters.
Natural Hazards Review,
4, 176-185.
Lindell, M.K. and D.J. Whitney, 2000: Correlates of household seismic hazard
adjustment adoption. Risk Analysis, 20, 13-25.
Lindell, M.K., J. Lu, and C.S. Prater, 2005: Household decision making and evacuation
in response to Hurricane Lili. Natural Hazards Review, 6(4), 171-179.
Linnerooth-Bayer, J. and R. Mechler, 2009: Insurance Against Losses from Natural
Disasters in Developing Countries. United Nations Department of Economic and
Social Affairs, New York, NY and Geneva, Switzerland.
Linnerooth-Bayer, J., C. Bals, and R. Mechler, 2010: Insurance as part of a climate
adaptation strategy. In: Making Climate Change Work for Us [Hulme, M. and H.
Neufeldt (eds.)]. Cambridge University Press, Cambridge, UK, pp. 340-366.
Lloyds of London, 2008: Coastal Communities and Climate Change: Maintaining
Future Insurability. 360 Degree Risk Project Report, Lloyd’s of London, London, UK.
Lonsdale, K.G., T. Downing, R. Nicholls, D. Parker, M.C. Uyarra, R. Dawson, and J. Hall,
2008: Plausible responses to the threat of rapid sea-level rise for the Thames
estuary. Climatic Change, 91(1-2), 145-169.
Luna, E., 2001: Disaster mitigation and preparedness: The case of NGOs in the
Philippines. Disasters, 25(3), 216-226.
Lunch, N. and C. Lunch, 2006: Insights into Participatory Video: A Handbook for The
Field. Insight, Oxford, UK.
Lund, J.R. and R.U. Reed, 1995: Drought water rationing and transferable rations.
Journal of Water Resources Planning and Management, 121(6), 429-437.
Lundqvist, L.J. and C.V. Borgstede, 2008: Whose responsibility? Swedish local
decision makers and the scale of climate change abatement. Urban Affairs
Review, 43(3), 299-324.
Macchi, M., G. Oveido, S. Gotheil, K. Cross, A. Boedhihartono, C. Wolfangel, and M.
Howell, 2008: Indigenous and Traditional Peoples and Climate Change. IUCN
Issues Paper, IUCN, Gland, Switzerland.
Maibach, E., C. Roser-Renouf, and A. Leiserowitz, 2008: Communication and marketing
as climate change - intervention assets: A public health perspective. American
Journal of Preventive Medicine, 35(5), 488-500.
Manyena, S.B., M. Fordham, and A. Collins, 2008: Disaster resilience and children:
Managing food security in Zimbabwe’s Binga District. Children, Youth and
Environments, 18(1), 303-331.
Marx, S., E. Wever, B. Orlove, A. Leiserowitz, D. Krantz, C. Roncoli, and J. Phillips, 2007:
Communication and mental processes: Experiential and analytic processing of
uncertain climate information. Global Environmental Change, 17, 47-58.
Masozera, M., M. Bailey, and C. Kerchner, 2007: Distribution of impacts of natural
disasters across income groups: A case study of New Orleans. Ecological
Economics
, 63(2-3), 299-306.
May, P. and R. Burby, 1998: Making sense out of regulatory reform. Law and Policy,
20(2), 158-182.
McAdoo, B., A. Moore, and J. Baumwoll, 2009: Indigenous knowledge and the near
field population. Natural Hazards, 48, 73-82.
McCall, M., 2008: Participatory Mapping and Participatory GIS (PGIS) for CRA,
Community DRR and Hazard Assessment. ProVention Consortium, CRA Toolkit,
Participation Resources, Geneva, Switzerland.
McCall, M.K., 2010: Local participation in mapping, measuring and monitoring for
community carbon forestry. In: Community Forest monitoring for the Carbon
Market: Opportunities Under REDD [Skutch, M. (ed.)]. Routledge, London, UK,
pp. 31-44.
McCormick, S., 2009: From “politico-scientists” to democratizing science movements:
The changing climate of citizens and science. Organization and Environment,
22(1), 34-51.
McCormick, S., 2010: Hot or not? Obstacles to the emergence of climate-induced
illness movements. In: Social Movements and Health Care in the United States
[Zald, J.B.-.a.S.L.M. (ed.)]. Oxford University Press, Oxford, UK.
McCormick, S., 2011: Damming the Amazon: Local movements and transnational
struggles over water. Society & Natural Resources, 24(1), 34-48.
McGee, T.K. and S. Russell, 2003: “It’s just a natural way of life...” an investigation of
wildfire preparedness in rural Australia. Environmental Hazards, 5(1-2), 1-12.
McGregor, D., D. Simon, and D. Thompson (eds.), 2006: The Peri-Urban Interface:
Approaches to Sustainable Natural and Human Resource Use. Earthscan,
London, UK.
McGuire, L., E. Ford, and C. Okoro, 2007: Natural disasters and older US adults with
disabilities: implications for evacuation. Disasters, 31(1), 49-56.
McKinsey Group, 2009: Shaping Climate Resilient Development: A Framework for
Decision Making. Climate Works Foundation, Global Environment Facility, European
Commission, McKinsey & Company, Rockefeller Foundation, and Swiss Re, 164 pp.
McLeman, R. and B. Smit, 2006: Migration as an adaptation to climate change.
Climatic Change, 76, 31-53.
McMichael, A.J., R.E. Woodruff, and S. Hales, 2006: Climate change and human
health: Present and future risks. Lancet, 367, 859-869.
McMichael, A.J., S. Friel, A. Nyong, and C. Corvalan, 2008: Global environmental
change and health: Impacts, inequalities, and the health sector. British Medical
Journal, 336, 191-194.
McTaggart-Cowan, R., L.F. Brosart, C.A. Davis, E.H. Atallah, J.R. Gyakum, and K.A.
Emanuel, 2006: Analysis of Hurricane Catarina (2004).
Monthly Weather
Review, 134(11), 3029-3053.
Chapter 5Managing the Risks from Climate Extremes at the Local Level
333
Meehl, G.A., T.F. Stocker, W.D. Collins, P. Friedlingstein, A.T. Gaye, J.M. Gregory, A.
Kitoh, R. Knutti, J.M. Murphy, A. Noda, S.C.B. Raper, I.G. Watterson, A.J. Weaver,
and Z-. Zhao, 2007: Global climate projections. In: Climate Change 2007: The
Physical Science Basis. Contribution of Working Group I to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change
[Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor
and H.L. Miller (eds.)]. Cambridge University Press, Cambridge, UK, and New
York, NY, pp. 747-845.
Méheux, K., D. Dominey-Howes, and K. Lloyd, 2010: Operational challenges to
community participation in post-disaster damage assessments: Observations
from Fiji. Disasters, 34(4), 1102-1122.
Mercer, J., I. Kelman, K. Lloyd, and S. Suchet-Pearson, 2008: Reflections on use of
participatory research for disaster risk reduction. Area, 40(2), 172-183.
Middlekoop, K., L. Meyer, J. Smit, R. Wood, and L. Bekker, 2006: Design and
evaluation of a drama-based intervention to promote voluntary counseling and
HIV testing in a South African community. Sexually Transmitted Diseases,
33(8), 524-526.
Milbert, E., 2006: Slums, slum dwellers and multilevel governance. European Journal
of Development Research, 18(2), 299-318.
Mileti, D.S., 1999: Disasters by Design: A Reassessment of Natural Hazards In the
United States. Joseph Henry Press, Washington, DC.
Mileti, D.S. and C. Fitzpatrick, 1993: The Great Earthquake Experiment: Risk
Communication and Public Action. Westview Press, Boulder, CO.
Mileti, D.S. and P.W. O’Brien, 1992: Warnings during disaster: Normalizing
communicated risk. Social Problems, 39, 40-57.
Mileti, D.S. and L.A. Peek, 2002: Understanding individual and social characteristics
in the promotion of household disaster preparedness. In: New Tools for
Environmental Protection: Education, Information, and Voluntary Measures
[Dietz, T. and P.C. Stern (eds.)]. The National Academies Press, Washington, DC,
pp. 125-140.
Mileti, D.S. and J.H. Sorensen, 1990: Communication of Emergency Public Warnings:
A Social Science Perspective and State-Of-The-Art Assessment. Oak Ridge
National Laboratory, Oak Ridge, TN.
Mills, M.A., D. Edmonson, and C.L. Park, 2007: Trauma and stress response among
Hurricane Katrina evacuees. American Journal of Public Health, 97(S1), S116-S123.
Mirza, M.M.Q., 2003: Climate change and extreme weather events: Can developing
countries adapt? Climate Policy, 3, 233-248.
Mitchell, J.K., 1988: Confronting natural disasters: An international decade for
natural hazard reduction. Environment, 30(2), 25-29.
Mitchell, J.K. (ed.), 1999: Crucibles of Hazard: Mega-cities and Disasters in
Transition. United Nations University Press, Tokyo, Japan.
Mitchell, J.K., 2003: European river floods in a change world. Risk Analysis, 23,
567-574.
Mitchell, T. and M. van Aalst, 2008: Convergence of Disaster Risk Reduction and
Climate Change Adaptation. UK Department for International Development,
London, UK.
Mitchell, T., K. Haynes, W. Choong, N. Hall, and K. Oven, 2008: The role of children
and youth in communicating disaster risk. Children, Youth and Environments,
18(1), 254-279.
Moench, M. and A. Dixit (eds.), 2004: Adaptive Capacity and Livelihood Resilience:
Adaptive Strategies for Responding To Floods and Droughts in South Asia.
Institute for Social and Environmental Transition, Kathmandu, Nepal.
Moench, M. and Risk to Resilience Study Team, 2008: From Risk to Resilience:
Understanding the Costs and Benefits of Disaster Risk Reduction Under
Changing Climatic Conditions. Working Paper 9, ProVention and ISET, Geneva,
Switzerland.
Montz, B.E. and G.A. Tobin, 2008: Living large with levees: Lessons learned and lost.
Natural Hazards Review, 9(3), 150-157.
Morgan, J., 1994: Sudanese refugees in Koboko. Gender and Development, 2(1),
41-44.
Morrow, B.H. and E. Enarson, 1996: Hurricane Andrew through women’s eyes: Issues
and recommendations. International Journal of Mass Emergencies and
Disasters, 14(1), 5-22.
Moser, S.C., 2009: Whether our are levers are long enough and the fulcrum strong?
Exploring the soft underbelly of adaptation decisions and actions. In: Adapting
to Climate Change: Thresholds, Values, Governance [Adger, W.N., I. Lorenzoni,
and K. O’Brien (eds.)]. Cambridge University Press, Cambridge, UK, pp. 313-334.
Moser, S. and L. Dilling, 2007: Creating a Climate for Change: Communicating
Climate Change and Facilitating Social Change. Cambridge University Press,
Cambridge, UK.
Moser, S.C. and J.A. Ekstrom, 2010: A framework to diagnose barriers to climate
change adaptation. Proceedings of the National Academy of Sciences, 107(51),
22026-22031.
Moynihan, D.P., 2009: The network governance of crisis response: Case studies of
incident command systems. Public Administration Review, 19(4), 895-915.
Mukheibir, P. and G. Ziervogel, 2007: Developing a municipal adaptation plan (MAP)
for climate change: The city of Cape Town. Environment and Urbanization,
19(1), 143-158.
Munich Re Group, 2004: Megacities–Megarisks: Trends and Challenges for
Insurance and Risk Management. Munich Re Group, Munich, Germany.
Mustafa, D., 1998: Structural causes of vulnerability to flood hazard in Pakistan.
Economic Geography, 74(3), 289-305.
Mutton, D. and C.E. Haque, 2004: Human vulnerability, dislocation and resettlement:
Adaptation processes of river-bank erosion-induced displacees in Bangladesh.
Disasters, 28(1), 41-62.
Myers, C.A., T. Slack, and J. Singlemann, 2008: Social vulnerability and migration in
the wake of disaster: The case of Hurricanes Katrina and Rita. Population and
Environment, 29, 271-291.
Nakagawa, Y. and R. Shaw, 2004: Social capital: A missing link to disaster recovery.
International Journal of Mass Emergencies and Disasters, 22(1), 5-34.
Nelson, E., G. Mendoza, J. Regetz, S. Polasky, H. Tallis, D.R. Cameron, K.M.A. Chan,
G.C. Daily, J. Goldstein, P.M. Kareiva, E. Lonsdorf, R. Naidoo, T.H. Ricketts, and
M.R. Shaw, 2009: Modeling multiple ecosystem services, biodiversity conservation,
commodity production, and tradeoffs at landscape scales. Frontiers in Ecology
and Environment, 7(1), 4-11.
Neria, Y., S. Galea, and F.H. Norris (eds.), 2009: Mental Health and Disasters.
Cambridge University Press, Cambridge, UK.
Neumayer, E. and T. Plümper, 2007: The gendered nature of natural disasters: The
impact of catastrophic events on the gender gap in life expectancy, 1981-2002.
Annals of the Association of American Geographers, 97(3), 551-566.
Ngo, E.B., 2001: When disasters and age collide: Reviewing vulnerability of the
elderly. Natural Hazards Review, 2(2), 80-89.
Nicholls, R., R.S.J. Tol, and A. Vafeidis, 2008: Global estimates of the impact of a
collapse of the west Antarctic ice sheet: An application of FUND. Climatic
Change, 91(1-2), 171-190.
Nicholson-Cole, S., 2004: Representing climate change futures: A critique on the
use of images for visual communication. Computers, Environment and Urban
Systems, 29, 255-273.
Niepold, F., D. McConville, and D. and Herring, 2008: The role of narrative and
geospatial visualization in fostering climate literate citizens. Physical
Geography, 29, 255-273.
NOAA, 2007: Construction Setbacks. National Oceanic and Atmospheric
Administration, Office of Ocean and Coastal Resource Management,
Washington, DC.
Nordås, R. and N.P. Gleditsch, 2007: Climate change and conflict. Political
Geography, 26, 627-638.
Nordhaus, W.D., 2007: Review: A review of the “Stern review on the economics of
climate change.” Journal of Economic Literature, 45(3), 686-702.
Nordstrom, K.F., 2000: Beaches and Dunes of Developed Coasts. Cambridge
University Press, Cambridge, UK.
Nordstrom, K.F., 2008: Beach and Dune Restoration. Cambridge University Press,
Cambridge, UK.
Norgaard, K., 2009: Cognitive and Behavioral Challenges in Responding to Climate
Change. The World Bank, Washington, DC.
North, D.C., 1990: Institutions, Institutional Change and Economic Performance.
Cambridge University Press, Cambridge, UK.
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
334
Noy, I. and T. Vu, 2010: The economics of natural disasters in a developing country:
The case of Vietnam. Journal of Asian Economics, 21, 345-354.
NRC, 1989: Improving Risk Communication. National Research Council, National
Academies Press, Washington, DC.
NRC, 2006: Facing Hazards and Disasters: Understanding Human Dimensions.
National Research Council, National Academy Press, Washington, DC.
NRC, 2007: Tools and Methods for Estimating Populations at Risk from Natural
Disasters and Complex Humanitarian Crises. National Research Council,
National Academies Press, Washington, DC.
Oba, G., 2001: The importance of pastoralists’ indigenous coping strategies for
planning drought management in the arid zone of Kenya. Nomadic Peoples,
5(1), 89-119.
Obasi, G.O.P., 2005: The impacts of ENSO in Africa. In: Climate Change and
Africa [Low, P.S. (ed.)]. Cambridge University Press, Cambridge, UK, pp. 218-
230.
O’Brien, G., T. Devisscher, and P. O’Keefe, 2011: The Adaptation Continuum:
Groundwork for the Future. Lambert Academic Publishing, Germany.
O’Brien, K., R. Leichenko, U. Kelkar, H. Venema, G. Aandahl, H. Tompkins, A. Javed, S.
Bhadwal, S. Barg, L. Nygaard, and J. West, 2004: Mapping vulnerability to
multiple stressors: Climate change and globalization in India. Global
Environmental Change, 14(4), 303-313.
O’Brien, P. and D.S. Mileti, 1992: Citizen participation in emergency response following
the Loma Prieta earthquake. International Journal of Mass Emergencies and
Disasters, 10(1), 71-89.
Oettlé, N., A. Arendse, B. Koelle, and A. Van Der Poll, 2004: Community exchange and
training in the Suid Bokkeveld: A UNCCD pilot project to enhance livelihoods
and natural resource management. Environmental Monitoring and Assessment,
99(1-3), 115-125.
O’Keefe, P., K. Westgate, and B. Wisner, 1976: Taking the naturalness out of natural
disasters. Nature, 260(5552), 566-567.
Okuyama, Y., 2004: Modeling spatial economic impacts of an earthquake: Input-
output approaches. Disaster Prevention and Management, 13, 297-306.
Oloruntoba, R., 2005: A wave of destruction and the waves of relief: Issues,
challenges and strategies. Disaster Prevention and Management, 14(4),
506-521.
Olsen, G.R., N. Carstensen, and K. Høyen, 2003: Humanitarian crises: What determines
the level of emergency assistance? Media coverage, donor interests and the aid
business. Disasters, 27(2), 109-126.
Olshansky, R.B. and J.D. Kartez, 1998: Managing land use to build resilience. In:
Cooperating with Nature: Confronting Natural Hazards with Land Use Planning
for Sustainable Communities [Burby, R. (ed.)]. Joseph Henry Press, Washington,
DC, pp. 167-201.
Olson, R., 2000: Towards a politics of disaster: Losses, values, agendas and blame.
International Journal of Mass Emergencies and Disasters, 18, 265-287.
O’Neill, M.S. and K.L. Ebi, 2009: Temperature extremes and health: Impacts of
climate variability and change in the United States. Journal of Occupational and
Environmental Medicine, 51(1), 13-25.
Opperman, J.J., G.E. Galloway, J. Fargione, J.F. Mount, B.D. Richter, and S. Secchi,
2009: Sustainable floodplains through large-scale reconnection to rivers.
Science, 326, 1487-1488.
Osman-Elasha, B., 2006a: Project AF14,-Assessments of Impacts and Adaptations
to Climate Change. Environmental Strategies to Increase Human Resilience to
Climate Change: Lessons for Eastern and Northern Africa, Final Report.
International START Secretariat, Washington, DC.
Osman-Elasha, B., 2006b: Human Resilience to Climate Change: Lessons for Eastern
and Northern Africa. International START Secretariat, Washington, DC.
Osman-Elasha, B. and T.E. Downing, 2007: Lessons Learned in Preparing National
Adaptation Programmes of Action in Eastern and Southern Africa. SEI Oxford
Working Paper, Stockholm Environment Institute, Oxford, UK.
Osman-Elasha, B. and A. El Sanjak, 2009: Global climate changes: Impacts on water
resources and human security in africa. In: Environment and Conflict in Africa:
Reflections on Darfur [Leroy, M. (ed.)]. University for Peace/Africa Programme,
Addis Ababa, Ethiopia, pp. 406-427.
Osman-Elasha, B., N. Goutbi, E. Spanger-Siegfried, W. Dougherty, A. Hanafi, S.
Zakieldeen, A. Sanjak, H.A. Atti, and H.M. Elhassan, 2006: Adaptation Strategies
to Increase Human Resilience Against Climate Variability and Change: Lessons
from the Arid Regions of Sudan. AIACC, International START Secretariat,
Washington, DC.
Oxfam, 2009: Exposed: Social Vulnerability and Climate Change in the US Southeast.
Oxfam America, Boston, MA.
Pahl-Wostl, C., J. Sendzimir, P. Jeffrey, J. Aerts, G. Berkamp, and K. Cross, 2007:
Managing change toward adaptive water management through social learning.
Ecology and Society, 12(2), 23/06/2010-30.
Palen, L., S. Vieweg, J. Sutton, S.B. Liu, and A. Hughes, 2007: Crisis informatics:
Studying crises in a networked world. In: Third International Conference on
e-Social Science, Ann Arbor, MI, 7-9 October 2007.
Parati, G., R. Antonicelli, F. Guazzarotti, E. Paciaroni, and G. Mancia, 2001:
Cardiovascular effects of an earthquake: Direct evidence by ambulatory blood
pressure monitoring. Hypertension, 38(5), 1093-1095.
Parkinson, A.J. and J.C. Butler, 2005: Potential impacts of climate change on
infectious diseases in the Arctic. International Journal of Circumpolar Health,
64, 478-486.
Patiño, L. and D. Gauthier, 2009: Integrating local perspectives into climate change
decision making in rural areas of the Canadian prairies. International Journal of
Climate Change Strategies and Management, 1(2), 179-196.
Patt, A.G., 2009: Learning to crawl: How to use seasonal climate forecasts to build
adaptive capacity. In: Adapting to Climate Change: Thresholds, Values,
Governance [Adger, W.N., I. Lorenzoni, and K. O’Brien (eds.)]. Cambridge
University Press, Cambridge, UK.
Patt, A.G. and C. Gwata, 2002: Effective seasonal forecast applications: Examining
constraints for subsistence farming in Zimbabwe. Global Environmental
Change–Human and Policy Dimensions, 12, 185-195.
Patt, A. and D. Schröter, 2008: Perceptions of climate risk in Mozambique:
Implications for the success of adaptation and coping strategies. Global
Environmental Change, 18, 458-467.
Patt, A.G., N. Peterson, M. Carter, M. Velez, U. Hess, and P. Suarez, 2009: Making
index insurance attractive to farmers. Mitigation and Adaptation Strategies for
Global Change, 14, 737-757.
Patz, J.A., D. Campbell-Lendrum, T. Holloway, and J.A. Foley, 2005: Impact of regional
climate change on human health. Nature, 438, 310-317.
Paul, B.K. and S. Dutt, 2010: Hazard warnings and responses to evacuation orders:
The case of Bangladesh’s Cyclone Sidr. Geographical Review, 100(3), 336-355.
Paul, B.K., H. Rashid, M.S. Islam, and L.M. Hunt, 2010: Cyclone evacuation in
Bangladesh: Tropical Cyclones Gorky (1991) vs. Sidr (2007). Environmental
Hazards, 9(1), 89-101.
Paulson, D.D., 1993: Hurricane hazard in western Samoa. Geographical Review,
83(1), 43-53.
Peacock, W., B.H. Morrow, and H. Gladwin (eds.), 1997: Hurricane Andrew: Ethnicity,
Gender, and the Sociology of Disasters. Routledge, London, UK.
Pearce, D. and D. Moran, 1994: The economic value of biodiversity. Earthscan
Publications Ltd., London, UK, 172 pp.
Pelling, M., 2003: The Vulnerability of Cities: Natural Disasters and Social Resilience.
Earthscan, London, UK.
Pelling, M., 2007: Learning from others: The scope and challenges for participatory
disaster risk assessment. Disasters, 31(4), 373-385.
Pelling, M. and K. Dill, 2010: Disaster politics: Tipping points for change in the
adaptation of sociopolitical regimes. Progress in Human Geography, 34(1),
21-37.
Pelling, M. and C. High, 2005: Understanding adaptation: What can social capital offer
assessments of adaptive capacity? Global Environmental Change, 15, 308-319.
Pelling, M., C. High, J. Dearing, and D. Smith, 2008: Shadow spaces for social learning:
A relational understanding of adaptive capacity to climate change within
organisations. Environment and Planning A, 40(4), 867-884.
Penning-Rowsell, E.C., C. Johnson, and S.M. Tunstall, 2006: Signals from pre-crisis
discourse: Lessons from UK flooding for global environmental policy change?
Global Environmental Change,
16, 323-339.
Chapter 5Managing the Risks from Climate Extremes at the Local Level
335
Perch-Nielson, S.L., M.B. Battig, and D. Imboden, 2008: Exploring the link between
climate change and migration. Climatic Change, 91(3-4), 375-393.
Perry, R.W. and M.K. Lindell, 1991: The effects of ethnicity on evacuation decision
making. International Journal of Mass Emergencies and Disasters, 9(1), 47-68.
Perry, R.W. and M.K. Lindell, 1997: Principles for managing community relocation as
a hazard mitigation measure. Journal of Contingencies and Crisis Management,
5(1), 49-59.
Petal, M., R. Green, I. Kelman, R. Shaw, and A. Dixit, 2008: Community-based
construction for disaster risk reduction. In: Hazards and the Built Environment:
Attaining Built-In Resilience [Bosher, L. (ed.)]. Routledge, London, UK, pp. 191-217.
Pettenger, M.E., 2007: The Social Construction of Climate Change: Power,
Knowledge, Norms, Discourses. Ashgate Publishers, Hampshire, UK.
Pfefferbaum, B., J.B. Houston, K.F. Wyche, R.L. Van Horn, G. Reyes, H. Jeon-
Slaughter, and C.S. North, 2008: Children displaced by Hurricane Katrina: A
focus group study. Journal of Loss & Trauma, 13(4), 303-318.
Pielke Jr., R.A., J. Gratz, C.W. Landsea, D. Collins, M.A. Saunders, and R. Musulin,
2008: Normalized hurricane damage in the United States: 1900-2005. Natural
Hazards Review, 9, 29-42.
Plush, T., 2009: Amplifying children’s voices on climate change: The role of participatory
video. Participatory Learning and Action, 60, 119-128.
Podestá, G., D. Letson, C. Messina, F. Royce, R.A. Ferreyra, J. Jones, J. Hansen, I.
Llovet, M. Grondona, and J.J. O’Brien, 2002: Use of ENSO-related climate
information in agricultural decision making in Argentina: A pilot experience.
Agricultural Systems, 74, 371-392.
Polack, E., 2008: A right to adaptation: Securing the participation of marginalized
groups. IDS Bulletin, 39(4), 16-23.
Pomeroy, R.S., B.D. Ratner, S.J. Hall, J. Pimoljind, and V. Vivekanandan, 2006: Coping
with disaster: Rehabilitating coastal livelihoods and communities. Marine
Policy, 30(6), 786-793.
Portes, A. and P. Landolt, 1996: The downside of social capital. The American
Prospect, 26, 18-21.
Postel, S.L., G.C. Daily, and P.R. Ehrlich, 1996: Human appropriation of renewable
fresh water. Science, 271(5250), 785-788.
Prabhakar, S.V.R.K., A. Srinivasan, and R. Shaw, 2009: Climate change and local
level disaster reduction planning: Need, opportunities and challenges.
Mitigation and Adaptation Strategies for Global Change, 14, 7-33.
Prasad, N., F. Ranghieri, F. Shah, Z. Trohanis, E. Kessler, and R. Sinha, 2009: Climate
Resilient Cities: A Primer On Reducing Vulnerabilities. World Bank, Washington, DC.
Priest, S.J., M.J. Clark, and E.J. Treby, 2005: Flood insurance: The challenge of the
uninsured. Area, 37(3), 295-302.
Pulwarty, R.S., 2003: Climate and water in the west: Science, information and
decision making. Water Resources, 124, 4-12.
Pulwarty, R.S., 2007: Communicating Agroclimatological Information, Including
Forecasts for Agricultural Decision. Guide to Agrometeorological Practices.
World Meteorological Organization, Geneva, Switzerland.
Pulwarty, R.S. and T. Melis, 2001: Climate extremes and adaptive management on
the Colorado River. Journal of Environmental Management, 63, 307-324.
Pulwarty, R., T. Broad, and T. Finan, 2004: ENSO forecasts and decision making in
Peru and Brazil. In: Mapping Vulnerability: Disasters, Development and People
[Bankoff, G., G. Frerkes, and T. Hilhorst (eds.)]. Earthscan, London, UK, pp. 83-98.
Quarantelli, E.L. (ed.), 1998: What is a Disaster? A Dozen Perspectives on the
Question. Routledge, London, UK.
Ramachandran, N., 2006: Women and Food Security in South Asia. Current Issues
and Emerging Concerns. WIDER Research Paper, UN University, World Institute
for Development Economic Research, Helsinki, Finland, 19 pp.
Ranger, N., S. Hallegatte, S. Bhattacharya, M. Bachu, S. Priya, K. Dhore, F. Rafique, P.
Mathur, N. Naville, F. Henriet, C. Herweijer, S. Pohit, and J. Corfee-Morlot, 2011:
An assessment of the potential impact of climate change on flood risk in
Mumbai. Climatic Change, 104(1), 139-167.
Rao, K. and U. Hess, 2009: Scaling up with India: The public sector. In: Index Insurance
and Climate Risk: Prospects for Development and Disaster Management [Hellmuth,
M.E., D.E. Osgood, U. Hess, A. Moorhead, and H. Bhojwani (eds.)]. International
Research Institute for Climate and Society (IRI), New York, NY, pp. 87-89.
Raschky, P.A., 2008: Institutions and the losses from natural disasters. Natural
Hazards Earth Systems Science, 8, 627-634.
Rayner, S., H. Ingram, and D. Lach, 2005: Weather forecasts are for wimps: Why
water resource managers do not use climate forecasts. Climatic Change, 69,
197-227.
Reid, H., M. Alam, R. Berger, T. Cannon, S. Huq, and A. Milligan, 2009: Community-
based adaptation to climate change: An overview. Participatory Learning and
Action, 60, 11-38.
Ribot, J.C., 2003: Democratic decentralization of natural resources: Institutional
choice and discretionary power transfers in sub-Saharan Africa. Public
Administration and Development, 23(1), 53-65.
Ribot, J.C., A. Agrawal, and A.M. Larson, 2006: Recentralizing while decentralizing:
How national governments reappropriate forest resources. World Development,
34(11), 1864-1886.
Rist, S., 2000: Linking ethics and market – Campesino economic strategies in the
Bolivian Andes. Mountain Research and Development,
20(4), 310-315.
Rist, S., F. Delgado, and U. Wiesmann, 2003: The role of social learning processes in
the emergence and development of Aymara land use systems. Mountain
Research and Development, 23(3), 263-270.
Ritchie, L. and D. Gill, 2007: Social capital theory as an integrating theoretical
framework in technological disaster research. Sociological Spectrum, 27(1),
103-129.
RMS, 2009: Analyzing the Effects of the ‘my Safe Florida Home’ Program on Florida
Insurance Risk. RMS Special Report, Summary of Analysis Prepared for the
Florida Department of Financial Services, Risk Management Solutions, Newark,
CA.
Roberts, S., 2008: Effects of climate change on the built environment. Energy Policy,
36(12), 4552-4557.
Robledo, C., N. Clot, A. Hammill, and B. Riche, 2011: The role of forest ecosystems in
community-based coping strategies to climate hazards: Three examples from
rural areas in Africa. Forest Policy and Economics, doi:10.1016/j.forpol.
2011.04.006.
Rodriguez-Oreggia, E., A.d.l. Fuente, R.d.l. Torre, H. Moreno, and C. Rodriguez,
2010: The Impact of Natural Disasters on Human Development and Poverty at
the Municipal Level in Mexico. CID Working paper 43, Center for International
Development, Harvard University, Cambridge, MA.
Rofi, A., S. Doocy, and C. Robinson, 2006: Tsunami mortality and displacement in
Aceh Province, Indonesia. Disasters, 30(3), 340-350.
Rojas Blanco, A.V., 2006: Local initiatives and adaptation to climate change.
Disasters, 30(1), 140-147.
Ronan, K.R., K. Crellin, D.M. Johnston, J. Becker, K. Finnis, and D. Paton, 2008:
Promoting child and family resilience to disasters: Effects, interventions
and prevention effectiveness. Children, Youth and Environments, 18(1), 332-
353.
Roncoli, C., K. Ingram, and P. Kirshen, 2001: The costs and risks of coping with
drought: Livelihood impacts and farmers’ responses in Burkina Faso. Climate
Research, 19, 119-132.
Rose, A., 2004: Economic principles, issues, and research priorities in hazard loss
estimation. In: Modeling Spatial and Economic Impacts of Disasters [Okuyama,
Y. and S. Chang (eds.)]. Springer, Berlin, Germany, pp. 14-36.
Rose, A. and S. Liao, 2005: Modeling regional economic resilience to disasters: A
computable general equilibrium analysis of water service disruptions. Journal of
Regional Science, 45, 75-112.
Rose, A., J. Benavides, S. Chang, P. Szczesnjak, and D. Lim, 1997: The regional
economic impact of an earthquake: Direct and indirect effects of electricity
lifeline disruptions. Journal of Regional Science, 37, 437-458.
Rose, A., G. Oladosu, and S. Liao, 2007: Business interruption impacts of a terrorist
attack on the electric power system of Los Angeles: Customer resilience to a
total blackout. Risk Analysis, 27(3), 513-531.
Rosenzweig
, C., D.C. Major, K. Demong, C. Stanton, R. Horton, and C. Stults, 2007:
Managing climate change risks in New York City’s water system: Assessment
and adaptation planning. Mitigation and Adaptation Strategies for Global
Change, 12(8), 1391-1409.
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
336
Rosenzweig, C., W.D. Solecki, R. Blake, M. Bowman, C. Faris, V. Gornitz, R. Horton,
K. Jacob, A. LeBlanc, R. Leichenko, M. Linkin, D. Major, M. O’Grady, L. Patrick, E.
Sussman, G. Yohe, and R. Zimmerman, 2011: Developing coastal adaptation to
climate change in New York City infrastructure-shed: Process, approach, tools,
and strategies. Climatic Change, 106(1), 93-127.
Ruppert, T.K., 2008: Eroding long-term prospects for Florida’s beaches: Florida’s
coastal construction control line program. Sea Grant Law and Policy Journal,
1(1), 65-98.
Sagala, S., N. Okada, and D. Paton, 2009: Predictors of intention to prepare for
volcanic risks in Mt. Merapi, Indonesia. Journal of Pacific Rim Psychology, 3(2),
47-54.
Sahlins, M.D., 1962: Moala: Culture and Nature on a Fijian Island. University of
Michigan Press, Ann Arbor, MI.
Salick, J. and A. Byg (eds.), 2007: Indigenous Peoples and Climate Change. Tyndall
Centre for Climate Change Research, Oxford, UK.
Salick, J. and N. Ross, 2009: Traditional peoples and climate change. Global
Environmental Change, 19, 137-139.
Satterthwaite, D., S. Huq, M. Pelling, H. Reid, and P. Lankao, 2007: Adapting to
Climate Change in Urban Areas: The Possibilities and Constraints in Low-and
Middle-Income Nations. International Institute for Environment and
Development, London, UK.
Schneider, S., 2008: Who’s to blame? (mis)perceptions of the intergovernmental
response to disasters. Disasters, 38(4), 715-738.
Schroeder, R.A., 1987: Gender Vulnerability to Drought: A Case Study of the Hausa
Social Environment. Natural Hazards Research, University of Colorado, Boulder,
CO.
Scoones, I., 1998: Sustainable Rural Livelihoods: A Framework for Analysis. IDS,
University of Sussex, Brighton, UK.
Scott, J.C., 1976: The Moral Economy of the Peasant Rebellion and Subsistence in
Southeast Asia. Yale University Press, New Haven, CT.
Seck, P., 2007: Links between Natural Disasters, Humanitarian Assistance and
Disaster Risk Reduction: A Critical Perspective. United Nations Development
Programme, New York, NY.
Semenza, J.C., 2005: Building healthy cities: A focus on interventions. In: Handbook
of Urban Health [Galea, S. and D. Vlahov (eds.)]. Spring Science and Business
Media, New York, NY, pp. 459-502.
Sen, A., 1981: Poverty and Famines: An Essay on Entitlement and Deprivation.
Clarendon, Oxford, UK.
Shaw, A., S. Sheppard, S. Burch, D. Flanders, A. Wiek, J. Carmichael, J. Robinson, and
S. Cohen, 2009: Making local futures tangible–Synthesizing, downscaling, and
visualizing climate change scenarios for participatory capacity building. Global
Environmental Change, 19, 447-463.
Shaw, R., 2006: Community-based climate change adaptation in Vietnam: Inter-
linkages of environment, disaster, and human security. In: Multiple Dimensions
of Global Environmental Changes [Sonack, S. (ed.)]. TERI Publication, Energy
and Resources Institute, New Delhi, India, pp. 521-547.
Shaw, R., A. Sharma, and Y. Takeuchi, 2009: Indigenous Knowledge and Disaster Risk
Reduction: From Practice to Policy. Nova Science Publishers, Hauppauge, NY,
pp. 409.
Shoaf, K. and S. Rottmann, 2000: The public health impact of disasters. Australian
Journal of Emergency Management, 15(3), 58-63.
Silk, J., 2000: Caring at a distance: (im)partiality, moral motivation, and the ethics of
representation – introduction. Ethics, Place & Environment, 3(3), 303-309.
Sims, H. and K. Vogelmann, 2002: Popular mobilization and disaster management in
Cuba. Public Administration and Development, 22, 389-400.
Skees, J.R. and B.J. Barnett, 2006: Enhancing microfinance using index-based risk-
transfer products. Agricultural Finance Review, 66(2), 235-250.
Skees, J.R., B.J. Barnett, and A.G. Murphy, 2008: Creating insurance markets for
natural disaster risk in lower income countries: The potential role for securitization.
Agricultural Finance Review, 68, 151-157.
Smith, S.K. and C. McCarty, 2006: Florida’s 2004 hurricane season: Demographic
response and recovery. In: Proceedings of Southern demographic association
meeting, Durham, NC, 2-4 Nov 2006.
Someshwar, S., 2008: Adaptation as ‘Climate-smart’ development. Development,
51, 366-374.
SOPAC and UNEP, 2005: Building Resilience in SIDS: The Environmental Vulnerability
Index. South Pacific Applied Geoscience Commission, Suva, Fiji.
Sorensen, J.H., 2000: Hazard warning systems: Review of 20 years of progress.
Natural Hazards Review, 1(2), 119-125.
Sorensen, J.H. and B.V. Sorensen, 2007: Community processes: Warning and
evacuation. In: Handbook of Disaster Research [Rodriquez, H., E.L. Quarantelli,
and R.R. Dynes (eds.)]. Springer, New York, NY, pp. 183-199.
Sorensen, J.H., B.L. Shumpert, and B.M. Vogt, 2004: Planning for protective action
decision making: Evacuate or shelter-in-place. Journal of Hazardous Materials,
109(1-3), 1-11.
Spanger-Siegfried, E., W. Dougherty, and B. Osman-Elasha, 2005: Methodological
Framework an Internal Scoping Report of the Project Strategies for Increasing
Human Resilience in Sudan: Lessons for Climate Change Adaptation in North
and East Africa. AIACC Working Paper No. 18, Assessments of Impacts and
Adaptations to Climate Change, Washington, DC.
Spence, R., 2004: Risk and regulation: Can improved government action reduce
the impacts of natural disasters? Building Research & Information, 32(5), 391-
405.
Sperling, F. and F. Szekely, 2005: Disaster Risk Management in a Changing Climate.
Vulnerability and Adaptation Resource Group, Washington, DC.
Stern, N., 2007: The Economics of Climate Change: The Stern Review. Cambridge
University Press, Cambridge, UK.
Stern, P. and W. Easterling (eds.), 1999: Making Climate Forecasts Matter. National
Academies Press, Washington DC, 175 pp.
Stevenson, J.R., C.T. Emrich, J.T. Mitchtell, and S.L. Cutter, 2010: Using building
permits to monitor disaster recovery: A spatio-temporal case study of coastal
Mississippi following Hurricane Katrina. Cartography and Geographic
Information Science, 37(1), 57-68.
Strobl, E., 2008: The Economic Growth Impact of Hurricanes: Evidence from US
Coastal Counties. IZA Discussion Papers Series, Institute for the Study of Labor,
Bonn, Germany.
Suarez, P., F. Ching, G. Ziervogel, I. Lemaire, D. Turnquest, J.M. de Suarez, and B.
Wisner, 2008: Video-Mediated Approaches for Community-Level Climate
Adaptation. Institute of Development Studies, University of Sussex, Brighton,
UK.
Sudmeier-Rieux, K., H. Masundire, A. Rizvi, and S. Rietbergen, 2006: Ecosystems,
Livelihoods And Disasters: An Integrated Approach to Disaster Risk
Management. International Union for Conservation of Nature and Natural
Resources, Gland, Switzerland, 56 pp.
Sundar, N. and R. Jeffery, 1999: A New Moral Economy for Indian Forests?
Discourses of Community and Participation. Sage, New Delhi, India, 304 pp.
Susman, P., P. O’Keefe, and B. Wisner, 1983: Global disasters, a radical reinterpretation.
In: Interpretations of Calamity from the Viewpoint of Human Ecology [Hewitt,
K. (ed.)]. Allen and Unwin, Boston, MA, pp. 263-283.
Tacoli, C., 2009: Crisis or adaptation? Migration and climate change in a context of
high mobility. Environment and Urbanization, 21(2), 513-525.
Tanner, T.M. and T. Mitchell, 2008: Poverty in a changing climate. IDS Bulletin, 39(4),
1-5.
Tanner, T., T. Mitchell, E. Polack, and B. Guenther, 2009: Urban Governance for
Adaptation: Assessing Climate Change for Resilience in Ten Asian Cities.
Institute of Development Studies, Brighton, UK.
Tanser, F.C., B. Sharp, and D. le Sueur, 2003: Potential effect of climate change on
malaria transmission in Africa. The Lancet, 362(9398), 1792-1798.
Tearfund, 2006: Adapting to Climate Change: Challenges and Opportunities for the
Development Community. Institute of Development Studies, Brighton, UK.
Thacker, M.T.F., R. Lee, R.I. Sabogal, and A. Henderson, 2008: Overview of deaths
associated with natural events, United States, 1979-2004. Disasters, 32(2),
303-315.
Thieken, A.H., T. Petrow, and H. Kreibich, and B. Merz, 2006: Insurability and
mitigation of flood losses in private households in Germany. Risk Analysis,
26(2), 383-395.
Chapter 5Managing the Risks from Climate Extremes at the Local Level
337
Thomalla, F., T. Downing, E. Spanger-Siegfried, G. Han, and J. Rockström, 2006:
Reducing hazard vulnerability: Towards a common approach between disaster
risk reduction and climate adaptation. Disasters, 30(1), 39-48.
Thomas , D.S.G. and C. Twyman, 2005: Equity and justice in climate change adaptation
among natural-resource-dependent societies. Global Environmental Change,
15, 115-124.
Thomson, R., 2007: Cultural models and shoreline social conflict. Coastal
Management, 35, 211-237.
Tierney, K., 2006: Social inequality, hazards, and disasters. In: On Risk and Disaster:
Lessons from Hurricane Katrina [Daniels, R.J., D.F. Kettl, and H. Kunreuther
(eds.)]. University of Pennsylvania Press, Philadelphia, PA, pp. 109-128.
Tierney, K.J., M.K. Lindell, and R.W. Perry, 2001: Facing the Unexpected: Disaster
Preparedness and Response in the United States. National Academies Press,
Washington, DC.
Tierney, K., C. Bevc, and E. Kuligowski, 2006: Metaphors matter: Disaster myths,
media frames, and their consequences in hurricane Katrina. Annals of the
American Academy of Political and Social Science, 604(1), 57-81.
Titus, J.G., 2011: Rolling Easements. United States Environmental Protection Agency,
Washington, DC, 166 pp.
Titus, J., D. Hudgens, D. Trescott, M. Craghan, W. Nuckols, C. Hershner, J. Kassakian,
C. Linn, P. Merritt, J. McCue, J. O’Connell, J. Tanski, and J. Wang, 2009: State and
local governments’ plan for development of most land vulnerable to rising sea
level along the US Atlantic coast. Environmental Research Letters, 4(4),
044008.
Tobin, G.A., 1995: The levee love affair: A stormy relationship. Water Resources
Bulletin, 31, 359-367.
Tol, R.S.J., 2003: Is the uncertainty about climate change too large for expected cost-
benefit analysis? Climatic Change, 56(3), 265-289.
Tompkins, E.L., 2005: Planning for climate change in small islands: Insights from
national hurricane preparedness in the Cayman Islands. Global Environmental
Change A, 15(2), 139-149.
Torresan, S., A. Critto, M.D. Valle, N. Harvey, and A. Marcomini, 2008: Assessing
coastal vulnerability to climate change: Comparing segmentation at global and
regional scales. Sustainability Science, 3(1), 45-65.
Travis, W., 2010: Going to extremes: Propositions on the social response to severe
climate change. Climatic Change, 98, 1-19.
Trawick, P., 2001: The moral economy of water: Equity and antiquity in the Andean
commons. American Anthropologist, 103(2), 361-379.
Tschakert, P. and K. Dietrich, 2010: Anticipatory learning for climate change
adaptation and resilience. Ecology and Society, 15(2), 11.
Tsuchiya, S., H. Tatano, and N. Okada, 2007: Economic loss assessment due to
railroad and highway disruptions. Economic Systems Research, 19(2), 147-162.
Turner, N.J. and H. Clifton, 2009: “It’s so different today”: Climate change and
indigenous lifeways in British Columbia, Canada. Global Environmental
Change, 19(2), 180-190.
Turvey, C., 2001: Weather insurance and specific event risks in agriculture. Review
of Agricultural Economics, 23(2), 333-351.
Twigg, J., 2001: Sustainable Livelihoods and Vulnerability to Disasters. Benfield Greig
Hazard Research Centre, Disaster Management Working Paper 2/2001,
University College London, London, UK.
Twigg, J., 2007: Characteristics of a Disaster-Resilient Community: A Guidance Note.
DFID Disaster Risk Reduction Interagency Coordination Group, Benfield, UK.
U.S. Bureau of Labor Statistics, 2006: The labor market impact of Hurricane
Katrina: An overview. Monthly Labor Review, August 2006, 3.
UNDP, 2004: Reducing Disaster Risk: A Challenge for Development. United Nations
Development Programme, New York, NY.
UNDP, 2007: Human Development Report 2007/2008: Fighting Climate Change:
Human Solidarity in a Divided World. United Nations Development Programme,
Palgrave Macmillan, Hampshire, UK.
UNFPA, 2009: Facing a Changing World: Women, Population and Climate. United
Nations Population Fund, New York, NY.
UN-HABITAT, 2007: Global Report on Human Settlements, 2007: Enhancing Urban
Safety and Security. Earthscan, London, UK.
UNIFEM, 2011: Pakistan Floods 2010: Rapid Gender Needs Assessment of Flood
Affected Communities. UN Development Fund for Women, Geneva, Switzerland.
UNISDR, 2004: Living With Risk: A Global Review of Disaster Reduction Initiatives.
United Nations International Strategy for Disaster Reduction, Geneva,
Switzerland.
UNISDR., 2006: Global Survey of Early Warning Systems: An Assessment of
Capacities, Gaps and Opportunities Toward Building a Comprehensive Global
Early Warning System for all Natural Hazards. United Nations International
Strategy for Disaster Reduction, Geneva, Switzerland, 56 pp.
UNISDR, 2009: Global Assessment Report on Disaster Risk Reduction: Risk and
Poverty in a Changing Climate. United Nations International Strategy for
Disaster Reduction, Geneva, Switzerland.
UNISDR and UNOCHA, 2008: Disaster Preparedness for Effective Response,
Guidance and Indicator Package for Implementing Priority Five of the Hyogo
Framework. United Nations Secretariat of the International Strategy for Disaster
Reduction (UNISDR) and the United Nations Office for Coordination of
Humanitarian Affairs (UNOCHA), Geneva, Switzerland, 51 pp.
UNOCHA and IDMC, 2009: Monitoring Disaster Displacement in the Context of
Climate Change. Findings of a Study by the United Nations Office for the
Coordination of Humanitarian Affairs and the Internal Displacement Monitoring
Centre, Norwegian Refugee Council, Internal Displacement Monitoring Centre,
Geneva, Switzerland, 32 pp.
Valdes, H.M., 1997: Community-operated early warning system for floods. In: UN-
IDNDR & QUIPUNET Internet conference on Floods, Drought: Issues for the 21st
Century. 22 September – 24 October 1997.
Valente, T.W. and R.L. Davis, 1999: Accelerating the diffusion of innovations using
opinion leaders. Annals of the American Academy of Political and Social
Science, 566, 55-67.
Valente, T.W., B.R. Hoffman, A. Ritt-Olson, K. Lichtman, and C.A. Johnson, 2003:
Effects of a social-network method for group assignment strategies on peer-led
tobacco prevention programs in schools. American Journal of Public Health, 93,
1837-1843.
van Aalst, M.K., 2006: The impacts of climate change on the risk of natural disasters.
Disasters, 30(1), 5-18.
van Aalst, M., K. Maarten, T. Cannon, and I. Burton, 2008: Community level
adaptation to climate change: The potential role of participatory community
risk assessment. Global Environmental Change: Human and Policy Dimensions,
18(1), 165-179.
Van Willigen, M., T. Edwards, B. Edwards, and S. Hessee, 2002: Riding out the storm:
Experiences of the physically disabled during Hurricanes Bonnie, Dennis, and
Floyd. Natural Hazards Review, 3(3), 98-106.
Vellinga, P., E. Mills, G. Berz, L. Bouwer, S. Huq, L.A. Kozak, J. Palutikof, B.
Schanzenbacher, and G. Soler, 2001: Insurance and other financial services. In:
Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the
Third Assessment Report of the Intergovernmental Panel on Climate Change
[McCarthy, J.J., O.F. Canziani, N.A. Leary, D.J. Dokken, and K.S. White (eds.)].
Cambridge University Press, Cambridge, UK, pp. 417-450.
Venton, C.C. and P. Venton, 2004: Disaster Preparedness Programmes in India: A
Cost Benefit Analysis. HPN Network Paper No. 49, Overseas Development
Institute, London, UK.
Vescovi, L., A. Bourque, G. Simonet, and A. Musy, 2009: Transfer of climate knowledge
via a regional climate-change management body to support vulnerability,
impact assessments and adaptation measures. Climate Research, 40(2-3), 163-
173.
Vörösmarty, C., G. Green, J. Salisbury, and R. Lammers, 2000: Global water
resources: Vulnerability from climate change and population growth. Science,
289(5477), 284-289.
Wachtendorf, T., J.M. Kendra, H. Rodriquez, and J. Trainor, 2006: The social impacts
and consequences of the December 2004 Indian ocean tsunami: Observations
from India and Sri Lanka. Earthquake Spectra, 22(3), 693-714.
Waddell, E., 1989: Observations on the 1972 frosts and subsequent relief pro-
gramme among the Enga of the western highlands. Mountain Research and
Development, 9(3), 210-223.
Chapter 5 Managing the Risks from Climate Extremes at the Local Level
338
Walker, I.J. and R. Sydneysmith, 2008: British Columbia. In: From Impacts to
Adaptation: Canada in a Changing Climate 2007 [Lemmen, D.S., F.J. Warren, J.
Lacroix, and E. Bush (eds.)]. Government of Canada, Ottawa, Canada, pp.
358-362.
Walker, P., 1989: Famine Early Warning Systems: Victims and Destitution. Earthscan
Publications, London, UK.
Wamsler, C., 2007: Bridging the gaps: Stakeholder-based strategies for risk reduction
and financing for the urban poor. Environment and Urbanization, 19(1), 115-142.
Ward, P.J., W. Beets, L.M. Bouwer, J.C.J.H. Aerts, and H. Renssen, 2010: Sensitivity of
river discharge to ENSO. Geophysical Research Letters, 37(L12402),
doi:10.1029/2010GL043215.
Ward, R.E.T., C. Herweijer, N. Patmore, and R. Muir-Wood, 2008: The role of insurers
in promoting adaptation to the impacts of climate change. The Geneva Papers,
33, 133-139.
Warner, K., M. Hamza, A. Oliver-Smith, F. Renaud, and A. Julca, 2010: Climate
change, environmental degradation and migration. Natural Hazards, 55(3),
689-715.
Watts, M., 1983: Hazards and crises: A political economy of drought and famine in
northern Nigeria. Antipode, 15(1), 24-34.
Waugh, W.L.J. and G. Streib, 2006: Collaboration and leadership for effective
emergency management. Public Administration Review, 66, 131-140.
Weems, C.F. and S. Overstreet, 2008: Child and adolescent mental health research in
the context of Hurricane Katrina: A ecological needs-based perspective and
introduction to the special section. Journal of Clinical Child and Adolescent
Psychology, 37(3), 487-494.
Weibel, W.W., 1988: Combining ethnographic and epidemiologic methods in targeted
AIDS interventions: The Chicago model. In: Needle Sharing among Intravenous
Drug Abusers: National and International Perspectives [Battjes, R.J. and R.W.
Pickens (eds.)]. NIDA Research Monograph 80, National Institute on Drug
Abuse, Rockville, MD, pp. 137-150.
Weichselgartner, J. and R.E. Kasperson, 2010: Barriers in the science-policy-practice
interface: Toward a knowledge-action-system in global environmental change
research. Global Environmental Change, 20, 266-277.
Weisler, R.H., J.G. Barbee IV, and M.H. Townsend, 2006: Mental health and recovery
in the Gulf Coast after Hurricanes Katrina and Rita. Journal of the American
Medical Association, 296, 585-588.
Weissbecker, I., S.E. Sephton, M.B. Martin, and D.M. Simpson, 2008: Psychological
and physiological correlates of stress in children exposed to disaster: Current
research and recommendations for intervention. Children, Youth and
Environments, 18(1), 30-70.
Weitzman, M.L., 2007: A review of the Stern Review on the economics of climate
change. Journal of Economic Literature, 45, 703-724.
Wellman
, B. and S.D. Berkowitz, 1988: Social Structures: A Network Approach.
Cambridge University Press, Cambridge, UK.
Wenzel, F., F. Bendimerad, and R. Sinha, 2007: Megacities–megarisks. Natural
Hazards, 42(3), 481-491.
West, C.T. and D.G. Lenze, 1994: Modeling the regional impact of natural disasters
and recovery: A general framework and an application to Hurricane Andrew.
International Regional Science Review, 17, 121-150.
White, G.F., R.W. Kates, and I. Burton, 2001: Knowing better and losing even more:
The use of knowledge in hazards management. Global Environmental Change
Part B: Environmental Hazards, 3, 81-92.
Whitehead, J.C., B. Edwards, M. Van Willigen, J.R. Maiolo, K. Wilson, and K.T. Smith,
2000: Heading for higher ground: Factors affecting real and hypothetical
hurricane evacuation behavior. Environmental Hazards, 2(4), 133-142.
Wiest, R.E., J.S.P. Mocellin, and D.T. Motsisi, 1994: The Needs of Women in Disasters
and Emergencies. Prepared for the Disaster Management Training Programme
of the United Nations Development Programme and the Office of the United
Nations Disaster Relief Coordinator, Disaster Research Unit, University of
Manitoba, Winnipeg, Canada.
Wilby, R.L., J. Troni, Y. Biot, L. Tedd, B.C. Hewitson, D.C. Smith, and R.T. Sutton, 2009:
A review of climate risk information for adaptation and development planning.
International Journal of Climatology, 29(9), 1193-1215.
Wilhelmi, O.V. and D.A. Wiilhite, 2002: Assessing vulnerability to agricultural
drought: A Nebraska case study. Natural Hazards, 25(1), 37-58.
Williams, R., D.A. Alexander, D. Bolsover, and F.K. Bakke, 2008: Children, resilience
and disasters: Recent evidence that should influence a model of psychosocial
care. Current Opinion in Psychiatry, 21(4), 338-344.
Winsvold, M., K.B. Stokke, J.E. Klausen, and I. Saglie, 2009: Organizational learning
and governance in adaptation in urban development. In: Adapting to Climate
Change: Thresholds, Values and Governance [Adger, W.N., I. Lorenzoni, and K.
O’Brien (eds.)]. Cambridge University Press, Cambridge, UK, pp. 476-490.
Wisner, B., 2002: Disability and disaster: Victimhood and agency in earthquake risk
reduction. In: Earthquakes.. [Rodrigue, C. and E. Rovai (eds.)]. Routledge,
London, UK.
Wisner, B., 2003: Disaster risk reduction in megacities: Making the most of human
and social capital. In: Building Safer Cities: The Future of Disaster Risk [Kreimer,
A., M. Arnold, and A. Carlin (eds.)]. World Bank Books, Washington, DC, pp.
181-196.
Wisner, B. and J. Uitto, 2009: Life on the edge: Urban social vulnerability and
decentralized, citizen-based disaster risk reduction in four large cities of the
Pacific Rim. Facing Global Environmental Change: Hexagon Series on Human
and Environmental Security and Peace, 4(II), 215-231.
Wisner, B., P. Blaike, T. Cannon, and I. Davis, 2004: At Risk: Natural Hazards, People’s
Vulnerability and Disasters. Routledge, London, UK.
WMO, 2003: Integrated Flood Plain Management Case Study: Bangladesh Flood
Management. World Meteorological Organization, Geneva, Switzerland.
World Bank, 2010: Natural Hazards, Unnatural Disasters, the Economics of Effective
Prevention. World Bank, Washington, DC.
Wu, S., B. Yarnal, and A. Fisher, 2002: Vulnerability of coastal communities to sea-level
rise: A case study of Cape May county, New Jersey, USA. Climate Research, 22,
255-270.
Yang, S.L., J.D. Milliman, P. Li, and K. Xu, 2011: 50,000 dams later: Erosion of the
Yangtze River and its delta. Global and Planetary Change, 75(1), 14-20.
Yohe, G., J. Neumann, and H. Ameden, 1995: Assessing the economic cost of greenhouse
induced sea level rise: Methods and applications in support of a national
survey. Journal of Environmental Economics and Management, 29, 78-97.
Zahran, S., S.D. Brody, W.G. Peacock, A. Vedlitz, and H. Gover, 2008: Social
vulnerability and the natural and built environment: A model of flood casualties
in Texas. Disasters, 32(4), 537-560.
Ziervogel, G., 2004: Targeting seasonal climate forecasts for integration into
household level decisions: The case of smallholder farmers in Lesotho. The
Geographical Journal, 170, 6-21.
Ziervogel, G. and P. Ericksen, 2010: Adapting to climate change to sustain food
security. Wiley Interdisciplinary Reviews: Climate Change, 1(4),
doi:10.1002/wcc.56.
Ziervogel, G. and F. Zermoglio, 2009: Climate-change scenarios and the development
of adaptation strategies in Africa: Challenges and opportunities. Climate
Research, 40(2-3), 133-146.
Ziervogel, G., A.O. Nyong, B. Osman-Elasha, C. Conde, S. Cortés, and T. Downing,
2006: Climate Variability and Change: Implications for Household Food
Security. AIACC Working Paper No. 20, Assessments of Impacts and Adaptations
to Climate Change, Washington, DC.
Ziervogel, G., M. Shale, and M. Du, 2010: Climate change adaptation in a developing
country context: The case of urban water supply in Cape Town. Climate and
Development, 2, 94-110.
Zimmermann, M. and F. Stössel, 2011: Disaster Risk Reduction in International
Cooperation: Switzerland‘s Contribution to the Protection of Lives and
Livelihoods. Swiss Agency for Development and Cooperation, Berne,
Switzerland, 23 pp.
Chapter 5Managing the Risks from Climate Extremes at the Local Level
339
Coordinating Lead Authors:
Padma Narsey Lal (Australia), Tom Mitchell (UK)
Lead Authors:
Paulina Aldunce (Chile), Heather Auld (Canada), Reinhard Mechler (Germany), Alimullah Miyan
(Bangladesh), Luis Ernesto Romano (El Salvador), Salmah Zakaria (Malaysia)
Review Editors:
Andrew Dlugolecki (UK), Takuo Masumoto (Japan)
Contributing Authors:
Neville Ash (Switzerland), Stefan Hochrainer (Austria), Robert Hodgson (UK), Tarik ul Islam (Canada),
Sabrina McCormick (USA), Carolina Neri (Mexico), Roger Pulwarty (USA), Ataur Rahman (Bangladesh),
Ben Ramalingam (UK), Karen Sudmeier-Reiux (France), Emma Tompkins (UK), John Twigg (UK),
Robert Wilby (UK)
This chapter should be cited as:
Lal, P.N., T. Mitchell, P. Aldunce, H. Auld, R. Mechler, A. Miyan, L.E. Romano, and S. Zakaria, 2012: National systems for
managing the risks from climate extremes and disasters. In: Managing the Risks of Extreme Events and Disasters to Advance
Climate Change Adaptation [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach,
G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups I and II of the
Intergovernmental Panel on Climate Change (IPCC). Cambridge University Press, Cambridge, UK, and New York, NY, USA,
pp. 339-392.
6
National Systems
for Managing the Risks
from Climate Extremes
and Disasters
National Systems for Managing the Risks from Climate Extremes and Disasters
340
Executive Summary .................................................................................................................................341
6.1. Introduction..............................................................................................................................344
6.2. National Systems and Actors for Managing the Risks
from Climate Extremes and Disasters......................................................................................345
6.2.1. National and Sub-National Governments........................................................................................................................346
6.2.2. Private Sector Organizations ...........................................................................................................................................347
6.2.3. Civil Society and Community-Based Organizations.........................................................................................................348
6.2.4. Bilateral and Multilateral Agencies .................................................................................................................................348
6.2.5. Research and Communication..........................................................................................................................................349
6.3. Planning and Policies for Integrated Risk Management,
Adaptation, and Development Approaches .............................................................................349
6.3.1. Developing and Supporting National Planning and Policy Processes.............................................................................349
6.3.2. Mainstreaming Disaster Risk Management and Climate Change Adaptation into Sectors and Organizations .............355
6.3.3. Sector-Based Risk Management and Adaptation.............................................................................................................357
6.4. Strategies including Legislation, Institutions, and Finance .....................................................357
6.4.1. Legislation and Compliance Mechanisms ........................................................................................................................358
6.4.2. Coordinating Mechanisms and Linking across Scales......................................................................................................358
6.4.3. Finance and Budget Allocation ........................................................................................................................................360
6.5. Practices including Methods and Tools....................................................................................362
6.5.1. Building a Culture of Safety.............................................................................................................................................362
6.5.1.1. Assessing Risks and Maintaining Information Systems ....................................................................................................................363
6.5.1.2.
Preparedness: Risk Awareness, Education, and Early Warning Systems............................................................................................364
6.5.2. Reducing Climate-Related Disaster Risk...
....
...................................................................................................................366
6.5.2.1. Applying Technological and Infrastructure-Based Approaches..........................................................................................................366
6.5.2.2.
Human Development and Vulnerability Reduction ...........................................................................................................................368
6.5.2.3. Investing in Natural Capital and Ecosystem-Based Adaptation........................................................................................................370
6.5.3. Transferring and Sharing ‘Residual’ Risks...
....
.................................................................................................................371
6.5.4. Managing the Impacts .....................................................................................................................................................373
6.6. Aligning National Disaster Risk Management Systems
with the Challenges of Climate Change ..................................................................................375
6.6.1. Assessing the Effectiveness of Disaster Risk Management in a Changing Climate........................................................375
6.6.2. Managing Uncertainties and Adaptive Management in National Systems.....................................................................377
6.6.3. Tackling the Underlying Drivers of Vulnerability .............................................................................................................379
6.6.4. Approaching Disaster Risk, Adaptation, and Development Holistically..........................................................................379
References ...............................................................................................................................................381
Chapter 6
Table of Contents
341
This chapter assesses how countries are managing current and projected disaster risks, given knowledge of how risks
are changing with observations and projections of weather and climate extremes [Table 3-2, 3.3], vulnerability and
exposure [4.3], and impacts [4.4]. It focuses on the design of national systems for managing such risks, the roles
played by actors involved in the system, and the functions they perform, acknowledging that complementary actions
to manage risks are also taken at local and international level as described in Chapters 5 and 7.
National systems are at the core of countries’ capacity to meet the challenges of observed and projected
trends in exposure, vulnerability, and weather and climate extremes (high agreement, robust evidence).
Effective national systems comprise multiple actors from national and sub-national governments, private sector,
research bodies, and civil society, including community-based organizations, playing differential but complementary
roles to manage risk according to their accepted functions and capacities. These actors work in partnership across
temporal, spatial, administrative, and social scales, supported by relevant scientific and traditional knowledge. Specific
characteristics of national systems vary between countries and across scales depending on their socio-cultural, political,
and administrative environments and development status. [6.2]
The national level plays a key role in governing and managing disaster risks because national government
is central to providing risk management-related public goods as it commonly maintains financial and
organizational authority in planning and implementing these goods (high agreement, robust evidence).
National governments are charged with the provision of public goods such as ensuring the economic and social well-
being, safety, and security of their citizens from disasters, including the protection of the poorest and most vulnerable
citizens. They also control budgetary allocations as well as creating legislative frameworks to guide actions by other
actors. Often, national governments are considered to be the ‘insurer of last resort’. In line with the delivery of public
goods, national governments and public authorities ‘own’ a large part of current and future disaster risks (public
infrastructure, public assets, and relief spending). In terms of managing risk, national governments act as risk
aggregators and by pooling risk, hold a large portfolio of public liabilities. This provides governments responsibility to
accurately quantify and manage risks associated with this portfolio – functions that are expected to become more
important given projected impacts of climate change and trends in vulnerability and exposure. [6.2.1]
In providing such public goods, governments choose to manage disaster risk by enabling national systems
to guide and support stakeholders to reduce risk where possible, transfer risk where feasible, and manage
residual risk, recognizing that risks can never be totally eliminated (high agreement, robust evidence). The
balance between reducing risk and other disaster risk management strategies is influenced by a range of factors,
including financial and technical capacity of stakeholders, robustness of risk assessment information, and cultural
elements involving risk tolerance. [6.2.1, 6.2-6.5]
The ability of governments to implement disaster risk management responsibilities differs significantly
across countries, depending on their capacity and resource constraints (high agreement, robust evidence).
Smaller or economically less-diversified countries face particular challenges in providing the public goods associated
with disaster risk management, in absorbing the losses caused by climate extremes and disasters, and in providing
relief and reconstruction assistance. [6.4.3] However, there is limited evidence to suggest any correlation between the
type of governance system in a country (e.g., centralized or decentralized; unitary or federal) and the effectiveness of
disaster risk management efforts. There is robust evidence and high agreement to suggest that actions generated
within and managed by communities with supporting government policies are generally most effective since they are
specific and tailored to local environments. [6.4.2]
In the majority of countries, national systems have been strengthened by applying the principles of the
Hyogo Framework for Action to mainstream risk considerations across society and sectors, although
greater efforts are required to address the underlying drivers of risk and generate the political will to
invest in disaster risk reduction (high agreement, robust evidence). The Hyogo Framework for Action has
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
Executive Summary
342
encouraged countries to develop and implement a systematic disaster risk management approach, and in some cases
has led to strategic shifts in the management of disaster risks, with governments and other actors placing greater
attention on disaster risk reduction compared to more reactive measures. This has included improvements in coordination
between actors, enhanced early warning and preparedness, more rigorous risk assessments, and increased awareness.
However, there is limited evidence and low agreement to suggest improvements in integration between efforts to
implement the Hyogo Framework for Action, the United Nations Framework Convention on Climate Change, and
broader development and environmental policy frameworks. [6.4.2]
A set of factors can be identified that make efforts to systematically manage current disaster risks more successful (all
high agreement, robust evidence). Systems to manage current disaster risk are more successful if:
Risks are recognized as dynamic and are mainstreamed and integrated into development policies, strategies, and
actions, and into environmental management. [6.3.1]
Legislation for managing disaster risks is supported by clear regulations that are effectively enforced across
scales and complemented by other sectoral development and management legislations where risk considerations
are explicitly integrated. [6.4.1]
Disaster risk management functions are coordinated across sectors and scales and led by organizations at the
highest political level. [6.4.2]
They include considerations of disaster risk in national development and sector plans, and, if they adopt climate
change adaptation strategies, translating these plans and strategies into actions targeting vulnerable areas and
groups. [6.5.2]
Risk is quantified and factored into national budgetary processes, and a range of measures including budgeting for
relief expenditure, reserve funds, and other forms of risk financing have been considered or implemented. [6.4.3]
Decisions are informed by comprehensive information about observed changes in weather, climate, and vulnerability
and exposure, and historic disaster losses, using a diversity of readily available tools and guidelines. [6.5.1]
Early warning systems deliver timely, relevant, and accurate predictions of hazards, and are developed and made
operational in partnership with the public and trigger effective response actions. [6.5.1]
Strategies include a combination of hard infrastructure-based options responses and soft solutions such as
individual and institutional capacity building and ecosystem-based responses, including conservation measures
associated with, for example, forestry, river catchments, coastal wetlands, and biodiversity. [6.5.2]
While there is robust evidence and high agreement on efforts to tackle current disaster risks, the assessment
found limited evidence of national disaster risk management systems and associated risk management
measures explicitly integrating knowledge of and uncertainties in projected changes in exposure,
vulnerability, and climate extremes. The effectiveness of efforts to manage projected disaster risks at the national
level are dependent on a range of factors, including the effectiveness of the system for managing current risks, the
ability of the system to flexibly respond to new knowledge, the availability of suitable data, and the resources available
to invest in longer-term risk reduction and adaptation measures. Developed countries are better equipped financially
and institutionally to adopt explicit measures to effectively respond and adapt to projected changes in exposure,
vulnerability, and climate extremes than developing countries. Nonetheless, all countries face challenges in assessing,
understanding, and then responding to such projected changes. [6.3.2, 6.6]
Measures that provide benefits under current climate and a range of future climate change scenarios,
called low-regrets measures, are available starting points for addressing projected trends in exposure,
vulnerability, and climate extremes. They have the potential to offer benefits now and lay the foundation
for addressing projected changes (high agreement, medium evidence). The assessment considered such ‘low’
regrets options across a range of key sectors, with some of the most commonly cited measures associated with
improvements to early warning systems, health surveillance, water supply, sanitation, and drainage systems; climate
proofing of major infrastructure and enforcement of building codes; better education and awareness; and restoration
of degraded ecosystems and nature conservation. Many of these low-regrets strategies produce co-benefits; help
address other development goals, such as improvements in livelihoods, human well-being, and biodiversity conservation;
and help minimize the scope for maladaptation. [6.3.1, Table 6-1]
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
343
Ecosystem-based solutions in the context of changing climate risks can offer ‘triple-win’ solutions, as they
can provide cost-effective risk reduction, support biodiversity conservation, and enable improvements in
economic livelihoods and human well-being, particularly for the poor and vulnerable (high agreement,
robust evidence). The assessment found that such ecosystem-based adaptation strategies, including mangrove
conservation and rehabilitation, integrated catchment management, and sustainable forest and fisheries management,
also minimize the scope for maladaptation in developed and developing countries. In choosing amongst ecosystem-
based adaptation options, decisionmakers may need to make tradeoffs between particular climate risk reduction
strategies and other valued ecosystem services. [6.5.2]
Insurance-related instruments are key mechanisms for helping households, business, and governments
absorb the losses from disasters; but their uptake is unequally distributed across regions and hazards, and
often public-private partnerships are required (high agreement, robust evidence). Disaster insurance and
other risk transfer instruments covered about 20% of reported weather-related losses over the period 1980 to 2003.
Distribution, though, is uneven, with about 40% of the losses insured in high-income as compared to 4% of losses in
low-income countries. Existing national insurance systems differ widely as to whether policies are compulsory or
voluntary, and importantly in how systems allocate liability and responsibility for disaster risks across society. With
changing weather and extreme events, vulnerability, and exposure, extended and innovative private-public sector
partnerships are required to better estimate and price risk as well as to develop robust insurance-related products,
which may be supported in developing countries by development partner funds. [6.5.3]
Pooling of risk by and between national governments contributes to reducing the fiscal and socioeconomic
consequences of disasters (medium agreement, medium evidence). As national governments hold a large
portfolio of public liabilities (infrastructure, public assets, and the provision of disaster relief), risk aggregation and
pooling are expected to become more important given projected impacts of climate change and trends in vulnerability
and exposure. In addition, particularly for small, low-income, and highly exposed countries, risk transfer of public sector
assets and relief expenditure recently have become a cornerstone of disaster risk reduction. Key innovative and promising
applications recently implemented comprise sovereign insurance for hurricane risk, insurance for humanitarian assistance
following droughts, and intergovernmental risk pooling. [6.4.3, 6.5.3]
Flexible and adaptive national systems are better suited to manage projected trends and associated
uncertainties in exposure, vulnerability, and weather and climate extremes than static and rigid national
systems (high agreement, limited evidence). Adaptive management brings together different scientific, social, and
economic information, experiences, and traditional knowledge into decisionmaking through ‘learning by doing.’ Multi-
criteria analysis, scenario planning, and flexible decision paths offer options for taking action when faced with large
uncertainties or incomplete information. National systems for managing disaster risk can adapt to climate change and
shifting exposure and vulnerability by (i) frequently assessing and mainstreaming knowledge of dynamic risks;
(ii) adopting ‘low regrets’ strategies; (iii) improving learning and feedback across disaster, climate, and development
organizations at all scales; (iv) addressing the root causes of poverty and vulnerability; (v) screening investments for
climate change-related impacts and risks to minimize scope for maladaptation; and (vi) increasing standing capacity
for emergency response as climatic conditions change over time. [6.6.1, 6.6.2, 6.6.4]
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
344
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
6.1. Introduction
The socioeconomic impacts of disaster events can be significant in all
countries, but low- and middle-income countries are especially vulnerable,
and experience higher fatalities even when exposed to hazards of similar
magnitude (O’Brien et al., 2006; Thomalla et al., 2006; Ibarraran et al.,
2009; IFRC, 2010). The number of deaths per cyclone event in the last
several decades, for example, was highest in low-income countries even
though a higher proportion of population exposed to cyclones lives in
countries with higher income; 11% of the people exposed to hazards
live in low human development countries, but they account for more
than 53% of the total recorded deaths resulting from disasters (UNDP,
2004a). At the same time, while in absolute terms the direct economic
losses from disasters are far greater in high-income countries, middle-
and low-income states bear the heaviest burden of these costs in terms
of damage relative to annual gross domestic product (GDP: UNDP,
2004a; DFID, 2005; O’Brien et al., 2006; Kellenberg et al., 2008; Pelham
et al., 2011). This burden has been increasing in the middle-income
countries, where the asset base is rapidly expanding and losses over the
period from 2001 to 2006 amounted to about 1% of GDP. For the low-
income group, losses totaled an average of 0.3% and for the high-
income countries amounted to less than 0.1% of GDP (Cummins and
Mahul, 2009). In some particularly exposed countries, including many
small island developing states, these wealth losses expressed as a
percentage of GDP can be considerably higher, with the average costs
over disaster and non-disaster years close to 10%, such as reported for
Grenada and St. Lucia (World Bank and UN, 2010). In extreme cases, the
costs of individual events can be as high as 200% of the annual GDP as
experienced in the Polynesian island nation of Niue following cyclone
Heta in 2004, or in the Hurricane Ivan event affecting Grenada in 2004
(McKenzie et al., 2005).
In terms of the macroeconomic and developmental consequences of
high exposure to disaster risk, a growing body of literature has shown
significant adverse effects in developing countries (Otero and Marti,
1995; Charveriat, 2000; Crowards, 2000; Murlidharan and Shah, 2001;
ECLAC, 2002, 2003; Mechler, 2004; Hochrainer, 2006; Noy, 2009). These
include reduced direct and indirect tax revenue, dampened investment,
and reduced long-term economic growth through their negative effect
on a country’s credit rating and an increase in interest rates for external
borrowing. Among the reasons behind limited coping capacity of
individuals, communities, and governments are reduced tax bases and
high levels of indebtedness, combined with limited household income and
savings, a lack of disaster risk transfer and other financing instruments,
few capital assets, and limited social insurance.
This body of evidence emphasizes that disasters can cause a setback for
development, and even a reversal of recent development gains in the
short- to medium-term, emphasizing the point that disaster risk
management is a development issue as much as a humanitarian one.
Poor development status of communities and countries increases their
sensitivity to disasters. Disaster impacts can also force households to fall
below the basic needs poverty line, further increasing their vulnerability
to other shocks (Owens et al., 2003; Lal, 2010). Consequently, disasters
are seen as barriers for development, requiring ex-ante disaster risk
reduction policies that also target poverty and development (del Ninno
et al., 2003; Owens et al., 2003; Skoufias, 2003; Benson and Clay, 2004;
Hallegatte et al., 2007; Raddatz, 2007; Cardona et al., 2010; IFRC, 2010).
However, some literature suggests that disasters may not always have
a negative effect on economic growth and development and for
some countries disasters may be regarded as a problem of, and not for
development (Albala-Bertrand, 1993; Skidmore and Toya, 2002; Caselli
and Malthotra, 2004; Hallegatte and Ghil, 2007). Disasters have also
been considered to increase economic growth in the short term as well
as spur positive economic growth and technological renewal in the
longer term, depending on the domestic capacity of nations to rebuild
and the inflow of international assistance (Skidmore and Toya, 2002).
This observation may be partially attributable to national accounting
practices, which positively record reconstruction efforts but do not
account for the immediate destruction of assets and wealth in some
cases (Skidmore and Toya, 2002).
To better respond to the impacts of disasters on human livelihoods,
environment, and economies, national disaster risk management systems
have evolved in recent years, guided in some cases by international
instruments, particularly the Hyogo Framework for Action (HFA) 2005-2015
and more recently as part of the adaptation agenda under the United
Nations Framework Convention on Climate Change (UNFCCC; see
Section 7.3). Increasing knowledge, understanding, and experiences in
dealing with disaster risks have gradually contributed to a paradigm
shift globally that recognizes the importance of reducing risks by
addressing underlying drivers of vulnerability and exposure, such as
targeting poverty, improving human well-being, better environmental
management, and adaptation to climate change as well as responding
to and rebuilding after disaster events (Yodmani, 2001; IFRC, 2004,
2010; Thomalla et al., 2006; UNISDR, 2008a; Venton and LaTrobe, 2008;
Pelham et al., 2011). While governments cannot act alone, the majority
are well placed and equipped to support communities and the private
sector to address disaster risks. Yet recent reported experiences suggest
that countries vary considerably in their responses, and concerns remain
about the lack of integration of disaster risk management into sustainable
development policies and planning as well as insufficient implementation
at different levels (CCCD, 2009; UNFCCC, 2008b).
It is at the national level that overarching development policies and
legislative frameworks are formulated and implemented to create
appropriate enabling environments to guide other stakeholders to
reduce, share, and transfer risks, albeit in different ways (Carter, 1992;
Freeman et al., 2003). National-level governments in developed countries
are often the de facto ‘insurers of last resort’ and used to be considered
the most effective insurance instruments of society (Priest, 1996).
Governments also have the ability to mainstream risks associated with
climate variability and change into existing disaster risk management
and sectoral development, policies, and plans, albeit to differing degrees
depending on their capacity. These include initiatives to assess risks and
uncertainties, manage these across sectors, share and transfer risks, and
345
establish baseline information and research priorities (Freeman et al.,
2003; Mechler, 2004; Prabhakar et al., 2009). Ideally, national-level
institutions are best able to respond to the challenges of climate
extremes, particularly given that when disasters occur they often
surpass people and businesses’ coping capacity (OAS, 1991; Otero and
Marti, 1995; Benson and Clay, 2002a,b). National governments are also
better placed to appreciate key uncertainties and risks and take
strategic actions, particularly based on their power of taxation (see
Sections 6.4.3 and 6.5.3), although particularly exposed developing
countries may be financially challenged to attend to the risks and
liabilities imposed by natural disasters (Mechler, 2004; Cummins and
Mahul, 2009; UNISDR, 2011a).
Changes in weather and climate extremes and related impacts pose
new challenges for national disaster risk management systems, which
in many instances remain poorly adapted to the risks posed by existing
climatic variability and extremes (Lavell, 1998; McGray et al., 2007;
Venton and La Trobe, 2008; Mitchell et al., 2010b). Nonetheless, valuable
lessons for advancing adaptation to climate change can be drawn from
existing national disaster risk management systems (McGray et al., 2007;
Mitchell et al., 2010b). Such national systems are comprised of actors
operating across scales, fulfilling a range of roles and functions, guided
by an enabling environment of institutions, international agreements,
and experience of previous disasters (Carter, 1992; Freeman et al., 2003).
These systems vary considerably between countries in terms of their
capacities and effectiveness and in the way responsibilities are distributed
between actors. Countries also put differential emphasis on integration
of disaster risk management with development processes and tackling
vulnerability and exposure, compared with preparing for and responding
to extreme events and disasters (Cardona et al., 2010).
Recent global assessments of disaster risk management point to a
general lack of integration of disaster risk management into sustainable
development policies and planning across countries and regions,
although progress has been made especially in terms of passing
legislation, in setting up early warning systems, and in strengthening
disaster preparedness and response (Amendola et al., 2008; UNISDR,
2011b; Wisner, 2011). Closing the gap between current provision and
what is needed for tackling even current climate variability and disaster
risk is a priority for national risk management systems and is also a crucial
aspect of countries’ responses to projected climate change. With a history
of managing climatic extremes, involving a large number of experienced
actors across scales and levels of government and widespread instances
of supporting legislation and cross-sectoral coordinating bodies
(Section 6.4.2), national disaster risk management systems offer a
promising avenue for supporting adaptation to climate change and
reducing projected climate-related disaster risks.
Accordingly, this chapter assesses the literature on national systems for
managing disaster risks and climate extremes, particularly the design of
such systems of functions, actors, and roles they play, emphasizing the
importance of government and governance for improved adaptation to
climate extremes and variability. Focusing particularly on developing
country challenges, the assessment reflects on the adequacy of existing
knowledge, policies, and practices globally and considers the extent to
which the current disaster risk management systems may need to
evolve to deal with the uncertainties associated with and the effects of
climate change on disaster risks. Section 6.2 characterizes national
systems for managing existing climate extremes and disaster risk by
focusing on the actors that help create the system – national and sub-
national government agencies, bilateral and multilateral organizations,
the private sector, research agencies, civil society, and community-based
organizations. Drawing on a range of examples from developed and
developing countries, Sections 6.3 through 6.5 describe what is known
about the status of managing current and future risk, what is desirable
in an effective national system for adapting to climate change, and what
gaps in knowledge exist. The latter parts of the chapter are organized
by the set of functions undertaken by the actors discussed in Section 6.2.
The functions are divided into three main categories – those associated
with planning and policies (Section 6.3), strategies (Section 6.4), and
practices, including methods and tools (Section 6.5), for reducing
climatic risks. Section 6.6 reflects on how national systems for managing
climate extremes and disaster risk can become more closely aligned to
the challenges posed by climate change and development – particularly
those associated with uncertainty, changing patterns of risk and
exposure, and the impacts of climate change on vulnerability and
poverty. Aspects of Section 6.6 are further elaborated in Chapter 8.
6.2. National Systems and Actors
for Managing the Risks from
Climate Extremes and Disasters
Managing climate-related disaster risks is a concern of multiple actors,
working across scales from international, national, and sub-national and
community levels, and often in partnership, to ultimately help individuals,
households, communities, and societies to reduce their risks (Twigg,
2004; Schipper, 2009; Wisner, 2011). Comprising national and sub-
national governments, the private sector, research bodies, civil society,
and community-based organizations and communities, effective national
systems would ideally have each actor performing to their accepted
functions and capacities. Each actor would play differential but
complementary roles across spatial and temporal scales (UNISDR,
2008a; Schipper, 2009; Miller et al., 2010) and would draw on a mixture
of scientific and local knowledge to shape their actions and their
appreciation of the dynamic nature of risk (see Figure 6-1). Given that
national systems are at the core of a country’s capacity to meet the
challenges of observed and projected trends in exposure, vulnerability,
and weather and climate extremes, this section assesses the literature
on the roles played by different actors working within such national
systems.
Figure 6-1 encapsulates the discussions to follow on the interface and
interaction between different levels of actors, roles, and functions, with
the centrality of national organizations and institutions engaging at the
international level and creating enabling environments to support
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
346
actions across the country, supported by scientific information and
traditional knowledge.
6.2.1. National and Sub-National Governments
The national level plays a key role in governing and managing disaster
risks because national governments are central to providing risk
management-related public goods as they maintain organizational and
financial authority in planning and providing such goods. National
governments have the moral and legal responsibility to ensure
economic and social well-being, including safety and security of their
citizens from disasters (UNISDR, 2004). It is also argued that it is
government’s responsibility to protect the poorest and most vulnerable
citizens from disasters, and to implement disaster risk management that
reaches all (McBean, 2008; O’Brien et al., 2008; CCCD, 2009). In terms
of risk ownership, government and public disaster authorities ‘own’ a
large part of current and future extreme event risks and are expected to
govern and regulate risks borne by other parts of society (Mechler,
2004). Various normative literature sources support this. As one
example, literature on economic welfare theory suggests that national
governments are exposed to natural disaster risk and potential losses
due to their three main functions: provision of public goods and services
(e.g., education, clean environment, and security); the redistribution of
income; and stabilizing the economy (Musgrave, 1959; Twigg, 2004;
White et al., 2004; McBean, 2008; Shaw et al., 2009). The risks faced by
governments include losing public infrastructure, assets, and national
reserves. National-level governments also redistribute income across
members of society and thus are called upon when those are in need
(Linnerooth-Bayer and Amendola, 2000), or when members of society
are in danger of becoming poor, and in need of relief payments to
sustain a basic standard of living, especially in countries with low per
capita income and/or that have large proportions of the population in
poverty (Cummins and Mahul, 2009). Finally, it can be argued that
governments are expected to stabilize the economy, for example, by
supply-side interventions when the economy is in disequilibrium.
National-level governments are often called ‘insurers of last resort’ as
the governments are often the final entity that private households and
firms turn to in case of need, although the degree of compliance and
ability to honor those responsibilities by governments differs significantly
across countries. Nonetheless, in the context of a changing climate, it is
argued that governments have a particularly critical role to play in relation
to not only addressing the current gaps in disaster risk management but
also in response to uncertainties and changing needs due to increases
in the frequency, magnitude, and duration of some climate extremes
(Katz and Brown, 1992; Meehl et al., 2000; Christensen et al., 2007; also
refer to Chapter 3).
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
Activities
Projects
Goal
• Bilateral and multilateral
partners
• Intergovernmental organizations
• National government and
statutory agencies
• Civil society organizations
• Private sector
• Research and communication
bodies
• Local government agencies
• Individuals, households, and
communities
• Private sector
• Community-based organizations
• Faith-based organizations
“BOTTOM-UP” Functions “TOP-DOWN” FunctionsPrimary Actors
INTERNATIONAL
Principles
Agreements
Commitment
Financial resources
INTERNATIONAL
GLOBAL
NATIONAL / SUB-NATIONAL
Policies
Strategies
Legislation & other instruments
Financial resources
NATIONAL
Vision
Development goal
Sectoral objectives
NATIONAL
LOCAL
Needs
Aspiration
Culture
LOCAL
LOCAL
Global Climate
Projections
Regional / National
Climate Projections
Scientific and Local
Experiential
Knowledge
Vulnerability,
Risk, and Adaptation
Assessments
Figure 6-1 | National system of actors and functions for managing disaster risk and adapting to climate change.
347
Different levels of governments – national, sub-national, and local
level – as well as respective sectoral agencies play multiple roles in
addressing drivers of vulnerability and managing the risk of extreme
events, although their effectiveness varies within a country as well as
across countries. They are well placed to create multi-sectoral platforms
to guide, build, and develop policy, regulatory, and institutional
frameworks that prioritize risk management (Handmer and Dovers, 2007;
UNISDR, 2008b; OECD, 2009); integrate disaster risk management with
other policy domains like development or environmental management,
which often are separated in different ministries (UNISDR 2004, 2009c;
White et al., 2004; Tompkins et al., 2008); and address drivers of
vulnerability and assist the most vulnerable populations (McBean, 2008;
CCCD, 2009). Governments across sectors and levels also provide many
public goods and services that help address drivers of vulnerability as
well as those that support disaster risk management (White et al., 2004;
Shaw et al., 2009) through education, training, and research (Twigg,
2004; McBean, 2008; Shaw et al., 2009).
Governments also allocate financial and administrative resources for
disaster risk management, as well as provide political authority (Spence,
2004; Twigg, 2004; Handmer and Dovers, 2007; CCCD, 2009). Evidence
suggests that successful disaster risk management is partly contingent
on resources being made available at all administration levels, but to
date, insufficient policy and institutional commitments have been made
to disaster risk management in many countries, particularly at the local
government level (Twigg, 2004; UNISDR, 2009d). It is argued that
governments also have an important role to guide and support the private
sector, civil society organizations, and other development partners in
playing their differential roles in managing disaster risk (O’Brien et al.,
2008; Prabhakar et al., 2009).
6.2.2. Private Sector Organizations
The private sector plays a small, but increasingly important role in
disaster risk management and adaptation, and some aspects of disaster
risk management may be suitable for nongovernmental stakeholders to
implement, albeit this would often effectively be coordinated within a
framework created and enabled by governments. Three avenues for private
sector engagement may be identified: (1) corporate social responsibility
(CSR); (2) public-private partnerships (PPP); and (3) businesses model
approaches. CSR involves voluntary advocacy and raising awareness by
businesses for disaster risk reduction as well as involving funding
support and the contribution of volunteers and expertise to implement
risk management measures. PPPs focus on enhancing the provision of
public goods for disaster risk reduction in joint undertakings between
public and private sector players. The business model approach pursues
the integration and alignment of disaster risk reduction with operational
and strategic goals of an enterprise (Warhurst, 2006; Roeth, 2009).
While CSR and PPP have received substantial attention, business model
approaches remain rather untouched areas, one very important exception
being the insurance industry as a supplier of tools for transferring and
sharing disaster risks and losses.
In terms of business model approaches, insurance is a key sector. In
exchange for pre-disaster premium payments, disaster insurance and
other risk transfer instruments in 2010 covered about 30% of disaster
losses overall (Munich Re, 2011). In terms of weather-related events, for
the period 1980 to 2003, insurance overall covered about 20% of the
losses, yet the distribution according to country income groups is
uneven, with about 40% of the losses insured in high-income as
compared to 4% in low-income countries (Mills, 2007). In developing
countries, despite complexities and uncertainties involved in both
supply and demand for risk transfer, risk financing mechanisms have
been found to demonstrate substantial potential for absorbing the
financial burden of disasters (Pollner, 2000; Andersen, 2001; Varangis et
al., 2002; Auffret, 2003; Dercon, 2005; Hess and Syroka, 2005;
Linnerooth-Bayer et al., 2005; Skees et al., 2005; World Bank, 2007;
Cummins and Mahul, 2009; Hazell and Hess, 2010). There is, though,
some uncertainty as to the extent to which the private sector would
continue to play this role in the context of a changing environment due
to uncertainty and imperfect information, missing and misaligned
markets, and financial constraints (Smit et al., 2001; Aakre et al., 2010).
Private insurers are concerned about changes in risks and associated risk
ambiguity, that is, the uncertainty about the changes induced by climate
change in terms of potentially modified extreme event intensity and
frequency. Accordingly, as climate change, and other drivers such as
changes in vulnerability and exposure (see Chapters 1, 2, and 3), are
projected to lead to changes in frequency and intensity of some
weather risks and extremes, insurers may be less prepared to underwrite
insurance for extreme event risks. Innovative private-public sector
partnerships may thus be required to better estimate and price risk as
well as develop robust insurance-related products, which may be
supported in developing countries by development partner funds as
well (see Section 6.5.3 and Case Study 9.2.13).
Professional societies (such as builders and architects) and trade
associations also play a key role in developing and implementing
standards and practices for disaster risk reduction. These practices may
include national and international standards and model building codes
that are adopted in the regulations of local, state, and national
governments. Although the potential for private sector players in
disaster risk reduction in sectors such as engineering and construction,
information communication technology, media and communication, as
well as utilities and transportation seems large, limited evidence of
successful private sector activity has been documented, owing to a
number of reasons (Roeth, 2009). The business case for private sector
involvement in disaster risk reduction remains unclear, hampering private
sector engagement. Companies may also be averse to reporting activities
that are fundamental to their business; and, in more community-
focused projects, companies often work with local nongovernmental
organizations (NGOs) and do not often report such efforts. Considering
climate variability and change within the business model, companies may
be an important entry point for disaster risk reduction, particularly in
terms of guaranteeing global value chains in the presence of potentially
large-scale disruptions triggered by climate-related disasters. For example,
the economic viability of the Chinese coastal zone – the economic
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
348
heartland of China and home to many multinational companies producing
a large share of consumer goods globally – is highly exposed to typhoon
risk and will increasingly depend on well-implemented disaster risk
reduction mechanisms (Roeth, 2009).
6.2.3. Civil Society and Community-Based Organizations
At the national level, civil society organizations (CSOs) and community-
based organizations (CBOs) play a significant role in developing initiatives
to respond to disasters, reduce the risk of disasters, and, recently, adapt
to climate-related hazards (see Section 5.1 for a discussion of ‘local’ and
‘community’ and Section 5.4.1 for the role of CBOs at the local level).
CSOs and CBOs are referred to here as the wide range of associations
around which society voluntarily organizes itself, with CBO referring to
those associations primarily concerned with local interests and ties. CSO
and CBO initiatives in the field of disaster risk management, which may
usually begin as a humanitarian concern, often evolve to also embrace
the broader challenge of disaster risk reduction following community-
focused risk assessment, including specific activities targeting education
and advocacy; environmental management; sustainable agriculture;
infrastructure construction; and increased livelihood diversification
(McGray et al., 2007; CARE International, 2008; Oxfam America, 2008;
Practical Action Bangladesh, 2008; SEEDS India, 2008; Tearfund, 2008;
World Vision, 2008).
Recently in some high-risk regions there has been rapid development of
national platforms of CSOs and CBOs that have been working together
in order to push for the transformation of policies and practices related
to disaster risk reduction. This is true in the case of Central America,
where at least four platforms are functioning in the same number of
countries, involving more than 120 CSOs and CBOs (CRGR, 2007a). The
efforts of these platforms have been aimed at advocacy, training,
research, and capacity building in disaster risk reduction. In Central
America, the experience is that advocacy on climate policy construction
has become a new feature of such platforms since 2007 (CRGR, 2009).
While beyond the scope of this chapter, on balance the majority of CSOs
and CBOs focus efforts at the local level, trying to link disaster risk
management with local development goals associated with water,
sanitation, education, and health, for example (GNDR 2009; Lavell,
2009). Faith-based organizations are also influential in assisting local
communities in disaster risk management, not only providing pastoral
care in times of disasters but also playing an important role in raising
awareness and training, with many international development partners
often working with local church groups to build community resilience
(see, for example, ADPC 2007; Gero et al., 2011; Tearfund, 2011).
In several countries in Latin America, CSOs and CBOs are considered, by
law, as part of national systems for civil protection (Lavell and Franco,
1996; CRGR, 2007b) though participation, with the exception of National
Red Cross/Red Crescent Societies, remains patchy (UNISDR, 2008c). In
some countries where governments are not able or willing to fulfill
certain disaster risk management functions, such as training, supporting
food security, providing adequate housing, and preparedness, CSOs and
CBOs have stepped in (Benson et al., 2001). While CSOs often face
challenges in securing resources for replicating successful initiatives and
scaling out geographically (CARE International, 2008; Oxfam America,
2008; Practical Action Bangladesh, 2008; SEEDS India, 2008; Tearfund,
2008; World Vision, 2008); sustaining commitment to work with local
governments and stakeholders over the long term and maintaining
partnerships with local authorities (Oxfam America, 2008); and
coordinating and linking local-level efforts with sub-national government
initiatives and national plans during the specific project implementation
(SEEDS India, 2008), they are particularly well positioned to draw links
between disaster risk reduction and climate change adaptation given
that such organizations are currently among the few to combine such
expertise (Mitchell et al., 2010b).
6.2.4. Bilateral and Multilateral Agencies
In developing countries, particularly where the government is weak and
has limited resources, bilateral and multilateral agencies play a
significant role in supplying financial, technical, and in some cases
strategic support to government and nongovernment agencies to
tackle the multifaceted challenges of disaster risk management and
climate change adaptation in the context of national development
goals (e.g., AusAid, 2009; DFID, 2011). Multilateral agencies are referred
to here as international institutions with governmental membership
that have a significant focus on development and aid recipient countries.
Such agencies can include United Nations agencies, regional groupings
(e.g., some European Union agencies), and multilateral development
banks (e.g., World Bank, Asian Development Bank). Bilateral agencies
(e.g., United Kingdom Department for International Development) are
taken here as national institutions that focus on the relationship
between one government and another. In the development sphere,
this is often in the context of a richer government providing support
to a poorer government. The role of international institutions,
including bilateral and multilateral agencies, is discussed extensively in
Section 7.3.
Bilateral and multilateral agencies have been key actors in advancing
mainstreaming of disaster risk reduction and climate change adaptation
into development planning (Eriksen and Næss, 2003; Klein et al., 2007;
see Section 6.3). This has primarily been driven by a concern that
development investments are increasingly exposed to climate- and
disaster-related risks and that climate change poses security concerns
(Harris, 2009; Persson and Klein, 2009). As a result, such agencies are
influencing development policy and implementation at a national level
as they require disaster and climate risk assessments and environmental
screening to be conducted at different points in the project approval
process and in some cases retrospectively when projects are already
underway (Klein et al., 2007; OECD, 2009; Hammill and Tanner, 2010). A
range of tools and methods have been developed, primarily by bilateral
and multilateral agencies, to support such processes (Klein et al., 2007;
Hammill and Tanner, 2010).
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
349
While significant progress has been made in developing appropriate
tools and methods for assessing and screening risk, many bilateral and
multilateral agencies continue to address disaster risk management and
climate change adaptation separately, and link with respective regional
and national agencies in the context of distinct international instruments
(Mitchell and Van Aalst, 2008; Mitchell et al., 2010b; Gero et al., 2011).
However, recent assessments suggest that the situation is improving,
partially attributable to the process of authoring this Special Report and
in the focus on risk management in the text of the Bali Action Plan
(2007) and Cancun Agreement (2010) (Mitchell et al., 2010b; see
Section 7.3.2.2 for more detail).
The diversity of national contexts requires bilateral and multilateral
agencies to adopt different modalities to maximize the effectiveness of
technical, financial, and strategic support. For example, in the Pacific and
the Caribbean, regional bodies (e.g., the Caribbean Disaster Emergency
Management Agency) commonly operate as an intermediary, channeling
resources to island countries where it is not efficient for international
agencies to establish a permanent adaptation or risk management-
focused presence (Hay, 2009; Gero et al., 2011). In countries with weak
national institutions, bilateral and multilateral agencies commonly choose
to channel resources through civil society organizations with the intention
of ensuring that resources reach the poorest and most vulnerable
(Wickham et al., 2009). In such situations, coordination between
agencies can be challenging and in certain circumstances can further
reduce the risk management capacity of government organizations
(Wickham et al., 2009). However, the broad trend is to maximize the
support to national governments by seeking to improve national ownership
of risk management and adaptation processes and in that respect
support national governments to lead national systems (GFDRR, 2010;
DFID, 2011).
6.2.5. Research and Communication
The effectiveness of national systems for managing climate extremes and
disaster risks is highly dependent on the availability and communication
of robust and timely scientific data and information (Sperling and
Szekely, 2005; Thomalla et al., 2006; CACCA, 2010) and traditional
knowledge (Mercer et al., 2007; Kelman et al., 2011; see Box 5-7) to inform
not only community-based decisions and policymakers who manage
national approaches to disaster risk and climate change adaptation, but
also researchers who provide further analytical information to support
such decisions.
Scientific and research organizations range from specialized research
centers and universities, to regional organizations, to national research
agencies, multilateral agencies, and CSOs playing differential roles, but
generally continue to divide into disaster risk management or climate
change adaptation communities. Scientific research bodies play important
roles in managing climate extremes and disaster risks by: (a) supporting
thematic programs to study the evolution and consequences of past
hazard events, such as cyclones, droughts, sandstorms, and floods;
(b) analyzing time-and-space dependency in patterns of weather-related
risks; (c) building cooperative networks for early warning systems,
modeling, and long-term prediction; (d) actively engaging in technical
capacity building and training; (e) translating scientific evidence into
adaptation practice; (f) collating traditional knowledge and lessons
learned for wider dissemination; and (g) translating scientific information
into user-friendly forms for community consumption (Sperling and
Szekely, 2005; Thomalla et al., 2006; Aldunce and González, 2009).
Disaster practitioners largely focus on making use of short-term weather
forecasting and effective dissemination and communication of hazard
information and responses (Thomalla et al., 2006). Such climate change
expertise can typically be found in meteorological agencies, environment
or energy departments, and in academic institutions (Sperling and Szekely,
2005), while disaster risk assessments have been at the core of many
multilateral and civil society organizations and national disaster
management authorities (Sperling and Szekely, 2005; Thomalla et al.,
2006). Although progress has been reported in the communication and
availability of scientific information, there is still a lack of, for example,
sufficient local or sub-national data on hazards and risk assessments to
underpin area-specific disaster risk management (Chung, 2009; UNISDR,
2009c).
6.3. Planning and Policies for Integrated
Risk Management, Adaptation, and
Development Approaches
Given that learning will come from doing and in spite of differences,
there are many ways that countries can learn from each other in
prioritizing their climate and disaster risks; in mainstreaming climate
change adaptation and disaster risk management into plans, policies,
and processes for development; and in securing additional financial and
human resources needed to meet increasing demands (UNDP, 2002;
Thomalla et al., 2006; Schipper, 2009). This subsection will address
frameworks for national disaster risk management and climate change
adaptation planning and policies (Section 6.3.1), the mainstreaming of
plans and policies nationally (Section 6.3.2), and the various sectoral
disaster risk management and climate change adaptation options
available for national systems (Section 6.3.3), recognizing the range of
actors engaged in these processes as described in Section 6.2.
6.3.1. Developing and Supporting National
Planning and Policy Processes
National and sub-national government and statutory agencies have
a range of planning and policy options to help create the enabling
environments for departments, public service agencies, the private
sector, and individuals to act (UNDP, 2002; Heltberg et al., 2009; OECD,
2009; ONERC, 2009; Hammill and Tanner, 2010). When considering
disaster risk management and adaptation to climate change actions, it
is often the scale of the potential climate and disaster risks and impacts,
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
350
the capacity of the governments or agencies to act, the level of certainty
about future changes, the timeframes within which these future impacts
and disasters will occur, and the costs and consequences of decisions
that play an important role in their prioritization and adoption (Heltberg
et al., 2008; World Bank, 2008; Wilby and Dessai, 2010).
The complexity and diversity of adaptation to climate change situations
implies that there can be no single recommended approach for assessing,
planning, and implementing adaptation options (Füssel, 2007; Hammill
and Tanner, 2010; Lu, 2011). When the planning horizons are short and
adaptation decisions only impact the next one or two decades, adaptation
to recent climate variability and observed trends may be sufficient
(Hallegatte, 2009; Wilby and Dessai, 2010; Lu, 2011). For long-lasting
risks and decisions, the timing and sequencing of adaptation options
and incorporation of climate change scenarios become increasingly
important (Hallegatte, 2009; OECD, 2009; Wilby and Dessai, 2010).
Studies suggest that the most pragmatic adaptation and disaster risk
management options depend on the timeframes under consideration and
the adaptive capacity and ability of the country or sectoral agencies to
effectively integrate information on climate change and its uncertainties
(McGray et al., 2007; Biesbroek et al., 2010; Krysanova et al., 2010;
Wilby and Dessai, 2010; Juhola and Westerhoff, 2011). Given the various
uncertainties at decisionmaking scales, studies suggest that adaptation
actions based on information on the observed climate and its trends
may be preferable in some cases while, in other cases with long-term
irreversible decisions, climate change scenario-guided adaptation
actions will be required (Auld, 2008b; Hallegatte, 2009; OECD, 2009;
Krysanova et al., 2010; Wilby and Dessai, 2010). Climate change scenarios
provide needed guidance for adaptation options when the direction of
the climate change impacts are known and when the decisions involve
long-term building infrastructure, development plans, and actions to
avoid catastrophic impacts from more intense extreme events
(Haasnoot et al., 2009; Hallegatte, 2009; Wilby and Dessai, 2010).
In dealing with climate change and disaster risk uncertainties, many
national studies identify gradations or categories of adaptation and
disaster risk management planning and policy options (Dessai and Hulme,
2007; Auld, 2008b; Hallegatte, 2009; Kwadijk et al., 2010; Mastrandrea
et al., 2010; Wilby and Dessai, 2010). These gradations in options range
from climate vulnerability or resilience approaches, sometimes
described as ‘bottom-up’; vulnerability, tipping point, critical threshold, or
policy-first approaches to climate modeling, impact-based approaches,
sometimes described as ‘top-down’; model or impacts-first; science-
first; or classical approaches (as illustrated in Figure 6-2 and outlined in
the sectoral option headings of Table 6-1 and described in Section
6.3.3). Although the bottom-up and top-down terms sometimes refer to
scale, subject matter, or policy (e.g., national versus local, physical to
socioeconomic systems), the terms are used here to describe the
sequences or steps needed to develop adaptation and disaster risk
management plans and policies at the national level. When dealing with
long-term future climate change risks, the main differences between the
scenarios-impacts-first and vulnerability-thresholds-first approaches lie
in the timing or sequencing of the stages of the analyses, as shown
in Figure 6-2 (Kwadijk et al., 2010; Ranger et al., 2010). Although this
difference appears subtle, it has significant implications for the
management of uncertainty, the timing of adaptation options, and the
efficiency of the policymaking (Dessai and Hulme, 2007; Auld, 2008b;
Kwadijk et al., 2010; Wilby and Dessai, 2010; Lu, 2011). For example,
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
Begin with the questions:
“Where are the
sensitivities, thresholds,
and priorities considering
climate variabilities?”
“What can communities
cope with?”
Input climate change
projections and other
relevant information
about underlying drivers
Begin with the question
“What if climate extremes
change according to
scenarios, x, y, z?”
Start with climate change
models, scenarios,
impacts, assessments,
reports, etc.
“Climate Models, Scenarios, Impacts-First” “Vulnerability, Thresholds-First”
Identify development context, hazards, and
vulnerability problems
Identify vulnerabilities, sensitivities, thresholds;
propose adaptation measures
Assess adaptation measures and timing for
action against climate change scenarios
Assess tradeoffs between adaptation options
Evaluate outcomes
Structure impacts problem
Assess relevant climatic changes from climate
change models, downscaling
Assess relevant impacts based on projected
climate changes
Design and assess adaptation options for
relevant impacts
Evaluate outcomes
Figure 6-2 | Top-down scenario, impacts-first approach (left panel) and bottom-up vulnerability, thresholds-first approach (right panel) – comparison of stages involved in identifying
and evaluating adaptation options under changing climate conditions. Adapted from Kwadijk et al. (2010) and Ranger et al. (2010).
351
when the lifespan of a decision, policy, or measure has implications for
multiple decades or the decision is irreversible and sensitive to climate,
the performance of adaptation and risk reduction options across a
range of climate change scenarios becomes critical (Auld, 2008b;
Kwadijk et al., 2010; Wilby and Dessai, 2010).
Vulnerability thresholds-based approaches start at the level of the
decisionmaker, identify desired system objectives and constraints,
consider how resilient or robust a system or sector is to changes in climate,
assess adaptive capacity and critical ‘tipping points’ or threshold points,
then identify the viable adaptation strategies that would be required to
improve resilience and robustness under future climate scenarios
(Auld, 2008b; Urwin and Jordan, 2008; Hallegatte, 2009; Kwadijk et al.,
2010; Mastrandrea et al., 2010; Wilby and Dessai, 2010).
Vulnerability-
thresholds
approaches can be independent of any specific future climate
condition.
Options that are known as ‘no regrets’ and ‘low regrets’ provide benefits
under any range of climate change scenarios, although they may not
be optimal for every future scenario, and are recommended when
uncertainties over future climate change directions and impacts are
high (Dessai and Hulme, 2007; Auld, 2008b; Hallegatte, 2009; Kwadijk
et al., 2010). These ‘low regrets’ adaptation options typically include
improvements to coping strategies or reductions in exposure to known
threats (Auld, 2008b; Kwadijk et al., 2010; Wilby and Dessai, 2010), such
as better forecasting and warning systems, use of climate information
to better manage agriculture in drought-prone regions, flood-proofing
of homesteads, or interventions to ensure up-to-date climatic design
information for engineering projects. The vulnerability-thresholds-first
approaches are particularly useful for identifying priority areas for
action now, assessing the effectiveness of specific interventions when
current climate-related risks are not satisfactorily controlled, when
climatic stress factors are closely intertwined with non-climatic factors,
planning horizons are short, resources are very limited (i.e., expertise,
data, time, and money), or uncertainties about future climate impacts
are very large (Agrawala and van Aalst, 2008; Hallegatte, 2009;
Prabhakar et al., 2009; Wilby and Dessai, 2010).
Vulnerability-thresholds-first approaches have sometimes been critiqued
for the time required to complete a vulnerability assessment, for their
reliance on experts, and for their largely qualitative results and limited
comparability across regions (Patt et al., 2005; Kwadijk et al., 2010).
Vulnerability-thresholds approaches can sometimes prove less suited
for guiding future adaptation decisions if coping thresholds change, or
if climate change risks emerge that are outside the range of recent
experiences (e.g., successive drought years could progressively reduce
coping thresholds of the rural poor by increasing indebtedness)
(McGray et al., 2007; Agrawala and van Aalst, 2008; Auld, 2008b;
Hallegatte, 2009; Prabhakar et al., 2009; Wilby and Dessai, 2010).
The scenarios-impact-first approaches typically start with several climate
change modeling scenarios and socioeconomic scenarios, evaluate the
expected impacts of climate change, and subsequently identify adaptation
and risk reduction options to reduce projected risks (Kwadijk et al.,
2010; Mastrandrea et al., 2010; Wilby and Dessai, 2010).
The scenarios-
impacts-first approaches are most useful to raise awareness of the
problem, to explore possible adaptation strategies, and to identify
research priorities,
especially when current climate and disaster risks
can be effectively controlled, when sufficient data and resources are
available to produce state-of-the-art climate scenarios at the spatial
resolutions relevant for adaptation, and when future climate impacts
can be projected reliably (Kwadijk et al., 2010; Wilby and Dessai, 2010).
Scenarios-impacts approaches
depend strongly on the chosen climate
change scenarios and downscaling techniques, as well as the assumptions
about scientific and socioeconomic uncertainties (OECD, 2009; Kwadijk
et al., 2010). Pure scenarios-impacts approaches may not be available at
the spatial scales relevant to the decisionmaker, may not be applicable
for the purpose of the decisionmaker, and
usually give less consideration
to current risks from natural climate variability, to non-climatic stressors,
and to key uncertainties along with their implications for robust
adaptation policies (Füssel, 2007; Wilby and Dessai, 2010). In practice,
there are very limited examples of actual adaptation policies being
developed and planned adaptation decisions being implemented based
on scenarios-impacts approaches only (Füssel, 2007; Biesbroek et al.,
2010; Wilby and Dessai, 2010).
Increasingly, studies are recognizing that the scenarios-impacts and
vulnerability-thresholds approaches are complementary and need to be
integrated and that both can benefit from the addition of stakeholder
and scientific input to determine critical thresholds for climate change
vulnerabilities (Auld, 2008b; Haasnoot et al., 2009; Kwadijk et al., 2010;
Mastrandrea et al., 2010; Wilby and Dessai, 2010). Critical thresholds (or
adaptation tipping points) help in answering the basic adaptation
questions of decision- and policymakers – namely, what are the first
priority issues that need to be addressed as a result of increasing disaster
risks under climate change and when might these critical thresholds be
reached (Auld, 2008b; Haasnoot et al., 2009; Kwadijk et al., 2010;
Mastrandrea et al., 2010). The integration of scenarios-impacts and
vulnerability-thresholds approaches provides guidance on the sensitivity
of sectors and durability of options under different climate change
scenarios (Haasnoot et al., 2009; Kwadijk et al., 2010; Mastrandrea et
al., 2010). Integrated approaches that link changes in climate variables to
decisions and policies and express uncertainties in terms of timeframes
over which a policy or plan may be effective (i.e., roughly when will the
critical threshold be reached) also provide valuable information for
plans and policies and their implementation (Haasnoot et al., 2009;
Kwadijk et al., 2010; Mastrandrea et al., 2010).
Regardless of the approaches used, it is important that uncertainty over
future climate change risks not become a barrier to climate change risk
reduction actions (Auld, 2008b; Hallegatte, 2009; Krysanova et al., 2010;
Wilby and Dessai, 2010). In cases where climate change uncertainties
remain high, countries may choose to increase or build on their capacity
to cope with uncertainty, rather than risk maladaptation from use of
ambiguous impact studies or no action (McGray et al., 2007; Hallegatte,
2009; Wilby and Dessai, 2010). In order to reduce the risk of maladaptation
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
352
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
Natural
Ecosystems
and Forestry
Synergies between UNFCCC and
Rio Conventions; avoid actions
that interfere with goals of
other UN conventions
3
Research on climate change–
ecosystem–forest links, climate
and ecosystem prediction
systems, climate change
projections; monitor ecosystem
and climate trends
3
Incorporate ecosystem
management into National
Adaptation Programmes of
Action and disaster risk
reduction plans
3
Adaptation to climate change
interventions to maintain
ecosystem resilience; corridors,
assisted migrations; plan EbA for
climate change
4
Seed, genetic banks; new
genetics; tree species
improvements to maintain
ecosystem services in future;
adaptive agroforestry
4
Changed timber harvest
management, new technologies
for adaptation to climate
change, new uses to conserve
forest ecosystem services
4
Micro-finance and
insurance to
compensate for lost
livelihoods
5
Investments in
additional insurance,
government reserve
funds for increased risks
due to loss of protective
ecosystem services
5
Replace lost ecosystem
services through additional
hard engineering, health
measures
6
Restore loss of damaged
ecosystems
6
Sustainable afforestation (for
robust forests), reforestation,
conservation of forests, wetlands
and peatlands, sustainable and
increased biomass; land use,
land-use change, and forestry;
reducing emissions from
deforestation
7
Incentives for sustainable
sequestration of carbon;
sustainable bio-energy; energy
self-sufficiency
7
Agriculture
and Food
Security
Food security via sustainable land and
water management, training; efficient
water use, storage; agro-forestry;
protection shelters, crop and livestock
diversification; improved supply of climate
stress tolerant seeds; integrated pest,
disease management
8
Climate monitoring; improved weather
predictions; disaster management, crop
yield and distribution models and
predictions
9
Increased agriculture-climate
research and development
10
Research on climate tolerant
crops, livestock; agrobiodiversity
for genetics
10
Integration of climate change
scenarios into national
agronomic assessments
11
Diversification of rural economies
for sensitive agricultural
practices
10
Adaptive agricultural and
agroforestry practices for new
climates, extremes
12
New and enhanced agricultural
weather, climate prediction
services
11
Food emergency planning;
distribution and infrastructure
networks
12
Diversify rural economies
12
Improved access to crop,
livestock, and income
loss insurance (e.g.,
weather derivatives)
13
Micro-financing and
micro-insurance
13
Subsidies, tax credits
13
Changed livelihoods and
relocations in regions with
climate sensitive practices
12
Secure emergency stock and
improve distribution of food
and water for emergencies
12
Energy efficient and sustainable
carbon sequestering practices;
training; reduced use of chemical
fertilizers
14
Use of bio-gas from agricultural
waste and animal excreta
14
Agroforestry
14
Coastal Zone
and Fisheries
EbA; integrated coastal zone
management (ICZM); combat salinity;
alternate drinking water availability; soft
and hard engineering
15
Strengthen institutional, regulatory, and
legal instruments; setbacks; tourism
development planning
16
Marine protected areas, monitoring fish
stocks, alter catch quantities, effort,
timing; salt-tolerant fish species
17
Climate risk reduction planning; hazard
delineation; improve weather forecasts,
warnings, environmental prediction
16
Climate change projections for
coastal management planning;
develop modeling capacity for
coastal zone-climate links;
climate-linked ecological and
resource predictions; improved
monitoring, geographic and
other databases for coastal
management
18
Monitor fisheries; selective
breeding for aquaculture, fish
genetic stocks; research on salt-
tolerant crop varieties
19
Incorporate adaptation to
climate change, sea level rise
into ICZM, coastal defenses
18
Hard and ‘soft’ engineering for
adaptation to climate change;
sustainable tourism
development planning; resilient
vessels and coastal facilities
16
Manage for changed fisheries,
invasives
20
Inland lakes: alter transportation
and industrial practices, soft and
hard engineering
20
Enhance insurance for
coastal regions and
resources; fisheries
insurance
21
Government reserve
funds
21
Enhance emergency
preparedness measures for
more frequent and intense
extremes, including more
evacuations
16
Relocations of communities,
infrastructure
16
Exit fishing; provide alternate
livelihoods
19
Use of sustainable renewable
energy; conservation, energy
self-sufficiency (especially for
islands, coastal regions)
22
Offshore renewable energy for
alternate incomes and
aquaculture habitat
22
Continued next page
Sector/
Response
‘No regrets’ and ‘low regrets’
actions for current and future
risks
(‘No/low regrets’ options
plus…) Preparing for
climate change risks by
reducing uncertainties
(building capacity)
(“Preparing for climate
change” risks plus...)
Reduce risks from future
climate change
Risk transfer
Accept and deal with
increased and
unavoidable (residual)
risks
‘Win-win’ synergies for
GHG reduction,
adaptation, risk
reduction, and
development benefits
Use of Ecosystem-based Adaptation (EbA)
or ‘soft engineering’; integrate disaster
risk reduction and climate into integrated
coastal zone and water resources
management, forest management, and
land use management; conserve, enhance
resilience of ecosystems; restore
protective ecosystem services
1
Adaptive forest management; forest fire
management, controlled burns;
agroforestry; biodiversity
2
Reduce forest degradation, unsustainable
harvests, and provide incentives for
alternate livelihoods, eco-tourism
6
Table 6-1 | National policies, plans, and programs: a selection of disaster risk reduction and adaptation to climate change options by selected sectors.
353
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
Water
resources
Implement Integrated Water Resource
Management (IWRM), national water
efficiency, storage plans
20
Effective surveillance, prediction, warning
and emergency response systems; better
disease and vector control, detection and
prediction systems; better sanitation;
awareness and training on public health
24
Adequate funding, capacity for resilient
water infrastructure and water resource
management; Improved institutional
arrangements, negotiations for water
allocations, joint river basin management
23
Develop prediction, climate
projection, and early warning
systems for flood events and
low water flow conditions;
research and downscaling for
hydrological basins
24
Multi-sectoral planning for
water; selective decentralization
of water resource management
(e.g., catchments and river
basins); joint river basin
management (e.g., bi-national)
23
National water policy
frameworks, robust integrated
and adaptive water resource
management for adaptation to
climate change
25
Investments in hard and soft
infrastructure considering
changed climate; river
restoration
25
Improved weather, climate,
hydrology-hydraulics, water
quality forecasts for new
conditions
24
Public-private
partnerships; Economics
for water allocations
beyond basic needs
26
Mobilize financial
resources and capacity
for technology and EbA
26
Insurance for
infrastructure
26
Enhance national
preparedness and evacuation
plans for greater risks
24
Enhance health infrastructure
for more failures
24
Alter transport, engineering;
increases to temporary
consumable water taking
permits
24
Enhance food , water
distribution for emergencies,
plan for alternate livelihoods
24
Integrated and sustainable water
efficiency and renewable hydro
power for adaptation to climate
change
23
Infra -structure,
Housing, Cities,
Transportation,
Energy
Improved downscaling of
climate change information;
maintain climate data networks,
update climatic design
information; increased
safety/uncertainty factors in
codes and standards; develop
adaptation to climate change
tools
28
Research on climate, energy,
coastal, and built environment
interface, including flexible
designs, redundancy; forensic
studies of failures (adaptation
learning); improved
maintenance
27
Investments for sustainable
energy development;
cooperation on trans-boundary
energy supplies (e.g., wind
energy at times of peak wind
velocity)
29
Codes, standards for changed
extremes
30
Publicly funded infrastructure,
coastal development and post-
disaster reconstruction to
include adaptation to climate
change
30
New materials, engineering
approaches; flexible design and
use structures; asset
management for adaptation to
climate change
30
Hazard mapping; zoning and
avoidance; prioritized retrofits,
abandon the most vulnerable;
soft engineering services
30
Design energy generation,
distribution systems for
adaptation; switch to less risky
energy systems, mixes; embed
sustainable energy in disaster
risk reduction and adaptation to
climate change planning
29
Infrastructure insurance
and financial risk
management
29
Insurance for energy
facilities, interruption
29
Innovative risk sharing
instruments
29
Government reserve
funds
29
More relocations
28
Enhance evacuation,
transportation, and energy
contingency planning for
increases in extreme events
28
Increase climate -resilient
shelter construction
28
Implement energy- and water-
efficient GHG reductions, disaster
risk reduction and adaptation to
climate change synergies
29
Scale up, market penetration for
sustainable renewable energy
production; increased
hydroelectric potential;
sustainable biomass; ‘greener’
distributed community energy
systems
29
Continued next page
Sector/
Response
‘No regrets’ and ‘low regrets’
actions for current and future
risks
(‘No/low regrets’ options
plus…) Preparing for
climate change risks by
reducing uncertainties
(building capacity)
(“Preparing for climate
change” risks plus...)
Reduce risks from future
climate change
Risk transfer
Accept and deal with
increased and
unavoidable (residual)
risks
‘Win-win’ synergies for
GHG reduction,
adaptation, risk
reduction, and
development benefits
Building codes, standards with updated
climatic values; climate- resilient
infrastructure (and energy) designs;
training, capacity, inspection,
enforcement; monitoring for priority
retrofits (e.g., permafrost); maintenance
27
Legal alternatives to informal settlements,
sanitation
27
Strengthen early warning systems, hazard
awareness; improved weather warning
systems; disaster-resilient building
components (rooms) in high-risk areas;
tourism development planning;
heat-health responses
28
Integrate urban planning, engineering,
maintenance
27
Diversified energy systems; maintenance;
self-sufficiency, clean energy technologies
for national energy plans, international
agreement goals (biogas, solar cooker);
use of renewable energy in remote and
vulnerable regions; use of appropriate
energy mixes nationally
29
Energy security; distributed energy
generation and distribution
29
Table 6-1 (continued)
354
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
Health
Community/urban and coastal zone
planning, building standards and
guidelines; cooling shelters; safe health
facilities; retrofits for vulnerable
structures; health facilities designed using
updated climate information
31
Strengthen surveillance, health
preparedness; early warning weather-
climate-health systems, heat alerts and
responses; capacity for response to early
warnings; prioritize disaster risks; disaster
prevention and preparedness; public
education campaigns; food security
31
Strengthen disease surveillance and
controls; improve health care services,
personal health protection; improve water
treatment/sanitation; water quality
regulations; vaccinations, drugs,
repellants; development of rapid
diagnostic tests
31
Monitor air and water quality;
regulations; urban planning
31
Better land and water use management
to reduce health risks
31
New food and water security,
distribution systems; air quality
regulations, alternate fuels
32
New warning and response
systems; predict and manage
health risks from landscape
changes; target services for most
at risk populations
32
Climate proofing, refurbish/
maintain national health facilities
and services
32
Address needs for additional
health facilities and services
32
Extend and expand
health insurance
coverage to include
new and changed
weather and climate
risks
33
Government reserve
funds
33
National plan for heat and
extremes emergencies
32
New disease detection and
management systems
32
Enhanced prediction and
warning systems for new
risks
32
Use of clean and sustainable
renewable energy and water
sources; increase energy
efficiency; air quality
regulations; clean energy
technologies to reduce harmful
air emissions (e.g., cooking
stoves)
34
Design sustainable infrastructure
for climate change and health
30
Education, disaster prevention
and preparedness
31
Research on climate-health
linkages and adaptation to
climate change options; develop
new health prediction systems
for emerging risks; research on
landscape changes, new
diseases, and climate; urban
weather-health modeling
31
Sector/
Response
‘No regrets’ and ‘low regrets’
actions for current and future
risks
(‘No/low regrets’ options
plus…) Preparing for
climate change risks by
reducing uncertainties
(building capacity)
(“Preparing for climate
change” risks plus...)
Reduce risks from future
climate change
Risk transfer
Accept and deal with
increased and
unavoidable (residual)
risks
‘Win-win’ synergies for
GHG reduction,
adaptation, risk
reduction, and
development benefits
References: 1. Adger et al., 2005; Barbier, 2009; Colls et al., 2009; FAO, 2008a; MEA, 2005; SCBD, 2009; Shepherd, 2004, 2008; UNEP, 2009; UNISDR, 2009d; World Bank, 2010. 2. FAO, 2007; Neufeldt et al., 2009; Shugart et al., 2003; Spittlehouse and Stewart, 2003, Weih,
2004. 3. Colls et al., 2009; FAO, 2008a; OECD, 2009; Rahel and Olden, 2008; Robledo et al., 2005; SCBD, 2009; UNEP, 2009; UNFCCC, 2006a. 4. Berry, 2007; FAO, 2007, 2008a,b; Leslie and McLeod, 2007; OECD, 2009; SCBD, 2009. 5. CCCD, 2009; Colls et al., 2009; FAO,
2008b; ProAct Network, 2008; UNFCCC, 2006a. 6. Chhatre and Agrawal, 2009; FAO, 2008b; Mansourian et al., 2009; Reid and Huq, 2005; SCBD, 2009. 7. FAO, 2008a; Reid and Huq, 2005; SCBD, 2009; UNEP, 2006; Venter et al., 2009. 8. Arnell, 2004; Branco et al., 2005;
Campbell et al., 2008; Easterling et al., 2007; FAO, 2008a, 2009; Howden et al., 2007; McGray et al., 2007; Neufeldt et al., 2009; SCBD, 2009; UNISDR, 2009d; World Bank, 2009. 9. Easterling et al., 2007; FAO, 2007, 2010; Hammer et al., 2003; McCarl, 2007; UNFCCC,
2006a; World Bank, 2009. 10. Campbell et al., 2008; CCCD, 2009; Easterling et al., 2007; FAO, 2007, 2010; World Bank, 2009. 11. Easterling et al., 2007; FAO, 2007, 2010; World Bank, 2009. 12. Butler and Oluoch-Kosura, 2006; Butt et al., 2005; CCCD, 2009; Davis, 2004;
FAO, 2006, 2008a; Howden et al., 2007; McCarl, 2007; World Bank, 2009. 13. CCCD, 2009; FAO, 2007; UNISDR, 2009d; World Bank, 2009. 14. FAO, 2007, 2008a; Rosenzweig and Tubiello, 2007. 15. Adger et al., 2005; FAO, 2008c; Kay and Adler, 2005; Kesavan and Swamina-
than, 2006. 16. Adger et al., 2005; FAO, 2008c; Kesavan and Swaminathan, 2006; Klein et al., 2001; Nicholls, 2007; Nicholls et al., 2008; Romieu et al., 2010; UNFCCC, 2006a. 17. FAO, 2007, 2008c; Rahel and Olden, 2008; UNFCCC, 2006a. 18. Adger et al., 2005; Dolan
and Walker, 2006; FAO, 2008b; Nicholls, 2007; Thorne et al., 2006; UNFCCC, 2006b; World Bank, 2010. 19. FAO, 2008c; Kesavan and Swaminathan, 2006; Rahel and Olden, 2008. 20. FAO, 2007, 2008c; IIED, 2009; Romieu et al., 2010. 21. FAO, 2007, 2008c; Nicholls, 2007.
22. FAO, 2008c; UNFCCC, 2006a. 23. Branco et al., 2005; CCCD, 2009; Hedger and Cacouris, 2008; ICHARM, 2009; Klijn et al., 2004; Krysanova et al., 2010; Mills, 2007; Olsen, 2006; Rahaman and Varis, 2005; World Bank, 2009; WSSD, 2002; WWAP, 2009. 24. Arnell and
Delaney, 2006; Auld et al., 2004; CCCD, 2009; DaSilva et al., 2004; Hedger and Cacouris, 2008; Krysanova et al., 2010; Mills, 2007; Muller, 2007; Thomalla et al., 2006; UNFCCC, 2006b, 2009a; WHO, 2003; WWAP, 2009; WWC, 2009. 25. CCCD, 2009; Crabbé and Robin,
2006; Hedger and Cacourns, 2008; Krysanova et al., 2010; Rahaman and Varis, 2005; WWAP, 2009. 26. Few et al., 2006; Kirshen, 2007; Mills, 2007; Rahaman and Varis, 2005; Warner et al., 2009; WWAP, 2009. 27. Auld, 2008b; Haasnoot et al., 2009; Hodgson and Carter,
1999; Lowe, 2003; Mills, 2007; Nicholls et al., 2008; NRTEE, 2009; ProVention, 2009; Rossetto, 2007; Wamsler, 2004; Wilby and Dessai, 2010; World Bank, 2000, 2008; WWC, 2009. 28. Auld, 2008a,b; Haasnoot et al., 2009; Hallegatte, 2009; Lewis and Chisholm, 1996; Mills,
2007; Neumann, 2009; Nicholls et al., 2008; ProVention, 2009; Rosetto, 2007; UNFCCC, 2006a. 29. Auld, 2008b ; Islam and Ferdousi, 2007; Kagiannas et al., 2003; Maréchal, 2007; Mills, 2007; Neumann, 2009; Robledo et al., 2005; Van Buskirk, 2006; Warner et al., 2009;
Younger et al., 2008. 30. Auld, 2008b; Freeman and Warner, 2001; Mills, 2007; Neumann, 2009; NRTEE, 2009; ProVention, 2009; Stevens, 2008; Younger et al., 2008. 31. Auld et al., 2004; Auld, 2008a; CCCD, 2009; Curriero et al., 2001; DaSilva et al., 2004; Ebi et al., 2006;
Haines et al., 2006; Patz et al., 2000, 2005; UNFCCC, 2006a; WHO, 2003, 2005; World Bank, 2003. 32. CCCD, 2009; Ebi et al., 2006; Ebi, 2008; Haines et al., 2006; Patz et al., 2005; UNFCCC, 2006a; WHO, 2003, 2005; Younger et al., 2008. 33. Mills, 2005, 2006. 34. Haines
et al., 2006; Younger et al., 2008.
Table 6-1 (continued)
355
into the future, some studies recommend the use of pro-adaptation and
robust options to deal with climate change uncertainties (Auld, 2008b;
Hallegatte, 2009; Wilby and Dessai, 2010). These robust options include
actions that are reversible, flexible, less sensitive to future climate
conditions (i.e., no and low regret), and can incorporate safety margins
(e.g., infrastructure investments), employ ‘soft’ solutions (e.g., ecosystem
services), and are mindful of actions being taken by others to either
reduce greenhouse gases (GHGs) or adapt to climate change in other
sectors (Hallegatte, 2009; Wilby and Dessai, 2010). Flexible options are
those that provide benefits under a variety of climate conditions or
reduce stress on affected systems to increase their flexibility (e.g.,
reducing pollution or demand on resources) (Auld, 2008b; Hallegatte,
2009; Wilby and Dessai, 2010).
Options that allow for incremental changes in, for example, infrastructure
over time, or allow incorporation of future change, for example, support
more flexible systems (Auld, 2008b; Hallegatte, 2009; OECD, 2009).
Uncertainties over future risks can also be accounted for through ‘safety
margin’ or over-design strategies to reduce vulnerability and increase
resiliency at low and sometimes null costs (Auld, 2008b; Hallegatte, 2009).
These safety margin strategies have been used to manage future risks for
sea level rise and coastal defenses, for water drainage management,
and for investments in other infrastructure (Hallegatte, 2009). Given
uncertainties, national policies may need to become more adaptable
and flexible, particularly where national plans and policies currently
operate within a limited range of conditions and are based on certainty
(McGray et al., 2007; Wilby and Dessai, 2010). Without flexibility, rigid
national policies may become disconnected from evolving climate risks and
bring unintended consequences or maladaptation (Sperling and Szekely,
2005; Hallegatte, 2009). Rigid plans and policies that are irreversible
and based on a specific climate scenario that does not materialize can
result in future maladaptation and imply wasted investments or harm to
people and ecosystems that can prove unnecessary.
Several studies indicate that national plans and policies for adaptation
to climate change and disaster risk management tend to favor options
that deal with the current or near-term climate risks and ‘win-win’
options that satisfy multiple synergies for GHG reduction, disaster risk
management, climate change adaptation, and development issues
(World Bank, 2008; Heltberg et al., 2009; Ribeiro et al., 2009;
Fankhauser, 2010; Mitchell and Maxwell, 2010). Many of these ‘win-
win’ options include ecosystem-based adaptation actions, sustainable
land and water use planning, carbon sequestration, energy efficiency,
and energy and food self-sufficiency. For example, the ecosystem
management practices of afforestation, reforestation, and conservation of
forests offer co-benefits for disaster risk reduction from floods, landslides,
avalanches, coastal storms, and drought while contributing to adaptation
to future climates, economic opportunities, increased biomass and
carbon sequestration, energy efficiency, energy savings, as well as energy
and food self sufficiency (Thompson et al., 2009).
Disaster risk transfer options offer a viable adaptation response to current
and future climate risks and include instruments such as insurance,
micro-insurance, and micro-financing; government disaster reserve
funds; government-private partnerships involving risk sharing; and new,
innovative insurance mechanisms (Linnerooth-Bayer and Mechler, 2006;
EC, 2009; World Bank, 2010). Risk transfer options can provide much
needed, immediate liquidity after a disaster, allow for more effective
government response, provide some relief from the fiscal burden placed
on governments due to disaster impacts, and constitute critical steps in
promoting more proactive risk management strategies and responses
(Arnold, 2008). Case Study 9.2.13 and Section 6.5.3 provide more detail
on risk transfer options.
Even with risk transfer instruments and adaptation to climate change
options in place, residual losses can be realized when extreme events –
well beyond those typically expected – result in high impacts. In spite of
the evidence, decisions to ignore increasing future risks and even
current risks remain common, particularly when uncertainties over the
directions of future climate change impacts are high, when capacity is
initially very limited, adaptation options are not available, or when the
risks of future impacts are considered to be very low (Linnerooth-Bayer
and Mechler, 2006; Heltberg et al., 2009; World Bank, 2010).The losses
from deferring adaptation and disaster risk reduction actions are borne
by all actors.
Table 6-1 outlines some of the adaptation to climate change and disaster
risk management policy and planning options available nationally for
selected sectors and described in the literature. Many of these options
are incremental actions that complement and reinforce each other. The
actions are organized using the gradations of planning and policy
options described in this section.
6.3.2. Mainstreaming Disaster Risk Management
and Climate Change Adaptation into
Sectors and Organizations
National adaptation to climate change will involve stand-alone adaptation
policies and plans as well as the integration or mainstreaming of
adaptation measures into existing activities (OECD, 2009). Mainstreaming
of adaptation and disaster risk management actions implies that national,
sub-national, and local authorities adopt, expand, and enhance measures
that factor disaster and climate risks into their normal plans, policies,
strategies, programs, sectors, and organizations (Few et al., 2006; UNISDR,
2008a; OECD, 2009; Biesbroek et al., 2010; CACCA, 2010).
In reality, it can be challenging to provide clear pictures of what
mainstreaming is, let alone how it can be made operational, supported,
and strengthened at the various national and sub-national levels (Olhoff
and Schaer, 2010). Some studies indicate that the real challenge to
mainstreaming adaptation is not planning but implementation
(Biesbroek et al., 2010; Krysanova et al., 2010; Tompkins et al., 2010).
Some of the barriers to implementation include lack of funding, limited
budget flexibility, lack of relevant information or expertise, lack of
political will or support, and institutional silos (Krysanova et al., 2010;
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
356
Preston et al., 2011). Studies indicate that effective plans, policies, and
programs for adaptation to climate change and disaster risk management
need to go beyond identifying potential options to include better
inventories of existing assets and liabilities for managing risk and
specific actions for overcoming adaptation barriers (Haasnoot et al.,
2009; Preston et al., 2011).
Recent studies investigating the success of existing adaptation plans
and policies for Australia, the United States, countries in Europe, and
major river basins in Africa and Asia, for example, indicate that there is
a need for mainstreaming of adaptation into existing national policies
and plans and a priority for capitalizing on ‘win-win’ or options that
take advantage of synergies with other national objectives (Biesbroek et
al., 2010; Tompkins et al., 2010; Preston et al., 2011). The studies found
that many strategies and institutions were focused to a greater extent
on lower-risk actions dealing with science and outreach (knowledge
acquisition) and capacity building rather than moving forward on
specific, more costly and difficult to implement adaptation and disaster
risk management actions and managing at-risk public goods (Tompkins
et al., 2010; Preston et al., 2011).
Preston et al. (2011) found in their studies from Australia, the United
States, and the United Kingdom that most national adaptation strategies
were based on vulnerability assessments informed by broad international
and national climate change guidance, rather than any consistent or
systematic use of scenarios, and favored bottom-up approaches for
coordination across sectors and multiple government scales. Biesbroek
et al. (2010) noted similar results for nine countries in Europe. Tompkins
et al. (2010) and Krysanova et al. (2010) found that the sectors with the
highest levels of adaptation implementation in the United Kingdom were
those that tended to be most affected by current weather variability and
extremes and that specific government initiatives had been successful
in stimulating adaptation and disaster risk reduction (e.g., mandatory
planning for flood-prone areas, ISO 14001). Tompkins et al. (2010) also
found that successful implementation frequently resulted from multiple
triggers, that few of these adaptation actions were solely initiated in
response to climate change, and that the relative impact of weather on
core business and organizational culture encouraged an ability and
willingness to proactively act on climate change information.
Adaptation to climate change and disaster risk management needs to
typically identify more adaptation options than most countries can
reasonably implement in the short term due to resource constraints,
requiring that actions be prioritized (OECD, 2009; Krysanova et al., 2010).
Initially, actions that remove the existing barriers to managing disaster
risks from today’s climate variability can help to reduce the even greater
barriers to managing future climate risks (UNDP, 2002, 2004a; CCCD,
2009; Prabhakar et al., 2009; Tompkins et al., 2010). As a result, a key
challenge, and an opportunity for mainstreaming adaptation and disaster
risk management, lies in building bridges between current disaster risk
management actions for existing climate vulnerabilities and the additional
revised efforts needed for future vulnerabilities (Few et al., 2006; Krysanova
et al., 2010; Olhoff and Schaer, 2010; Wilby and Dessai, 2010).
An important prerequisite for informed decisions on adaptation to
climate change and disaster risk management is that they should be
based upon the best available information (OECD, 2009; Biesbroek et al.,
2010; Lu, 2011). Preston et al. (2011) noted that many of the specific
adaptation plans from Australia, the United States, and the United
Kingdom indicated a need for improved gathering and sharing of climate
and climate change science information prior to or in conjunction with
the delivery of adaptation actions, perhaps reflecting a preference for
delaying adaptation actions until greater certainty or better information
on different adaptation actions was known. As noted in Chapter 3
(Section 3.2.3 and Box 3-2), many extreme events occur at small
temporal and spatial scales, where climate change models, even when
downscaled, cannot provide simulations at such spatial and temporal
resolutions. A number of studies also contend that increased and better
information on climate change scenarios and projections and potential
impacts will accomplish little on their own to mainstream and alter
on-the-ground decisions, policies, and plans unless the information
provided can directly meet decisionmakers’ needs (Stainforth et al.,
2007; Auld, 2008b; Haasnoot et al., 2009; Krysanova et al., 2010;
Mastrandrea et al., 2010; Wilby and Dessai, 2010). Users require
relevant climate risk information that is accessible, can be explained in
understandable language, provides straightforward estimates of
uncertainties, and is relevant or tailored to their management
functions (Stainforth et al., 2007; Mastrandrea et al., 2010; Lu, 2011).
Increasingly, studies are showing that this is best accomplished
through sustained interactions between scientists and stakeholders and
policymakers, usually maintained through years of relationship- and
trust-building (Mastrandrea et al., 2010; Wilby and Dessai, 2010; Lu,
2011).
Studies generally indicate that the most essential means for effectively
mainstreaming both adaptation and disaster risk management nationally
involve ‘whole of government’ coordination across different levels and
sectors of governance, including the involvement of a broad range of
stakeholders (Few et al., 2006; Thomalla et al., 2006; OECD, 2009; also
Section 6.4.2). In spite of the strong interdependencies, governments have
tended to manage these issues in their ‘silos’ with environment or energy
authorities and scientific institutions typically responsible for climate
change adaptation while disaster risk management authorities may
reside in a variety of national government departments and national
disaster management offices (Sperling and Szekely, 2005; Thomalla et
al., 2006; Prabhakar et al., 2009). Progress in planning for adaptation and
developing and implementing strategies within government agencies
usually depends on political commitment, institutional capacity, and, in
some cases, on enabling legislation, regulations, and financial support
(Few et al., 2006; OECD, 2009; Krysanova et al., 2010; see Section 6.4).
Nationally, studies indicate that it may be important to clearly identify
a lead for disaster and climate risk reduction efforts where that lead has
influence on budgeting and planning processes (Few et al., 2006; OECD,
2009). In some cases, countries and regions may be able to build on
phases of raised awareness and increased attention to disaster risk in
order to develop and strengthen their responsible institutions (Few et
al., 2006; Krysanova et al., 2010).
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
357
While developed countries may be more financially equipped to meet
many of the challenges of mainstreaming adaptation and disaster risk
reduction into national plans and policies, the situation is often more
challenging in developing countries (Krysanova et al., 2010). Nonetheless,
there are examples from developing countries where adaptation to
climate change and disaster risk management mainstreaming issues
have been priorities for many years and significant progress in
mainstreaming has been noted (e.g., the Caribbean Mainstreaming
Adaptation to Climate Change project, which was implemented from 2004
to 2007; Case Studies 9.2.9 and 9.2.12). In other cases, international
funding mechanisms such as the Least Developed Countries (LDC) Fund,
the Special Climate Change Fund, the Multi-donor Trust Fund on Climate
Change, and the Pilot Programme for Climate Resilience under the
Climate Investment Fund are making funding and resources available to
developing countries to pilot and mainstream changing climate risks
and resilience into core development and as an incentive for scaled-up
action and transformational change, although needs exceed availability
of funds (O’Brien et al., 2008; Krysanova et al., 2010; see Sections
7.4.3.3 and 7.4.2 for additional discussion).
6.3.3. Sector-Based Risk Management and Adaptation
The challenge for countries is to manage short-term climate variability
while also ensuring that different sectors and systems remain resilient
and adaptable to changing extremes and risks over the long term
(Füssel, 2007; Wilby and Dessai, 2010). The requirement is to balance the
short-term and the longer-term actions needed to resolve the underlying
causes of vulnerability and to understand the nature of changing climate
hazards (UNFCCC, 2008a; OECD, 2009). Achieving adaptation and
disaster risk management objectives while attaining human development
goals requires a number of cross-cutting, interlinked sectoral and
development processes, as well as effective strategies within sectors
and coordination between sectors (Few et al., 2006; Thomalla et al.,
2006; Biesbroek et al., 2010). Climate change is far too big a challenge
for any single ministry of a national government to undertake (CCCD,
2009; Biesbroek et al., 2010).
Sector-based organizations and departments play a central role in
national decisionmaking and are a logical focus for adaptation actions
(McGray et al., 2007; Biesbroek et al., 2010). The impacts of changing
climate risks in one sector, such as tourism, can affect other sectors and
scales significantly, especially since sectoral linkages operate both
vertically and horizontally. Sector plans, policies, and programs are
linked vertically from national to local levels within the same sector as
well as horizontally across different sectors at the same level (Urwin
and Jordan, 2008; UNFCCC, 2008b; CCCD, 2009; Biesbroek et al., 2010).
While the case and need for integration within sectors and levels may
be clear, the issue of how to integrate or mainstream nationally across
multiple sectors and multiple levels still remains challenging, requiring
governance mechanisms and coordination that can cut across
governments and sectoral organizations (UNISDR, 2005; UNFCCC,
2008b; CCCD, 2009; ONERC, 2009; Biesbroek et al., 2010). Typically,
multi-sector integration tends to deal with the broader national scale
(e.g., entire economy or system) and aims to be as comprehensive as
possible in covering several affected sectors, regions, and issues (UNFCCC,
2008b). Studies from organizations and academia indicate that effective
adaptation and risk reduction coordination between all sectors may
only be realized if all areas of government are coordinated from the
highest political and organizational level (Schipper and Pelling, 2006;
UNFCCC, 2008b; CCCD, 2009; Prabhakar, 2009). Even when ‘political
champions’ at the highest levels encourage mainstreaming across sectors
and departments, competing national priorities will remain an impediment
to progress.
Table 6-1 (Section 6.3.1) outlines adaptation to climate change and
disaster risk management options for several selected sectors. As the
table indicates, adaptation and disaster risk management approaches
for many development sectors benefit jointly from ecosystem-based
adaptation and integrated land, water, and coastal zone management
actions. For example, conservation and sustainable management of
ecosystems, forests, land use, and biodiversity have the potential to create
win-win disaster risk protection services for agriculture, infrastructure,
cities, water resource management, and food security. They can also
create synergies between climate change adaptation and mitigation
measures (SCBD, 2009; CCCD, 2009), as well as produce many co-
benefits that address other development goals, including improvements
in livelihoods and human well being, particularly for the poor and
vulnerable, and biodiversity conservation, and are discussed further in
Section 6.5.2.3 and in Case Studies 9.2.3, 9.2.4, 9.2.5, 9.2.7, 9.2.8, and
9.2.9. Likewise, water resource, land, and coastal zone management
options deal with many sectors and issues and jointly provide disaster
risk management and adaptation solutions, as mentioned in Case Studies
9.2.6 and 9.2.8 (WHO, 2003; Urwin and Jordan, 2008; UNFCCC, 2008b;
CCCD, 2009; WWAP, 2009). Human health is a cross-cutting issue
impacted by actions taken in many sectors, as indicated in Table 6-1 and
discussed in Case Studies 9.2.2 and 9.2.7.
6.4. Strategies including Legislation,
Institutions, and Finance
National systems for managing the risks of extreme events and disasters
are shaped by legislative provision, compliance mechanisms, the nature
of cross-stakeholder bodies, and financial and budgetary processes that
allocate resources to actors working at different scales. These elements
help to create the technical architecture of national systems and are often
led by national government agencies. However non-technical dimensions
of good governance, such as the distribution and decentralization of
power and resources, processes for decisionmaking, transparency, and
accountability are woven into the technical architecture and are significant
factors in determining the effectiveness of risk management systems
and actions (UNDP, 2004b, 2009). These technical and non-technical
aspects of risk governance vary between countries as governance
capacity varies (and, as detailed in Section 6.3, are critical in shaping
investment in particular adaptation and disaster risk management
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
358
options). Accordingly, risks can be addressed through both formal and
informal governance modes and institutions in all countries (Jaspars
and Maxwell, 2009), but a clear correlation between particular risk
governance models and specific political-administrative contexts is
difficult to identify (UNISDR, 2011a). The balance between formal or
informal, or technical and non-technical, risk governance strategies
depends on the economic, political, and environmental contexts of
individual countries or scales within countries, and the culture of
managing risks (Menkhaus, 2007; Kelman, 2008).
6.4.1. Legislation and Compliance Mechanisms
Disaster risk management legislation commonly establishes organizations
and their mandates, clarifies budgets, provides (dis)incentives, and
develops compliance and accountability mechanisms (UNDP, 2004b;
Llosa and Zodrow, 2011). Creating and improving legislation for disaster
risk reduction was included as a priority area in the HFA (UNISDR, 2005)
and the majority of countries – in excess of 80% – now have some form
of disaster risk management legislation (UNISDR, 2005; Bhavnani et al.,
2008). Legislation continues to be considered as an important component
of effective national disaster risk management systems (UNDP, 2004b;
UNISDR, 2011a) as it creates the legal context of the enabling environment
in which others, working at different scales, can act, and it helps to
define people’s rights to protection from disasters, assistance, and
compensation (Pelling and Holloway, 2006). Multi-stakeholder, cross-
sector bodies for coordinating disaster risk management actions and
implementing the HFA, known commonly as National Platforms, are
seen as key advocacy routes for achieving new and improved legislation
(UNISDR, 2005, 2007b). Where National Platforms are less prevalent or
less well organized, literature suggests that regional disaster management
bodies are viewed as responsible for advancing legislation (Pelling and
Holloway, 2006; UNISDR, 2007b). With new information on the impacts
of climate change, legislation on managing disaster risk may need to be
modified and strengthened to reflect changing rights and responsibilities
and to support the uptake of adaptation options (UNDP, 2009; Llosa and
Zodrow, 2011; see Case Study 9.2.12 on legislation).
There have been few detailed cross-comparative studies that assess the
extent to which legislation in different countries is oriented toward
managing uncertainty and reducing disaster risk compared with disaster
response (Llosa and Zodrow, 2011). Limited evidence suggests that
legislation in some countries (such as the United Kingdom, the United
States, and Indonesia) has led to a focus on building institutional capacity
to help create resilience to disasters at different scales, but even in such
cases a strongly reactive culture is retained when observing the system
as a whole (O’Brien and Read, 2005, O’Brien, 2006, 2008; UNDP, 2009;
O’Brien and O’Keefe, 2010). This has been attributed to lack of political
will and insufficient financial and human resources for disaster risk
reduction (O’Brien 2006, 2008). Additionally, few studies have assessed
whether disaster risk management legislation includes provision for the
impact of climate change on disaster risk or whether aspects of managing
disaster risk are included in other complementary pieces of legislation
(Case Study 9.2.12; Llosa and Zodrow, 2011), though there are also a very
limited number of normative studies on these aspects (Llosa and Zodrow,
2011). However, where reforms of disaster management legislation
have occurred, they have tended to: (a) demonstrate a transition from
emergency response to a broader treatment of managing disaster risk;
(b) recognize that protecting people from disaster risk is at least partly
the responsibility of governments; and (c) promote the view that reducing
disaster risk is everyone’s responsibility (Case Study 9.2.12; UNDP,
2004b; Llosa and Zodrow, 2011).
Vietnam has taken steps to integrate disaster risk management into
legislation across key development sectors, including its Land Use Law
and Law on Forest Protection. Vietnam’s Poverty Reduction Strategy
Paper also included a commitment to reduce by 50% those falling back
into poverty as a result of disasters and other risks (Pelling and Holloway,
2006). Case Study 9.2.12, in examining legislation development
processes in the Philippines and South Africa, highlights a number of
components of effective disaster risk management legislation. An
act needs to be: (a) comprehensive and overarching; (b) establish
management structures and secure links with development processes at
different scales; and (c) establish participation and accountability
mechanisms that are based on information provision and effective
public awareness and education. Box 6-1 supplements these cases with
reflections on the process that led to the creation of disaster risk
management legislation in Indonesia.
Where risk management dimensions are a feature of national legislation,
positive changes are not always guaranteed (UNDP, 2004b). A lack of
financial, human, or technical resources and capacity constraints present
significant obstacles to full implementation, especially as experience
suggests that legislation should be implemented continuously from the
national to local level and is contingent on strong monitoring and
enforcement frameworks and adequate decentralization of responsibilities
and human and financial resources at every scale (UNDP, 2004b; Pelling
and Holloway, 2006). In some countries, building codes, for instance, are
often not implemented properly because of a lack of technical capacity
and political will of officials concerned (UNDP, 2004b). Where enforcement
is unfeasible, accountability for disaster risk management actions is
extremely challenging; this supports the need for an inclusive, consultative
process for discussing and drafting the legislation (UNDP, 2004b;
UNISDR, 2007b). Effective legislation includes benchmarks for action, a
procedure for evaluating actions, integrated planning to assist coordination
across geographical or sectoral areas of responsibility, and a feedback
system to monitor risk reduction activities and their outcomes (UNISDR,
2005; Pelling and Holloway, 2006).
6.4.2. Coordinating Mechanisms and Linking across Scales
As the task of managing the risks of changing climate conditions and
climate extremes and disasters cuts across the majority of sectors and
involves a wide range of actors, multi-stakeholder and cross-government
mechanisms are commonly cited as preferred way to ‘organize’ disaster
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
359
risk management systems at the national level (UNISDR, 2005, 2007b;
see Section 6.3.3), as well as for addressing the challenges associated
with adaptation to climate change (ONERC, 2009). The HFA terms these
‘National Platforms,’ defined as a “generic term for national mechanisms
for coordination and policy guidance on disaster risk reduction that are
multi-sectoral and inter-disciplinary in nature, with public, private and civil
society participation involving all concerned entities within a country”
(UNISDR, 2005). In some countries such coordinating mechanisms are
referred to by other names (Hay, 2009; Gero et al., 2011) but essentially
perform the same function. Guidelines on establishing National Platforms
suggest that they need to be built on existing relevant systems and
should include participation from different levels of government, key
line ministries, disaster management authorities, scientific and academic
institutions, civil society, the Red Cross/Red Crescent, the private sector,
opinion shapers, and other relevant sectors associated with disaster
risk management (UNISDR, 2007b). Evaluations and reflections on the
effectiveness of National Platforms for delivering results on the HFA and
on disaster risk management more broadly indicate widely varying
results (GTZ/DKKV, 2007; UNISDR, 2007c, 2008c; UNISDR/DKKV/Council
of Europe, 2008; Sharma, 2009). An assessment in Asia found
National Platforms struggling to obtain the legal mandate to secure full
participation of stakeholders, particularly NGOs, difficulty in obtaining
sustainable funding sources, and challenges associated with translating
intent into implementation (Sharma, 2009). On the other hand, pockets of
evidence exist where National Platforms have succeeded in generating
senior political commitment for disaster risk reduction, in strengthening
integration of disaster risk reduction into national policy and development
plans, and in establishing institutions and programs on disaster risk
management with engagement from academia, media, and the private
sector (UNISDR, 2008b; Sharma, 2009). This assessment found only a
limited number of genuinely independent studies on the effectiveness
of National Platforms, with evidence particularly weak in Africa and
elsewhere.
While the evidence again suggests significant differences between
countries, on balance, national coordination mechanisms for adaptation to
climate change and disaster risk management remain largely disconnected,
although evidence suggests that the trajectory is one of improvement
(National Platform for Kenya, 2009; Mitchell et al., 2010b; discussed in
Chapter 1). Benefits of improved coordination between adaptation to
climate change and disaster risk management bodies, and development
and disaster management agencies, include the ability to (i) explore
common tradeoffs between present and future action, including
addressing human development issues and reducing sensitivity to
disasters versus addressing post-disaster vulnerability; (ii) identify
synergies to make best use of available funds for short- to longer-term
adaptation to climate risks as well as to tap into additional funding
sources; (iii) share human, information, technical, and practice resources;
(iv) make best use of past and present experience to address emerging
risks; (v) avoid duplication of project activities; and (vi) collaborate on
reporting requirements (Mitchell and Van Aalst, 2008). Barriers to
integrating disaster risk management and adaptation coordination
mechanisms include the underdevelopment of the ‘preventative’
component of disaster risk management, the paucity of projects that
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
Box 6-1 | Enabling Disaster Risk Management Legislation in Indonesia
Indonesia: Disaster Management Law (24/2007)
The legislative reform process in Indonesia that resulted in the passing of the 2007 Disaster Management Law (24/2007) created a
stronger association between disaster risk management and development planning processes. The process was considered successful
due to the following factors:
Strong, visible professional networks – Professional networks born out of previous disasters meant a high level of trust and
willingness to coordinate and became pillars of the legal reform process. The political and intellectual capital in these networks,
along with leadership from the MPBI (The Indonesian Society for Disaster Management), was instrumental in convincing the
lawmakers about the importance of disaster management reform.
Civil society leading the advocacy – Civil society leading the advocacy for reform has resulted in CSOs being recognized by the
Law as key actors in implementing disaster risk management in Indonesia.
The impact of the 2004 South Asian tsunami helping to create a supportive political environment – The reform process was
initiated in the aftermath of the tsunami that highlighted major deficiencies in disaster management. However, the direction of the
reform (from emergency management toward disaster risk reduction) was influenced by the international focus, through the HFA,
on disaster risk reduction.
An inclusive drafting process – Consultations on the new Disaster Management Law were inclusive of practitioners and civil
society, but were not so far-reaching as to delay or lose focus on the timetable for reform.
Consensus that passing an imperfect law is better than no law at all An imperfect law can be supplemented by additional
regulations, which helps to maintain interest and focus.
Source: UNDP (2004b, 2009); Pelling and Holloway (2006).
360
integrate climate change in the context of disaster risk management,
disconnects between different levels of government, and the weakness
of both disaster risk management and adaptation to climate change in
national planning and budgetary processes (Few et al., 2006; Mitchell
and Van Aalst 2008; Mitchell et al., 2010b) (see Box 6-2).
While national level coordination is important and the majority of risks
associated with disasters and climate extremes are owned by national
governments and are managed centrally (see Section 6.2.1), sources
suggest that decentralization can be an effective risk management
strategy, especially in support of community-based disaster risk
management processes (Mitchell and Van Aalst, 2008; GNDR, 2009;
Scott and Tarazona, 2011). However, there are few studies that critically
examine the effectiveness of decentralization of disaster risk management
in detail (Twigg, 2004; Tompkins et al., 2008; Scott and Tarazona, 2011).
One such study of four countries – Colombia, Mozambique, Indonesia,
and South Africa – found that effective decentralization of disaster risk
reduction can be constrained by (a) low capacity at the local level;
(b) funds dedicated to disaster risk reduction often being channeled
elsewhere; (c) the fact that decentralization does not automatically lead
to more inclusive decisionmaking processes; (d) an appreciation that
decentralized systems face significant communications challenges; and
(e) knowledge that robust measures for ensuring accountability and
transparency are vital for effective disaster risk management but are
often missing (Scott and Tarazona, 2011). It appears that motivation for
management at a particular scale promises to influence how well the
impacts of disasters and climate change are managed, and therefore
affect disaster outcomes (Tsing et al., 1999). Decisions made at one scale
may have unintended consequences for another (Brooks and Adger,
2005), meaning that governance decisions will have ramifications across
scale and contexts. In all cases, the selection of a framework for
governance of disasters and climate change-related risks may be issue-
or context-specific (Sabatier, 1986).
6.4.3. Finance and Budget Allocation
Governments in the past have ignored catastrophic risks in decisionmaking,
implicitly or explicitly exhibiting risk-neutrality (Mechler, 2004). This is
consistent with the Arrow Lind theorem (Arrow and Lind, 1970), according
to which a government may be well equipped to efficiently (i) pool risks
as it possesses a large number of independent assets and infrastructure
so that the aggregate risk converges to zero, and/or (ii) spread risk
across the taxpaying population base, so that per capita risk accruing to
risk-averse households converges to zero. In line with this theorem, due
to their ability to spread and diversify risks, governments are sometimes
termed ‘the most effective insurance instrument of society’ (Priest 1996).
Accordingly, it has been deduced that, although individuals are risk-
averse [to disasters risk], governments can take a risk-neutral approach.
However, the experiences of highly exposed countries suggest otherwise
and have led to a recent paradigm shift, with governments changing
from being ‘risk neutral’ to being risk averse and managing disaster risks.
Many highly exposed developing and developed countries (especially in
the wake of the recent financial crisis) have very limited economic
means, rely on small and exhausted tax bases, have high levels of
indebtedness, and are unable to raise sufficient and timely capital to
replace or repair damaged assets and restore livelihoods following
major disasters. This can lead to increased impacts of disaster shocks on
poverty and development (OAS, 1991; Linnerooth-Bayer et al., 2005;
Hochrainer, 2006; Mahul and Ghesquiere, 2007; Cummins and Mahul,
2009). Exposed countries thus have had to rely on donors to ‘bail’ them
out after events, although ex-post assistance usually only provides
partial relief and reconstruction funding, and such assistance is also often
associated with substantial time lags (Pollner, 2000; Mechler, 2004).
Furthermore, extreme events that are associated with large losses may
lead to important downstream economic effects (see Section 4.5),
causing depressed incomes and reduced ability to share the losses.
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
Box 6-2 | National and Sub-National Coordination for Managing Disaster Risk
in a Changing Climate: Kenya
Kenya’s National Platform is situated under the Office of the President and has made significant achievements in coordinating multiple
stakeholders, but is constrained by limited resources and lack of budgets for disaster risk reduction in line ministries (National Platform
for Kenya, 2009). Some key constraints of the national system are recognized as being difficulties in integrating disaster risk reduction in
planning processes in urban and rural areas and lack of data on risks and vulnerabilities at different scales (Few et al., 2006). In this regard,
Nairobi has experienced periods of drought and heavy rains in the last decade, prompting action to reduce exposure and vulnerability to
what is perceived as changing hazard trends (ActionAid, 2006). Increasing exposure and vulnerability has resulted from a rapid expansion
of poor people living in informal settlements around Nairobi, leading to houses of weak building materials being constructed immediately
adjacent to rivers and blocking natural drainage areas. While data and coordination systems are still lacking, the Government of Kenya
has established the Nairobi Rivers Rehabilitation and Restoration Programme (African Development Bank Group, 2010), designed to
install riparian buffers, canals, and drainage channels, while also clearing existing channels. The Programme also targets the urban poor
with improved water and sanitation, paying attention to climate variability and change in the location and design of wastewater
infrastructure and environment monitoring for flood early warning (African Development Bank Group, 2010). This demonstrates the kind
of options for investments that can be achieved in the absence of a fully fledged nationally coordinated disaster management system
and in the absence of complete multi-hazard, exposure, and vulnerability data sets.
361
Consequently, a risk-neutral stance in dealing with catastrophic risk
(implying that the consideration of risk broadly in terms of means – the
statistical expectation – is sufficient) may not be suitable for exposed
developing countries with limited diversification of their economies or
small tax bases. Accordingly, assessing and managing risks over the
whole spectrum of probabilities is gaining momentum (Cardenas et al.,
2007; Cummins and Mahul, 2009). As the Organization of American States
suggests: “Government decisions should be based on the opportunity
costs to society of the resources invested in the project and on the loss
of economic assets, functions and products. In view of the responsibility
vested in the public sector for the administration of scarce resources,
and considering issues such as fiscal debt, trade balances, income
distribution, and a wide range of other economic and social, and
political concerns, governments should not act risk-neutral” (OAS,
1991). Also, in more developed economies, less-pronounced but still
considerable effects imposed by events linked to climate variability can
be identified. This has been shown by the Austrian political and fiscal
crisis in the aftermath of large-scale flooding that led to billions of
Euros in losses in 2002 (Mechler et al., 2010).
Budget and resource planning for extremes is not an easy proposition.
Governments commonly plan and budget for direct liabilities, that is,
liabilities that manifest themselves through certain and annually recurrent
events. Those liabilities are of explicit (as recognized by law or contract),
or implicit nature (moral obligations) (see Table 6-2). Yet, governments
are not good at planning for contingencies even for probable events, let
alone improbable events. Explicit, contingent liabilities deal with the
reconstruction of infrastructure destroyed by events, whereas implicit
obligations are associated with providing relief – commonly considered
as a moral liability for governments (Polackova Brixi and Mody, 2002).
In many countries, governments do not explicitly plan for contingent
liabilities, and rely on reallocating their resources following disasters
and raising capital from domestic and international donations to meet
infrastructure reconstruction needs and costs.
More recently, some developing and transition countries that face large
contingent liabilities in the aftermath of extreme events and associated
financial gaps have begun to plan for and consider contingent natural
events (also see Section 6.5.3). Mexico, Colombia, and many Caribbean
countries now include contingent liabilities in their budgetary process
and eventually even transfer some of these risks (Cardenas et al., 2007;
Linnerooth-Bayer and Mechler, 2007; Cummins and Mahul, 2009; see
Box 6-3). Similarly, many countries have also started to focus on
improving human development conditions as an adaptation strategy for
climate change and extreme events, particularly with the help of
international agencies such as the World Bank. These deliberations are
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
Table 6-2 | Government liabilities and disaster risk. Modified from Polackova Brixi and
Mody (2002).
Liabilities
Direct: obligation in
any event
Contingent: obligation if
a particular event occurs
Explicit: Government
liability recognized by
law or contract
Foreign and domestic
sovereign borrowing,
expenditures by budget law
and budget expenditures
State guarantees for non-
sovereign borrowing and
public and private sector
entities, reconstruction of
public infrastructure
Implicit: A ‘moral’
obligation of the
government
Future recurrent costs of
public investment projects,
pensions, and health care
expenditure
Default of sub-national
government as public or
private entities provide disaster
relief
Reduce Risks
Manage Residual Risks and Uncertainties
Risks Acceptance
Threshold
Mutual and reserve
funds
Financial insurance
Social networks and
social capital
• Alternative forms of
risk transfer
Mutual and reserve
funds
Financial insurance
Social networks and
social capital
• Alternative forms of
risk transfer
Early warning and
communication
Evacuation plan
Humanitarian: relief
supplies
• Post-disaster livelihood
support and recovery
Early warning and
communication
Evacuation plan
Humanitarian: relief
supplies
• Post-disaster livelihood
support and recovery
Flexibility in
decisionmaking
Adaptive learning and
management
Improved knowledge
and skills
• Systems transformation
over time
Flexibility in
decisionmaking
Adaptive learning and
management
Improved knowledge
and skills
• Systems transformation
over time
Mainstream risk
management into
development processes
Building codes and
retrofitting
Defensive infrastructure
and environmental
buffers
Land use planning
Catchment and other
ecosystem management
• Incentive mechanisms
for individual actions to
reduce exposure
Mainstream risk
management into
development processes
Building codes and
retrofitting
Defensive infrastructure
and environmental
buffers
Land use planning
Catchment and other
ecosystem management
• Incentive mechanisms
for individual actions to
reduce exposure
Poverty reduction
• Health improvements
Access to services and
productive assets
enhanced
• Livelihood diversification
• Access to
decisionmaking
increased
• Community security
improved
Reduce Hazards
and Exposure
Pool, Transfer, and
Share Risks
Prepare and
Respond Effectively
Increase Capacity to
Cope with “Surprises”
Reduce
Vulnerability
Figure 6-3 |
Complementary response measures for observed and projected disaster risks supported by respective institutional and individual capacity for making informed decisions.
362
in line with the described ‘no’ and ‘low regrets’ strategies discussed in
Section 6.3.1.
6.5. Practices including Methods and Tools
With some success and with many challenges, countries are increasingly
adopting a diverse range of approaches, methods, and tools to manage
disaster risk and adapt to a changing climate, with the intention of
building a safe, secure society. This section discusses efforts made in
building a culture of safety (Section 6.5.1), which includes methods
associated with assessing and communicating risk; reducing climate-
related disaster risks (Section 6.5.2); transferring and sharing residual
risks (Section 6.5.3); and managing the impacts of disasters holistically
(Section 6.5.4), as disaster risks can never be reduced to zero.
Accordingly, it is important to recognize that the approaches, methods,
and tools discussed here are complementary, often overlapping, and
can be pursued simultaneously. Whereas the Summary for Policymakers
includes a visual representation of the range of such approaches (see
Figure SPM-2), Figure 6-3 on the previous page is tailored to incremental
action at the national level.
Figure 6-3 characterizes the range of risk management and adaptation
options open to stakeholders involved in national systems for managing
disaster risk. Such options exist along a continuum of action, with
choices between different options being dependent on the quality
of information and how it is communicated, the findings of risk
assessments, the culture of risk management/acceptability of risk, and
on capacities and resources. In practice, different options will likely be
pursued simultaneously and will have a high degree of co-dependence.
6.5.1. Building a Culture of Safety
Building a culture of safety involves several strategies and activities that
start with the assessment of risk factors and development of information
systems that provide relevant information for critical decisionmaking. It
also involves understanding the large variety of beliefs and core value
systems, which will help determine decisions made by different actors
and stakeholders. A key ingredient is appropriate public education and
awareness raising, and as such, early warning systems play an important
role in managing residual risk as they can provide timely warnings to
exposed communities and thus can promote action for a quick
response. Time series empirical data used to generate risk assessments,
including those contributing to early warnings, are also critical for long-
term planning because of their relevance in generating appropriate
information about adequate land use planning, for example, to reduce
climatic risks. As examples, in the same sense, analyzed information
about climate-adapted infrastructure, enhanced human development,
ecosystems protection, risks transfer, and sharing and managing the
impacts of climate-related disasters can play a fundamental role in
building a culture and practice of human safety.
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
Box 6-3 | Mexico’s Fund for Natural Disasters, FONDEN
Mexico is exposed to natural hazards due to its location within one of the world’s most active seismic regions and in the path of
hurricanes and tropical storms originating in the Caribbean Sea and Atlantic and Pacific Oceans. There have been many disaster events
and in the past severe hurricane disasters (in addition to earthquakes) have created large fiscal liabilities and imbalances (Cardenas et
al., 2007). Given high perceived financial vulnerability to disasters, in 1994, the Mexican Government passed a law that required federal,
state, and municipal public assets to be insured. This was intended to relieve the central government of the obligation of having to pay
for the reconstruction of public infrastructure, although the adequate level of insurance, particularly for very large events remained a
concern. As a next step, in 1996, the national government established a system of allocating resources for disaster spending (FONDEN)
in order to enhance the country’s financial preparedness for disaster losses and prevent imbalances in the federal government finances
derived from outlays caused by catastrophes. FONDEN serves as last-resort funding for uninsurable losses, such as emergency response
and disaster relief expenditures. In addition to this budgetary program, in 1999, a reserve fund was created that accumulates the surplus
of the previous year’s FONDEN budget item (Cardenas et al., 2007).
After the initial phase that was characterized by spending in line with requirements caused by disaster events, one concern for the
disaster management authorities became the regular demands on the funds in non-disaster years. As a consequence, budgeted FONDEN
resources were declining while demands on FONDEN’s resources were becoming more volatile, and outlays often exceeded budgeted
funds causing the reserve fund to decline. In 2005, after the severe hurricane season affecting large parts of coastal Mexico, the fund
was finally exhausted. This forced the Mexican Government to look at alternative risk financing strategies, which included hedging
against disaster shocks, government agencies at all levels providing their insurance protection independent of FONDEN, and that
FONDEN itself should only indemnify losses that exceed the financial capacity of the federal, local, or municipal government agencies. In
2006, Mexico became the first transition country to transfer part of its public sector catastrophe risk to the international reinsurance and
capital markets, and, in 2009, the transaction was renewed for another three years, covering both hurricane and earthquake risk
(Cardenas et al., 2007).
363
6.5.1.1. Assessing Risks and Maintaining Information Systems
As discussed widely in Chapter 1, the first key step in managing risk is
to assess and characterize it. In terms of risk factors, disaster risk is
commonly defined by three elements: the hazard, exposure of elements,
and vulnerability (Swiss Re, 2000; Kuzak, 2004; Grossi and Kunreuther,
2005; CACCA, 2010). Thus, understanding risk involves observing and
recording hazards and hazard analysis, studying exposure and drivers of
vulnerability, and vulnerability assessment. Responding to risks is
dependent on the way risk-based information is framed in the context
of public perception and risk management needs.
Given the ‘public good’ nature of much of disaster-related information
(Benson and Clay, 2004), governments have a fundamental role in
providing good-quality and context-specific risk information about, for
example, the geographical distribution of people, assets, hazards, risks, and
disaster impacts and vulnerability to support disaster risk management
(McBean, 2008). Good baseline information and robust time series
information are key for long-term risk monitoring and assessments, not
only for hazards, but also for evaluating the evolution of vulnerability and
exposure (McEntire and Myers, 2004; Aldunce and León, 2007). Regular
updating of information about hazards, exposure, and vulnerability are
also necessary because of the dynamic nature of disaster risk, especially
due to the effects of climate change and the associated uncertainty this
creates (UNISDR, 2004; Prabhakar et al., 2009; CACCA, 2010).
A key component in the risk assessment process is to determine exposed
elements at risk. This may relate to persons, buildings, infrastructure
(e.g., water and sewer facilities, roads, and bridges), agricultural assets,
livelihoods, ecosystems, natural infrastructure, and ecosystem services
in harm’s way that can be impacted in case of a disaster event. For
national level assessments, their aggregate values are of interest.
Ideally, this would be based on national asset inventories, national
population census, and other national information. In practice, collecting
an inventory on assets and their values often proves very difficult and
expensive due to the heterogeneity and sheer number of the examined
elements (see Cummins and Mahul, 2009). In addition, risk management
processes require identifying those elements of the social process that
also contribute to vulnerability – such as organizational and economic
capacities, human development status of communities at risk, and
capacity to respond to disasters (Lavell, 1996; Cardona et al., 2010) – as
well as assessing the impacts following disaster events (ECLAC, 2003;
Benson and Clay, 2004). Considerable progress has been made in the
generation and use of such information including in some developing
countries (Benson and Clay, 2004; UNISDR 2009c). Nevertheless, in
many countries this is not a regular practice and efforts to document
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
Activities Examples of information needs
Cross -cutting
Climate change modeling Time series information on climate variables – air and sea surface temperatures, rainfall and precipitation measures, wind, air
circulation patterns, and greenhouse gas levels
Hazard zoning and ‘hot spot’ mapping Georeferenced inventories of landslide, flood, drought, and cyclone occurrence and impacts at local, sub-national and
national levels
Human development indicators Geospatial distribution of poverty, livelihood sources, access to water and sanitation
Disbursement of relief payments Household surveys of resource access, social well-being, and income levels
Seasonal outlooks for preparedness
planning
Seasonal climate forecasts; sea surface temperatures; remotely sensed and in situ measurements of snow cover/depth, soil
moisture, and vegetation growth; rainfall-runoff; crop yields; epidemiology
A system of risk indicators reflecting
macro and financial health of nation,
social and environmental risks, human
vulnerability conditions, and strength
of governance (Cardona et al., 2010)
Macroeconomic and financial indicators (Disaster Deficit Index)
Measures of social and environmental risks
Measures of vulnerability conditions reflected by exposure in disaster-prone areas, socioeconomic fragility, and lack of social
resilience in general
Measures of organizational, development, and institutional strengths
Flood risk
management
Early warning systems for fluvial,
glacial, and tidal hazards
Real-time meteorology and water-level telemetry; rainfall, stream flow, and storm surge; remotely sensed snow, ice, and lake
areas; rainfall-runoff model and time series; probabilistic information on extreme wind velocities and storm surges
Flooding hot spots, and structural and
non-structural flood controls
Rainfall data, rainfall-runoff, stream flow, floods, and flood inundation maps
Inventories of pumps, stream gauges, drainage and defense works; land use maps for hazard zoning; post-disaster plan;
climate change allowances for structures; floodplain elevations
Artificial draining of proglacial lakes Satellite surveys of lake areas and glacier velocities; inventories of lake properties and infrastructure at risk; local hydro-
meteorology
Drought
management
Traditional rain and groundwater
harvesting, and storage systems
Inventories of system properties including condition, reliable yield, economics, ownership; soil and geological maps of areas
suitable for enhanced groundwater recharge; water quality monitoring; evidence of deep-well impacts
Long-range reservoir inflow forecasts Seasonal climate forecast model; sea surface temperatures; remotely sensed snow cover; in situ snow depths; multi-decadal
rainfall-runoff series
Water demand management and
efficiency measures
Integrated climate and river basin water monitoring; data on existing systems’ water use efficiency; data on current and future
demand metering and survey effectiveness of demand management
Table 6-3 | Information requirements for selected disaster risk management and adaptation to climate change activities. Adapted from Wilby (2009).
364
impacts are started only after major disasters (Prabhakar et al., 2009).
Regular monitoring of vulnerability is also at a nascent stage (Dilley,
2006; Cardona et al., 2010). Table 6-3 on the previous page shows a
sample of the kinds of information required for effective disaster risk
management and adaptation to climate change activities.
Country- and context-specific information on disaster impacts and losses,
including baseline data about observations (different types of losses,
weather data) from past events, is often very limited and of mixed
quality (see Embrechts et al., 1997; Carter et al., 2007). Data records at
best may date back several decades, and thus often would provide
only one reference data point for extreme events, such as a 100-
year event (see Section 3.2.1). Data on losses from extremes can also be
systematically biased due to high media attention (Guha-Sapir and
Below, 2002). At times the data on losses are incomplete, as in the
Pacific small island developing states, because of limited capacity to
systematically collect information at the time of disaster, or because of
inconsistent methodologies and the costs of measures used (Chung,
2009; Lal, 2010).
International disaster impact databases are available, such as the EM-
DAT database of the Centre for the Epidemiology of Disasters (CRED) in
Brussels, Desinventar maintained by a network of scientists involved
in studying disasters in Latin America (Red de Estudios Sociales en
Prevención de Desastres en América Latina – LA RED), as well as
databases of reinsurers such as Munich Re. Comparisons of international
and national disaster loss databases have shown significant variations
in documented losses due to inconsistencies in the definition of key
parameters and estimation methods used. This emphasizes the need to
standardize parameter definitions and estimation methods (Guha-Sapir
and Below, 2002; ECLAC, 2003; Tschoegl et al., 2006). For some countries,
a reasonable quality and quantity of information may exist on the direct
impacts, particularly where the reinsurance industry, consulting firms,
and multi-lateral financial institutions have worked together with the
research communities. Limited information is generally available on
socially relevant effects of disasters, such as the incidence of health
effects after a disaster as well as the impacts on ecosystems, which
have not been well studied (Benson and Twigg 2004). Furthermore, the
assessment of indirect disaster impacts on social or economic systems,
such as on income-generating sectors and national savings, needs greater
attention (ECLAC, 2003; Benson and Clay, 2004). Such information can
often also be very useful in order to assess risks by using statistical
estimation techniques (Embrechts et al., 1997) or catastrophe modeling
approaches (Grossi and Kunreuther, 2005).
6.5.1.2. Preparedness: Risk Awareness, Education,
and Early Warning Systems
National governments create the environment and communication
channels to develop and disseminate different kinds of information on, for
example, the hazards that affect different populations and preparedness
for disaster response. Numerous studies indicate that up-to-date and
robust early warning systems play a critical role in reducing the impacts
of potential disasters and enable populations to protect lives and some
property and infrastructure (White et al., 2004; Aldunce and León, 2007;
McBean, 2008; Rogers and Tsirkunov, 2010), and as illustrated in Case
Study 9.2.11.
Traditionally, early warning systems have been interpreted narrowly as
technological instruments for detecting and forecasting impending hazard
events and for issuing alerts (Rogers and Tsirkunov, 2010). However, this
interpretation does not clarify whether warning information is received
by or helpful to the population it serves or is actually used to reduce
risks (Basher, 2006; Rogers and Tsirkunov, 2010). As noted in Case Study
9.2.11, the HFA 2005-2015 stated a need for more than just accurate
predictions, stressing that early warning systems should be “people
centered” and that warnings need to be “timely and understandable to
those at risk” and include “guidance on how to act upon warnings”
(UNISDR, 2005).
Governments also maintain early warning systems to warn their citizens
and themselves about impending creeping climate- and weather-related
hazards. For example, ‘early warnings’ of potentially poor seasons have
been successful at informing key actions for agricultural planning on
longer time scales and for producing proactive responses (Meinke et al.,
2006; Vogel and O’Brien, 2006). Case Study 9.2.11 provides examples of
early warning systems for short-response hazards as well as for creeping
hazards operating on time scales from weeks to seasonal. This case study
also highlights the possibility of using weather and climate predictions
for timeframes longer than a few days to provide advanced warning of
extreme conditions, which has been only a very recent development.
Studies indicate that successful early warning systems are reliant on
close inter-institutional collaboration between national meteorological
and hydrological services and the agencies that directly intervene in
rural areas, such as extension services, development projects, and civil
society organizations (Meinke et al., 2006; Vogel and O’Brien, 2006;
Rogers and Tsirkunov, 2010).
An effective early warning system delivers accurate, timely, and meaningful
information, with its success dependent on whether the warnings
trigger effective responses (UNISDR, 2005; Basher, 2006; Gwimbi,
2007; Auld, 2008a; van Aalst, 2009; Rogers and Tsirkunov, 2010).
Warnings fail in both developing and developed countries for a number
of reasons, including inaccurate weather and climate forecasting,
public ignorance of prevailing conditions of vulnerability, failure to
communicate the threat clearly or in time, lack of local organization,
and failure of the recipients to understand or believe in the warning or
to take suitable action (UNISDR, 2006; Auld, 2008a; Rogers and
Tsirkunov, 2010). To be effective and complete, an early warning system
typically is composed of four interacting elements (Basher, 2006;
UNISDR, 2006): (1) generation of risk knowledge including monitoring
and forecasting; (2) surveillance and warning services; (3) dissemination
and communication; and (4) response capability. Warnings are received
and understood by the target audience and are most relevant when the
communications have meaning that is shared between those who issue
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
365
the forecasts, local knowledge, and the decisionmakers they are intended
to inform (Basher, 2006; UNISDR, 2006; Auld, 2008a; Case Study 9.2.2).
Because emergency responders, the media, and the public often are
unable to translate the scientific information on forecast hazards in
warnings into risk levels and responses, early warning systems are most
effective when their users can identify and interpret the general warning
messages into simple and relevant local impacts and actions (e.g., flash
flood warning and the need to evacuate areas at risk), prioritize the most
dangerous hazards, assess potential contributions from cumulative and
sequential events to risks, and identify thresholds linked to escalating
risks for infrastructure, communities, and disaster response (UNISDR,
2006; Auld, 2008a).
Different hazards and different sectors often require unique preparedness,
warnings, and response strategies (Basher, 2006; UNISDR, 2006). For
example, the needs and responses behind a warning of a drought, a
tornado, a cyclone, or a fire are very different. Some hazards may
represent singular extreme events, sequences, or compound combinations
of hazards while other hazards can be described as ‘creeping’ or
accumulations of events (or non-events). For example, the World
Meteorological Organization (WMO), national meteorological and
hydrological services, the World Health Organization (WHO), the Food and
Agriculture Organization (FAO), and others recognize that combinations
of weather and climate hazards can result in complex emergency
response situations and are working to establish multi-hazard early
warning systems for complex risks such as heat waves and vector-borne
diseases (UNISDR, 2006; WMO, 2007) and early warnings of pests and
food safety threats and disease outbreaks (e.g., prediction of a potential
desert locust crisis) (WMO, 2004, 2007; FAO, 2010). Other ‘creeping’
hazards can evolve over a period of days to months; floods and
droughts, for example, can result from cumulative or sequential multi-
hazard events, especially when accompanied by an already existing
vulnerability, while other hazards such as accumulated precipitation can
lead to critical infrastructure failure (Basher, 2006; Auld, 2008a; Rogers
and Tsirkunov, 2010). Section 3.1.3 provides more detail on compound,
multiple, and creeping hazards.
Studies indicate that an understanding by the public and community
organizations of their risks and vulnerabilities are critical but insufficient
for risk management and that early warning systems need to be
complemented by preparedness programs as well as public education
and awareness programs (Basher, 2006; UNISDR, 2006; Gwimbi, 2007;
Rogers and Tsirkunov, 2010). This requires systematic linkages and
integration between early warning systems and contingency planning
processes (Pelham et al., 2011). For example, a significant long-term
social protection program known as the Productive Safety Net
Programme (PSNP) was implemented in Ethiopia in 2007 in response to
experiences from a series of drought-related disaster responses during
the late 1990s and early 2000s (Pierro and Desai, 2008; Conway and
Schipper, 2011). The aim of the PSNP was to shift institutional approaches
away from just emergency responses and into more sustainable livelihood
approaches involving asset protection and food security. Under this
program, millions of people in ‘chronically’ food-insecure households in
rural Ethiopia received resources from the PSNP through cash transfers
or food payments for their participation in labor-intensive public works
projects with a particular focus on environmental rehabilitation
(Conway and Schipper, 2011). The case study on drought (Case Study
9.2.3) also emphasizes the importance of proactive steps in the form of
drought preparedness and mitigation, and improved monitoring and
early warning systems.
Some studies indicate that public awareness and support for disaster
prevention and preparedness are often high immediately after a major
disaster event and that such moments can be capitalized on to strengthen
and secure the sustainability of, for example, early warning systems
(Basher, 2006; Rossetto, 2007). It should be noted that such windows
require the pre-existence of a social basis for cooperation that, in turn,
supports a collaborative framework between research and management
(Rossetto, 2007; Tompkins et al., 2008; Pelham et al., 2011).
The timing and form of climatic information (including forecasts and
projections), and access to trusted guidance to help interpret and
implement the information and projections in decisionmaking processes,
may be more important to individual users than improved reliability and
forecast skill (Pulwarty and Redmond, 1997; Rayner et al., 2005;
Gwimbi, 2007; Rogers and Tsirkunov, 2010). Decisionmakers typically
manage risks holistically, while scientific information is generally
derived using reductionist approaches (Meinke et al., 2006). The net
outcome can be a ‘disconnect’ between scientists and decisionmakers
with the result that climate and hydro-meteorological information can
be developed that, although scientifically sound, may lack relevance to
the decisionmaker (Cash and Buizer, 2005; Meinke et al., 2006; Vogel and
O’Brien, 2006; Averyt, 2010). Perceptions of irrelevance, inconsistency,
confusion, or doubt can delay action (NRC, 2009). Some studies (Lowe,
2003; Glantz, 2005; Meinke et al., 2006; Feldman and Ingram, 2009)
advise scientists and practitioners to work together to produce trustworthy
knowledge that combines scientific excellence with social relevance, a
point also emphasized in Case Study 9.2.2 on fire. These studies suggest
that decision support activities should be driven by users’ needs, not by
scientific research priorities, and that these user needs are not always
known in advance, but should be identified collaboratively and iteratively
in ongoing two-way communication between knowledge producers and
decisionmakers (Cash and Buizer, 2005; NRC, 2009). It has been
suggested that this ongoing interaction, two-way communication, and
collaboration allows scientists and decisionmakers to get to know each
other, to develop an understanding of what decisionmakers need to
know and what science can provide, to build trust, and, over time,
develop highly productive relationships as the basis for effective
decision support (Feldman and Ingram, 2009; NRC, 2009; Averyt, 2010).
Since early warning information systems are multi-jurisdictional and
multidisciplinary, they usually require anticipatory coordination across a
spectrum of technical and non-technical actors. National governments
can play an important role in setting the high-level policies and supporting
frameworks involving multiple organizations, in adopting multi-hazard
and multi-stakeholder approaches, and in promoting community-based
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
366
early warning systems (Pulwarty et al., 2004; Basher, 2006, UNISDR,
2010). National governments can also interact with regional and
international governments and agencies to strengthen early warning
capacities and to ensure that warnings and related responses are directed
toward the most vulnerable populations (Basher, 2006; UNISDR, 2010).
At the same time, national governments can also play an important role
in supporting regions and sub-national governments in developing
operational and local response capabilities (Basher, 2006; UNISDR,
2010; see Section 6.5.4). In Japan and the Mekong region, for example,
in addition to using an early warning system based on extensive flood
modeling exercise, the emergency basin-level management relies on the
flood mitigation capacity of paddy fields (Masumoto et al., 2006, 2008).
6.5.2. Reducing Climate-Related Disaster Risk
National climate-related disaster risk reduction activities include a
broad range of options that vary from safe infrastructure and building
codes to those aimed to enhance and protect natural ecosystems,
support human development and even ‘build back better’ following a
disaster. Each of these strategies can prove minimally effective in
isolation but highly effective in combination. These and other different
options, along with their limitations (e.g. lack of information and
understanding, human resource capacity, scientific requirements,
financing) are addressed in the following subsections, noting how risk
reduction measures are increasingly being considered as good practices
to promote adaptation to climate change.
6.5.2.1. Applying Technological
and Infrastructure-Based Approaches
Climate change has the potential to directly and indirectly impact the
safety of existing infrastructure and to alter engineering and maintenance
practices, and will require changes in building codes and standards
where they exist (Bourrelier et al., 2000; Füssel, 2007; Wilby, 2007; Auld,
2008b; Stevens, 2008; Hallegatte, 2009). The changing climate also has
the potential regionally to increase premature deterioration and
weathering impacts on the built environment, exacerbating vulnerabilities
to climate extremes and disasters and negatively impacting the expected
and useful life spans of structures (Auld, 2008b; Larsen et al., 2008;
Stewart et al., 2011). As noted in Case Study 9.2.8, people living with
un-adapted and inadequate infrastructure and housing will be more at
risk from climate change.
With projected increases in the magnitude and/or frequency of some
extreme events in many regions (see Chapter 3), small increases in climate
extremes above thresholds or regional infrastructure ‘tipping points’
have the potential to result in large increases in damages to all forms of
existing infrastructure nationally and to increase disaster risks (Coleman,
2002; Munich Re, 2005; Auld, 2008b; Larsen et al., 2008; Kwadijk et al.,
2010; Mastrandrea et al., 2010). Since infrastructure systems, such as
buildings, water supply, flood control, and transportation networks
often function as a whole or not at all, an extreme event that exceeds
an infrastructure design or ‘tipping point’ can sometimes result in
widespread failure and a potential disaster (Ruth and Coelho, 2007;
Haasnoot et al., 2009). For example, a break in a water main, dike, or
bridge can impact other systems and sectors and render the regional
system incapable of providing needed services (Ruth and Coelho, 2007).
These infrastructure thresholds or adaptation ‘tipping points’ become
important when considering sensitivities to climate change and
adaptation and disaster risk reduction options for the future (see
Section 6.6.1 for further discussion on thresholds and management of
climate change uncertainties). Infrastructure thresholds refer here to the
critical climate conditions where acceptable technical, economic, spatial,
or societal limits are exceeded and the current built environment system
is no longer “future climate proof (i.e., it fails, requiring proactive
adaptation actions and changes in infrastructure codes, standards, and
management processes) (Auld, 2008b; Haasnoot et al., 2009; Kwadijk et
al., 2010; Mastrandrea et al., 2010).
The need to address the risk of climate extremes and disasters in the
built environment and urban areas, particularly for low- and middle-
income countries, is one that is not always fully appreciated by many
national governments and development and disaster specialists
(Rossetto, 2007; Moser and Satterthwaite, 2008). Low- and middle-
income countries, which account for close to three-quarters of the
world’s urban populations, are at greatest risk from extreme events and
also have far less capacity than do high-income countries, largely due to
backlogs in protective infrastructure and services and limitations in urban
government (Satterthwaite et al., 2007; Moser and Satterthwaite, 2008).
Rapid growth and expansion in urban areas, particularly in developing
countries, can outpace infrastructure development and lead to a lack of
infrastructure services for housing, sewer systems, effective transportation,
and emergency response and increased vulnerability to weather and
climate extremes (Satterthwaite et al., 2007; Birkmann et al., 2011).
These impacts from the changing climate will be particularly severe for
populations living in poor-quality housing on illegally occupied land,
where there is little incentive for investments in more resilient buildings
or infrastructure and service provision (Freeman and Warner, 2001;
Satterthwaite et al., 2007; Birkmann et al., 2011). Case Study 9.2.8
provides further discussion on the best adaptation and risk management
practices for cities and their built environment.
An inevitable result of potentially increased damages to infrastructure
will be a dramatic increase in the national resources needed to restore
infrastructure and assist the poor affected by damaged infrastructure
(Freeman and Warner, 2001). A study by the Australian Academy of
Technological Sciences and Engineering concluded that national retrofit
measures will be needed to safeguard existing infrastructure in Australia
and new adaptation approaches and national codes and standards will
be required for construction of new infrastructure (Stevens, 2008).
Recommendations reported from this study call for research to fill gaps
on the future climate risks, comprehensive risk assessments for existing
critical climate-sensitive infrastructure, development of information and
supporting tools (e.g., non-stationary extreme value analysis methods)
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
367
about future climate change events, investigation of the links between
soft and hard engineering solutions, and strengthened research efforts
to improve the modeling of small-scale climate events (Wilby, 2007;
Auld, 2008b; Stevens, 2008).
The recommended national adaptation options to deal with projected
impacts to the built environment range from deferral of actions pending
development of new climate change information to modification of
infrastructure components according to national guidance, acceptance of
residual losses, reliance on insurance and other risk transfer instruments,
formalized asset management and maintenance, mainstreaming into
environmental assessments, new structural materials and practices,
improved emergency services, and retrofitting and replacement of
infrastructure elements (Bourrelier et al., 2000; Auld, 2008b; Stevens,
2008; Haasnoot et al., 2009; Hallegatte, 2009; Neumann, 2009; Kwadijk
et al., 2010; Wilby and Dessai, 2010).
Strategic environmental assessment approaches, such as those
recommended by the Organisation for Economic Cooperation and
Development (OECD) and many national environmental assessment
agencies, offer an effective means for ensuring that adaptation to climate
change and disaster risk management, as well as GHG reduction practices,
are mainstreamed into policies and planning for new programs on
infrastructure and systems (OECD, 2006; Benson, 2007). Environmental
impact assessment approaches can reduce the risks of environmental
degradation from a project and reduce future disaster risks from current
and changing climate conditions (Benson, 2007). For long-lived
infrastructure or networks, studies recommend consideration of likely
climate change impacts that will potentially affect the planned useful
life of the infrastructure system (e.g., seasonal variability in water flows,
temperatures, incidence of extreme weather events) (OECD, 2006;
Bosher et al., 2007; Auld, 2008b; Larsen et al., 2008; Neumann, 2009;
NRTEE, 2009).
The implementation of adequate national building codes that incorporate
up-to-date regionally specific climate data and analyses can improve
resilience of infrastructure for many types of weather-related risks
(Auld, 2008b; WWC, 2009; Wilby et al., 2009). Typically, infrastructure
codes and standards in most countries use historical climate analyses to
climate-proof new structures, assuming that the past climate can be
extrapolated to represent the future. For example, water-related
engineering structures, including both disaster-proofed infrastructure
and services infrastructure (e.g., water supply, irrigation and drainage,
sewerage, and transportation), are typically designed using analysis of
historical rainfall records (Ruth and Coelho, 2007; Auld, 2008b;
Haasnoot et al., 2009; Hallegatte, 2009; Wilby and Dessai, 2010). Since
infrastructure is built for long life spans and the assumption of climate
stationarity will not hold for future climates, it is important that national
climate change guidance, tools, and consistent adaptation options be
developed to ensure that climate change can be incorporated into
infrastructure design (Auld, 2008b; Stevens, 2008; Hallegatte, 2009;
Wilby et al., 2009). While some government departments responsible for
building regulations and the insurance industry are taking the reality of
climate change very seriously, challenges remain about how to
incorporate the uncertainty of future climate projections into engineering
risk management and into codes and standards, especially for climate
elements such as extreme winds and extreme precipitation and their
various phases (e.g., short- and long-duration rainfalls, freezing rain,
snowpacks) (Sanders and Phillipson, 2003; Auld, 2008b; Haasnoot et al.,
2009; Hallegatte, 2009; Kwadijk et al., 2010; Wilby and Dessai, 2010; Lu,
2011). Recent advances in characterizing the uncertainties of climate
change projections, in regionalization of climate model outputs, and in
the application and mainstreaming of integrated top-down, bottom-up
approaches for assessing impacts and adaptation options (Sections
6.3.1 and 6.3.2) will help to ensure that infrastructure and technology
can be better adapted to a changing climate. Sections 3.2.3, 3.3, and 3.4
provide further details on scientific advances for the construction,
assessment, and communication of climate change projections, including
a discussion on recent advances in the development of regionalization
or downscaling techniques and approaches used to quantify uncertainties
in climate change model outputs.
Some implementation successes are emerging. In one example, discussed
in Case Study 9.2.10, the Canadian Standards Association (CSA) and its
National Permafrost Working Group developed a Technical Guide, CSA
Plus 4011-10, on Infrastructure in Permafrost: A Guideline for Climate
Change Adaptation, that directly incorporated climate change temperature
projections from an ensemble of climate change models. This CSA Guide
considered climate change projections of temperature and precipitation
and incorporated risks from warming and thawing permafrost to
foundations over the planned life spans of the structure (Hayley and
Horne, 2008; NRTEE, 2009; CSA, 2010a; Smith et al., 2010; Grosse et al.,
2011). The guide suggested possible adaptation options, taking into
account the varying levels of risks and the consequences of failure for
foundations of structures, whether buildings, water treatment plants,
towers, tank farms, tailings ponds, or other infrastructure (NRTEE, 2009;
CSA, 2010a; see Case Study 9.2.10). Similarly, working with the Canadian
meteorological service, engineering associations, and national water
stakeholder associations, the CSA has also developed an initial rainfall
Intensity-Duration-Frequency Guideline for water practitioners with
adaptation guidance (CSA, 2010b).
In developing countries, structures are often built using prevalent local
practices, which may not reflect best practices from disaster risk reduction
or adaptation perspectives. These prevalent local practices usually do
not include the use of national building standards or adequately
account for local climate conditions (Rossetto, 2007).
While the perception
in some developing countries is that national building codes and standards
are too expensive, experience in the implementation of incremental
hazard-proof measures in building structures has proven in some countries
to be relatively inexpensive and highly beneficial in reducing losses
(Rossetto, 2007; ProVention, 2009). In reality, the most expensive
components of codes and standards are usually the cost to implement
national policies for inspections, knowledge transfer to trades, and
national efforts for their uptake and implementation (Rossetto, 2007).
Bangladesh, for example, has implemented simple modifications to
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
368
improve the cyclone resistance of (non-masonry) kutcha or temporary
houses, with costs that amounted to only 5% of the construction costs
(Lewis and Chisholm, 1996; Rossetto, 2007). Bangladesh is also
developing national policies requiring that houses built following
disasters include a small section of the replacement house that meets
‘climate proofing’ standards and acts as a household shelter in the next
disaster. In many countries, climate-proofing guidelines and standards
are applied to structures that are used as emergency shelters and for
structures that form the economic and social lifeline of a society, such
as its communications links, hospitals, and transportation networks
(Rossetto, 2007).
Many studies advocate that technical and infrastructure solutions are
not the only way of adapting to changing climates and that ‘soft
solutions’ such as financial tools, land use planning, and ecosystem
conservation or soft engineering approaches are also needed (Adger et
al., 2007; Auld, 2008b; Nicholls et al., 2008; Hallegatte, 2009; McEvoy et
al., 2010). Land and water use planning, use of bioshields as natural
buffers, soft defenses, and green or ‘soft engineering’ are complementary
adaptation options, described further in Section 6.5.2.3 and in Case
Studies 9.2.1 and 9.2.8.
6.5.2.2. Human Development and Vulnerability Reduction
Vulnerabilities to climate-related hazards and the options to reduce
them vary between and within countries due to factors such as poverty,
social positioning, geographic location, gender, age, class, ethnicity,
ecosystem condition, community structure, community decisionmaking
processes, and political issues (Yodmani, 2001; Yamin et al., 2005;
Halsnaes and Traerup, 2009). Overall, studies indicate that the extent of
the vulnerability to climate variability and climate change is shaped by
both the dependence of the national economy and livelihoods on climate-
sensitive natural resources and the resilience or robustness of the
country’s social institutions to equitable distribution of resources under
climate change (Ikeme, 2003; Brooks et al., 2009; Virtanin et al., 2011).
The poorest regions are often characterized by vulnerable housing, weak
emergency services and infrastructure, and a dependence on agriculture
and other natural resources (Ikeme, 2003; Manuel-Navarrete et al.,
2007; Reid et al., 2010).
Many vulnerable communities already suffer greater water stress, food
insecurity, disease risks, and loss of livelihoods, which have the potential
to increase under climate change (Manuel-Navarrete et al., 2007; Brooks
et al., 2009; Halsnaes and Traerup, 2009; Virtanin et al., 2011). For
example, climate change may increase the risk of waterborne diseases,
requiring targeted assistance for health and water sanitation issues
(Curriero et al., 2001; Brooks et al., 2009). Small island states and low-
lying countries may require support to relocate vulnerable groups to
safer locations or other countries, all requiring a complex set of actions
at the national and international levels (Manuel-Navarrete et al., 2007;
McGranahan et al., 2007). Other studies indicate that resilient housing and
safe shelters will remain a key adaptation action to protect vulnerable
people from disasters and climate extremes, requiring national guidelines
to ensure that new or replacement structures are built with flexibility to
accommodate future changes (Ikeme, 2003; Manuel-Navarrete et al.,
2007; Rossetto, 2007; Auld, 2008b). Under climate change, it is expected
that food security issues among vulnerable populations will become
more impacted by climate variability, erratic rainfall, and more frequent
extreme events (Ikeme, 2003; IRI, 2006; Brooks et al., 2009; Halsnaes
and Traerup, 2009; and regional studies through global partnerships,
such as the Consultative Group on International Agricultural Research).
When faced with food scarcity, vulnerable populations sometimes adopt
maladaptive coping strategies such as overgrazing, deforestation, and
unsustainable extraction of water resources that aggravate long-term
disaster risks (Brooks et al., 2009; Bunce et al., 2010).
Studies indicate that the greatest losses in suitable agricultural cropland
due to climate change are likely to be in Africa, particularly sub-Saharan
Africa (Ikeme, 2003; FAO, 2010). Assessing food security issues in this
vulnerable area requires consideration of multiple socioeconomic and
environmental variables, including climate (Verdin et al., 2005; Virtanin
et al., 2011). In sub-Saharan Africa, where large and widely dispersed
populations depend on rain-fed agriculture and pastoralism, climate
monitoring and forecasting are important inputs to food security analysis
and assessments. Since conventional climate and hydro-meteorological
networks in these areas are sparse and often report with significant delays,
there is a growing need for increased capacity in rainfall observations,
forecasting, data management, and modeling applications (Verdin et
al., 2005; Heltberg et al., 2009; FAO, 2010). Studies indicate a need for
rainfall observation networks to be expanded and to incorporate satellite
information; for data management systems to be improved; for tailored
forecast information to be disseminated and used by decisionmakers;
and for more effective early warning systems that can integrate seasonal
forecasts with drought projections as inputs for hazards, food security,
and vulnerability analysis (Verdin et al., 2005; Heltberg et al., 2009; FAO,
2010). Other short-term but limited strategies to minimize food security
risks include diversifying livelihoods to spread risk, farming in different
ecological niches, building social networks, productive safety net and
social protection schemes, and risk pooling at the regional or national
level to reduce financial exposure (Brooks et al., 2009; Halsnaes and
Traerup, 2009; Heltberg et al., 2009; FAO, 2010). Specific longer-term
strategies to address the increasing risks, particularly given uncertainties,
include land rehabilitation, terracing and reforestation, measures to
enhance water catchment and irrigation techniques, improvements to
infrastructure quality for better access to markets, and the introduction
of drought-resistant crop varieties (Halsnaes and Traerup, 2009;
Heltberg et al., 2009).
In the longer term, studies indicate that increasing food security risks
under climate change will require higher agricultural productivity, reduced
production variability, and agricultural systems that are more resilient
to disruptive events (Cline, 2007; Stern, 2007; Halsnaes and Traerup,
2009; FAO, 2010). This implies transformations in the management of
natural resources; new climate-smart agriculture policies, practices, and
tools; better use of climate science information in assessing risks and
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
369
vulnerability; and financing for food security (Brooks et al., 2009;
Ericksen et al., 2009; FAO, 2010). Other coping strategies may include
increased non-farm incomes, migration, government and other financial
assistance, microfinance, social protection, other safety nets, and
various insurance products (Barrett et al., 2007; Heltberg et al., 2009;
FAO, 2010). The Sustainable Livelihoods Approach or Framework has been
used internationally for rural and coastal development to holistically
describe the variables that impact livelihoods locally and to define the
capacity, assets (both natural and social), and policies required for
sustainable living, poverty reduction, and recovery from disasters
(Brocklesby and Fisher, 2003; Yamin et al., 2005). Sections 2.3, 2.4.3,
2.6.1, 5.2.3, and 5.4, and Case Studies 19.2.1 and 19.2.2, also discuss
sustainable livelihood approaches that can be considered in
building adaptive capacity and resilience to climate hazards and climate
change.
Early identification of populations at risk can enable timely and
appropriate actions needed to avert widespread impacts. Reliable and
detailed information on the current and future climates and their
impacts can play an important role in the recognition of the need to
adapt and the successful evolution of effective adaptation strategies
(Ikeme, 2003; Verdin et al., 2005; Heltberg et al., 2009; Wilby et al., 2009;
and as discussed in Section 6.5.1). Some studies claim that one of the
potential barriers for identifying the most vulnerable regions and people
in developing countries under future climate change is the limited
human resource capacity regionally to downscale global and regional
climate projections to a scale suitable to support national-level planning
and programming processes (Ikeme, 2003; Verdin et al., 2005; CCCD,
2009; Wilby et al., 2009). Not all of the climate variables of importance
for development can be projected and downscaled with confidence,
particularly given that many development activities are especially
sensitive to changes in climate extremes (Agrawala and van Aalst,
2008). Even when downscaled results are available, their use can be
limited by a lack of understanding and interpretation of how these
downscaled projections can be translated to highlight vulnerabilities
with certainty (Agrawala and van Aalst, 2008; Heltberg et al., 2009).
Agrawala and van Aalst (2008) argue that development practitioners
and climate scientists should join forces to make climate information
more accessible, relevant, and usable.
Because the risks posed by climate change can affect the long-term
efficiency with which development resources can be invested and
development objectives achieved, studies indicate that it remains
important to integrate or mainstream disaster risk management and
climate change adaptation into a range of development activities
(Agrawala and van Aalst, 2008; Halsnaes and Traerup, 2009; Heltberg
et al., 2009; Mitchell et al., 2010a). Lack of awareness within the
development community of the many implications of climate change
and limitations on resources for implementation are frequently cited
reasons for difficulties in mainstreaming adaptation and disaster risk
management (Agrawala and van Aalst, 2008; Heltberg et al., 2009; also
see Section 6.3.2). Adaptation to climate change and disaster risk
management actions can be considered to be successfully mainstreamed
when they reduce the vulnerability of susceptible populations to existing
climate variability and are also able to strengthen the capacity of the
population to prepare for and respond to further changes (Yamin et al.,
2005; Manuel-Navarrete et al., 2007; Mertz et al., 2009). Studies indicate
that national policies can increase this capacity (Ikeme, 2003; Heltberg
et al., 2009). Policies and measures such as the establishment of an
LDC fund, Special Climate Fund, Adaptation Fund, climate change Multi-
Donor Trust Fund, etc., have all been developed to address the special
adaptation and risk reduction issues of vulnerable countries (see
Sections 7.4.2 and 7.4.3.3 for more details).
In spite of recommendations to target assistance to the most vulnerable
in the developing world, practical ‘on the ground’ examples have been
limited (Yamin et al., 2005; Ayers and Huq, 2009; Heltberg et al.,
2009). Nonetheless, some developing countries have implemented
successful policies and plans. Nationally, good progress is being made
in strengthening some disaster reduction capacities for disaster
preparedness and early warning and response systems and in addressing
some of the underlying risk drivers in many developing country regions
and sectors (Manuel-Navarrete et al., 2007; UNISDR, 2009c). For example,
social safety nets and other similar national-level programs, particularly
for poverty reduction and attainment of the Millennium Development
Goals, have helped the poorest to reduce their exposure to current and
future climate hazards (Yamin et al., 2005; Tanner and Mitchell 2008;
Heltberg et al., 2009). Some examples of social safety nets are cash
transfers to the most vulnerable, versions of weather-indexed crop
insurance, employment guarantee schemes, and asset transfers (Yamin
et al., 2005; CCCD, 2009; also see Section 6.6.3). A national policy to
help the vulnerable build assets should incorporate climate screening in
order to remain resilient under a changing climate (UNISDR, 2004;
Tanner and Mitchell, 2008; Heltberg et al., 2009). Other measures, such
as social pensions that transfer cash from the national level to vulnerable
people, provide some buffers against climate hazards (Davies et al., 2008;
Heltberg et al., 2009). However, lack of capacity and good governance
has remained a major barrier to efficient and effective delivery of
assistance to the most vulnerable (Yamin et al., 2005; CCCD, 2009;
Heltberg et al., 2009; Warner et al., 2009).
National Adaptation Programme of Actions (NAPA) under the UNFCCC
process have helped least-developed countries assess the climate-
sensitive sectors and prioritize projects to address the most urgent
adaptation issues of the most vulnerable regions, communities, and
populations. The NAPA process has proven instrumental in increasing
awareness of climate change and its potential impacts in the poorest
countries. The proposed adaptation projects under the NAPA usually
cover small areas and address a few components within a given sector
with a view to addressing urgent and immediate needs. The choice of
projects is based on the urgency of the actions as well as cost-effectiveness
in cases where delays would increase the costs of later addressing the
issue. Assessment of completed NAPAs show different national and
regional priority sectors such as health, food security, infrastructure,
coastal zone and marine ecosystem, insurance, early warning and disaster
management, terrestrial ecosystem, education and capacity building,
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
370
tourism, energy, water resources, and cross-sectoral areas. The NAPA
process forms a good basis for developing medium- and long-term
adaptation plans and policies. The capacity within NAPA teams and the
subsequent networks that are created are proving very useful in the
design of broader national adaptation plans (UNFCCC, 2011a,b).
6.5.2.3. Investing in Natural Capital
and Ecosystem-Based Adaptation
Ecosystem-based adaptation, which integrates the use of biodiversity
and ecosystem services into an overall adaptation strategy, can be a cost-
effective strategy for responding to the effects of weather and climate
extremes (SCBD, 2009). It is generally agreed that investment in sustainable
ecosystems and environmental management has the potential to also
provide improved livelihoods and increased biodiversity conservation
(Bouwer, 2006; UNEP, 2006, 2010; McGray et al., 2007; Colls et al.,
2009; SCBD, 2009; Sudmeier-Rieux and Ash, 2009; World Bank, 2009).
Healthy, natural or modified, ecosystems (see Section 6.3.1 and Box 6-4)
have a critical role to play in reducing risks of climate extremes and
disasters (Sidle et al., 1985; Dorren et al., 2004; Phillips and Marden,
2005; Reid and Huq, 2005; UNISDR, 2005, 2007a,b, 2009a,b; Bebi et al.,
2009; Colls et al., 2009; SCBD, 2009; Sudmeier-Rieux and Ash, 2009;
UNEP, 2009; Lal, 2010). Although the scientific evidence base relating
to the role of ecosystem services in reducing the sensitivity of natural
systems to weather and climate extremes and reducing vulnerabilities
to many disasters is nascent, investment in natural ecosystem
management has long been used to reduce risks of disasters (see Box
6-4). Forests, for example, have been used in the Alps and elsewhere as
effective risk-reducing measures against avalanches, rockfalls, and
landslides since the 1900s (Sidle et al., 1985; Dorren et al., 2004; Phillips
and Marden, 2005; Bebi et al., 2009). The damage caused by wildfires,
wind erosion, drought, and desertification are reported to have been
buffered by forest management, shelterbelts, greenbelts, hedges, and
other ‘living fences’ (ProAct Network, 2008; Dudley et al., 2010).
Mangrove replanting has been used as a buffer against cyclones and
storm surges, with reports of a 70 to 90% reduction in energy from
wind-generated waves in coastal areas (UNEP, 2006) and reduction in
the number of deaths from cyclones (Das and Vincent, 2009), depending
on the health and extent of the mangroves. Many sectoral examples
are provided in Table 6-1 that also provide evidence of the value of
ecosystem services in disaster risk reduction and adaption to climate
change (see also Section 6.5.2.1).
The extent to which ecosystems support such benefits, though, depends
on a complex set of dynamic interactions among ecosystem-related
factors, as well as the intensity of the hazard (UNEP, 2006; Sudmeier-
Rieux and Ash, 2009) and institutional and governance arrangements
(see case studies in Angelsen et al., 2009). Scientific understanding of
the relationship between ecosystem structure and function and the
reduction of risks associated with weather and climate extremes is
limited, though growing.
Investment in natural ecosystems also contributes significantly to
reduction in GHG emissions, through practices such as those associated
with Land Use, Land-Use Change, and Forestry (LULUCF) and through
Reduced Carbon Emissions from Deforestation and Forest Degradation
(REDD) or REDD+, which additionally includes the value of conservation
from sustainable management of forests and enhancement of forest
carbon stocks (UNEP, 2006; SCBD, 2009). Mangrove ecosystems, for
example, are important for carbon sequestration and storage,
containing among the highest carbon pools: 1,060-2,020 t CO
2
ha
-1
or an annual carbon sequestration of 6.32 t CO
2
ha
-1
(Murray et al.,
2010), as well as providing the buffers against weather and climate
extremes, biodiversity values, and livelihood benefits discussed above.
Investment in natural ecosystems, through REDD and REDD+ related
strategies, can generate alternative sources of income for local
communities and provide much needed financial incentives to prevent
deforestation (Reid and Huq, 2005; Angelsen et al., 2009; SCBD, 2009;
Sudmeier-Rieux and Ash, 2009; Murray et al., 2010), as well as provide
additional livelihood benefits from the conservation and restoration of
forest ecosystems and the services they support (Longley and Maxwell,
2003; MEA, 2005; SEEDS India, 2008; Sudmeier-Rieux and Ash, 2009;
Murray et al., 2010).
Some countries have begun to explicitly consider ecosystem-based
solutions for climate change mitigation and/or adaptation to risks
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
Box 6-4 | Value of Ecosystem Services in
Disaster Risk Management:
Some Examples
In the Maldives, degradation of protective coral reefs
necessitated the construction of artificial breakwaters at a
cost of US$ 10 million per kilometer (SCBD, 2009).
In Vietnam, the Red Cross began planting mangroves in
1994 with the result that, by 2002, some 12,000 hectares of
mangroves had cost US$1.1 million for planting but saved
annual levee maintenance costs of US$ 7.3 million, shielded
inland areas from a significant typhoon in 2000, and
restored livelihoods in planting and harvesting shellfish
(Reid and Huq, 2005; SCBD, 2009).
In the United States, wetlands are estimated to reduce
flooding associated with hurricanes at a value of US$ 8,250
per hectare per year, and US$ 23.2 billion a year in storm
protection services (Costanza et al., 2008).
In Orissa, India, a comparison of the impact of the 1999
super cyclone on 409 villages in two tahsils with and
without mangroves showed that villages that had healthy
stands of mangroves suffered significantly less loss of lives
than those without (or limited areas) healthy mangroves,
even though all villages had the benefit of early warnings
and accounting for other social and economic variables
(Das and Vincent, 2009).
371
associated with weather and climatic extremes as an integral element
of national and sectoral development decisions (see Box 6-5).
Ecosystem-based adaptation strategies, often considered as part of ‘soft’
options, are a widely applicable approach to climate change adaptation
because they can be applied at regional, national, and local levels, at
both project and programmatic levels, and benefits can be realized over
short and long time scales. They can be a more cost-effective adaptation
strategy than hard infrastructure and engineering solutions, as also
discussed in Section 6.5.2.1, and produce multiple benefits, and are also
considerably more accessible to the rural poor than measures based on
hard infrastructure and engineering solutions (Sudmeier-Rieux and Ash,
2009). Communities are also able to integrate and maintain traditional
and local knowledge and cultural values in their risk reduction efforts
(SCBD, 2009).
In the choice of ecosystem-based adaptation options, decisionmakers
may at times require making judgements about the tradeoffs between
particular climatic risk reduction services and other ecosystem services
also valued by humans. Such decisions benefit from information resulting
from risk assessments, scenario planning, and adaptive management
approaches that recognize and incorporate these potential tradeoffs.
This might be the case, for example when deciding to use wetlands for
coastal protection that requires emphasis on silt accumulation and
stabilization possibly at the expense of wildlife values and recreation
(SCBD, 2009), particularly when achieving a full complement of
biodiversity values is highly complex and long-term in nature (UNEP,
2006).
However, countries would need to overcome many challenges if they are
to be successful in increasing investment in ecosystem-based solutions,
including for example:
Insufficient recognition of the economic and social benefits of
ecosystem services under current risk situations, let alone under
potential changes in climate extremes and disaster risks (Vignola
et al., 2009).
Lack of interdisciplinary science and implementation capacity for
making informed decisions associated with complex and dynamic
systems (Leslie and McLeod, 2007; OECD, 2009).
Ability to estimate economic values of different ecosystem services
supported by nature (TEEB, 2009).
Lack of capacity to undertake careful cost and benefit assessments
of alternative strategies to inform choices at the local level. Such
assessments could provide the total economic value of the full range
of disaster-related ecosystem services, compared with alternative
uses of the forested land such as for agriculture (see, e.g., Balmford
et al., 2002).
Where they exist, data on and monitoring of ecosystem status and
risk are often dispersed across agencies at various scales and are
not always accessible at the sub-national or municipal level where
land use-planning decisions are made (UNISDR, 2009a).
The mismatch in geographic scales and mandates between the
administration and responsibilities for disaster reduction, and that
of ecosystem extent and functioning, such as in water basins
(Leslie and McLeod, 2007; OECD, 2009).
6.5.3. Transferring and Sharing ‘Residual’ Risks
Not all risk can be reduced, and a residual, often sizeable risk will
remain. Mechanisms for sharing and transferring residual risks for
households and businesses have been introduced in Section 5.5.2.2 in
the context of managing local-level impacts and risks. Chapter 5 also
discusses the incentive and disincentive aspects provided by insurance
for risk management and adaptation to climate change at the local level.
This section sets out the role of national-level institutions, especially
governments, in enabling and regulating practices at national scales. It
also discusses the need on the part of some governments to transfer
their own risks.
Markets offer risk-sharing and transfer solutions, most prominently
property and asset insurance for households and businesses, and crop
insurance for farmers. Insurance markets are generally segregated and
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
Box 6-5 | Some Examples of Ecosystem
Based Adaptation Strategies and
Disaster Risk Management
Interventions Taking into Account
the Role of Ecosystem Services
Vietnam has applied strategic environmental assessments to
land use-planning projects and hydropower development for
the Vu Gia-Thu Bon River basin, including climatic disaster
risks (OECD, 2009; SCBD, 2009).
European countries affected by severe flooding, notably the
United Kingdom, The Netherlands, and Germany, have made
policy shifts to ‘make space for water’ by applying more
holistic river basin management plans and integrated
coastal zone management (DEFRA, 2005; Wood and Van
Halsema, 2008; EC, 2009; ONERC 2009).
At the regional level, the Caribbean Development Bank has
integrated weather and climatic disaster risks into its
environmental impact assessments for new development
projects (CDB and CARICOM, 2004; UNISDR, 2009c).
Under the Amazon Protected Areas Program, Brazil has
created a more than 30 million hectare mosaic of
biodiversity-rich forests reserve of state, provincial, private,
and indigenous land, resulting in a potential reduction in
emissions estimated at 1.8 billion tonnes of carbon through
avoided deforestation (World Bank, 2009).
Swiss Development Cooperation’s four-year project in
Muminabad, Tajikistan adopted an integrated approach to
risk through reforestation and integrated watershed
management (SDC, 2008).
372
regulated nationally. Existing national insurance systems commonly
offer a wide variety of choice in providing protection for property and
assets against natural hazards. National insurance systems differentially
include hazards, such as storms, hail, floods, earthquake, and also
landslides or subsidence. Risks may be covered separately or bundled
with a fire policy or covered under an ‘all hazards’ policy. The contracts
differ in the extent of cover offered, as well as indemnity limits, and
whether the policies are compulsory, bundled, or voluntary. Importantly,
they differ institutionally with regard to the involvement of the public
authorities and private insurers and how they allocate liability and
responsibility for disaster losses across individual households, businesses,
and taxpayers (Schwarze and Wagner, 2004; Aakre et al., 2010).
Yet, insurance coverage is limited and globally only about 20% of the
losses from weather-related events have been insured over the period
1980 to 2003 (also see Section 6.2.2). In many instances, insurance
providers even in industrialized countries have been reluctant to offer
region- or nationwide policies covering flood and other hazards
because of the systemic nature of these risks, as well as problems of
moral hazard and adverse selection (Froot, 2001; Aakre et al., 2010). In
some highly exposed countries, such as The Netherlands for flood risk,
insurance is even non-existent and government relief is dispensed in
lieu (Botzen et al., 2009). In many developing countries, there is little in
terms of insurance for disaster risks, yet novel index-based micro-
insurance solutions have been developed and are starting to show
results (Hazell and Hess, 2010; see also Sections 5.6.3 and Case Study
9.2.13 on risk financing). Market mechanisms may work less well in
developing countries, particularly because there is often limited risk
assessment information, limited scope for risk pooling, and little or no
supply of insurance instruments. In such circumstances, governments
may need to create enabling environments by helping to estimate risk,
helping to develop training programs for insurer’s staff, and generally
promoting awareness among the population at risk (Linnerooth-Bayer
et al., 2005; Hoeppe and Gurenko, 2006; Cummins and Mahul, 2009;
Hazell and Hess, 2010).
Employing insurance and other risk-financing instruments for helping to
manage the vagaries of nature may often involve the building of PPPs
in developing and in developed countries in order to tackle market
failure, adverse selection, and the sheer non-availability of such
instruments (see Aakre et al., 2010). Because of such reasons, there is a
role for governments to not only create an enabling environment for
private sector engagement, but also to regulate its activities. In the
development context, Hazell and Hess (2010) distinguish between
protection and promotion models, while acknowledging that in many
instances hybrid combinations may contain elements of both.
Protection relates to governments helping to protect themselves,
individuals, and businesses from destitution and poverty by providing
ex-post financial assistance, which, however, is taken out as an ex-ante
instrument as insurance before disasters. The promotion model relates
to the public sector promoting more stable livelihoods and higher
income opportunities by better helping businesses and households
access risk financing, including micro-financing.
Private insurers are often not willing to fully underwrite the risks and
many countries, including Japan, France, the United States, Norway, and
New Zealand, have therefore instituted public-private national insurance
systems, where participation of the insured is mandatory or voluntary
and single hazards may be insured or comprehensive insurance offered
(Linnerooth-Bayer and Mechler, 2007). Further, specific strategies may
be employed to increase market penetration of risks that are not
easily covered by regular avenues. As one example, in India, pro-poor
regulation stipulates that insurers within their regular business segment
reserve a certain quota of insurance policies for the poor and thus cross-
subsidize fledgling low-income micro-insurance policies (Mechler et al.,
2006a).
As well, governments may insure their liabilities through sovereign
insurance. Liabilities arise as governments own a large portfolio of
public infrastructure and other assets that are exposed to disaster risks.
Moreover, most governments accept their role as provider of post-
disaster emergency relief and assistance to vulnerable and affected
households and businesses. In wealthy countries, government (sovereign)
insurance hardly exists at the national level and, in Sweden, insurance
for public assets is illegal (Linnerooth-Bayer and Amendola, 2000). On
the other hand, states in the United States, Canada, and Australia,
although regulated not to incur budget deficits, often carry cover for
their public assets (Burby, 1991). As discussed earlier (see Section 6.4.3),
this is consistent with the Arrow and Lind Theorem, which suggests that
governments can efficiently spread and share risk over their citizens
without buying sovereign insurance policies.
Yet, realizing the shortcomings of after-the-event approaches for coping
with disaster losses for small, low-income or highly exposed countries
with over-stretched tax bases and highly correlated infrastructure risks
(OAS, 1991; Pollner, 2000; Mechler, 2004; Cardona, 2006; Linnerooth-
Bayer and Mechler, 2007; Mahul and Ghesquiere, 2007), sovereign
insurance for public sector assets and relief expenditure has become a
recent cornerstone for tackling the substantial and increasing effects of
disasters (Mahul and Ghesquiere, 2007). As a general statement, the
strategy involves transferring a layer of risks ranging from infrequent
risk (such as events with a return period of more than 10 years) up to risks
associated with 150-year return periods, beyond which it will become
very costly to insure (Cummins and Mahul, 2009). One key element is to
define the financial vulnerability indicating the inability to bear losses
with a certain return period (Mechler et al., 2010).
Key applications have been implemented in Mexico in 2006, which
insured its government emergency relief expenditure, and in the
Caribbean with the Caribbean Catastrophe Risk Insurance Facility in
2007 (Ghesquiere, et al., 2006; Cardenas et al., 2007). Like national
governments, donor organizations, exposed indirectly through their
relief and assistance programs, also have been considering similar
transactions; the World Food Programme in 2006, for example, purchased
‘humanitarian insurance’ for its drought exposure in Ethiopia through
index-based reinsurance (see Section 9.2.13). These transactions set
innovative and promising precedents in terms of protecting highly
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
373
exposed developing and transition government portfolios against the
risks imposed by disasters.
6.5.4. Managing the Impacts
Even in the rare circumstances where efforts outlined previously are all
in place, there still needs to be investment in capacities to manage
potential disaster impacts as risk cannot be reduced to zero (Pelling,
2003; Wisner et al., 2004; Coppola, 2007). The scale of the disaster
impact should ideally dictate the level and extent of response. Individual
household capacities to respond to disasters may be quickly overwhelmed,
requiring local resources to be mobilized (del Ninno, 2001). When
community-level responses are overwhelmed, regional or central
governments are called upon (Coppola, 2007). Some events may
overwhelm national government capacities too, and may require
mobilization of the international community of humanitarian responders
(Fagen, 2008; Harvey, 2009). International responses pose the most
complex management challenges for national governments, because of
the diversity of actors that are involved and the multiple resources flows
that are established (Borton, 1993; Bennett et al., 2006; Ramalingam et
al., 2008; ALNAP, 2010a). However, although humanitarian principles
call for a proportionate and equitable response, in practice there are a few
high-profile disasters that are over-resourced, with many more that are
‘forgotten or neglected emergencies’ (Slim, 2006). Despite the definition of
international or national disasters as those where immediate capacities
are overwhelmed, evaluations routinely find that most of the vital life-
saving activities happen at the local level, led by households, communities,
and civil society (see Sections 5.1 and 5.2; Smillie, 2001; Hilhorst, 2003;
ALNAP, 2005; Telford and Cosgrave, 2006).
In terms of how responses are managed nationally, there are different
models to consider (ALNAP, 2010b). Many countries now have some
standing capacity to manage disaster events (Interworks, 1998) and this
should be considered distinct from national systems for managing
disaster risk, commonly associated with ‘national platforms’ detailed in
Section 6.4.2. Examples of standing disaster management capacity
include the Federal Emergency Management Agency in the United
States, Public Safety in Canada, the National Commission for Disaster
Reduction in China, the National Disaster Management Authorities in
India and Indonesia, National Disaster Management Offices (NDMO) in
many Pacific island countries, and the Civil Contingencies Secretariat in
the United Kingdom. Comparative analysis of these structures shows
that there are a number of common elements (Interworks, 1998;
Coppola, 2007). Countries with formal disaster management structures
typically operate a system comprised of a National Disaster Committee,
which works to provide high-level authority and ministerial coordination,
alongside an NDMO to lead the practical implementation of disaster
preparedness and response (Interworks, 1998). National Committees
are typically composed of representatives from different ministries and
departments as well as the Red Cross/Red Crescent. They might also
include donor agencies, NGOs, and the private sector. The committee
works to coordinate the inputs of different institutions to provide a
comprehensive approach to disaster management. NDMOs usually act
as the executive arm of the national committee. Focal points for disaster
management are usually professional disaster managers. NDMOs may
be operational, or in large countries they may provide policy and
strategic oversight to decentralized operational entities at federal or
local levels. Where formal structures do not exist, national ministerial
oversight is provided to the efforts of the NDMO in times of national
disasters.
Government ownership of the national disaster management function
can vary, with three models evident: it may reside with the presidential
or prime ministerial offices; it may sit within a specific ministry; or it may
be distributed across ministries (Interworks, 1998). The way in which the
international community is engaged in major emergencies is shaped by
existing national capabilities and social contracts, with four possible
response approaches (Chandran and Jones 2008; ALNAP, 2010b; see
Table 6-4). Analysis based on these broad categories helps clarify the
ways in which international agencies are mobilized to manage disaster
impacts, following from national structure and capabilities.
There may be states where there is an existing or emerging social
contract with its citizens, by which the state undertakes to assist and
protect them in the face of disasters, and there is a limited role for
international agencies, focusing on advocacy and fundraising. By
comparison, there are states that have a growing capacity to respond
and request international agencies to supplement their effort in specific
locally owned ways, through filling gaps in national capacities or
resources. Next, there are states that have limited capacity and resources
to meet their responsibilities to assist and protect their citizens in the
face of disasters, and which request international assistance to cope
with the magnitude of a disaster, resulting in a fully fledged international
response. Finally, there are states that lack the will to negotiate a
resilient social contract, including assisting and protecting their citizens
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
Pre-disaster
Immediate
post -disaster
Recovery
Public education
Awareness raising
Warning and
evacuation plans
Pre-positioning of
resources and
supplies
Last minute
alleviation and
preparedness
measures
Search and rescue
Emergency medical
treatment
Damage and Needs
Assessment
Provision of services
water, food, health,
shelter, sanitation,
social services,
security
Resumption of critical
infrastructure
Coordination of
response
Coordination /
Management of
development partner
support
Transitional shelter in form of
temporary housing or long-
term shelter
Demolition of critically
damaged structures
Repair of less seriously
damaged structures
Clearance, removal, and
disposal of debris
Rehabilitation of infrastructure
New construction
Social rehabilitation
‘Building back better’ to reduce
future risk
Employment schemes
Reimbursement for losses
Reassessment of risks
Table 6-4 | Activities associated with managing the impacts of disasters. Adapted
from Coppola (2007) and ALNAP (2010a).
374
in times of disaster. These pose significant challenges and involve a
combination of direct delivery and advocacy. Across all four categories
of response, there are challenges around resources availability,
proportionality of distribution, coordination, and leadership (ALNAP,
2010a).
Box 6-6 outlines details of the disaster management systems of two
countries, which were chosen to illustrate the different stages of
disaster management development that are evident across states.
Although level of response and actors involved can vary considerably
between disasters and countries (ALNAP, 2010a), the basic actions
taken to manage disaster impacts remain broadly the same across
countries, and correspond closely to the different stages of the disaster
timeline (see Table 6-4; Coppola, 2007). In general, disaster management
employs immediate humanitarian activities, needs assessments, and the
delivery of goods and services to meet requirements. The demand for
water, food, shelter, sanitation, healthcare, security, and – later on –
education, employment, reconstruction, and so on is balanced against
available resources (Wisner and Adams, 2002).
Despite the existence of evidence that climate change is not responsible
for the vast majority of the increasing trend in disaster losses (see SPM
and Section 4.5.3.3), climate change-related disasters are still widely, if
incorrectly, seen by particularly the humanitarian community as playing
a major role in increasing the overall human impact of disasters.
Numerous trends in disaster events are commonly attributed to climate
change (IASC, 2009a; IFRC, 2009), and, as such, climate change is often
cited as a reason for enhancing both national and international disaster
management capacities (HFP, 2007; Oxfam, 2007; IASC, 2009a,b).
Consequently, climate change-related considerations are increasingly
featuring in literature on disaster management (Barrett et al., 2007;
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
Box 6-6 | National Disaster Preparedness, Prevention, and Management Systems: China and Kenya
China
The Government’s disaster management process, developed as National Integrated Disaster Reduction, is a comprehensive system bringing
together a number of central and local government sectors and covering the different phases of disasters preparedness, response, and
recovery/rehabilitation. China has put in place over 30 laws and regulations regarding disaster management. The Emergency Response
Law was adopted on 30 August 2007, as the central legal document governing all disaster-related efforts in China.
Under the related law and regulations, the Government has established an emergency response system consisting of three levels:
The National Master Plan for Responding to Public Emergencies – a framework to be used throughout government to ensure public
security and cope with public emergency events, including all disaster response activities.
Five national thematic disaster response plans that outline the detailed assignment of duties and arrangements for major disaster
response categories – disaster relief, flood and drought, earthquakes, geological disasters, and very severe forest fires.
Emergency response plans for 15 central government departments and their detailed implementation plans and operation norms
(UNESCAP, 2009).
Kenya
The government is working toward a national disaster management policy with the intention of preventing disasters and minimizing the
disruption they cause through taking steps to reduce risks. The policy will help enhance existing capacities by building resilience to hazard
events, building institutional capacity, developing a well-managed disaster response system, reducing vulnerability, and ensuring that
disaster policy is integrated with development policy and poverty reduction and takes a multi-sectoral, multi-level approach. The Ministry
of State for Special Programmes will be responsible for the coordination of the disaster management policy, will promote integration
and coordination of disaster management, and will establish a national institute for disaster research to improve systematic monitoring
and promotion of research.
The draft policy published in 2009 stressed the central role of climate change in any future sustainable planned and integrated National
Strategy for Disaster Management. It sets out principles for effective disaster management, codes of conduct of different stakeholders, and
provides for the establishment of an institutional framework that is legally recognized and embedded within the government structures.
It stresses the importance of mobilizing resources to enable the implementation of the policy, with provision of 2% of the annual public
budget to a National Disaster Management Fund.
At the time of writing, this policy has not reached Parliament for discussion and approval (MOSSP, 2010).
375
McGray et al., 2007; Mitchell and Van Aalst, 2008; Venton and La Trobe,
2008; IASC, 2009a). As presented in this report, evidence is available for
the influence of climate change on some extreme weather events but
not for others (see Chapter 3), and, perhaps because of this, challenges
remain in how climate change-related information can be used as a
direct guide to decisionmaking in the humanitarian sector (IASC, 2009a).
The challenges of climate change call for institutional changes in
approaches to managing disasters that are far from trivial (Salter 1998),
with such challenges including more appropriate policies and legislation;
decentralization of capacities and resources; greater budgetary allocation;
improved capacity building at the local level; and the political will to
bridge the divide between disaster risk reduction activities and the
humanitarian action associated with managing disasters (Sanderson,
2000; UNISDR, 2005). Recent analyses of the need for greater innovation
in international humanitarian responses (Ramalingam et al., 2009)
present these shifts as among the most significant and important
reforms the international system must undergo.
6.6. Aligning National Disaster Risk
Management Systems with the
Challenges of Climate Change
As mentioned, climate change presents multidimensional challenges for
national systems for managing the risks of climate extremes and disaster
risks, including potential changes in the way society views, treats, and
responds to risks and projected impacts on hazards, exposure, and
vulnerability. As climate change is altering the frequency and magnitude
of some extreme events (see Chapter 3) and contributing to trends in
exposure and vulnerability (see Chapter 4), the efficacy of national systems
of disaster risk management requires review and realignment with the
new challenges (UNISDR, 2009c; Mitchell et al., 2010a; Polack, 2010; see
FAQ 6.1). Literature suggests that the effectiveness of national systems
for managing disaster risk in a changing climate will be improved if they
integrate assessments of changing climate extremes and disasters into
current investments, strategies, and activities; seek to strengthen the
adaptive capacity of all actors; and address the causes of vulnerability
and poverty recognizing climate change as one such cause (Schipper,
2009; UNISDR 2009c; Mitchell et al., 2010a). In practice, this might
require: (i) new alliances and hybrid organizations across government
and potentially across countries; (ii) different actors to join the national
system; (iii) new cross-sector relationships; (iv) reallocation of functions,
responsibilities, and resources across scales; and (v) new practices (Hedger
et al., 2010; Mitchell et al., 2010a; Polack, 2010). As a complement to the
available data, information, and knowledge about the impact of climate
change and disaster risk presented in Chapters 2, 3, and 4, this section
seeks to elaborate the key areas where realignment of national systems
could occur – in assessing the effectiveness of disaster risk management
in a changing climate (Section 6.6.1), managing uncertainty and adaptive
management (Section 6.6.2), in tackling poverty, vulnerability, and their
structural causes (Section 6.6.3); and commenting on the practicalities of
approaching such changes holistically (Section 6.6.4).
6.6.1. Assessing the Effectiveness of Disaster Risk
Management in a Changing Climate
In order to align disaster risk management with the challenges presented
by climate change, it is necessary to assess the effectiveness and
efficiency of management options in a changing climate based on the
best available information, recognizing that that information remains
patchy at best. Adopting an economic assessment framework, different
approaches have been used to comment on the effectiveness or efficiency
of adaptation options. Many climate adaptation studies have focused on
the national-level costs of adaptation rather than comparing costs and
benefits (i.e., examining the benefits of adaptation or reduced disaster
impacts and damage costs) (see Nordhaus, 2006; EEA, 2007; UNFCCC,
2007a; Agrawala and Fankhauser, 2008; World Bank, 2008; ECA, 2009;
Parry et al., 2009). National-level adaptation assessments have been
conducted, among others, in the European Union, the United Kingdom,
Finland, The Netherlands, and Canada, as well as in a number of
developing countries using the NAPA approach (UNDP, 2004c; MMM, 2005;
DEFRA, 2006; UNFCCC, 2007b; Lemmen et al., 2008; De Bruinet al., 2009a).
Other approaches include assessments of disaster risk management
with risk assessment at the core, and focusing on economic efficiency of
management responses (see World Bank, 1996; Benson and Twigg, 2004;
Mechler, 2004). Using such a rationale, the World Bank, for example,
goes as far as suggesting that governments should in many instances
prioritize allocating their resources on early warning (such as for
floods), critical infrastructure, such as water and electricity lifelines, and
supporting environmental buffers such as mangroves, forests, and
wetlands, of which the latter should be treated with caution (World
Bank and UN, 2010). Another report suggests taking an adaptation cost
curve approach to selecting adaptation options (ECA, 2009); this
approach organizes adaptation options around their cost-benefit ratios,
similar to mitigation cost curves. Interestingly, many of the options
considered efficient in this analysis are ‘soft’ options, such as reviving
reefs, using mangroves as barriers, and nourishing beaches.
It is, however, difficult to make conclusive assessments about the
effectiveness of disaster risk management in a changing climate, as
overall the evidence base used to determine economic efficiency – that
is, benefits net of costs of adaptation – remains limited and fragmented
(Adger et al., 2007; UNFCCC, 2007a; Agrawala and Fankhauser, 2008).
In addition to the rather small number of studies available, there are
important limitations of these assessments as well. These relate to the
types of hazards examined as well as treatment of extreme events and
risk, affecting the robustness of the results. Another key limitation, relevant
for this report, is that only very few national level studies assessing
economic efficiency of options have focused explicitly on disaster risk,
and in most instances the hazards examined have been gradual, such as
sea level rise and slower onset impacts, such as drought, on agriculture
(see UNFCCC, 2007a; Agrawala and Fankhauser, 2008). Where extreme
events and disaster risks have been considered, studies have often
adopted deterministic impact metrics, when disaster risk associated
with frequency and variability of extreme events can change. Where
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
disaster risks have been accounted for, the robustness of future projections
of risk is also uncertain (Bouwer, 2010).
Furthermore, many of the economic cost assessments faced key
methodological challenges, including the difficulty in estimating economic
values of intangible effects of disasters, such as impact on human life,
suffering, and ecological services, different rates of time preferences or
discounting the future, as well as the difficulties associated with properly
accounting for the distribution of costs and benefits across different
sectors of society (Parry et al., 2009). Such challenges suggest that the
value of tools, such as cost-benefit analysis, for the assessment of
economic efficiency, even with risk considerations, may lie in the
usefulness of the analytical process rather than the numeric outcomes
per se. They suggest that in the context of climate adaptation, such
tools may be most usefully employed as a heuristic tool in the context
of iterative stakeholder decisionmaking processes (Moench et al., 2007).
376
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
FAQ 6.1 | What can a government do to better prepare its people for changing
climate-related disaster risks?
In almost all countries, governments create the enabling environment of policies, regulations, institutional arrangements, and coordination
mechanisms to guide and support the efforts of all agencies and stakeholders involved in managing disaster risks at different scales.
Such risks are increasing and changing because of population growth, migration, climate change, and a range of other factors. National
systems for managing disaster risk need to act on these changes in order to build resilience in the short and long term. Accordingly, the
following measures can be considered:
Generate and communicate robust information about the dynamic nature of disaster risk: Given the dynamic and changing
nature of disaster risks in the context of climate change, regular updates on changes in the level of risk will further strengthen such
systems if the information is acted upon. Not possessing information about changing disaster risks or not integrating the information
into decisions about longer-term investments can lead to increases in the exposure and vulnerability of people and assets and may
increase risk over time. An example could be non-drought-tolerant monoculture agriculture in an area likely to experience increased
frequency and/or longer durations of drought conditions, or water harvesting tanks installed in houses or communities that lack the
capacity to supply water during longer periods of drought, or roads not raised sufficiently above future projected flood levels.
Knowledge about dynamic risks can be generated from scientific observations and models, combined with analysis of patterns of
vulnerability and exposure and from the experiences of local communities (see Section 6.5.1).
Even without robust information, consider ‘no or
low regrets’ strategies, including ecosystem-based
adaptation: Countries have started to adopt ‘no or low
regrets’ strategies that generate short-term benefits as
well as help to prepare for projected changes in disaster
risks, even when robust information is not available (see
Section 6.3.1). Included in these ‘no or low regrets’
strategies are ecosystem-based strategies that not only
help reduce current vulnerabilities and exposure to
hazards under a range of climatic conditions, but also
produce other co-benefits such as improved livelihoods
and poverty reduction that help reduce vulnerability to
projected changes in climate. Table 6-5, a considerably
reduced version of Table 6-1, shows a summary of these
options. Such ‘no or low regrets’ practices also tend to
include measures to tackle the underlying drivers of
disaster risk and are effective irrespective of projected
changes in extremes of weather or climate (see Section
6.5.2). Where better information is available, this can be
mainstreamed across line ministries and other agencies
to shape practices that help to build resilience to
projected changes in disaster risk over the longer term.
These are highlighted in the right-hand column of
Table 6-5.
Table 6-5 | Range of practices to demonstrate comparison between ‘no or low regrets’
measures and those integrating projected changes in disaster risk.
Advances in human development and
poverty reduction, through, for example,
social protection, employment, and
wealth creation measures, taking
future exposure to weather and climate
extremes into account (very high
confidence)
‘No or low regrets’ practices
with demonstrated evidence of
having integrated observed
trends in disaster risks to reduce
the effects of disasters
Practices that enhance
resilience to projected changes
in disaster risk
Effective early warning systems and
emergency preparedness (very high
confidence)
Integrated water resource management
(high confidence)
Rehabilitation of degraded coastal and
terrestrial ecosystems (high confidence)
Robust building codes and standards
reflecting knowledge of current disaster
risks (high confidence)
Ecosystem-based/nature-based
investments, including ecosystem
conservation measures (high confidence)
Micro-insurance, including weather-
indexed insurance (medium confidence)
Vulnerability-reducing measures such as
pro-poor economic and human
development, through for example
improved social services and protection,
employment, wealth creation (very high
confidence)
Crop improvement for drought tolerance
and adaptive agricultural practices,
including responses to enhanced
weather and climate prediction services
(high confidence)
Integrated coastal zone management
integrating projections of sea level risk
and weather/climate extremes (medium
confidence)
National water policy frameworks and
water supply infrastructures,
incorporating future climate extremes
and demand projections (medium-high
confidence)
Strengthened and enforced building
codes, standards for changed climate
extremes (medium confidence)
Continued next page
377
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
Use risk-sharing and transfer mechanisms to protect financial security: To effectively support communities and protect the
financial security of the country, governments are increasingly using a range of financial instruments for transferring costs of disaster
losses through risk-sharing mechanisms. Key risk transfer instruments include financial insurance, micro-insurance, and micro-
financing, investment in social capital, government disaster reserve funds, and intergovernmental risk sharing. The latter two help
to provide much needed relief and immediate liquidity after a disaster in regions where individual countries, because of their size
and lack of diversity, cannot have viable risk insurance schemes. Such mechanisms can allow for more effective government
response, provide some relief of the fiscal burden placed on governments due to disaster impacts, and constitute critical steps in
promoting more proactive risk management strategies and responses (see Section 6.5.3).
Not all disaster risk can be eliminated, so act to manage residual risk too: Even with effective disaster risk reduction policies
and practices in place, it is impossible to reduce all disaster risks to zero and some residual risks will remain. With disaster risks
increasing in many countries, steps could be taken to strengthen governments’ ability to effectively manage residual risks effectively,
and in doing so will need to strengthen partnerships with other actors and stakeholders to enable quick and effective humanitarian
response that includes measures to ‘build back better’ and build resilience over time (for example, using rapid climate risk assessments
to position critical infrastructure or relief camps in safer locations during relief and reconstruction phases). Many governments are
also already working to enhance their disaster preparedness and early warning systems, focusing on the accuracy and timeliness of
warnings, increasing public awareness, working with communities to ensure messages are communicated and transmitted effectively,
and enhancing preparedness measures, such as first aid training, providing swimming lessons, encouraging households to have a
disaster plan and an emergency kit, securing and indicating evacuation routes and shelters, and enhancing the skills of relief workers
in child protection, for example (see Section 6.5.4).
Review resilience-building efforts: Given competing priorities and development goals, governments are forced to balance
resource allocation across development goals. The decision to bear residual losses is always a risk management option due to
financial and other constraints. Many governments decide to accept the full risk of very low probability and surprise events, but
new information on the impacts of climate change on such events may lead to such decisions being reviewed. Even in such cases
where risk reduction and risk transfer is not a viable management option, investments in reducing vulnerability and enhancing early
warning, preparedness, and standing capacity for emergency response can lead to positive returns. Furthermore, given uncertainties
associated with disasters, efforts to promote flexible institutions, cross-scale learning, improved knowledge and awareness, and
redundancies in response systems (in case one part of the system is badly impacted) can all help to promote resilience to very low
probability and surprise events. Many governments are also encouraging maintenance and strengthening of social cohesiveness
and social networks as a form of insurance enabling families and friends to support each other in times of disasters (see Sections
6.6.2 and 6.6.3).
A limited number of studies have used other tools such as multi-criteria
analysis and other variants, which do not rely on just quantitative values,
to help in the stakeholder-based adaptation decisionmaking (De Bruin
et al., 2009a; Debels et al., 2009; Cardona et al., 2010). Debels et al.
(2009) developed a multi-purpose index for a quick evaluation of
adaptation practices in terms of proper design, implementation, and
post-implementation evaluation and applied it to cases in Latin
America. Mechler et al. (2006b) developed a metric for measuring fiscal
vulnerability to natural hazards, capturing the relationship between the
economic and fiscal losses that a country could experience when a
catastrophic event occurs and the availability of funds to address the
situation. Cardona et al. (2010), building on this, constructed the Disaster
Deficit Index and applied it across a range of Latin American countries
to support governmental decisionmaking in disaster risk management
over time. De Bruin et al. (2009b) describe a hybrid approach based on
qualitative and quantitative assessments of adaptation options for
flood risk in The Netherlands. For the qualitative part, stakeholders
selected options in terms of their perceived importance, urgency, and
other elements. In the quantitative assessment, costs and benefits of
key adaptation options are determined. Finally, using priority ranking
based on a weighted sum of the qualitative and quantitative criteria
suggests that in The Netherlands, for example, an integrated portfolio of
nature and water management with risk-based policies has particularly
high potential and acceptance for stakeholders. Overall, the assessment
of adaptation explicitly considering the risk-based nature of extreme
events remains fragmented and incipient, and more work will be necessary
to improve the robustness of results and confidence in assessments.
6.6.2. Managing Uncertainties and
Adaptive Management in National Systems
Disasters associated with climate extremes are inherently complex,
involving socioeconomic as well as environmental and meteorological
uncertainty (Hallegatte et al., 2007; see Chapter 3). Population, social,
economic, and environmental change all influence the way in which
hazards are experienced, through their impact on levels of exposure and
on people’s sensitivity to hazards (Pielke Jr. et al., 2003; Aldunce et al.,
2008). Uncertainty about the magnitude, frequency, and severity of
climate extremes is managed, to an extent, through the development of
predictive models and early warning systems (see Section 3.2.3 and
Box 3-2; Section 9.2.11). Early warning systems are also based on
models and consequently there is always a probability of their success (or
failure) in predicting events accurately, although the failure to heed early
warning systems is also a function of social factors, such as perception
of risk, trust in the information-providing institution, previous experience
of the hazard, degree of social exclusion, and gender (see, e.g., Drabek,
1986, 1999). Enhanced scientific modeling and interdisciplinary
approaches to early warning systems can address some of these
uncertainties provided good baseline and time series information are
available (see Section 3.2.3 and Box 3-2). Even where such information
is available, there remain other unresolved questions that influence the
outcome of hazards. These relate to the capacity of ecosystems to provide
buffering services, and the ability of systems to recover. Management
approaches that take these issues into account include adaptive
management and resilience, yet these approaches are not without their
challenges (also see Section 8.6.3.1).
Adaptive management, as defined in Chapter 8 (Section 8.6.3.1), is “a
structured process for improving management policies and practices by
systemic learning from the outcomes of implemented strategies, and by
taking into account changes in external factors in a proactive manner”
(Pahl-Wostl et al., 2009; Pahl-Wostl, 2009). It has come to also mean
bringing together interdisciplinary science, experience, and traditional
knowledge into decisionmaking through ‘learning by doing’ by individuals
and organizations (Walters, 1997). Decisionmakers, under adaptive
management, are expected to be flexible in their approach, and accept
new information as it become available, or when new challenges emerge,
and not be rigid in their responses. Proponents argue that effective
adaptive management contributes to more rapid knowledge acquisition
and better information flows between policymakers, and ensures that
there is shared understanding of complex problems (Lee, 1993).
In most cases, adaptive management has been implemented at the local
or regional scale and there are few examples of its implementation at
the national level. Examples of adaptive management abound in
ecosystem management (Johnson, 1999; Ladson and Argent, 2000) and
in disaster risk management (Thompson and Gaviria, 2004; Tompkins,
2005; see Box 6-7). Nearly 40 years of research, after the seminal paper
was published by Holling in 1973, have produced evidence of the impacts
of aspects of resilience policy (notably adaptive management) on forests,
coral reefs, disasters, and adaptation to climate change; however, most
of this has been at the local or ecosystem scale.
One of the main unresolved issues in adaptive management is how to
ensure that scientists and engineers tasked with investigating adaptation
and disaster risk management processes are able to learn from each
other and from practitioners and how this learning can be integrated to
inform policy and management practices. In the case of the restoration
of the Florida Everglades, a limiting factor to effective management
observed was the unwillingness of some parts of society to accept short-
term losses for longer-term sustainability of ecosystem services (Kiker et
al., 2001). Investment in hurricane preparedness in New Orleans prior to
Hurricane Katrina provides a contemporary example of science not being
included in disaster risk decisionmaking and planning (Laska, 2004;
Congleton, 2006). The Cayman Islands hurricane management, on the
other hand, demonstrates a success story in a flexible disaster management
committee being prepared to change its strategies and measures from
experience, and essentially learning by doing (Box 6-7).
Spare capacity within institutions has been argued to increase the ability
of socio-ecological systems to address surprises or external shocks (Folke
et al., 2005). McDaniels et al. (2008), in their analysis of hospital resilience
to earthquake impacts, agreed with this finding, concluding that key
features of resilience include the ability to learn from previous experience,
careful management of staff during hazard, daily communication, and
willingness of staff to address specific system failures. The latter can be
achieved through creating overlapping institutions with shared delivery
of services/functions, and providing redundant capacity within these
institutions thereby allowing a sharing of the risks (Low et al., 2003).
Such redundancy increases the chances of social memory being retained
within the institution (Ostrom, 2005). However, if not carefully managed,
costs of this approach can include fragmented policy, high transactions
costs, duplication, inconsistencies, and inefficiencies (Imperial, 1999).
378
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
Box 6-7 | Building Resilience to Disasters in the Cayman Islands
Key aspects that are relevant to building disaster resilience are flexibility, learning, and adaptive governance (Adger et al., 2005; Berkes,
2007), and the Cayman Islands (Tompkins et al., 2008) illustrate how such factors help to successfully manage their disaster risks. For
example, in 2004, Hurricane Ivan (which was similar in magnitude to Hurricane Katrina that hit New Orleans in 2005) only caused two
fatalities in the island, largely due to the activities of the National Hurricane Committee (NHC), which manages hurricane disaster risk
reduction in the Cayman Islands and is responsible for preparedness, response, and recovery. The NHC is a learning-based organization.
It learns from its successes, but more importantly from mistakes made. Each year the disaster managers actively assess the previous
year’s risk management successes and failures. Every year the National Hurricane Plan is revised to incorporate this learning and to
ensure that good practices are institutionalized. Evidence of adaptive governance can be observed, for example, in the changing
composition of the NHC, its structure, network arrangements, funding allocation, and responsibilities. Policymakers are encouraged to
design and to implement new initiatives, to make adjustments, and take motivated actions. Creating such space for experimentation,
innovation, learning, and institutional adjustment is crucial for disaster resilience.
379
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
‘Learning by doing’ in disaster risk management can only be undertaken
effectively if the management institutions are scaled appropriately,
where necessary at the local level, or at multiple scales with effective
interaction (Gunderson and Holling, 2002; Eriksen et al., 2011). For the
management of climate extremes, the appropriate scale is influenced by
the magnitude of the hazard and the affected area, including biological
diversity. Research suggests that increasing biological diversity of
ecosystems allows a greater range of ecosystem responses to hazards,
and this increases the resilience of the entire system (Elmqvist et al.,
2003). Other research has shown that reducing non-climate stresses on
ecosystems can enhance their resilience to climate change. This is the
case for coral reefs (Hughes et al., 2003; Hoegh-Guldberg et al., 2008)
and rainforests (Malhi et al., 2008). Managing the resources at the
appropriate scale, for example, water catchment or coastal zone instead
of managing smaller individual tributaries or coastal sub-systems (such
as mangroves), is becoming more urgent (Sorensen, 1997; Parkes and
Horwitz, 2009).
Climate resilience as a development objective is, however, difficult to
implement, particularly as it is unclear as to what resilience means
(Folke, 2006). Unless resilience is clearly defined and broadly understood,
with measurable indicators designed to fit different local contexts and
to show the success, the potential losers from this policy may go
unnoticed, causing problems with policy implementation and legitimacy
(Eakin et al., 2009). See the Glossary for this report’s definition of
resilience, and more details regarding uncertainty and resilience related
to extreme events in the light of climate change are given in Section
3.2.3, Box 3-2, and Section 8.5.1.
6.6.3. Tackling the Underlying Drivers of Vulnerability
This assessment has found that future trends in exposure, vulnerability,
and climate extremes may further alter disaster risk and associated impacts.
Future trends in climate extremes will be affected by anthropogenic
climate change in addition to natural climate variability, and exposure and
vulnerability will be influenced by both climatic and non-climatic factors
(SPM; Sections 2.2, 2.3, and 2.5). Accordingly, reducing vulnerability and
its underlying drivers is a considered a critical aspect of addressing both
observed and projected changes in disaster risk (UNISDR 2009c, 2011b;
Figure 6-3). Section 6.5.2.2 discussed the centrality of human development
and vulnerability reduction to the goal of disaster risk reduction. As
an extension, literature focused on aligning national disaster risk
management systems to the challenges posed by climate change and
other dynamic drivers of disaster risk places considerable importance on
addressing the underlying drivers of vulnerability as one of the most
effective ‘low or no regrets’ measures (see Figure 6-3 and Table 6-5 in
FAQ 6.1; Tanner and Mitchell, 2008; Davies et al., 2008; CCCD, 2009;
UNISDR, 2009c; Mitchell et al., 2010a). Such underlying drivers of
vulnerability include inequitable development; poverty; declining
ecosystems; lack of access to power, basic services, and land; and weak
governance (Wisner et al., 2004; Schipper 2009; UNISDR 2009c, 2011b).
An approach to managing disaster risk in the context of a changing
climate highlights that disaster risk management efforts should seek to
develop partnerships to tackle vulnerability drivers by focusing on
approaches that promote more socially just and economic systems;
forge partnerships to ensure the rights and entitlements of people to
access basic services, productive assets, and common property
resources; empower communities and local authorities to influence the
decisions of national governments, NGOs, and international and private
sector organizations and to promote accountability and transparency;
and promote environmentally sensitive development (Hedger et al.,
2010; Mitchell et al., 2010a; Polack, 2010).
To date, strategies for tackling the risks of climate extremes and
disasters, in practice, have tended to focus on treating the symptoms of
vulnerability, and with it risk, rather than the underlying causes, partly
due to disaster risk management still not being a core component of
sustainable development (Schipper, 2009). The mid-term review of the HFA
indicates that insufficient effort is being made to tackle the conditions that
create risk (UNISDR, 2011b), and other studies have found a continued
disconnect between disaster risk management and development
processes that tackle the structural causes of poverty and vulnerability
and between knowledge and implementation at all scales (CCCD, 2009;
UNISDR, 2009c). The impacts of climate change, both on disaster risk
and on vulnerability and poverty, are viewed by some as a potential
force that will help to forge a stronger connection between disaster risk
reduction measures and poverty and vulnerability reduction measures,
also partly as a result of increased availability of financial resources and
renewed political will (Soussan and Burton, 2002; Schipper, 2009;
Mitchell et al., 2010a). A recent and growing body of literature has
focused on the potential for strengthening the links among particular
forms of social protection, disaster risk reduction, and climate change
adaptation measures as a way to simultaneously tackle the drivers of
vulnerability, poverty, and hence disaster risk (see Section 8.3.1; Davies
et al., 2008; Heltberg et al., 2009). With increasing levels of exposure to
disaster risk in middle-income countries (see Section 6.1; UNISDR,
2009c, 2011a), reducing vulnerability of poor people and their assets in
such locations is becoming a focus for those governments and for CSOs
and CBOs (Tanner and Mitchell, 2008).
6.6.4. Approaching Disaster Risk, Adaptation,
and Development Holistically
As this chapter has demonstrated, climate change poses diverse and
complex challenges for actors in national disaster risk management
systems and for disaster risk management policies and practices more
broadly. These challenges include changes in the magnitude and
frequency of some hazards in some regions, impacts on vulnerability
and exposure, new agreements and resource flows, and the potential of
climate change to alter value systems and people’s perceptions. As Table
SPM.1 highlights, it is the complexity resulting from the combination of
these factors, in addition to the uncertainty generated, that means
national disaster risk management systems and broader national
strategies may need to be realigned to maintain and improve their
effectiveness. There is high agreement but limited evidence to suggest
that a business-as-usual approach to disaster risk management that
fails to take the impacts of climate change into account will become
increasingly ineffective. Section 6.6 and other parts of this chapter
have assessed evidence on the different elements involved in such a
realignment. A selection of these elements is briefly summarized here.
As discussed in Section 6.6.2, there is high agreement but limited
evidence to suggest that flexible and adaptive national systems for
disaster risk management, based on the principle of learning by doing,
are better suited to managing the challenges posed by changes in
exposure, vulnerability, and weather and climate extremes than static
and rigid systems (see Section 8.6). This ability to be flexible will be
tested by a systems’ capacity to act on new knowledge generated by
the frequent assessment of dynamic risk needed to capture trends in
exposure vulnerability and weather and climate extremes and by
information on how the costs and benefits of different response measures
change as a result (Section 6.6.1). The accuracy of these assessments
will be based on the quality of available data (Section 6.5.2.1). Where
such assessments generate uncertainty for decisionmakers, tools such
as multi-criteria analysis, scenario planning, and flexible decision paths
offer ways of supporting informed action (Section 6.6.1).
There is high agreement and robust evidence to demonstrate that the
mainstreaming of disaster risk management processes into development
planning and practice leads to more resilient development pathways. By
extension, with climate change and other development processes having
an impact on disaster risk, these changes then need to be factored into
development and economic planning decisions at different scales. This
suggests an ideal national system for managing the risks from climate
extremes and disasters would be designed to be fully integrated with
economic and social development, environmental, poverty reduction,
and humanitarian dimensions to create a holistic approach. The nature
of transformational changes in thinking, analysis, planning, approaches,
strategies, and actions is the subject of Chapter 8 (particularly Section
8.2.2).
While there is limited evidence that some countries have begun to factor
climate change into the way disaster risks are assessed and managed
(see Sections 6.3 and 6.6.1), few countries appear to have adopted a
comprehensive approach – for example, by addressing projected changes
in exposure, vulnerability, and extremes as well as adopting a learning-
by-doing approach to decisionmaking embedded in the context of
national development planning processes. Incremental efforts toward
implementing suitable strategies for mainstreaming climate change
responses into national development planning and budgetary processes,
and climate proofing at the sector and project levels (Sections 6.2 and
6.3) in the context of disaster risk management appear to be the most
likely approach adopted by many countries. None of these measures will
be easy to implement, as actors and stakeholders at all levels of society
are being asked to embrace a dynamic notion of risk as an inherent part
of their decisions, and continuously learn and modify policies, decisions,
and actions taking into account new traditional and scientific knowledge
as it emerges.
The knowledge base for understanding changing climate-related disaster
risks and for the way national systems are acting on this understanding
through modifying practices, altering the nature of relationships
between different actors, and adopting new strategies and policies is
fragmented and incomplete. As this chapter has illustrated, incomplete
information and knowledge gaps do not need to present blockages to
action. As FAQ 6.1 and Section 6.3.1 highlight, there is considerable
experience of governments and other actors investing in measures to
respond to existing climate variability and disaster risk that can be
considered as ‘no or low regrets’ options when taking into account the
uncertainty associated with future climate. However, in conducting this
assessment, some knowledge gaps have emerged that, if filled, would
aid the creation of enduring national risk management systems for
tackling observed and projected disaster risk. These gaps include the
need for more research on:
The extent to which efforts to build disaster risk management
capacities at different scales prepare people and organizations for
the challenges posed by climate change.
Whether the current trend of decentralizing disaster risk management
functions to sub-national and local governments and communities
is effective, given the level of information and capacity requirements,
changing risks, and associated uncertainties presented by climate
change.
How the function, roles, and responsibilities of different actors
working within national disaster risk management systems are
changing, given the impacts of climate change at the national and
sub-national level.
The characteristics of flexibility, learning-by-doing, and adaptive
management in the context of national disaster risk management
systems in different governance contexts.
How decisions on disaster risk management interventions are made
at different scales if there is limited context-specific information.
The costs and benefits of different risk management interventions
if the impacts of climate change and other dynamic drivers of risk
are factored in.
The benefits and tradeoffs of creating integrated programs and
policies that seek to manage disaster risk, mitigate GHGs, adapt to
climate change, and reduce poverty simultaneously.
380
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
381
References
A digital library of non-journal-based literature cited in this chapter that
may not be readily available to the public has been compiled as part of
the IPCC review and drafting process, and can be accessed via either the
IPCC Secretariat or IPCC Working Group II web sites.
Aakre, S., I. Banaszak, R. Mechler, D. Rübbelke, A. Wreford, and H. Kalirai, 2010:
Financial adaptation to disaster risk in the European Union. Mitigation and
Adaptation Strategies for Global Change, 15(7), 721-736.
ActionAid, 2006: Climate Change, Urban Flooding and the Rights of the Urban Poor
in Africa: Key findings from six African cities. ActionAid International, London,
UK and Johannesburg, South Africa.
Adger, W.N., T.P. Hughes, C. Folke, S.R. Carpenter, and J. Rockstrom, 2005: Social-
ecological resilience to coastal disasters. Science, 5737(309), 1036-1039.
Adger, W.N., S. Agrawala, M.M.Q. Mirza, C. Conde, K. O’Brien, J. Pulhin, R. Pulwarty,
B. Smit, and K. Takahashi, 2007: Assessment of adaptation practices, options,
constraints and capacity. In: Climate Change 2007. Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani,
J.P. Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 717- 743.
ADPC, 2007: Community Self-Resilience and Flood Risk Reduction. ADB TA 4574-
CAM Report, Asian Disaster Preparedness Centre, Asian Development Bank,
Manila, Philippines.
African Development Bank Group, 2010: Nairobi Rivers Rehabilitation and
Restoration Program: Sewerage Improvement Project. Project Appraisal Report,
African Development Bank, www.afdb.org/fileadmin/uploads/afdb/Documents/
Project-and-Operations/Kenya%20-%20Nairobi%20rivers%20rehabilitation
%20and%20restoration%20program%20-%20sewerage%20improvement
%20project.pdf. Last accessed on: 26th Sept 2011.
Agrawala, S. and S. Fankhauser, 2008: Economic Aspects of Adaptation to Climate
Change. Costs, benefits and policy instruments. Organisation for Economic
Cooperation and Development, Paris, France.
Agrawala, S. and M. van Aalst, 2008: Adapting development cooperation to adapt
to climate change. Climate Policy, 8(3), 183-193.
Albala-Bertrand, J.M., 1993: Political Economy of Large Natural Disasters with
Special Reference to Developing Countries. Clarendon Press, Oxford, UK.
Aldunce, P. and M. González, 2009: Desastres Asociados al Clima en la Agricultura y
Medio Rural de Chile. Universidad de Chile, Santiago, Chile.
Aldunce, P. and A. Leon, 2007: Opportunities for improving disaster management in
Chile: A case study. Disaster Prevention and Management, 16(1), 33-41.
Aldunce, P, C. Neri, and C. Szlafsztein (eds.), 2008: Hacia la Evaluación de Prácticas
de Adaptación ante la Variabilidad y el Cambio Climático. Nucleo de Meio
Ambiente/Universidad Federal do Para, Belém, Brazil.
ALNAP, 2005: Capacity Building amid Humanitarian Response. Active Learning Network
for Accountability and Performance in Humanitarian Action, London, UK.
ALNAP, 2010a: The State of the Humanitarian System: assessing performance and
progress. Active Learning Network for Accountability and Performance in
Humanitarian Action, London, UK.
ALNAP, 2010b: The Role of National Governments in International Humanitarian
Response to Disasters. Meeting Background Paper, 26th ALNAP Meeting, Active
Learning Network for Accountability and Performance in Humanitarian Action,
London, UK, www.alnap.org/pool/files/26-meeting-background-paper.pdf.
Amendola, A., J. Linnerooth-Bayer, N. Okada, and P. Shi, 2008: Towards integrated
disaster risk management: case studies and trends from Asia. Natural Hazards,
44, 163-168.
Andersen, T.J., 2001: Managing Economic Exposures of Natural Disasters. Exploring
Alternative Financial Risk Management Opportunities and Instruments. Inter-
American Development Bank, Washington, DC.
Angelsen, A., M. Brockhaus, M. Kanninen, E. Sills, W.D. Sunderlin, and S. Wertz-
Kanounnikoff, 2009: Realising REDD: National strategy and policy options.
Center for International Forestry Research, Indonesia.
Arnell, N., 2004: Climate change and global water resources: SRES emissions and
socio-economic scenarios. Global Environmental Change, 14(1), 31-52.
Arnell, N.W. and E.K. Delaney, 2006: Adapting to climate change: Public water
supply in England and Wales. Climatic Change, 78(2-4), 227- 255.
Arnold, M., 2008: The Role of Risk Transfer and Insurance in Disaster Risk Reduction
and Climate Change Adaptation. Policy brief, Commission on Climate Change
and Development, Stockholm, Sweden.
Arrow, K. and R. Lind, 1970: Uncertainty and the evaluation of public investment
decisions. The American Economic Review, 60, 364-378.
Auffret, P., 2003: Catastrophe Insurance Market in the Caribbean Region: Market
Failures and Recommendations for Public Sector Interventions. World Bank
Policy Research Working Paper 2963, World Bank, Washington DC.
Auld, H., 2008a: Disaster risk reduction under current and changing climate conditions.
World Meteorological Organization Bulletin, 57(2), 118-125.
Auld, H.E., 2008b: Adaptation by design: The impact of changing climate on
infrastructure. Journal of Public Works and Infrastructure, 3, 276-288.
Auld, H., D. MacIver, and J. Klaassen, 2004: Heavy rainfall and waterborne disease
outbreaks: The Walkerton example. Journal of Toxicology and Environmental
Health, Part A, 67(20-22), 1879-1887.
AusAid, 2009: Investing in a Safer Future: A Disaster Risk Reduction policy for the
Australian aid program. AusAid, Canberra, Australia, www.ausaid.gov.au/
publications/pdf/disasterriskreduction.pdf.
Averyt, K., 2010: Are we successfully adapting science to climate change? Bulletin
of the American Meteorological Society, 91(6), 723-726.
Ayers, J.M. and S. Huq, 2009: The value of linking mitigation and adaptation: A case
study of Bangladesh. Environmental Management, 43(5), 753-764.
Balmford, A., A. Bruner, P. Cooper, R. Costanza, S. Farber, R.E. Green, M. Jenkins, P.
Jefferiss, V. Jessamy, and J. Madden, 2002: Economic reasons for conserving
wild nature. Science, 297(5583), 950-953.
Barbier, E.B., 2009: Rethinking the Economic Recovery: A Global Green New Deal.
Report prepared for the Economics and Trade Branch, Division of Technology,
Industry and Economics, United Nations Environment Programme (UNEP),
Geneva, Switzerland.
Barrett, E., S. Murfitt, and P. Venton, 2007: Mainstreaming the Environment into
Humanitarian Response: An Exploration of Opportunities and Issues
.
Environmental Resource Management Limited, London, UK.
Basher, R., 2006: Global early warning systems for natural hazards: Systematic and
people-centered. Philosophical Transactions of the Royal Society A, 364(1845),
2167-2182.
Bebi, P., D. Kulakowski, and C. Rixen, 2009: Snow avalanche disturbances in forest
ecosystems – State of research and implications for management. Forest
Ecology and Management, 257(9), 1883-1892.
Bennett, J., W. Bertrand, C. Harkin, S. Samarasinghe, and H. Wickramatillake, 2006:
Coordination of International Humanitarian Assistance in Tsunami Affected
Countries. Tsunami Evaluation Coalition, London, UK.
Benson, C., 2007: Tools for Mainstreaming Disaster Risk Reduction: Environmental
Assessments. Guidance Note 7, ProVention Consortium, Geneva, Switzerland.
Benson, C. and E. Clay, 2002a: Understanding the Economic and Financial Impacts
of Natural Disasters. World Bank, Washington, DC.
Benson, C. and E. Clay, 2002b: Disasters, vulnerability, and the global economy. In:
Building Safer Cities: The Future of Disaster Risk [Kreimer, A., M. Arnold, and A.
Carlin (eds.)]. Disaster Risk Management Series No. 3, World Bank, Washington,
DC, pp. 3-32.
Benson, C. and E. Clay, 2004: Understanding the Economic and Financial Impacts of
Natural Disasters. Disaster Risk Management Series 4, World Bank,
Washington, DC.
Benson, C. and J. Twigg, 2004: Measuring Mitigation: Methodologies for Assessing
Natural Hazard Risks and the Net Benefits of Mitigation – A Scoping Study. The
International Federation of Red Cross and Red Crescent Societies/The
ProVention Consortium, Geneva, Switzerland.
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
382
Benson, C., J. Twigg, and M. Myers, 2001: NGO initiatives in risk reduction: An
overview. Disasters, 25(3), 199-215.
Berkes, F., 2007: Understanding uncertainty and reducing vulnerability: Lessons from
resilience thinking. Natural Hazards, 41(2), 283-295.
Berry, P., 2007: Adaptation Options on Natural Ecosystems. Report to the Financial
and Technical Support Division, United Nations Framework Convention on
Climate Change, Bonn, Germany.
Bhavnani, R., M. Owor, S. Vordzorgbe, and F. Bousquet, 2008: Status of Disaster Risk
Reduction in the Sub-Saharan Africa Region. World Bank, Washington, DC.
Biesbroek, G.R., R.J. Swart, T.R. Carter, C. Cowan, T. Henrichs, H. Mela, M.D.
Morcecroft, and D. Rey, 2010: Europe adapts to climate change: Comparing
National Adaptation strategies. Global Environmental Change, 20(3), 440-450.
Birkmann, J., M. Garschagen, F. Kraas, and N. Quang, 2011: Adaptive urban
governance: new challenges for the second generation of urban adaptation
strategies to climate change. Sustainability Science, 5, 185-206.
Borton, J., 1993: Recent trends in the international relief system. Disasters, 17(3),
187-201.
Bosher, L., P. Carillo, A. Dainty, J. Glass, and A. Price, 2007: Realising a resilient and
sustainable built environment: Towards a strategic agenda for the United
Kingdom. Disasters, 31, 236-255.
Botzen, W.J.W., J.C.J.H. Aerts, and J.C.J.M. van den Bergh, 2009: Willingness of
homeowners to mitigate climate risk through insurance. Ecological Economics,
68(8-9), 2265-2277.
Bourrelier, P.H., B.G. Deneufbourg, and B. de Vanssay, 2000: IDNDR objectives: French
technical sociological contributions. Natural Hazards Review, 1(1), 18-26.
Bouwer, L.M., 2006: The benefits of disaster risk reduction and their effects on loss
trends. In: Climate Change and Disaster Losses: Understanding and Attributing
Trends and Projections [Hoppe, P. and R. Pielke (eds.)]. Centre for Science and
Technology Policy Research, Tyndall Centre, Munich Re and National Science
Foundation, Hamburg, Germany, pp 26-28.
Bouwer, L.M., 2010: Disasters and Climate Change: Analysis and Methods for
Projecting Future Losses from Extreme Weather. PhD thesis, Vrije Universiteit,
Amsterdam, The Netherlands, 141 pp., dare.ubvu.vu.nl/bitstream/1871/16355/
1/dissertation.pdf.
Branco, A., J. Suassuna, and S.A. Vainsencher, 2005: Improving access to water
resources through rainwater harvesting as a mitigation measure: The case of
the Brazilian semi-arid region. Mitigation and Adaptation Strategies for Global
Change, 10(3), 393-409.
Brocklesby, M. and E. Fisher, 2003: Community development in sustainable livelihoods
approaches – an introduction. Community Development Journal, 38, 185-198.
Brooks, N. and W.N. Adger, 2005: Adaptation policy frameworks for climate change:
developing strategies, policies and measures. In: Assessing and Enhancing
Adaptive Capacity
[Lim, B., E. Spanger-Siegfried, I. Burton, E.L. Malone, and S.
Huq (eds.)]. Cambridge University Press, Cambridge, UK, pp. 165-181.
Brooks, N., N. Grist, and K. Brown, 2009: Development futures in the context of climate
change: Challenging the present and learning from the past. Development
Policy Review, 27, 741-765.
Bunce, M., K. Brown, and S. Rosendo, 2010: Policy misfits, climate change and cross-
scale vulnerability in coastal Africa: how development projects undermine
resilience. Environmental Science and Policy, 13, 485-497.
Burby, R, 1991. Sharing Environmental Risks: How to Control Governments’ Losses
in Natural Disasters. Westview Press, Boulder, CO.
Butler, C.D. and W. Oluoch-Kosura, 2006: Linking future ecosystem services and
future human well-being. Ecology and Society, 11(1), 30.
Butt, T.A., B.A. Mccarl, J. Angerer, P.T. Dyke, and J.W. Stuth, 2005: The economic and
food security implications of climate change in Mali. Climatic Change, 68(3),
355-378.
CACCA, 2010: Approaches to Climate Change Adaptation. Committee on
Approaches to Climate Change Adaptation, Global Environment Bureau,
Ministry of Environment, Tokyo, Japan, www.env.go.jp/en/earth/cc/adapt_guide.
Campbell, A., V. Kapos, A. Chenery, S.I. Kahn, M. Rashid, J.P.W. Scharlemann, and B.
Dickson, 2008: The Linkages between Biodiversity and Climate Change
Mitigation. UNEP World Conservation Monitoring Center, Cambridge, UK.
Cardenas, V., S. Hochrainer, R. Mechler, G. Pflug, and J. Linnerooth-Bayer, 2007:
Sovereign financial disaster risk management: The case of Mexico.
Environmental Hazards, 7(1), 40-53.
Cardona, O., 2006: Measuring vulnerability to natural hazards: Towards disaster
resilient societies. In: A system of indicators for disaster risk management in the
Americas [Birkmann, J. (ed.)]. United Nations University Press, Tokyo, Japan.
Cardona, O.D., M.G. Ordaz, M.C. Marulanda, M.L. Carreno, and A.H. Barbat, 2010:
Disaster risk from a macroeconomics perspective: a metric for fiscal vulnerability
evaluation. Disasters, 34(4), 1064-1083.
CARE International, 2008: Community preparedness for emergencies helps poverty.
In: Linking Disaster Risk Reduction and Poverty: Good Practices and Lessons
Learnt [United Nations International Strategy for Disaster Reduction (ed.)].
United Nations International Strategy for Disaster Reduction, Geneva,
Switzerland, pp. 6-10.
Carter, T.R., R.N. Jones, X. Lu, S. Bhadwal, C. Conde, L.O. Mearns, B.C. O’Neill,
M. Rounsevell, and M.B. Zurek, 2007: New assessment methods and the
characterisation of future conditions. In: Climate Change 2007. Impacts,
Adaptation and Vulnerability. Contribution of Working Group II to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change [Parry,
M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)].
Cambridge University Press, Cambridge, UK, pp.133-171.
Carter, W.N., 1992: Disaster Management: A Disaster Manager’s Handbook. Asian
Development Bank, Manila, Philippines.
Caselli, F. and P. Malthotra, 2004: Natural Disasters and Growth: From Thought
Experiment to Natural Experiment. International Monetary Fund, Washington,
DC.
Cash, D.W. and J. Buizer, 2005: Knowledge-Action Systems for Seasonal to
Interannual Climate Forecasting: Summary of a Workshop. Roundtable on
Science and Technology for Sustainability Policy and Global Affairs Division,
National Research Council of the National Academies, The National Academies
Press, Washington DC.
CCCD, 2009: Closing the Gaps: Disaster Risk Reduction and Adaptation to Climate
Change in Developing Countries: Final Report. Commission on Climate Change
and Development, Ministry for Foreign Affairs, Stockholm, Sweden.
CDB and CARICOM, 2004: Sourcebook on the Integration of Natural Hazards into
Environmental Impact Assessment (NHIA-EIA Sourcebook). Caribbean
Development Bank, Bridgetown, Barbados.
Chandran, R. and B. Jones, 2008: Concepts and Dilemmas of State Building in Fragile
Situations: from fragility to resilience. OECD/DAC Discussion Paper,
Organisation for Economic Cooperation and Development, Paris, France.
Charveriat, C., 2000: Natural disasters in Latin America and the Caribbean: an
overview of risk. Inter-American Development Bank, Washington, DC.
Chhatre, A. and A. Agrawal, 2009: Trade-offs and synergies between carbon storage
and livelihood benefits from forest commons. Proceedings of the National
Academy of Sciences, 106(42), 17667-17670.
Christensen, J.H., B. Hewitson, A. Busuioc, A Chen, A. Gao, I. Held, R. Jones, R.K. Koli,
W.T. Kwon, R. Laprise, V.M. Rueda, L. Mearns, C.G. Menéndez, J. Räisänen, A.
Rinke, A. Sarr, and P. Whetton, 2007: Regional climate predictions. In: Climate
Change 2007: The Physical Science Basis. Contribution of Working Group I to
the Fourth Assessment Report of the Intergovernmental Panel on Climate
Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M.
Tignor and H.L. Miller (eds.)]. Cambridge University Press, Cambridge, UK, and
New York, NY, pp. 847-940.
Chung, J., 2009: Situation analysis for strengthening disaster information management
systems in Fiji: baseline data preparedness and assessment methodologies. A
report prepared for the Fiji Government and Pacific Disaster Risk Management
Partnership, United Nations International Strategy for Disaster Reduction and
South Pacific Applied Geoscience Commission, Suva, Fiji.
Cline, W.R., 2007: Global Warming and Agriculture: Impact Estimates by Country.
Center for Global Development and Peterson Institute for International
Economics, Washington, DC.
Coleman, T., 2002: The Impact of Climate Change on Insurance against
Catastrophes. Insurance Australia Group, Melbourne, Australia.
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
383
Colls, A., N. Ash, and N. Ikkala, 2009: Ecosystem-based Adaptation: A Natural
Response to Climate Change. International Union for Conservation of Nature,
Gland, Switzerland.
Congleton, R.D., 2006: The story of Katrina: New Orleans and the political economy
of catastrophe. Public Choice, 127(1), 5-30.
Conway, D. and L. Schipper, 2011: Adaptation to climate change in Africa:
Challenges and opportunities identified from Ethiopia. Global Environmental
Change, 21(1), 227-237.
Coppola, D.P, 2007: Introduction to International Disaster Management.
Butterworth-Heinemann, Oxford, UK.
Costanza, R., O. Pérez-Maqueo, M.L. Martinez, P. Sutton, S.J. Anderson, and K.
Mulder, 2008: The value of coastal wetlands for hurricane protection. Ambio,
37(4), 241-248.
Crabbé, P. and M. Robin, 2006: Institutional adaptation of water resource infrastructures
to climate change in eastern Ontario. Climatic Change, 78(1), 103 -133.
CRGR, 2007a: Lineamientos generales de organización y funcionamiento de la
Concertación Regional de Gestión de Riesgos. Políticas, Prácticas y Gestión del
Riesgo, 1(7), 4-7.
CRGR, 2007b: Reforma del Estado y gestión del riesgo de desastres. Políticas,
Prácticas y Gestión del Riesgo, 2(8), 3-11.
CRGR, 2009: Centroamérica frente a los retos del cambio climático. Posicionamiento
político de la sociedad civil y pueblos indígenas organizados. Políticas, Prácticas
y Gestión del Riesgo, 3(12), 12-15.
Crowards, T., 2000: Comparative Vulnerability to Natural Disasters in the Caribbean.
Caribbean Development Bank, Bridgetown, Barbados.
CSA, 2010a: Infrastructure in Permafrost: A Guideline for Climate Change Adaptation.
A Technical Guide. Canadian Standards Association, Mississauga, Canada.
CSA, 2010b: Development, interpretation and use of rainfall intensity-duration-
frequency (IDF) information: Guideline for Canadian water resources practitioners.
Technical Guide, Canadian Standards Association, Mississauga, Canada.
Cummins, J., and O. Mahul, 2009: Catastrophe Risk Financing In Developing
Countries: Principles For Public Intervention. World Bank, Washington, DC.
Curriero, F.C., J.A. Patz, J.B. Rose, and S. Lele, 2001: The association between extreme
precipitation and waterborne disease outbreaks in the United States, 1948-
1994. American Journal of Public Health, 91(8), 1194-1199.
Das, S. and J.R. Vincent, 2009: Mangroves protected villages and reduced death tolls
during Indian super cyclone. Proceedings of the National Academy of Sciences,
106(18), 7357-7360.
DaSilva, J., B. Garanganga, V. Teveredzi, S.M. Marx, S.J. Mason, and S.J. Connor, 2004:
Improving epidemic malaria planning, preparedness and response in Southern
Africa. Malaria Journal
, 3(1), 37.
Davies, M., B. Guenther, J. Leavy, T. Mitchell, and T. Tanner, 2008: Adaptive social
protection synergies for poverty reduction. Institute of Development Studies
(IDS) Bulletin, 39(4), 105-112.
Davis, K. (ed.), 2004: Technology Dissemination among Small-Scale Farmers in Meru
Central District of Kenya: Impact of Group Participation. USA: University of
Florida.
De Bruin, K., R. Dellink, and S. Agrawala, 2009a: Economic Aspects of Adaptation to
Climate Change: Integrated Assessment Modeling of Adaptation Costs and
Benefits. Environment Working Paper No. 6, Organisation for Economic
Cooperation and Development, Paris, France.
De Bruin, K., R.B. Dellink, A. Ruijs, L. Bolwidt, A. van Buuren, J. Graveland, R.S. de
Groot, P.J. Kuikman, S. Reinhard, R.P. Roetter, V.C. Tassone, A. Verhagen, and E.C.
van Ierland, 2009b: Adapting to climate change in The Netherlands: an
inventory of climate adaptation options and ranking of alternatives. Climatic
Change, 95, 23-45.
Debels, P., C. Szlafsztein, P. Aldunce, C. Neri, Y. Carvajal, M. Quintero-Angel, A. Celis, A.
Bezanilla, and D. Martinez, 2009: IUPA: A tool for the evaluation of the general
usefulness of practices for adaptation to climate change and variability. Natural
Hazards, 50(2), 211-233.
DEFRA, 2005: Making Space for Water. Taking Forward a New Government Strategy
for Flood and Coastal Erosion Risk Management in England. Department for
Environment, Food and Rural Affairs, London, UK.
DEFRA, 2006: The UK’s Fourth National Communication under the United Nations
Framework Convention on Climate Change. Department for Environment, Food
and Rural Affairs, London, UK.
del Ninno, C., 2001: The 1998 Floods in Bangladesh: Disaster Impacts, Household
Coping Strategies, and Response. Research Report 122, International Food
Policy Research Institute, Washington, DC.
del Ninno, C., P.A. Dorosh, and L.C. Smith, 2003: Public policy, markets and household
coping strategies in Bangladesh; Avoiding a food security crisis following the
1998 floods. World Development, 31(7), 1221-1238.
Dercon, S., 2005: Risk, insurance, and poverty: A review. In: Insurance Against
Poverty. A study prepared by the World Institute for Development Economics
Research of the United Nations University (UNU-WIDER)[ Dercon, S. (ed.)].
Oxford University Press, Oxford, UK, pp. 9-37.
Dessai, S., and M. Hulme, 2007: Assessing the robustness of adaptation decisions to
climate change uncertainties: A case study on water resources management in
the East of England. Global Environmental Change, 17, 59-72.
DFID, 2005: Natural Disaster and Disaster Risk Reduction Measures: A Desk Review
of Costs and Benefits. UK Department for International Development, The
Stationery Office, London, IK.
DFID, 2011: Multilateral Aid Review: Ensuring Maximum Value for Money for UK Aid
through Multilateral Organisations. UK Department for International
Development, London, UK.
Dilley, M., 2006: Risk identification: A critical component of disaster risk management.
World Meteorological Organisation Bulletin, 55(1), 13-20.
Dolan, A.H. and I.J. Walker, 2006: Understanding vulnerability of coastal communities to
climate change related risks. In: Proceedings of the 8th Int’l Coastal Symposium.
Itajaí, Brazil, 14-18 March 2004. Journal of Coastal Research, SI 39
, 1316-1323.
Dorren, L.K.A., F. Berger, A.C. Imeson, B. Maier, and F. Rey, 2004: Integrity, stability
and management of protection forests in the European Alps. Forest Ecology and
Management, 195(1-2), 165-176.
Drabek, T.E., 1986: Human System Responses To Disaster: An Inventory Of
Sociological Finding. Springer-Verlag New York, NY.
Drabek, T.E., 1999: Understanding disaster warning responses. The Social Science
Journal, 36, 515-523.
Dudley, N., S. Stolton, A. Belokurov, L. Krueger, N. Lopoukhine, K. Mackinnon, T.
Sandwith, and N. Sekhran, 2010: Natural Solutions: Protected Areas Helping
People Cope with Climate Change. IUCN, WCPA, TNC, UNDP, WCS, World Bank
and WWF, Gland, Switzerland, Washington, DC and New York, NY.
Eakin, H., E.L. Tompkins, D.R. Nelson, and J.M. Anderies, 2009: Adapting to climate
change: Thresholds, values, governance. In: Hidden Costs and Disparate
Uncertainties: Trade-Offs Involved in Approaches to Climate Policy [Adger, W.N.,
I. Lorenzoni, and L.O.B. Karen (eds.)]. Cambridge University Press, Cambridge,
UK, pp. 212-226.
Easterling, W.E., P.K. Aggarwal, P. Batima, K.M. Brander, L. Erda, S.M. Howden, A.
Kirilenko, J. Morton, J.-F. Soussana, J. Schmidhuber, and F.N. Tubiello, 2007: Food,
fibre and forest products. In: Climate Change 2007. Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani,
J.P. Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 273-313.
Ebi, K.L., 2008: Adaptation costs for climate change-related cases of diarrhoeal
disease, malnutrition, and malaria in 2030. Globalization and Health, 6, 9,
doi:10.1186/1744-8603.
Ebi, K.L., R.S. Kovats, and B. Menne, 2006. An approach for assessing human health
vulnerability and public health interventions to adapt to climate change.
Environmental Health Perspectives, 114, 12, doi:10.1289/Ehp.8430.
EC, 2009: Technical Report – 2009 – 040, Guidance Document No. 24, River Basin
Management. In: A Changing Climate, Common Implementation Strategy for
the Water Framework. Directive 2000/60/EC. European Commission, Brussels,
Belgium.
ECA, 2009: Shaping Climate-Resilient Development: A Framework For Decision-
Making Study. Economics of Climate Adaptation Working Group, World Bank,
Washington, DC.
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
384
ECLAC, 2002: Handbook for Establishing the Socio-Economic and Environmental
Effects of Natural Disasters. United Nations, ECLAC and World Bank, Mexico
City, Mexico.
ECLAC, 2003: Handbook for Establishing the Socio-Economic and Environmental
Effects of Natural Disasters. United Nations, Economic Commission for Latin
America and the Caribbean, and World Bank, Santiago, Chile.
EEA, 2007: Climate change: The Cost of Inaction and the Cost of Adaptation.
Technical Report, European Environment Agency, Copenhagen, Denmark.
Elmqvist, T., C. Folke, M. Nyström, G. Peterson, J. Bengtsson, B. Walker, and J.
Norberg, 2003: Response diversity, ecosystem change, and resilience. Frontiers
in Ecology and the Environment, 1(9), 488-494.
Embrechts, P., C. Klüppelberg, and T. Mikosch, 1997: Modelling Extremal Events for
Insurance And Finance. Springer, Berlin, Germany.
Ericksen, P., J. Ingram, and D. Liverman, 2009: Food security and global environmental
change: emerging challenges. Environmental Science and Policy, 12, 373-377.
Eriksen, S. and L.O. Næss, 2003: Pro-Poor Climate Adaptation: Norwegian
Development Cooperation and Climate Change Adaptation – An Assessment of
Issues, Strategies and Potential Entry Points. Report 2003:02. Center for
International Climate and Energy Research – Oslo (CICERO), Oslo, Norway,
75pp.
Eriksen, S., P. Aldunce, C. Bahinipati, R. D’Almeida, J. Molefe, C. Nhemachena, K.
O’Brien, F. Olorunfemi, J. Park, L. Sygna, and K. Ulsrud, 2011: When not every
response to climate change is a good one: Identifying principles for sustainable
adaptation. Climate Change and Development, 3(1), 7-20.
Fagen, P., 2008: Natural Disasters in Latin America and the Caribbean: National,
Regional and International Interactions. HPG Working Paper, Overseas
Development Institute, London, UK.
Fankhauser, S., 2010: The costs of adaptation. Wiley Interdisciplinary Reviews:
Climate Change, 1(1), 23-30.
FAO, 2006: Third session of the sub-committee on aquaculture: Committee on
Fisheries (CoFI). India, New Delhi, 4-8 September. Food and Agriculture
Organization, Rome, Italy.
FAO, 2007: Building Adaptive Capacity to Climate Change: Policies to Sustain
Livelihoods and Fisheries. New Directions in Fisheries – A Series of Policy Briefs
on Development Issues, 0816. Food and Agriculture Organization, Rome, Italy.
FAO, 2008a: Climate change adaptation and mitigation in the food and agriculture
sector: In: Technical background document from the expert consultation held
on 5 to 7 March 2008 (Paper presented at Climate change, energy and food).
Food and Agriculture Organization, Rome, Italy.
FAO, 2008b: Challenges for Sustainable Land Management for Food Security in
Africa. 25th Regional Conference for Africa, Nairobi, Kenya. Food and
Agriculture Organization, Rome, Italy.
FAO, 2008c: Climate Change for Fisheries and Aquaculture. (Technical background
document from the expert consultation). Food and Agriculture Organization,
Rome, Italy.
FAO, 2009: Seed Security for Food Security in the Light of Climate Change and
Soaring Food Prices: Challenges and Opportunities (Report of the twenty first
session of the Committee on Agriculture). Food and Agriculture Organization,
Rome, Italy.
FAO
, 2010: “Climate-Smart” Agriculture: Policies, Practices and Financing for Food
Security, Adaptation and Mitigation. Food and Agriculture Organization, Rome,
Italy.
Feldman, D.L. and H.M. Ingram, 2009: Making science useful to decision makers:
Climate forecasts, water management, and knowledge networks. Weather,
Climate and Society, 1, 9-20.
Few, R., H. Osbahr, L.M. Bouwer, D. Viner, and F. Sperling, 2006: Linking Climate
Change Adaptation and Disaster Risk Management for Sustainable Poverty
Reduction (Synthesis report). Vulnerability and Adaptation Resource Group,
Washington, DC.
Folke, C., 2006: Resilience: The emergence of a perspective for social-ecological
systems analyses. Global Environmental Change, 16(3), 253-267.
Folke, C., T. Hahn, P. Olsson, and J. Norberg, 2005: Adaptive governance of social-
ecological systems. Annual Review of Environment and Resources, 30, 441-473.
Freeman, P.K. and Warner, K., 2001: Vulnerability of Infrastructure to Climate
Variability: How Does this Affect Infrastructure Lending Policies? World Bank
and ProVention Consortium, Washington, DC.
Freeman, P.K, L.A. Martin, J. Linnerooth-Bayer, R. Mechler, G. Pflug, and K. Warner,
2003: Disaster Risk Management: National Systems for the Comprehensive
Management of Disaster Risk and Financial Strategies for Natural Disaster
Reconstruction. Sustainable Development Department, Inter-American
Development Bank, Washington, DC.
Froot, K.A., 2001: The market for catastrophe risk: A clinical examination. Journal of
Financial Economics, 60, 529-571.
Füssel, H., 2007: Adaptation planning for climate change: Concepts, assessment
approaches, and key lessons. Sustainability Science, 2(2), 265-275.
Gero, A., K. Méheux, and D. Dominey-Howes, 2011: Integrating community based
disaster risk reduction and climate change adaptation: examples from the
Pacific. Natural Hazards and Earth System Sciences, 11, 101-113.
GFDRR, 2010: Partnership Charter. Global Facility for Disaster Risk Reduction,
Washington, DC, gfdrr.org/gfdrr/sites/gfdrr.org/files/publication/GFDRR_
Partnership_Charter_2010.pdf.
Ghesquiere, F., O. Mahul, M. Forni, and R. Gartley, 2006: Caribbean Catastrophe
Risk Insurance Facility: A solution to the short-term liquidity needs of small
island states in the aftermath of natural disasters. World Bank, Washington, DC,
siteresources.worldbank.org/PROJECTS/Resources/Catastrophicriskinsurance
facility.pdf.
Glantz, M.H., 2005: Hurricane Katrina rekindles thoughts about fallacies of a so-
called “natural” disaster. Sustainability: Science, Practice and Policy, 1(2),
1-4.
GNDR, 2009: Clouds but Little Rain. Views from the Frontline: A Local Perspective of
Progress towards Implementation of the Hyogo Framework for Action. Global
Network of Civil Society Organisations for Disaster Reduction, Teddington, UK.
Grosse, G., V. Romanovsky, T. Jorgenson, K.W. Anthony, J. Brown, and P.P. Overduin,
2011: Vulnerability and feedbacks of permafrost to climate change. EOS,
Transactions AGU, 92(9), 73-80.
Grossi
, P., and H. Kunreuther (eds.), 2005: Catastrophe Modeling: A New Approach
to Managing Risk. Springer, New York, NY.
GTZ/DKKV, 2007: National Platforms for Disaster Reduction: Study on current status
of disaster reduction, institutional arrangements and potential for national
platforms for disaster reduction 3 South and South East Asian Countries.
Gesellschaft für Technische Zusammenarbeit/ German Committee for Disaster
Risk Reduction, Bonn, Germany.
Guha-Sapir, D. and R. Below, 2002: Quality and Accuracy of Disaster Data: A
Comparative Analysis of Three Global Datasets. Working document prepared for
Disaster Management Facility, World Bank, Washington DC.
Gunderson, L. and C.S. Holling, 2002: Panarchy Synopsis: Understanding
Transformations in Human and Natural Systems. Island Press, Washington, DC.
Gwimbi, P., 2007: The effectiveness of early warning systems for the reduction of
flood disasters: Some experiences from cyclone induced floods in Zimbabwe.
Journal of Sustainable Development in Africa, 9(4), 152-169.
Haasnoot, M., H. Middelkoop, E. van Beek, and W.P.A. van Deursen, 2009: A method
to develop sustainable water management strategies for an uncertain future.
Sustainable Development, 19, doi:10.1002/sd.438.
Haines, A., R.S. Kovats, D. Campbell-Lendrum, and C. Corvalan, 2006: Climate change
and human health: impacts, vulnerability and public health. Public Health,
120(7), 585-596.
Hallegatte, S., 2009: Strategies to adapt to an uncertain climate change. Global
Environmental Change, 19(2), 240-247.
Hallegatte, S. and M. Ghil, 2007: Endogenous Business Cycles and the Economic
Response to Exogenous Shocks. Fondazione Eni Enrico Mattei, Milan, Italy.
Hallegatte, S., J.-C. Hourcade, and P. Dumas, 2007: Why economic dynamics matter
in assessing climate change damages: Illustration on extreme events. Ecological
Economics, 62, 330-340.
Halsnaes, K., and S. Traerup, 2009: Development and climate change: A mainstreaming
approach for assessing economic, social, and environmental impacts of adaptation
measures. Environmental Management, 43, 765-778.
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
385
Hammer, K., N. Arrowsmith, and T. Gladis, 2003: Agrobiodiversity with emphasis on
plant genetic resources. Naturwissenschaften 2003, 90(6), 241-250.
Hammill, A. and T. Tanner, 2010: Climate risk screening and assessment tools:
making sense of a crowded field. In: Meeting of the OECD DAC-EPOC Joint Task
Team on Climate Change and Development Co-operation, Amsterdam, The
Netherlands, 12-13 October 2010.
Handmer, J. and S. Dovers, 2007: Handbook of Disaster and Emergency Policies and
Institutions. Earth Scan, London, UK.
Harris, P. (ed.), 2009: Climate Change and Foreign Policy: Case Studies from East to
West. Routledge, London, UK.
Harvey, P., 2009: Towards Good Humanitarian Government: The Role of the Affected
State in Disaster Response. Humanitarian Policy Group, Overseas Development
Institute, London, UK.
Hay, J., 2009: Institutional and Policy Analysis of Disaster Risk Reduction and Climate
Change Adaptation in Pacific Island Countries. A report prepared for the United
Nations International Strategy for Disaster Reduction (UNISDR) and the United
Nations Development Programme (UNDP). UNISDR and UNDP, Suva, Fiji.
Hayley, D.W. and B. Horne, 2008: Rationalizing climate change for design of structures
on permafrost: a Canadian perspective. In: Proceedings of Ninth International
Conference on Permafrost [Kane, D.L. and K.M. Hinkle (eds.)]. University of
Alaska Fairbanks, Fairbanks, AK, 1, 681-686.
Hazell, P. and U. Hess, 2010: Drought insurance for agricultural development and
food security in dryland areas. Food Security, 2, 395-405.
Hedger, M. and J. Cacouris, 2008: Separate Streams? Adapting Water Resources
Management to Climate Change. Tearfund, Teddington, UK.
Hedger, M., A. Singha, and M. Reddy, 2010: Building Climate Resilience at State
Level: Disaster Risk Management and Rural Livelihoods in Orissa.
Strengthening Climate Resilience Discussion Paper 5, Institute of Development
Studies, Brighton, UK.
Heltberg, R., S.L. Jorgensen, and P.B. Siegel, 2008: Climate Change, Human
Vulnerability, and Social Risk Management. World Bank, Washington, DC.
Heltberg, R., P.B. Siegel, and S.L. Jorgensen, 2009: Addressing human vulnerability
to climate change: toward a ‘no-regrets’ approach. Global Environmental
Change, 19(1), 89-99.
Hess, U. and H. Syroka, 2005: Weather-based Insurance in Southern Africa. The Case
of Malawi. World Bank, Washington, DC.
HFP, 2007: Dimensions of Crisis Impacts: Humanitarian Needs by 2015. Working
Paper, Humanitarian Futures Programmes, Kings College London, London,
UK.
Hilhorst, D., 2003: Unlocking disaster paradigms: An actor-oriented focus on disaster
response. Abstract submitted for Session 3 of the Disaster Research and Social
Crisis Network Panels of the 6th European Sociological Conference, Murcia,
Spain 23-26 September 2003.
Hochrainer, S., 2006: Macroeconomic Risk Management against Natural Disasters.
Deutscher Universita Ts-Verlag, Wiesbaden, Germany.
Hodgson, R. and M. Carter, 1999: Some lessons for a national approach to building
for safety in Bangladesh. In: Natural Disaster Management [Ingleton, J. (ed.)].
Tudor Rose, Leicester, UK, pp. 160-162.
Hoegh-Guldberg, O., P.J. Mumby, A.J. Hooten, R.S. Steneck, P. Greenfield, E. Gomez,
D.R. Harvell, P.F. Sale, A.J. Edwards, K. Caldeira, N. Knowlton, C.M. Eakin, R.
Iglesias-Prieto, N. Muthiga, R.H. Bradbury, H. Dubi, and M.E. Hatziolos, 2008:
Coral adaptation in the face of climate change. Science, 320(5874), 315.
Hoeppe, P. and E.N. Gurenko, 2006: Scientific and economic rationales for innovative
climate insurance solutions. Climate Policy, 6(6), 607-620.
Holling, C.S., 1973: Resilience and stability of ecological systems. Annual Review of
Ecology and Systematics, 4(1), 1-23.
Howden, S.M., J.F. Soussana, F.N. Tubiello, N. Chhetri, M. Dunlop, and H. Meinke,
2007: Adapting agriculture to climate change. Proceedings of the National
Academy of Sciences, 104(50), 19691-19696.
Hughes, T.P., A.H. Baird, D.R. Bellwood, M. Card, S.R. Connolly, C. Folke, R. Grosberg,
O. Hoegh-Guldberg, J.B.C. Jackson, and J. Kleypas, 2003: Climate change,
human impacts, and the resilience of coral reefs. Science, 301(5635),
929.
IASC, 2009a: Addressing the Humanitarian Challenges of Climate Change Regional
and National Perspectives. Preliminary Findings from the IASC Regional and
National Level Consultations, Inter-Agency Standing Committee, Geneva,
Switzerland, www.humanitarianinfo.org/iasc/pageloader.aspx?page=content-
news-newsdetails&newsid=134.
IASC, 2009b: Letter from the IASC Principals to Yvo de Boer, Executive Secretary of
the United Nations Framework Convention on Climate Change. Inter-Agency
Standing Committee, Geneva, Switzerland, www.humanitarianinfo.org/iasc/
pageloader.aspx?page=content-news-newsdetails&newsid=134.
Ibarraran, M.E., M. Ruth, S. Ahmad, and M. London, 2009: Climate change and
natural disasters: Macroeconomic performance and distributional impacts.
Environment, Development and Sustainability, 11, 549-569.
ICHARM, 2009: Global Trends in Water-Related Disasters: An Insight for
Policymakers. International Centre for Water Hazard and Risk Management,
World Water Assessment Programme, UNESCO, Paris, France.
IFRC, 2004: World Disasters Report 2004: Focus on Community Resilience. Int’l
Federation of the Red Cross and Red Crescent Societies, Geneva, Switzerland.
IFRC, 2009: World Disasters Report: Focus on Early Warning, Early Action.
International Federation of the Red Cross and Red Crescent Societies, Geneva,
Switzerland.
IFRC, 2010: World Disasters Report 2010: Focus on Urban Risk. International
Federation of the Red Cross and Red Crescent Societies, Geneva, Switzerland.
IIED, 2009: Climate Change and the Urban Poor: Risk and Resilience in 15 of the
World’s Most Vulnerable Cities. International Institute for Environment and
Development, London, UK.
Ikeme, J., 2003: Climate change adaptational deficiencies in developing countries:
The case of sub-Saharan Africa. Mitigation and Adaptation Strategies for Global
Change, 8
, 29-52.
Imperial, M.T., 1999: Institutional analysis and ecosystem-based management: The
institutional analysis and development framework. Environmental
Management, 24(4), 449-465.
Interworks, 1998: Model for a National Disaster Management Structure,
Preparedness Plan, and Supporting Legislation, UNOCHA Disaster Management
Training Programme. UN Office for the Coordination of Humanitarian Affairs,
ocha.unog.ch/drptoolkit/PreparednessTools/IL%20Frameworks/Models%20for
%20Institutional%20structures%20and%20legislation,%20DMTP,%20InterW
orks,1998.pdf.
IRI, 2006: A Gap Analysis for the Implementation of the Global Climatic Observing
System Programme in Africa. IRI Technical Report No. IRI-TR/06/01,
International Research Institute for Climate and Society, Columbia University,
New York, NY.
Islam, T., and S. Ferdousi, 2007: Renewable energy development – challenges for
Bangladesh. Energy and Environment, 18, 421-430.
Jaspars, S. and D. Maxwell, 2009: Food Security and Livelihoods Programming in
Conflict: A Review. HPN Network Paper 65. Humanitarian Practice Network,
Overseas Development Institute, London, UK.
Johnson, B.L., 1999: Introduction to the special feature: Adaptive management –
scientifically sound, socially challenged. Conservation Ecology, 3(1), 10.
Juhola, S., and L. Westerhoff, 2011: Challenges of adaptation to climate change
across multiple scales: a case study of network governance in two European
countries. Environmental Science and Policy, 14, 239-247.
Kagiannas, A., D. Askounis, K. Anagnostopoulos, and J. Psarras, 2003: Energy policy
assessment of the Euro-Mediterranean Cooperation. Energy Conversion and
Management, 44(16), 2665-2686.
Katz, R. and G.B. Brown, 1992: Extreme events in a changing climate: Variability is
more important than averages. Climatic Change, 21, 289-302.
Kay, R. and J. Adler, 2005: Coastal planning and management. Third Edition.
Routledge, New York, NY.
Kellenberg, D., K. Mobarak, and A. Mushfiq, 2008: Does rising income increase or
decrease damage risk from natural disasters? Journal of Urban Economics,
63(3), 788-802.
Kelman, I., 2008: Relocalising disaster risk reduction for urban resilience. Urban
Design and Planning, 161(4), 197-204.
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
386
Kelman, I., J. Lewis, J.C. Gaillard, and J. Mercer, 2011: Participatory action research
for dealing with disasters on islands. Island Studies Journal, 6, 59-86.
Kesavan, P.C. and Swaminathan, M.S., 2006: Managing extreme natural disasters in
coastal areas. Philosophical Transactions of the Royal Society A, 364(1845),
2191-2216.
Kiker, C.F., J.W. Milon, and A.W. Hodges, 2001: Adaptive learning for science-based
policy: The everglades restoration. Ecological Economics, 37(3), 403-416.
Kirshen, P., 2007: Adaptation Options and Costs in Water Supply. Report to the
UNFCCC Secretariat, Financial and Technical Support Division, United Nations
Framework Convention on Climate Change, Bonn, Germany.
Klein, R.J.T., R.J. Nicholls, S. Ragoonaden, M. Capobianco, J. Aston, and E.N. Buckley,
2001: Technological options for adaptation to climate change in coastal zones.
Journal of Coastal Research, 17(3), 531-543.
Klein, R.J.T., S. Eriksen, L.O. Næss, A. Hammill, C. Robledo, and K. O’Brien, 2007:
Portfolio screening to support the mainstreaming of adaptation to climate
change into development. Climatic Change, 84, 23-44.
Klijn, F., M. van Buuren, and S.A.M. van Rooij, 2004: Flood-risk management
strategies for an uncertain future: living with Rhine river floods in the
Netherlands? Ambio, 33, 141-147.
Krysanova, V., C. Dickens, J. Timmerman, C. Varela-Ortega, M. Schlüter, K. Roest, P.
Huntjens, F. Jaspers, H. Buiteveld, E. Moreno, J. de Pedraza Carrera, R. Slámová,
M. Martínková, I. Blanco, P. Esteve, K. Pringle, C. Pahl-Wostl, and P. Kabat, 2010:
Cross-comparison of climate change adaptation strategies across large river
basins in Europe, Africa and Asia. Water Resources Management, 24, 4121-
4160.
Kuzak, D., 2004: The application of probabilistic earthquake risk models in
managing earthquake insurance risks in Turkey. In: Catastrophe Risk and
Reinsurance: A Country Risk Management Perspective [Gurenko, E. (ed.)]. Risk
Books, London, UK.
Kwadijk, C.J., M. Haaasnoot, J.P.M. Hoogvliet, M.C. Marco, A.B.M. Jeuken, N.G.C. van
der Oostrom, H.A. Schelfhout, E.H. van Velzen, H. van Waveren, and M.J.M. de
Wit, 2010: Using adaptation tipping points to prepare for climate change and
sea level rise: a case study in the Netherlands. Wiley Interdisciplinary Reviews:
Climate Change, 1, 729-740.
Ladson, A.R. and R.M. Argent, 2000: Adaptive management of environmental flows:
Lessons for the Murray-Darling Basin from three large North American rivers.
Australian Journal of Water Resources, 5(51), 89-101.
Lal, P.N., 2010: Vulnerability to natural disasters: an economic analysis of the impact
of the 2009 floods on the Fijian sugar belt. Pacific Economic Bulletin, 25(2),
62-77.
Larsen, P.H., S. Goldsmith, O. Smith, M.L. Wilson, K. Strzepek, P. Chinowsky, and B.
Saylor, 2008: Estimating future costs for Alaska public infrastructure at risk from
climate change. Global Environmental Change, 18(3), 442-457.
Laska, S., 2004: What if hurricane Ivan had not missed New Orleans? Natural
Hazards Observer, 29(2), 5-6.
Lavell
, A., 1996. Environmental degradation, risks and urban disasters. In: Issues and
Concepts: Towards the Definition of a Research Agenda. La Red, Lima, Peru, pp.
19-58.
Lavell, A., 1998. Decision making and risk management. In: Furthering Cooperation
in Science and Technology for Caribbean Development. Port of Spain, Trinidad,
23-25 September 1998.
Lavell, A., 2009: Unpacking Climate Change Adaptation and Disaster Management:
Searching for the Links and Differences: A Conceptual and Epistemological
Critique and Proposal. FLACSO, Bogota, Colombia.
Lavell, A. and E. Franco, 1996: Estado, sociedad y gestiãn de loes desastres en
america latina: en busqueda del paradigma perdido. La Red, Bogota, Colombia.
Lee, K., 1993: Compass and Gyroscope. Island Press, California.
Lemmen, D., F. Warren, J. Lacroix, and E. Bush, 2008: From Impacts to Adaptation:
Canada in a Changing Climate 2007. Natural Resources Canada, Ottawa,
Canada.
Leslie, H.M. and K.L. Mcleod, 2007: Confronting the challenges of implementing
marine ecosystem-based management. Frontiers in Ecology and the
Environment, 5(10), 540-548.
Lewis, J. and M.P. Chisholm, 1996: Cyclone-resistant domestic construction in
Bangladesh. In: Implementing Proceedings of Hazard-Resistant Housing:
Proceedings of the First International Housing and Hazards Workshop to
Explore Practical Building for Safety Solutions [Hodgson, R.L.P., S.M. Seraj, and
J.R. Choudhury (eds.)]. Dhaka, Bangladesh, 3-5 December 1996.
Linnerooth-Bayer, J. and A. Amendola, 2000: Global change, catastrophic risk and
loss spreading. The Geneva Papers on Risk and Insurance, 25(2), 203-219.
Linnerooth-Bayer, J. and R. Mechler, 2006: Insurance for assisting adaptation to
climate change in developing countries: A proposed strategy. Climate Policy,
6(6), 621-636.
Linnerooth-Bayer, J. and R. Mechler, 2007: Disaster safety nets for developing
countries: Extending public-private partnerships. Environmental Hazards, 7,
54-61.
Linnerooth-Bayer, J., R. Mechler, and G. Pflug, 2005: Refocusing disaster aid.
Science, 309(5737), 1044-1046.
Llosa, S. and I. Zodrow, 2011: Disaster Risk Reduction Legislation as a Basis for
Effective Adaptation. Background Paper prepared for the 2011 Global
Assessment Report on Disaster Risk Reduction. United Nations International
Strategy for Disaster Reduction, Geneva, Switzerland.
Longley, C. and D. Maxwell, 2003: Livelihoods, chronic conflict and humanitarian
response: A review of current approaches. Natural Resource Perspectives, 89,
1-6.
Low, B., E. Ostrom, C. Simon, and J. Wilson, 2003: Navigating social-ecological
systems: Building resilience for complexity and change. In: Redundancy and
diversity: Do they influence optimal management? [Berkes, F., J. Colding, and C.
Folke (eds.)]. Cambridge University Press, Cambridge, UK, pp. 83-114.
Lowe, R., 2003: Preparing the built environment for climate change. Building
Research and Information,
31(3-4), 195-199.
Lu, X., 2011: Provision of climate information for adaptation to climate change.
Climate Research, 47, 83-94.
Mahul, O. and F. Ghesquiere, 2007: Sovereign Natural Disaster Insurance for
Developing Countries: A Paradigm Shift in Catastrophe Risk Financing. Policy
Research Working Paper 4345, World Bank, Washington, DC.
Malhi, Y., J.T. Roberts, R.A. Betts, T.J. Killeen, W. Li, and C.A. Nobre, 2008: Climate
change, deforestation and the fate of the Amazon. Science, 319(5860),
169-172.
Mansourian, S., A. Belokurov, and P. Stephenson, 2009: The role of forest protected
areas in adaptation to climate change. Unasylva, 60, 63-69.
Manuel-Navarrete, D., J. Gomez, and G. Gallopin, 2007: Syndromes of sustainability
of development for assessing the vulnerability of coupled human–environmental
systems: The case of hydrometeorological disasters in Central America and the
Caribbean. Global Environmental Change, 17, 207-217.
Maréchal, K., 2007: The economics of climate change and the change of climate in
economics. Energy Policy, 35(10), 5181- 5194.
Mastrandrea, M.D., N.E. Heller, T.L. Root, and S.H. Schneider, 2010: Bridging the gap:
linking climate-impacts research with adaptation planning and management.
Climatic Change, 100(1), 87-101.
Masumoto, T., T. Yoshida, and T. Kubota, 2006: An index for evaluating flood-prevention
function of paddies. Paddy and Water Environment, 4(4), 205-210.
Masumoto, T., H. Pham Thanh, and K. Shimizu, 2008: Impact of paddy irrigation levels
on floods and water use in the Mekong River Basin. Hydrological Processes,
22(9), 1321-1328.
McBean, G.A., 2008: Role of prediction in sustainable development and disaster
management. Springer- Hexagon Series on Human and Environmental Security
and Peace, 3, 929-938.
McCarl, B.A., 2007: Adaptation Options for Agriculture, Forestry and Fisheries.
A Report to the UNFCCC Secretariat Financial and Technical Support
Division, United Nations Framework Convention on Climate Change, Bonn,
Germany.
McDaniels, T., S. Chang, D. Cole, J. Mikawoz, and H. Longstaff, 2008: Fostering
resilience to extreme events within infrastructure systems: characterizing
decision contexts for mitigation and adaptation. Global Environmental Change:
Human and Policy Dimensions, 18, 310-318.
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
387
McEntire, D.A. and V. Myers, 2004: Preparing communities for disasters: Issues and
processes for government readiness. Disaster Prevention and Management,
13(2), 140-152.
McEvoy, D., P. Matczak, I. Banaszak, and A. Chorynski, 2010: Framing adaptation to
climate-related extreme events. Mitigation and Adaptation Strategies to Global
Change, 15, 779-795.
McGranahan, G., D. Balk, and B. Anderson, 2007: The rising tide: assessing the risks
of climate change and human settlements in low elevation coastal zones.
Environment and Urbanization, 19, 17-37.
McGray, H., A. Hammill, R. Bradley, E.L. Schipper, and J.E. Parry, 2007: Weathering
the Storm: Options for Framing Adaptation and Development. World Resources
Institute, Washington, DC.
McKenzie, E., B. Prasad, and A. Kaloumaira, 2005: Economic Impact of Natural
Disasters on Development in the Pacific. Volume I. SOPAC and USP, Suva, Fiji.
MEA, 2005: Ecosystems and Human Well-Being: Synthesis. Millennium Ecosystem
Assessment, Island Press, Washington, DC.
Mechler, R., 2004: Natural Disaster Risk Management and Financing Disaster Losses
in Developing Countries. Verlag für Versicherungswirtschaft, Germany.
Mechler, R., J. Linnerooth-Bayer, and D. Peppiatt, 2006a: Microinsurance for Natural
Disasters in Developing Countries: Benefits, Limitations and Viability.
ProVention Consortium, Geneva, Switzerland.
Mechler, R., J. Linnerooth-Bayer, S. Hochrainer, and G. Pflug, 2006b: Assessing
financial vulnerability and coping capacity: The IIASA CATSIM model. In:
Measuring Vulnerability and Coping Capacity to Hazards of Natural Origin.
Concepts and Methods [Birkmann, J. (ed.)]. United Nations University Press,
Tokyo, pp. 380-398.
Mechler, R., S. Hochrainer, A. Aaheim, Z. Kundzewicz, N. Lugeri, M. Moriondo, H.
Salen, M. Bindi, I. Banaszak, A. Chorynski, E. Genovese, H. Kalirai, J. Linnerooth-
Bayer, C. Lavalle, D. Mcevoy, P. Matczak, M. Radziejewski, D. Rübbelke, M.J.
Schelhaas, M. Szwed, and A. Wreford, 2010: Making climate change work for
us: European perspectives on adaptation and mitigation strategies. In: A Risk
Management Approach for Assessing Adaptation to Changing Flood and
Drought Risks in Europe [Hulme, H.N. (ed.)]. Cambridge University Press,
Cambridge, UK, pp. 200-229.
Meehl, G.A., F. Zwiers, J. Evans, T. Knutson, L. Mearns, and P. Whetton, 2000: Trends
in extreme weather and climate events: Issues related to modeling extremes in
projections of future climate change. Bulletin of the American Meteorological
Society, 81(3), 427-436.
Meinke, H., R. Nelson, P. Kokic, R. Stone, R. Selvaraju, and W. Baethgen, 2006:
Actionable climate knowledge: From analysis to synthesis. Climate Research,
33(1), 101.
Menkhaus, K., 2007: Governance without government in Somalia: Spoilers, state
building and the politics of coping. International Security, 31(3), 74-106.
Mercer, J., D. Dominey-Howes, I. Kelman, and K. Lloyd, 2007: The potential for
combining indigenous and western knowledge in reducing vulnerability to
environmental hazards in small island developing states. Environmental
Hazards, 7, 245-256.
Mertz, O., K. Halsnaes, J. Olesen, and R. Rasmussen, 2009: Adaptation to climate
change in developing countries: Special feature. Environmental Management,
43, 743-752.
Miller
, F., H. Osbahr, E. Boyd, F. Thomalla, S. Bharwani, G. Ziervogel, B. Walker, J.
Birkmann, S. van der Leeuw, J. Rockström, J. Hinkel, T. Downing, C. Folke, and D.
Nelson, 2010: Resilience and vulnerability: Complementary or conflicting
concepts? Ecology and Society, 15(3), 11.
Mills, E., 2005: Insurance in a climate of change. Science, 309(5737), 1040-1044.
Mills, E., 2006: Testimony to the National Association of Insurance Commissioners.
The Role of NAIC in responding to climate change. UCLA Journal of
Environmental Law and Policy, 26(1), 129-168.
Mills, E., 2007: Synergisms between climate change mitigation and adaptation: An
insurance perspective. Mitigation and Adaptation Strategies for Global Change,
12(15), 809-842.
Mitchell, T. and S. Maxwell, 2010: Defining Climate Compatible Development. Policy
Brief, Climate and Development Knowledge Network, London, UK.
Mitchell, T. and M. van Aalst, 2008: Convergence of Disaster Risk Reduction and
Climate Change Adaptation: A Review for DFID. Department for International
Development (DFID), London, UK.
Mitchell, T., M. Ibrahim, K. Harris, M. Hedger, E. Polack, A. Ahmed, N. Hall, K.
Hawrylyshyn, K. Nightingale, M. Onyango, M. Adow, and M. Sajjad, 2010a:
Climate Smart Disaster Risk Management. Strengthening Climate Resilience.
Institute of Development Studies, Brighton, UK.
Mitchell, T., M. van Aalst, and P. Villaneuva, 2010b: Assessing Progress on
Integrating Disaster Risk Reduction and Climate Change Adaptation in
Development Processes. Strengthening Climate Resilience Programme, Institute
of Development Studies, Brighton, UK.
MMM, 2005: Finland’s National Strategy for Adaptation to Climate Change. Ministry
of Agriculture and Forestry, Helsinki, Finland, www.mmm.fi/en/index/
frontpage/ymparisto/ilmastopolitiikka/ilmastomuutos.html.
Moench, M., R. Mechler, and S. Stapleton, 2007: Guidance note on the costs and
benefits of disaster risk reduction. In: UNISDR High Level Platform on Disaster
Risk Reduction, Geneva , Switzerland, 4-7 June 2007. United Nations
International Strategy for Disaster Reduction (UNISDR), Geneva, Switzerland.
Moser, C. and D. Satterthwaite, 2008: Towards Pro-Poor Adaptation to Climate
Change in the Urban Centers of Low- and Middle-Income Countries.
International Institute for Environment and Development, London, UK.
MOSSP, 2010: Draft National Policy for Disaster Management in Kenya. Ministry of
State of Special Programs, Kenya, www.kecosce.org/downloads/DRAFT_
DISASTER_MANAGEMENT_POLICY.pdf.
Muller, M., 2007: Adapting to climate change: Water management for urban
resilience. Environment and Urbanization, 19(1), 99-113.
Munich Re, 2005: Short Annual Report 2005: Paving the way for opportunities.
Munich Re Group, Munich, Germany.
Munich Re, 2011: Topics Geo Natural Catastrophes 2010: Analyses, Assessments,
Positions. Munich Re, Munich, Germany.
Murlidharan, T.L. and H.C. Shah, 2001: Catastrophes and macro-economic risk
factors: An empirical study. In: Conference on Integrated Disaster Risk
Management: Reducing Socio-Economic Vulnerability. International Institute
for Applied Systems Analysis (IIASA), Laxenburg, Austria.
Murray, B.C., W.A. Jenkins, S. Sifleet, L. Pendleton, and A. Baldera, 2010: Payment for
Blue Carbon: Potential for Protecting Threatened Coastal Habitats. Nicholas
Institute for Environmental Policy Solutions, Duke University, Durham, NC.
Musgrave, R.A., 1959: The Theory of Public Finance. McGraw Hill, New York, NY.
National Platform for Kenya, 2009: Interim National Progress Report on Hyogo
Framework for Action. Prepared for the UN Global Assessment Report 2009,
United Nations International Strategy for Disaster Reduction, Geneva,
Switzerland, preventionweb.net/files/7432_Kenya.pdf.
Neufeldt, H., A. Wilkes, R.J. Zomer, J. Xu, E. Nang’ole, C. Munster, and F. Place, 2009:
Trees on Farms: Tackling the Triple Challenges of Mitigation, Adaptation and
Food Security. Policy Brief 07, World Agroforestry Centre, Kenya.
Neumann, J., 2009: Adaptation to Climate Change: Revisiting Infrastructure Norms.
Resources for the Future, Washington, DC.
Nicholls, R.J., 2007: Adaptation Options for Coastal Areas and Infrastructure: An
Analysis for 2030. Report to the United Nations Framework Convention on
Climate Change (UNFCCC), UNFCCC, Bonn, Germany.
Nicholls, R.J., P.P. Wong, V. Burkett, C.D. Woodroffe, and J. Hay, 2008: Climate change
and coastal vulnerability assessment: scenarios for integrated assessment.
Sustainability Science, 3, 89-102.
Nordhaus, W., 2006: The Economics of Hurricanes in the United States. National
Bureau of Economic Research, Cambridge, MA.
Noy, I., 2009: The macroeconomic consequences of disasters. Journal of Development
Economics, 88(2), 221-231.
NRC (National Research Council), 2009: Informing Decisions in a Changing Climate.
Panel on Strategies and Methods for Climate-Related Decision Support, NRC,
National Academies Press, Washington, DC.
NRTEE, 2009: True North: Adapting Infrastructure to Climate Change in Northern
Canada. National Round Table on Environment and Economy, Ottawa,
Canada.
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
388
O’Brien, G., 2006: UK emergency preparedness – A step in the right direction?
Journal of International Affairs, 59(2), 63-85.
O’Brien, G., 2008: UK emergency preparedness – A holistic local response? Disaster
Prevention and Management, 17(2), 232-243.
O’Brien, G. and P. O’Keefe, 2010: Resilient responses to climate change and variability:
A challenge for public policy. International Journal of Public Policy, 6(3-4), 369-385.
O’Brien, G. and P. Read, 2005: Future UK emergency management: New wine, old
skin? Disaster Prevention and Management, 14(3), 353-361.
O’Brien, G., P. O’Keefe, J. Rose, and B. Wisner, 2006: Climate change and disaster
management. Disasters, 30(1), 64-80.
O’Brien, K., L. Sygna,, R. Leichenko, W.N. Adger, J. Barnett, T. Mitchell, L. Schipper, T.
Tanner, C. Vogel, and C. Mortreux, 2008: Disaster Risk Reduction, Climate
Change Adaptation and Human Security. Report prepared for the Royal
Norwegian Ministry of Foreign Affairs by the Global Environmental Change and
Human Security (GECHS) Project, GECHS Report 2008:3, Oslo, Norway.
OAS, 1991: Primer on Natural Hazard Management in Integrated Regional
Development Planning. Organization of American States, Washington, DC.
OECD, 2006: Applying Strategic Environmental Assessment: Good Practice Guidance
For Development Co-Operation. Organisation for Economic Cooperation and
Development, Paris, France.
OECD, 2009: Policy Guidance on Integrating Climate Change Adaptation into
Development Co-operation. Organisation for Economic Cooperation and
Development, Paris, France.
Olhoff, A. and C. Schaer, 2010: Screening Tools and Guidelines to Support the
Mainstreaming of Climate Change Adaptation into Development Assistance –
A Stocktaking Report. United Nations Development Programme, New York, NY.
Olsen, J.R., 2006: Climate change and floodplain management in the United States.
Climatic Change, 76(3-4), 407- 426.
ONERC, 2009: Climate Change Costs of Impacts and Lines of Adaptation. A report to
the Prime Minister and Parliament, Observatoire National sur les Effets du
Réchauffement Climatique, Paris, France.
Ostrom, E., 2005: Understanding Institutional Diversity. Princeton University Press,
Princeton, NJ.
Otero, R.C. and R.Z. Marti, 1995: The impacts of natural disasters on developing
economies: Implications for the international development and disaster
community. In: Disaster Prevention for Sustainable Development: Economic and
Policy Issues [Munasinghe, M. and C. Clarke (eds.)]. World Bank, Washington,
DC, pp. 11-40.
Owens, T., T. Hoddinott, and B. Kinsey, 2003: Ex-ante actions and ex-post public
responses to drought shocks: Evidence and simulations from Zimbabwe. World
Development, 31(7), 1239-1255.
Oxfam, 2007:
Climate Alarm: Disasters Increase as Climate Change Bites. Oxfam
Briefing Paper 108, Oxfam, Oxford, UK, www.oxfam.org.uk/resources/policy/
climate_change/downloads/bp108_weather_alert.pdf.
Oxfam America, 2008: National network for DRR helps curb poverty. In: Linking
Disaster Risk Reduction and poverty: Good practices and Lessons Learnt [United
Nations International Strategy for Disaster Reduction (UNISDR) (ed.)]. UNISDR,
Geneva, Switzerland, pp. 11-14.
Pahl-Wostl, C., 2009: A conceptual framework for analysing adaptive capacity and
multi-level learning processes in resource governance regimes. Global
Environmental Change, 19, 354-365.
Pahl-Wostl, C., J. Sendzimer, and P. Jeffrey, 2009: Resources management in transition.
Ecology and Society, 14(1), 46, www.ecologyandsociety.org/vol14/iss1/art46/.
Parkes, M.W and P. Horwitz, 2009: Water, ecology and health: Ecosystems as settings
for promoting health and sustainability. Health Promotion International, 24(1), 94.
Parry, M., N. Arnell, P. Berry, D. Dodman, S. Fankhauser, C. Hope, S. Kovats, R.
Nichollas, D. Satterthwaite, R. Tiffin, and R. Wheeler, 2009: Assessing the Costs
of Adaptation to Climate Change: A review of the UNFCCC and other recent
estimates. International Institute for Environment and Development and
Grantham Institute for Climate Change, London, UK.
Patt, A., R.J.T. Klein, and A. de la Vega-Leinert, 2005: Taking the uncertainty in climate-
change vulnerability assessment seriously. Comptes Rendus Geoscience,
337(4), 411-424.
Patz, J.A., M.A. Mcgeehin, S.M. Bernard, K.L. Ebi, P.R. Epstein, A. Grambsch, D.J.
Gubler, P. Reither, I. Romieu, J.B. Rose, J.M. Samet, and J. Trtanj, 2000: The potential
health impacts of climate variability and change for the United States:
Executive Summary of the report of the health sector of the U.S. National
Assessment. Environmental Health Perspectives, 108(4), 367-376.
Patz, J.A., D. Campbell-Lendrum, T. Holloway, and J.A. Foley, 2005: Impact of regional
climate change on human health. Nature, 438(7066), 310-317.
Pelham, L., E. Clay, and T. Braunholz, 2011: Natural Disasters: What is the Role for
Social Safety Nets? World Bank, Washington, DC.
Pelling, M., 2003: The Vulnerability of Cities: Natural Disasters and Social Resilience.
Earthscan, London, UK.
Pelling, M. and A. Holloway, 2006: Legislation for Mainstreaming Disaster Risk
Reduction. Tearfund, Teddington, UK.
Persson, Å. and R.J.T. Klein, 2009: Mainstreaming adaptation to climate change into
official development assistance: challenges to foreign policy integration. In:
Climate Change and Foreign Policy: Case Studies from East to West [P. Harris
(ed.)]. Routledge, London, UK, pp. 162-177.
Phillips, C. and M. Marden, 2005: Landslide hazard and risk. In: Reforestation
Schemes to Manage Regional Landslide Risk [Glade, T., M. Anderson, and M.J.
Crozier (eds.)]. John Wiley and Sons Ltd, Sussex, UK, pp. 517-548.
Pielke, Jr., R.A., J. Rubiera, C. Landsea, M.L. Fernández, and R. Klein, 2003: Hurricane
vulnerability in Latin America and the Caribbean: Normalized damage and loss
potentials. Natural Hazards Review,
4, 101-114.
Pierro, R. and B. Desai, 2008: Climate insurance for the poor: Challenges for
targeting and participation. Institute of Development Studies Bulletin, 39, 4.
Polack, E., 2010: Integrating Climate Change into Regional Disaster Risk
Management at the Mekong River Commission. Strengthening Climate
Resilience Discussion Paper 4, Institute of Development Studies, Brighton, UK.
Polackova Brixi, H. and A. Mody, 2002: Dealing with government fiscal risk: An
overview. In: Government at Risk [Polackova Brixi, H. and A. Schick (eds.)].
World Bank, Washington, DC.
Pollner, J., 2000: Managing Catastrophic Risks Using Alternative Risk Financing and
Insurance Pooling Mechanisms. World Bank, Washington, DC.
Prabhakar, S.V.R.K., A. Srinivasan, and R. Shaw, 2009: Climate change and local
level disaster risk reduction planning: Need, opportunities and challenges.
Mitigation and Adaptation Strategies for Global Change, 14(1), 7-33.
Practical Action Bangladesh, 2008: Risk reduction boosts livelihood security in
disaster-prone District. In: Linking Disaster Risk Reduction and Poverty: Good
Practices and Lessons Learnt [United Nations International Strategy for Disaster
Reduction (UNISDR) (ed.)]. UNISDR, Geneva, Switzerland, pp. 1-5.
Preston, B.L., R.M. Westaway, and E. Yuen, 2011: Climate adaptation planning in
practice: an evaluation of adaptation plans from three developed nations.
Mitigation and Adaptation Strategies for Global Change, 16(4), 407-438.
Priest, G.L., 1996. The government, the market, and the problem of catastrophic loss.
Journal of Risk and Uncertainty, 12(2), 219-237.
ProAct Network, 2008: Environmental Management, Multiple Disaster Risk
Reduction and Climate Change Adaptation Benefits for Vulnerable
Communities. ProAct Network, Switzerland.
ProVention, 2009: Cities and resilience. In: Climate Policy Brief from Cities and
Resilience Dialogue. Bangkok, Thailand, 28-29 September 2009.
Pulwarty, R.S., and K. Redmond, 1997: Climate and salmon restoration in the
Columbia River basin: The role and usability of seasonal forecasts. Bulletin of
the American Meteorological Society,78(3), 381-397.
Pulwarty, R.S., K. Broad, and T. Finan, 2004: El Niño events, forecasts and decision-
making. In: Mapping Vulnerability: Disasters, Development, and People
[Bankoff, G., G. Frerks, and T. Hilhorst (eds.)]. Earthscan, London, UK, pp. 83-98.
Raddatz, C., 2007: Are external shocks responsible for the instability of output in
low-income countries? Journal of Development Economics, 84, 155-187.
Rahaman, M. and O. Varis, 2005: Integrated water resources management:
Evolution, prospects and future challenges. Sustainability: Science, Practice and
Policy, 1(1), 15-21.
Rahel, F.J. and J.D. Olden, 2008: Assessing the effects of climate change on aquatic
invasive species. Conservation Biology, 22(3), 521-533.
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
389
Ramalingam, B., H. Jones, T. Reba, and J. Youngs, 2008: Exploring the Science of
Complexity: Ideas and Implications for International Development and
Humanitarian Efforts. ODI Working Paper 285, Overseas Development Institute,
London, UK.
Ramalingam, B., K. Scriven, and C. Foley, 2009: Innovations in International
Humanitarian Action. In: ALNAP 8th Review of Humanitarian Action. Active
Learning Network for Accountability and Performance in Humanitarian Action
(ALNAP), London, UK.
Ranger, N., A. Millner, S. Dietz, S. Fankhauser, A. Lopez, and G. Ruta, 2010:
Adaptation in the UK: a Decision Making Process. Policy Brief, Grantham
Research Institute on Climate Change and the Environment and Centre for
Climate change Economics and Policy, London, UK.
Ranger, N. and S.-L. Garbett-Shiels, 2011: How can Decision-Makers in Developing
Countries Incorporate Uncertainty about Future Climate Risks into Existing
Planning and Policymaking Processes? Policy Paper, Centre for Climate Change
Economics and Policy and Grantham Research Institute on Climate Change and
the Environment in collaboration with the World Resources Report, World
Resources Institute, Washington, DC, www.worldresourcesreport.org/files/wrr/
papers/wrr_ranger_uncertainty.pdf.
Rayner, S., D. Lach, and H. Ingram, 2005: Weather forecasts are for wimps: why
water resource managers do not use climate forecasts. Climatic Change,
69(31), 197-227.
Reid, H. and S. Huq, 2005: Tropical forests and adaptation to climate change: In
search of synergies. In: Climate Change – Biodiversity and Livelihood Impacts
[Robledo, C., M. Kanninen, and L. Pedroni, (eds.)]. Center for International
Forestry Research, Indonesia.
Reid, H., D. Dodman, R. Janssen, and S. Huq, 2010: Building capacity to cope with
climate change in the least developed countries. In: Changing Climates, Earth
Systems and Society [Dodson, J. (ed.)]. International Year of Planet Earth,
Springer, the Netherlands.
Ribeiro, M., C. Losenno, T. Dworak, E. Massey, R. Swart, M. Benzie, and C. Laaser,
2009: Design of Guidelines for the Elaboration of Regional Climate Change
Adaptations Strategies. Study for the European Commission, Ecologic Institute,
Vienna, Austria.
Robledo, C., M. Kanninen, and L. Pedroni (eds.), 2005: Tropical Forests and
Adaptation to Climate Change: In Search of Synergies. Center for International
Forestry Research, Indonesia.
Roeth, H., 2009: The Development of a Public Partnership Framework and Action
Plan for Disaster Risk Reduction in Asia. United Nations International Strategy
for Disaster Reduction, Bangkok, Thailand.
Rogers, D., and V. Tsirkunov, 2010: Costs and Benefits of Early Warning Systems.
Global Assessment Report on Disaster Risk Reduction. United Nations
International Strategy for Disaster Reduction and World Bank, Geneva,
Switzerland and Washington, DC.
Romieu, E., T. Welle, S. Schneiderbauer, M. Pelling, and C. Vinchon, 2010: Vulnerability
assessment within climate change and natural hazard contexts: revealing gaps
and synergies through coastal applications. Sustainability Science, 5, 159-170.
Rosenzweig, C. and F.N. Tubiello, 2007: Adaptation and mitigation strategies in
agriculture: An analysis of potential synergies. Mitigation and Adaptation
Strategies for Global Change, 12(5), 855-873.
Rossetto, T., 2007: Construction Design, Building Standards and Site Selection: Tools
for Mainstreaming Disaster Risk Reduction. ProVention Consortium, Geneva,
Switzerland.
Ruth, M., and D. Coelho, 2007: Understanding and managing the complexity of
urban systems under climate change. Climate Policy, 7, 317-336.
Sabatier, P.A., 1986: Top-down and bottom-up approaches to implementation
research: A critical analysis and suggested synthesis. Journal of Public Policy,
6(1), 21-48.
Salter, J., 1998. Risk management in the emergency management context.
Australian Journal of Emergency Management, 13, 22-28.
Sanders, C.H., and M.C. Phillipson, 2003: UK adaptation strategy and technical
measures: The impacts of climate change on buildings. Building Research and
Information, 31(3-4), 210-221.
Sanderson, I., 2000: Cities, disasters and livelihoods. Environment and Urbanization,
12(2), 93-102.
Satterthwaite, D., 2007. Adaptation Options for Infrastructure In Developing
Countries. A report to the UNFCCC Financial and Technical Support Division.
United Nations Framework Convention on Climate Change, Bonn, Germany.
Satterthwaite, D., S. Huq, M. Pelling, H. Reid, and P.R. Lankao, 2007: Adapting to
Climate Change in Urban Areas. International Institute for Environment and
Development, London, UK.
SCBD, 2009: Connecting Biodiversity and Climate Change Mitigation and Adaptation.
Report of the second ad hoc technical expert group on biodiversity and climate
change, Secretariat of the Convention on Biological Diversity, Montreal, Canada.
Schipper, L., 2009: Meeting at the crossroads? Exploring the linkages between climate
change adaptation and disaster risk reduction. Climate and Development, 1, 16-30.
Schipper, L., and M. Pelling, 2006: Disaster risk, climate change and international
development: Scope and challenges for integration. Disasters, 30, 19-38.
Schwarze, R. and G. Wagner, 2004. In the aftermath of Dresden: New directions in
German flood insurance.The Geneva Papers on Risk and Insurance, 29(2), 154-168.
Scott, Z. and M. Tarazona, 2011: Decentralization and Disaster Risk Reduction. Study
on disaster risk reduction, decentralization and political economy analysis for
UNDP contribution to the 2011 Global Assessment Report on Disaster Risk
Reduction, UNISDR, Geneva, Switzerland.
SDC, 2008: Natural disaster risk management in Muminabad. SDC Project Fact Sheet
CARITAS, Swiss Development Cooperation, Switzerland.
SEEDS India, 2008: Reducing risk in poor urban areas to protect shelters, hard-won
assets and livelihoods. In: Linking Disaster Risk Reduction and Poverty: Good
Practices and Lessons Learnt [United Nations International Strategy for Disaster
Reduction (UNISDR) (ed.)]. UNISDR, Geneva, Switzerland, pp. 20-22.
Sharma, A., 2009: Progress Review of National Platforms for DRR in the Asia and
Pacific Region. United Nations International Strategy for Disaster Reduction,
Bangkok, Thailand.
Shaw, R., Y. Takeuchi, and B. Rouhban, 2009: Education, capacity building and public
awareness for disaster reduction. In: Landslides – Disaster Risk Reduction
[Sassa, P. (ed.)]. Springer-Verlag, Berlin, Germany, pp. 499-515.
Shepherd, G., 2004: The Ecosystem Approach: Five Steps to Implementation. World
Conservation Union (IUCN), Gland, Switzerland.
Shepherd, G. (ed.), 2008: The Ecosystem Approach: Learning from Experience. World
Conservation Union (IUCN), Gland, Switzerland.
Shugart
, H., R. Sedjo, and B. Sohngen, 2003: Forests and Global Climate Change:
Potential Impacts on U.S. Forest Resources. Pew Center on Global Climate
Change, Arlington, VA.
Sidle, R.C., A.J. Pearce, C.L. O’Loughlin, and A.G. Union, 1985: Hillslope Stability and
Land Use. American Geophysical Union, Washington, DC.
Skees, J., P. Varangis, D.F. Larson, and P. Siegel, 2005: Can financial markets be tapped
to help poor people cope with weather risks? In: Insurance Against Poverty
[Dercon, S. (ed.)]. Oxford University Press, Oxford, UK, pp. 422-437.
Skidmore, M. and H. Toya, 2002: Do natural disasters promote long run growth?
Economic Inquiry, 40(4), 664-687.
Skoufias, E., 2003: Economic crises and natural disasters: Coping strategies and
policy implications. World Development, 13(7), 1087-1102.
Slim, H., 2006: Global welfare: A realistic expectation for the international
humanitarian system? In: ALNAP Review of Humanitarian Action: Evaluation
Utilisation [Active Learning Network for Accountability and Performance in
Humanitarian Action (ed.)]. Overseas Development Institute, London, UK,
www.alnap.org/pool/files/ch1-f1.pdf.
Smillie, I., 2001: Patronage or Partnership: Local capacity building in humanitarian
crises. Kumarian Press , Bloomfield, CT.
Smit, B., O. Pilifosova, I. Burton, B. Challenger, S. Huq, R. Klein, and G. Yohe, 2001:
Adaptation to climate change in the context of sustainable development and
equity. In: Climate Change 2007. Impacts, Adaptation and Vulnerability.
Contribution of Working Group II to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P.
Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 879-906.
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
390
Smith, S.L., V.E. Romanovsky, A.G. Lewkowicz, C.R. Burn, M. Allard, G.D. Clow, K.
Yoshikawa, and J. Throop, 2010: Thermal state of permafrost in North America:
A contribution to the International Polar Year. Permafrost and Periglacial
Processes, 21, 117-135.
Sorensen, J., 1997: National and international efforts at Integrated Coastal
Management: Definitions, achievements, and lessons. Coastal Management,
25, 3-41.
Soussan, J. and I. Burton, 2002: Adapt and thrive: Combining adaptation to climate
change, disaster mitigation, and natural resources management in a new
approach to the reduction of vulnerability and poverty. In: UNDP Expert Group
Meeting, Integrating Disaster Reduction and Adaptation to Climate Change,
Havana, Cuba, 17-19 June 2002. United Nations Development Program,
Havana, Cuba.
Spence, R., 2004: Risk and regulation: Can improved government action reduce
the impacts of natural disasters? Building Research and Information, 32(5),
391-402.
Sperling, F. and F. Szekely, 2005: Disaster risk management in a changing climate,
vulnerability and adaptation. Discussion paper prepared for the world conference
on disaster reduction – reprint with addendum on conference outcomes.
Vulnerability and Adaptation Resource Group, Washington, DC.
Spittlehouse, D.L. and R.B. Stewart, 2003: Adapting to climate change in forest
management. Journal of Ecosystems and Management, 4, 7-17.
Stainforth, D.A., T.E. Downing, R. Washington, A. Lopez and M. New, 2007: Issues in
the interpretation of climate model ensembles to inform decisions.
Philosophical Transactions of the Royal Society A, 365, 2163-2177.
Stern, N.H., 2007: The Stern Review: The Economics of Climate Change. Cambridge
University Press, Cambridge, UK.
Stevens, L., 2008: Assessment of Impacts of Climate Change on Australia’s Physical
Infrastructure. Academy of Technological Sciences and Engineering, Parksville,
Australia.
Stewart, M.G., X. Wang, and M.N. Nguyen, 2011: Climate change impact and risks
of concrete infrastructure deterioration. Engineering Structures, 33, 1326-1337.
Sudmeier-Rieux, K. and N. Ash, 2009: Environmental Guidance Note for Disaster
Risk Reduction. International Union for Conservation of Nature (IUCN), Gland,
Switzerland.
Swiss Re, 2000: Storm over Europe. An underestimated risk. Swiss Reinsurance
Company, Zurich, Switzerland.
Tanner, T.M. and T. Mitchell (eds.), 2008: Poverty in a changing climate. Institute of
Development Studies (IDS) Bulletin, 39, 4.
Tearfund, 2008: Livelihood initiatives helps poor women build community resilience.
In: Linking Disaster Risk Reduction and Poverty: Good Practices and Lessons
Learnt [United Nations International Strategy for Disaster Reduction (UNISDR)
(ed.)]. UNISDR, Geneva, Switzerland, pp. 45-50
Tearfund, 2011: Disasters and the Local Church: Guidelines for church leaders in
disaster-prone areas. Tearfund, Teddington, UK.
TEEB, 2009. The Economics of Ecosystems and Biodiversity for National and
International Policy Makers Summary: Responding to the Value of Nature. The
Economics of Ecosystems and Biodiversity, Geneva, Switzerland, teebweb.org/
ForPolicymakers/tabid/1019/Default.aspx.
Telford, J. and J. Cosgrave, 2006: Joint Evaluation of the International Response to
the Indian Ocean Tsunami: Synthesis Report. Tsunami Evaluation Coalition,
London, UK.
Thomalla, F., T. Downing, E. Spanger-Siegfried, G. Han, and J. Rockstrom, 2006:
Reducing hazard vulnerability: towards a common approach between disaster
risk reduction and climate adaptation. Disasters, 30(1), 39-48.
Thompson, I., B. Mackey, S. McNulty, and A. Mosseler, 2009: Forest Resilience,
Biodiversity, and Climate Change. A synthesis of the biodiversity/resilience/
stability relationship in forest ecosystems. Technical Series no. 43, Secretariat of
the Convention on Biological Diversity, Montreal, Canada.
Thompson, M. and I. Gaviria, 2004: Weathering the storm: Lessons in risk reduction
from Cuba. Oxfam, Boston, MA.
Thorne, C., E. Evans, and E. Penning-Rowsell (eds.), 2006: Future Flooding And
Coastal Erosion Risks. Thomas Telford, London, UK.
Tompkins, E.L., 2005: Planning for climate change in small islands: Insights from
national hurricane preparedness in the Cayman Islands. Global Environmental
Change Part A, 15(2), 139-149.
Tompkins, E.L., M.C. Lemos, and E. Boyd, 2008: A less disastrous disaster: Managing
response to climate-driven hazards in the Cayman Islands and NE Brazil. Global
Environmental Change, 18, 736-745.
Tompkins, E.L., W.N. Adger, E. Boyd, S. Nicholson-Cole, K. Weatherhead, and N.
Arnell, 2010: Observed adaptation to climate change: UK evidence of transition
to a well-adapting society. Global Environmental Change, 20, 627-635.
Tschoegl, L., R. Below, and D. Guha-Sapir, 2006: An Analytical Review of Selected
Data Sets on Natural Disasters and Impacts. Centre for Research on the
Epidemiology of Disasters, United Nations Development Programme, Brussels,
Belgium.
Tsing, A.L., J.P. Brosius, and C. Zerner, 1999: Assessing community-based natural-
resource management. Ambio, 28(2), 197-198.
Twigg, J., 2004: Project planning. In: Disaster Risk Reduction: Mitigation and
Preparedness in Development and Emergency Programming. Humanitarian
Practice Network, Overseas Development Institute, London, UK, pp. 31-60.
UNDP, 2002: A climate risk management approach to disaster reduction and
adaptation to climate change. In: Proceedings of UNDP Expert Group Meeting,
Havana, Cuba, 19-21 June 2002. United Nations Development Programme,
New York, NY.
UNDP, 2004a: Reducing Disaster Risk: A Challenge for Development. United Nations
Development Programme, New York, NY.
UNDP, 2004b: A Global Review: UNDP Support for Institutional and Legislative
Systems for Disaster Risk Management: Executive Summary. United Nations
Development Programme, New York, NY.
UNDP, 2004c: Adaptation Policy Frameworks for Climate Change. Developing
strategies, policies and measures
. Cambridge University Press, Cambridge, UK.
UNDP, 2009: Indonesia: Institutional and Legal Systems for Early Warning and
Disaster Risk Reduction. Regional Programme on Capacity Building for
Sustainable Recovery and Risk Reduction, United Nations Development
Programme, New York, NY.
UNEP, 2006: In the Front Line: Shoreline Protection and Other Ecosystem Services
from Mangroves And Coral Reefs. United Nations Environment Programme
World Conservation Monitoring Centre, Cambridge, UK.
UNEP, 2009. The Role of Ecosystem Management in Climate Change Adaptation and
Disaster Risk Reduction. Issues paper prepared for the global platform for
Disaster Risk Reduction. United Nations Environment Programme, Geneva,
Switzerland.
UNEP, 2010: Using Ecosystems to Address Climate Change – Ecosystem Based
Adaptation. UNEP Regional Seas Information Series, United Nations
Environment Programme, Nairobi, Kenya.
UNESCAP, 2009: Case Study: The National Disaster Management System Of China
And Its Response To The Wenchuan Earthquake. United Nations Economic and
Social Commission for Asia and the Pacific, Bangkok, Thailand,
www.unescap.org/Idd/Events/Cdrr-2009/Cdr_2e.Pdf.
UNFCCC, 2006a: Application of Environmentally Sound Technologies for Adaptation
to Climate Change. United Nations Framework Convention on Climate Change,
Bonn, Germany.
UNFCCC, 2006b. Technologies for Adaptation to Climate Change. United Nations
Framework Convention on Climate Change, Bonn, Germany.
UNFCCC, 2007a: Investment and Financial Flows to Address Climate Change. United
Nations Framework Convention on Climate Change, Bonn, Germany.
UNFCCC, 2007b: National Adaptation Programmes of Action (NAPAs). United
Nations Framework Convention on Climate Change, Bonn, Germany.
UNFCCC, 2008a: Disaster Risk Reduction Strategies and Risk Management Practices:
Critical Elements for Adaptation to Climate Change. United Nations Framework
Convention on Climate Change, Bonn, Germany.
UNFCCC, 2008b: Integrating Practices, Tools and Systems for Climate Risk
Assessment and Management and Strategies for Disaster Risk Reduction into
National Policies and Programmes. United Nations Framework Convention on
Climate Change, Bonn, Germany.
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
391
UNFCCC, 2009a: Potential Costs and Benefits of Adaptation Options: A Review of
Existing Literature. Technical Paper, United Nations Framework Convention on
Climate Change, Bonn, Germany.
UNFCCC, 2011a: NAPA Priorities Database. United Nations Framework Convention
on Climate Change, Bonn, Germany, unfccc.int/cooperation_support/least_
developed_countries_portal/napa_priorities_database/items/4583.php.
UNFCCC, 2011b: Frequently Asked Questions about LDCs, the LEG and NAPAs.
United Nations Framework Convention on Climate Change, Bonn, Germany,
unfccc.int/cooperation_support/least_developed_countries_portal/frequently
_asked_questions/items/4743.php.
UNISDR, 2004: Living with Risk: A Global Review of Disaster Reduction Initiative.
United Nations International Strategy for Disaster Reduction, Geneva, Switzerland.
UNISDR, 2005: Hyogo Framework for Action 2005–2015: Building the resilience of
nations and communities to disasters. United Nations International Strategy for
Disaster Reduction, Geneva, Switzerland.
UNISDR, 2006: Developing Early Warning Systems: A checklist. The conclusions of
the third international conference on early warning. Germany, Bonn, 27-29.
United Nations International Strategy for Disaster Reduction, Geneva, Switzerland.
UNISDR, 2007a: Building Disaster Resilient Communities: Good Practices and
Lessons Learned. United Nations International Strategy for Disaster Reduction,
Geneva, Switzerland.
UNISDR, 2007b: Words into Action: A Guide for Implementing the Hyogo Framework.
United Nations International Strategy for Disaster Reduction, Geneva, Switzerland.
UNISDR, 2007c: Acting with Common Purpose’. Proceedings of the first session of
the Global Platform on Disaster Risk Reduction. Geneva, 5-7 June 2007, United
Nations International Strategy for Disaster Reduction, Geneva, Switzerland.
UNISDR, 2008a: Linking Disaster Risk Reduction and Poverty Reduction: Good
Practices and Lessons Learnt. United Nations International Strategy for Disaster
Reduction, Geneva, Switzerland.
UNISDR, 2008b: Towards National Resilience: Good practices of National Platforms
for Disaster Risk Reduction. United Nations International Strategy for Disaster
Reduction, Geneva, Switzerland.
UNISDR, 2009a: Adaptation to Climate Change by Reducing Disaster Risks: Country
Practices and Lessons. United Nations International Strategy for Disaster
Reduction, Geneva, Switzerland.
UNISDR, 2009b: Applying Disaster Risk Reduction for Climate Change Adaptation:
Country Practices and Lessons. United Nations International Strategy for
Disaster Reduction, Geneva, Switzerland.
UNISDR, 2009c: Risk and Poverty in a Changing Climate: Global Assessment Report
on Disaster Risk Reduction. United Nations International Strategy for Disaster
Reduction, Geneva, Switzerland.
UNISDR, 2009d: Recommendations of National Platforms to the Chair and
Participants of the Second Session of the Global Platform for Disaster Risk
Reduction. United Nations International Strategy for Disaster Reduction,
Geneva, Switzerland.
UNISDR, 2010: Early Warning Practices can Save Many Lives: Good Practices and
Lessons Learned. United Nations International Strategy for Disaster Reduction,
Geneva, Switzerland.
UNISDR, 2011a: 2011 Global Assessment Report on Disaster Risk Reduction:
Revealing Risk, Redefining Development. United Nations International Strategy
for Disaster Reduction, Geneva, Switzerland.
UNISDR
, 2011b: Hyogo Framework for Action 2005-2015: Mid Term Review. United
Nations International Strategy for Disaster Reduction, Geneva, Switzerland.
UNISDR/DKKV/Council of Europe, 2008: Disaster risk reduction in Europe:
Overview of European National Platforms, Hyogo Framework for Action focal
points and regional organisations and institutions. United Nations International
Strategy for Disaster Reduction, German Committee for Disaster Risk
Reduction, and Council of Europe.
Urwin, K. and A. Jordan, 2008: Does public policy support or undermine climate
change adaptation? Exploring policy interplay across different scales of
governance. Global Environmental Change, 18(1), 180-191.
Van Buskirk, R., 2006: Analysis of long-range clean energy investment scenarios for
Eritrea, East Africa. Energy Policy, 34(14), 1807-1817.
Varangis, P., J.R. Skees, and B.J. Barnett, 2002: Weather indexes for developing
countries. In: Climate Risk and the Weather Market [Dischel, B. (ed.)]. Risk
Books, London, UK, pp. 279-294.
Venter, O., W.F. Laurance, T. Iwamura, K.A. Wilson, R.A. Fuller, and H.P. Possingham,
2009: Harnessing carbon payments to protect biodiversity. Science, 326(5958),
1368.
Venton, P. and S. La Trobe, 2008: Linking Climate Change Adaptation and Disaster
Risk Reduction. Tearfund, Tweddington, UK.
Verdin, J., C. Funk, G. Senay, and R. Choularton, 2005: Climate science and
famine early warning. Philosophical Transactions of the Royal Society B, 360,
2155-2168.
Vignola, R., B. Locatelli, C. Martinez, and P. Imbach, 2009: Ecosystem-based adaptation
to climate change: What role for policy-makers, society and scientists?
Mitigation and Adaptation Strategies for Global Change, 8(14), 691-696.
Virtanin, P., E. Palmujpki, and D. Gemechu, 2011: Global climate policies, local
institutions and food security in a pastoral society in Ethiopia. Consilience: The
Journal of Sustainable Development, 5, 96-118.
Vogel, C., and K. O’Brien, 2006: Who can eat information? Examining the effectiveness
of seasonal climate forecasts and regional climate-risk management strategies.
Climate Research, 33, 111-122.
Walters, C., 1997. Challenges in adaptive management of riparian and coastal
ecosystems. Conservation Ecology, 2(1), 1.
Wamsler, C., 2004: Managing urban risk: Perceptions of housing and planning as a
tool for reducing disaster risk. Global Built Environment Review, 4(2), 11-28.
Warhurst, A., 2006: Disaster Prevention: A Role for Business? Maplecroft and
ProVention Consortium, Geneva, Switzerland.
Warner, K., N. Ranger, S. Surminski, M. Arnold, J. Linnnerooth-Bayer, E. Michel-Kerjan,
P. Kovacs, and C. Herweijer, 2009: Adaptation to Climate Change: Linking
Disaster Risk Reduction and Insurance. United Nations International Strategy
for Disaster Reduction Secretariat, Geneva, Switzerland.
Weih, M., 2004: Intensive short rotation forestry in boreal climates: Present and
future perspectives. Canadian Journal of Forest Research, 34(7), 1369-1378.
White, P., M. Pelling, K. Sen, D. Seddon, S. Russell, and R. Few, 2004: Disaster Risk
Reduction: A Development Concern. A Scoping Study on Links between Disaster
Risk Reduction, Poverty and Development. Department for International
Development, London, UK.
WHO, 2003: Climate Change and Human Health - Risks and Responses: Summary.
World Health Organization, Geneva, Switzerland.
WHO, 2005: Health and Climate Change: The “Now and How”: A Policy Action Guide.
World Health Organization Regional Office for Europe, Copenhagen, Denmark.
Wickham, F., J. Kinch, and P.N. Lal, 2009: Institutional Capacity within Melanesian
Countries to Effectively Respond to Climate Change Impacts, with a Focus on
Vanuatu and the Solomon Islands. A report prepared for the Bishop Museum
and SPREP. Bishop Museum, Honolulu, Hawaii.
Wilby, R.L., 2007: A review of climate change impacts on the built environment. Built
Environment, 33(1), 31-45.
Wilby, R.L., 2009: Climate for Development in South Asia (ClimDev-SAsia): An
inventory of cooperative programmes and sources of climate risk information
to support robust adaptation. Report prepared on behalf of the UK Department
for International Development (DFID), DFID, London, UK.
Wilby, R.L. and S. Dessai, 2010: Robust adaptation to climate change. Weather,
65(7), 180-185.
Wilby, R.L., J. Troni, Y. Biot, L. Tedd, B.C. Hewitson, D.M. Smith, and R.T. Sutton, 2009:
A review of climate risk information for adaptation and development planning.
International Journal of Climatology, 29(9), 1193-1215.
Wilches-Chaux, G., 2008: ¿Qu-ENOS pasa? – Guía de LA RED para la gestión radical
de riesgos asociados con el fenómeno ENOS. LA RED, Bogotá, Colombia.
Wisner, B., 2011: Are we there yet? Reflections on integrated disaster risk management
after ten years. Journal of Integrated Disaster Risk Management, 1, 1-14.
Wisner, B. and J. Adams, 2002: Environmental Health in Emergencies and Disasters:
A Practical Guide. World Health Organization, Geneva, Switzerland.
Wisner, B., P. Blaikie, T. Cannon, and I. Davis, 2004: At Risk: Natural Hazards, People’s
Vulnerability And Disasters. 2nd ed. Routledge, London, UK.
Chapter 6 National Systems for Managing the Risks from Climate Extremes and Disasters
392
WMO, 2004: Report of the Expert Meeting on Meteorological Information for Locust
Control, 18-20 October 2004. World Meteorological Organization, Geneva,
Switzerland.
WMO, 2007: WMO to Provide Guidance for Heat Health Warning System. WMO Press
Release 781, World Meteorological Organization, Geneva, Switzerland.
Wood, A.P. and Van Halsema, G.E., 2008: Scoping agriculture, wetland interactions:
Towards a sustainable multiple-response strategy. FAO Water Reports. Rome:
Food and Agriculture Organisation (FAO).
World Bank, 1996. Argentina flood protection project. Staff Appraisal Report 15354,
World Bank, Washington, DC.
World Bank, 2000: Cities, Seas and Storms: Managing Change in Pacific Island
Economies. World Bank, Washington, DC.
World Bank, 2003: Protecting New Health Facilities from Natural Disasters:
Guidelines for the Promotion of Disaster Mitigation. World Bank, Washington,
DC.
World Bank, 2007: The Caribbean Catastrophe Risk Insurance Initiative. Results of
preparation work on the design of a Caribbean Catastrophe Risk Insurance
Facility, World Bank, Washington, DC.
World Bank, 2008: The Economics of Adaptation to Climate Change: Methodology
Report. World Bank, Washington, DC.
World Bank, 2009: Convenient Solutions to an Inconvenient Truth: Ecosystem-Based
Approaches to Climate Change. World Bank, Washington, DC.
World Bank, 2010: Mainstreaming Adaptation to Climate Change in Agriculture and
Natural Resources Management Projects. Washington D.C.: World Bank.
World Bank and UN, 2010: Natural Hazards, UnNatural Disasters: The Economics of
Effective Prevention. International Bank for Reconstruction and Development/
World Bank, Washington, DC, 254 pp., gfdrr.org/gfdrr/nhud-home.
World Vision, 2008: Reducing vulnerabilities and poverty through disaster mitigation.
In: Linking Disaster Risk Reduction and Poverty: Good Practices and Lessons
Learnt [United Nations International Strategy for Disaster Reduction (UNISDR)
(ed.)]. UNISDR, Geneva, Switzerland, pp. 15-19.
WSSD, 2002: Report of the World Summit on Sustainable Development,
Johannesburg, South Africa, 26 August – 4 September 2002.
WWAP, 2009: Water in a Changing World. The United Nations World Water
Development Report. World Water Assessment Programme, Earthscan, London,
UK.
WWC, 2009: Water Supply and Sanitation. World Water Council, France.
Yamin, F., A. Rahman, and S. Huq, 2005: Vulnerability, adaptation and climate disasters:
A conceptual overview. IDS Bulletin, 36(4), 1-14.
Yodmani, S., 2001: Disaster risk management and vulnerability reduction: Protecting
the poor. In: Proceedings of Asia and Pacific Forum on Poverty: Reforming
policies and institutions for poverty reduction. Manila, Philippines, 5-9 February
2001.
Younger, M., H.R. Morrow-Almeida, S.M. Vindigni, and A.L. Dannenberg, 2008: The
built environment, climate change, and health: Opportunities for co-benefits.
American Journal of Preventive Medicine, 35(5), 517-52.
Chapter 6National Systems for Managing the Risks from Climate Extremes and Disasters
393
Coordinating Lead Authors:
Ian Burton (Canada), O. Pauline Dube (Botswana)
Lead Authors:
Diarmid Campbell-Lendrum (Switzerland), Ian Davis (UK), Richard J.T. Klein (Sweden),
Joanne Linnerooth-Bayer (USA), Apurva Sanghi (USA), Ferenc Toth (Austria)
Review Editors:
Joy Jacqueline Pereira (Malaysia), Linda Sygna (Norway)
Contributing Authors:
Neil Adger (UK), Thea Dickinson (Canada), Kris Ebi (USA), Md. Tarik ul Islam (Canada / Bangladesh),
Clarisse Kehler Siebert (Sweden)
This chapter should be cited as:
Burton, I., O.P. Dube, D. Campbell-Lendrum, I. Davis, R.J.T. Klein, J. Linnerooth-Bayer, A. Sanghi, and F. Toth, 2012: Managing
the risks: international level and integration across scales. In: Managing the Risks of Extreme Events and Disasters to
Advance Climate Change Adaptation [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea,
K.J. Mach, G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups I and II of the
Intergovernmental Panel on Climate Change (IPCC). Cambridge University Press, Cambridge, UK, and New York, NY, USA,
pp. 393-435.
7
Managing the Risks:
International Level and
Integration across Scales
Managing the Risks: International Level and Integration across Scales
394
Executive Summary .................................................................................................................................396
7.1. The International Level of Risk Management..........................................................................398
7.1.1. Context and Background..................................................................................................................................................398
7.1.2. Related Questions and Chapter Structure .......................................................................................................................398
7.2. Rationale for International Action ...........................................................................................398
7.2.1. Systemic Risks and International Security.......................................................................................................................399
7.2.2. Economic Efficiency..........................................................................................................................................................399
7.2.3. Shared Responsibility.......................................................................................................................................................400
7.2.4. Subsidiarity ......................................................................................................................................................................401
7.2.5. Legal Obligations .............................................................................................................................................................401
7.2.5.1. Scope of International Law, Managing Risks, and Adaptation ...............................................
..........................................................401
7.2.5.2.
International Conventions ................................................................................................................................................................402
7.2.5.3. Customary Law and Soft Law Principles...........................................................................................................................................402
7.2.5.4. Non-Legally Binding Instruments......................................................................................................................................................402
7.3. Current International Governance and Institutions.................................................................403
7.3.1. The Hyogo Framework for Action ....................................................................................................................................403
7.3.1.1. Evolution and Description................................................................................................................................................
.................403
7.3.1.2. Status of Implementation .................................................................................................................................................................404
7.3.2. The United Nations Framework Convention on Climate Change...
....
.............................................................................406
7.3.2.1. Evolution and Description................................................................................................................................................
.................406
7.3.2.2. Status of Implementation .................................................................................................................................................................406
7.3.3. Current Actors ...
....
...........................................................................................................................................................408
7.3.3.1. International Coordination in Linking Disaster Risk Management and Climate Change Adaptation ...............................................408
7.3.3.2.
International Technical and Operational Support..............................................................................................................................409
7.3.3.3. International Finance Institutions and Donors..................................................................................................................................410
7.4. Options, Constraints, and Opportunities for Disaster Risk Management
and Climate Change Adaptation at the International Level....................................................411
7.4.1. International Law.............................................................................................................................................................411
7.4.1.1. Limits and Constraints of International Law.....................................................................................................................................411
7.4.1.2.
Opportunities for the Application of International Law....................................................................................................................412
7.4.2. International Finance ...
....
................................................................................................................................................412
7.4.3. Technology Transfer and Cooperation..............................................................................................................................414
7.4.3.1. Technology and Climate Change Adaptation...................................................
.................................................................................414
7.4.3.2.
Technologies for Extreme Events ......................................................................................................................................................416
7.4.3.3. Financing Technology Transfer ..........................................................................................................................................................417
7.4.4. Risk Sharing and Transfer...
....
..........................................................................................................................................418
7.4.4.1. International Risk Sharing and Transfer............................................................................................................................................418
7.4.4.2.
International Risk-Sharing and Transfer Mechanisms.......................................................................................................................418
7.4.4.3. Value Added by International Interventions .....................................................................................................................................420
7.4.5. Knowledge Acquisition, Management, and Dissemination ...
....
......................................................................................421
7.4.5.1. Knowledge Acquisition ................................................................................................................................................
.....................421
7.4.5.2. Knowledge Organization, Sharing, and Dissemination.....................................................................................................................422
Chapter 7
Table of Contents
395
7.5. Considerations for Future Policy and Research .......................................................................425
7.6. Integration across Scales .........................................................................................................426
7.6.1. The Status of Integration.................................................................................................................................................426
7.6.2. Integration at a Spatial Scale ..........................................................................................................................................427
7.6.3. Integration at a Temporal Scale.......................................................................................................................................427
7.6.4. Integration at a Functional Scale.....................................................................................................................................427
7.6.5. Toward More Integration .................................................................................................................................................427
References ...............................................................................................................................................428
Chapter 7 Managing the Risks: International Level and Integration across Scales
396
Chapter 7Managing the Risks: International Level and Integration across Scales
Increasing global interconnectivity, population, and economic growth, and the mutual interdependence of
economic and ecological systems, can serve both to reduce vulnerability and to amplify disaster risks (high
confidence). Global development pathways are becoming a more important factor in the management of
vulnerability and disaster risk. [7.2.1]
The international community has accumulated substantial experience in providing help for disasters and
risk management in the context of localized and short-term events associated with climate variability and
extremes. Experience in disaster risk management includes both bottom-up and top-down approaches, but most
often has developed from disasters considered first as local issues, then at the national level, and only at the
international level where needs exceed national capacity, especially in terms of humanitarian assistance and capacity
building. [7.2.4]
There are two main mechanisms at the international level that are purpose-built and dedicated to disaster
risk management and climate change adaptation. These are the United Nations International Strategy for
Disaster Reduction (UNISDR) and the United Nations Framework Convention on Climate Change (UNFCCC),
in particular in its adaptation components. This chapter focuses on these two bodies while recognizing that
there are many others that have an international role to play. Page limitations require a selective approach and a
comprehensive assessment of all relevant bodies is impractical. The UNISDR and the UNFCCC are very different
institutions with different mandates and scope and objectives, and with varying strengths and capacities (high
confidence). Up to the present this fact has made the integration of disaster risk management and climate change
adaptation difficult to achieve (medium confidence). [7.3] The evolution of disaster risk management has come
from various directions: from the top down where legislation has required safe practice at operational
levels and from the local level up to the national and international levels. The evolution of climate change
adaptation has been driven primarily by the recognition of the global issue of anthropogenic climate
change (high confidence). [7.3]
In addition to the UNISDR and the UNFCCC, other areas of international law and practice are being used
to address climate change adaptation and disaster risk management. The relationship between legal
aspirations and obligations in these areas of international action and management is complex and neither
is well understood or agreed upon (high confidence). Other areas include international refugee law, which has
been invoked to deal with the displacement of people that might be in part attributed to climate change; human
rights law as used by citizens against states for climate change impacting on the enjoyment of human rights; and the
attempts to expand existing legal doctrines such as the emerging ‘responsibility to protect’ doctrine to motive states
to act on climate change. Such attempts to use tools from other areas of international law to address climate change
adaptation and disaster risk reduction challenges have generally not been successful. [7.2.5]
International action on disaster risk reduction and climate change adaptation can be motivated both by
national interests and a concern for the common (global) public good. [7.2] The interdependence of the global
economy, the public good, and the transboundary nature of risk management, and the potential of regional risk pooling,
can make international cooperation on disaster risk reduction and climate change adaptation more economically efficient
than national or sub-national action alone. Notions of solidarity and equity motivate addressing disaster risk reduction
and climate change adaptation at the international level in part because developing countries are more vulnerable to
physical disasters. [7.2]
Closer integration at the international level of disaster risk reduction and climate change adaptation, and
the mainstreaming of both into international development and development assistance, could foster
efficiency in the use of available and committed resources and capacity (high confidence). [7.4] Neither
disaster risk reduction nor climate change adaptation is as well integrated as they could be into current development
policies and practices. Both climate change adaptation and disaster risk reduction might benefit from sharing of
Executive Summary
397
Chapter 7 Managing the Risks: International Level and Integration across Scales
knowledge and experience in a mutually supportive and synergistic way. Climate change adaptation could be factored
into all disaster risk management, and weather-related disasters are becoming an essential component of the adaptation
agenda. [7.4]
Opportunities exist to create synergies in international finance for disaster risk management and adaptation
to climate change, but these have not yet been fully realized (high confidence). International funding for
disaster risk reduction remains relatively low as compared to the scale of spending on international humanitarian
response. [7.4.2] Governments have committed to mobilize greater amounts of funding for climate change adaptation
and this may also help to support the longer-term investments necessary for disaster risk reduction. [7.4.2]
Expanded international financial support for climate change adaptation as specified in the Cancun
Agreements of 2010 and the Climate Change Green Fund will facilitate and strengthen disaster risk
management (medium confidence). The agreements to provide substantial additional finance at the international
level for adaptation to climate change have been formulated to include climate- and weather-related disaster risk
reduction. There is therefore some prospect that projects and planning for disaster risk reduction and climate change
adaptation can increasingly be combined and integrated at the national level (high confidence). [7.3.2.2, 7.4]
Technology transfer and cooperation under the United Nations Framework Convention on Climate Change
has until recently focused more on the reduction of greenhouse gas emissions than on adaptation (high
confidence). Technology for disaster risk management, especially to advance and strengthen forecasting and warning
systems and emergency response, is promoted through the Hyogo Framework for Action (HFA), but is widely dispersed
among many international and national-level organizations and is not closely linked to the UNISDR. Technology transfer
and cooperation to advance disaster risk reduction and climate change adaptation are important. Coordination on
technology transfer and cooperation between these two fields has been lacking, which has led to fragmented
implementation (high confidence). [7.4]
International financial institutions, bilateral donors, and other international actors have played a catalytic role in the
development of catastrophic risk transfer and other risk-sharing instruments in the more vulnerable countries.
Stronger products and methods for risk sharing and risk transfer are being developed as a relatively new
and expanding area of international cooperation to help achieve climate change adaptation and disaster
risk reduction (high confidence). [7.4] Established mechanisms include remittances, post-disaster credit, and
insurance and reinsurance. Partly in response to concerns about climate change, additional insurance instruments are
in various stages of development and expansion including international risk pools and weather index micro-insurance.
These processes and products are being developed by international financial institutions as well as by nongovernmental
organizations and the private sector. [7.4.4.2]
One lesson from disaster risk reduction and climate change adaptation is that stronger efforts at the
international level do not necessarily lead to substantive and rapid results on the ground and at the local
level. There is room for improved integration across scales from international to local (high confidence).
[7.6] The expansion of disaster risk reduction through the International Decade for Natural Disaster Reduction
(1990-1999), and the establishment of the UNISDR and the creation and adoption of the HFA have had results that
are difficult to specify or to quantify – but which may have contributed to some reduction in morbidity and mortality,
while enjoying much less success in the area of economic and property losses. The problems of disaster risk have
continued to grow due in large part to the relentless expansion in exposure and vulnerability even as the international
management capacity has expanded (medium confidence). [7.5, 7.6]
398
Chapter 7Managing the Risks: International Level and Integration across Scales
7.1. The International Level
of Risk Management
7.1.1. Context and Background
A need to cope with the risks associated with atmospheric processes
(floods, droughts, cyclones, and so forth) has always been a fact of human
life (Lamb, 1995). In more recent decades, extreme weather events have
increasingly come to be associated with large-scale disasters and an
increasing level of economic losses (Chapters 2 and 4). Considerable
experience has accumulated at the international (as well as local and
national) level on ways of coping with or managing the risks.
The same cannot be said for the risks associated with anthropogenic
climate change. These are new risks identified as possibilities or
probabilities (IPCC 1990, 1996, 2007).
Acceptance of climate change and its growing impacts has led to a
stronger emphasis on the need for adaptation, as exemplified, for
example, in the Bali Action Plan (adopted at the 13th Session of the
Conference of the Parties to the UNFCCC (UNFCCC, 2007a) and the
Cancun Agreements of December 2010.
The international community is thus faced with a contrast between a
long record of managing disasters and the risks of ‘normal’ climate
extremes, and the new problem of adaptation to anthropogenic climate
change and its associated changes in variability and extremes. It has
been asked how the comparatively new field of anthropogenic climate
change adaptation (CCA) can benefit from the longer experience in
disaster risk management (DRM). That question is a major focus of this
Special Report.
Climate extremes can have both negative and positive effects. The
occurrence of extreme events has raised consciousness of climate change
within the public and in policymakers. This can then help to enhance a
sense of priority to governmental action in terms of supporting DRM,
enhancing adaptation, and promoting mitigation (Adger et al., 2005).
An international framework for integration of climate-related DRM and
CCA in the development process could provide the potential for reducing
exposure and vulnerability (Thomalla et al., 2006; Venton and La Trobe,
2008). Collective efforts at the international level to reduce greenhouse
gases are a way to reduce long-term exposure to frequent and more
intense climate extremes. International frameworks designed to facilitate
adaptation with a deliberate effort to address issues of equity, technology
transfer, globalization, and the need to meet the Millennium Development
Goals (MDGs) can, when combined with mitigation, lead to reduced
vulnerability (Adger et al., 2005; Haines et al., 2006). The 2007/2008 Human
Development Report noted that if climate change is not adequately
addressed now, 40% of the world’s poorest (i.e., 2.6 billion people) will
be confined to a future of diminished opportunity (Stern, 2007; Watkins,
2007). The long-term potential to reducing exposure to climate risks lies
in sustainable development (O’Brien et al., 2008). Both seek to build
resilience through sustainable development (O’Brien et al., 2008).
Some claim that DRM and CCA could be realized through increased
awareness and use of synergies and differences, and by the provision of
a framework for integration in areas of overlap between the two
(Venton and La Trobe, 2008). The World Conference on Disaster
Reduction held in Kobe (UNISDR, 2005c), Hyogo Prefecture, Japan in
2005 and the Bali Action Plan both point to the need for incorporation
of measures that can reduce climate change impacts within the practice
of disaster risk reduction (DRR). Integration of the relevant aspects of
DRR and CCA can be facilitated by using the Hyogo Framework for
Action (2005-2015) as agreed by 168 governments in Kobe (UNISDR,
2005a).
7.1.2. Related Questions and Chapter Structure
Within the context of the overarching question – how can experience
with disaster risk management inform and help with climate change
adaptation? – there are a series of other related issues to be addressed
in this chapter in order to provide a basis for their closer integration. A
first question concerns the rationale for disaster risk management and
climate change adaptation at the international level. The issues of
systemic risks and international security, economic efficiency, solidarity,
and subsidiarity are addressed in Section 7.2.
A second topic concerns the nature and development of institutions and
capacity at the international level. This topic is explored in Section 7.3
concentrating on the Hyogo Framework for Action and the United
Nations Framework Convention on Climate Change.
A third issue concerns the opportunities for and constraints on disaster
risk management and climate change adaptation at the international
level. These include the matters of legal, financial, technology, risk
transfer, and cooperation, and the creation of knowledge and its
management and dissemination. All are addressed in Section 7.4.
Considerations of future policy and research are addressed in Section
7.5.
The challenge of bringing lessons from disaster risk reduction to climate
change adaptation takes on a different complexion at different temporal
and spatial scales. The question of integration across scales is taken up
in Section 7.6.
7.2. Rationale for International Action
This section provides a brief overview of selected concepts and principles
that have been invoked to justify (or restrain) financing, assistance,
regulation, and other types of international policy interventions for
disaster risk management and climate change adaptation. There is no
attempt to be comprehensive, and additional principles are discussed in
Section 7.2.5. Starting from the reality that risks of extreme weather and
risk management interventions cross national borders and transcend
399
single nation policies and procedures, this section discusses the
systemic nature of these risks and their effects on international security
before turning to a discussion of efficiency, shared responsibility, and
subsidiarity as these principles have shaped international discourse,
practices, and legal obligations within existing frameworks and
conventions.
7.2.1. Systemic Risks and International Security
The term ‘systemic risk’ refers to risks that are characterized by linkages
and interdependencies in a system, where the failure of a single entity
or cluster of entities can cause cascading impacts on other interlinked
entities. Because of greatly increased international interdependency,
shocks occurring in one country can potentially have major and bi-
directional systemic impacts on other parts of the world (Kleindorfer,
2009), although the full extent of these impacts is not well documented.
Moreover, major interlinked events, such as melting of glaciers, will bring
increased levels of hazard to specific areas, and the initial impacts of
such changes can extend to second- and third-order impacts (Alexander,
2006). This can apply to the contiguous zones of many countries, such as
shared basins with associated flood risks, which calls for transboundary,
international mechanisms (Linnerooth-Bayer et al., 2001).
Relationships and connections involving the movement of goods
(trade), finance (capital flows and remittances), and people (displaced
populations) can also have transboundary impacts as discussed below.
Moreover, actions in one country impact another, for example, clearing
forests in an upstream riparian country can increase flood risks
downstream. Chastened by the unexpected systemic cascading of the
2007-2008 financial crisis, firms with global supply chains are now
devoting significant resources to crisis management and disruption risk
management (Sheffi, 2005; Harrington and O’Connor, 2009).
A few examples can illustrate the cascading nature of the financial and
economic impacts from disaster. Due to Hurricane Katrina in 2005, the
International Energy Agency announced a coordinated drawdown of
European and Asian oil stocks totaling 60 million barrels (Bamberger
and Kumins, 2005), and reportedly oil prices rose not only in the United
States but also as far away as Canada and the United Kingdom.
Disasters also have an impact on international trade. Using a gravity
model across 170 countries (1962-2005), Gassebner et al. (2010)
conclude that an additional disaster reduces imports on average by
0.2% and exports by 0.1%. The main conditions determining the impact
of disastrous events on trade are the level of democracy and the
geographical size of the affected country.
Turning specifically to displaced persons as a cascading impact,
estimates of the numbers of current and future migrants due not only
to disasters but generally to environmental change are divergent and
controversial (Myers, 2001; Christian Aid, 2007). A middle-range
estimate puts the figure at 200 million by 2050 (Brown, 2008). Looking
only at extreme weather as a cause of migration, a recent report
estimates that over 20 million people were displaced due to sudden-
onset climate-related disasters in 2008 (OCHA/IDMC, 2009). This report
and others, however, acknowledge the difficulty of disentangling the
drivers of migration, including climate change risks, rising poverty,
spread of infectious diseases, and conflict (Castles, 2002; Myers, 2005;
Thomalla et al., 2006; Barnett and Adger 2007; CIENS, 2007; Dun and
Gemenne, 2008; Guzmán, 2009; Morrissey, 2009).
As opposed to abrupt displacement due to extreme weather events,
mobility and migration can also be an adaptation strategy to gradual
climatic change (Barnett and Webber, 2009), which normally leads to
slower migration shifts. However, the very poor and vulnerable will in
many cases be unable to move (Tacoli, 2009). To the extent that weather
extremes contribute to migration, it can result in a huge burden to the
destination areas (Barnett and Adger, 2007; Heltberg et al., 2008;
Morrissey, 2009; Tacoli, 2009; Warner et al., 2009a). As part of this
burden, the conflict potential of migration depends to a significant
degree on how the government and people in the transit, destination,
or place of return respond. Governance, the degree of political stability,
the economy, and whether there is a history of violence are generally
important factors (Kolmannskog, 2008).
The international impacts of climate-related disasters can extend
beyond financial consequences, international trade, and migration, and
affect human security more generally. O’Brien et al. (2008) report on the
intricate and systemic linkages between DRR, CCA, and human security,
and they emphasize the importance of confronting the societal context,
including development levels, governance, inequality, and cultural
practices. A further rationale for disaster risk reduction in the face of
climate change at the international scale thus places emphasis on ethical
issues and the growing connections among people and places in coupled
social-ecological systems.
7.2.2. Economic Efficiency
The public policy literature describes situations in which government
intervention is justified to address market deficiencies and inefficiencies,
a rationale that can also be applied to international interventions. Stern
(2007) makes the case that adaptation will not happen autonomously
because of inefficiencies in resource allocation brought about by missing
and misaligned markets. As a case in point, markets do not allocate
resources efficiently in the case of public goods, which are goods that
meet two conditions: the consumption of the good by one individual
does not reduce availability of the good for consumption by others; and
no one can be effectively excluded from using the good. Tompkins and
Adger (2005) and Berkhout (2005) discuss how some areas, such as
water resources, change from being public to private depending on
national regulations and circumstances. Nevertheless, the principles of
interdependence and public goods suggested by Stern and others (and
which lead to inefficient allocation of resources) are frequently noted in
the literature on international responsibility (Stern, 2007; Vernon, 2008;
Gupta et al., 2010; World Bank, 2010a).
Chapter 7 Managing the Risks: International Level and Integration across Scales
400
Early warning systems (as an example of a public good) can depend on
regional and international cooperation to make more efficient use of
climate data through its exchange. In the field of meteorology, many
years of discussion under the auspices of the World Meteorological
Organization (WMO) have led to formal agreements on the types of
data that are routinely exchanged (WMO, 1995; Basher, 2006). There
are similar levels of agreement in other hazard fields, for instance,
sharing resources and expertise in managing floods at the river basin
scale. As another example of enhanced efficiency through international
cooperation, many Caribbean countries have formed a catastrophe
insurance pool to reduce reinsurance premiums (see Sections 6.3.3 and
7.4, and Case Study 9.2.13).
7.2.3. Shared Responsibility
It is not only efficiency claims that can be invoked to justify international
interventions, but also considerations of shared responsibility and
solidarity, especially with those least able to cope with the impacts of
extreme events and changes in them due to climate change. This
subsection makes reference to selected principles found in the current
literature on adaptation to weather-related extremes; there is no
attempt to comprehensively assess the moral and ethical literature on
this topic.
In the words of the Millennium Declaration that was adopted by 189
nations in September 2000:“We recognize that, in addition to our
separate responsibilities to our individual societies, we have a collective
responsibility to uphold the principles of human dignity, equality and
equity at the global level. Global challenges must be managed in a way
that distributes the costs and burdens fairly in accordance with basic
principles of equity and social justice. Those who suffer or who benefit
least deserve help from those who benefit most” (UNGA, 2000).
In the poorest countries, people have a higher burden in terms of loss of
life per event and loss of their assets relative to their income. Based on
historical loss data from Munich Re, average fatalities for major disaster
events have been approximately 40 times higher in low-income as
compared to high-income countries (groupings according to the World
Bank), and direct asset losses as a percentage of gross national income
have averaged three times greater (Barnett et al., 2008; Linnerooth-
Bayer et al., 2010). Changes in frequency, magnitude, and spatial coverage
of some climate extremes (see Table 3-1) can result in losses that exceed
the capability of many individual countries to manage the risk
(Rodriguez et al., 2009). Many have concluded that without significant
international assistance the most vulnerable countries will have difficulty
in adapting to changes in extreme events and their impacts due to climate
change, as well as other impacts of climate change (Agrawala and
Fankhauser, 2008; Agrawala and van Aalst, 2008; Klein and Persson,
2008; Klein and Möhner, 2009; Gupta and van de Grijp, 2010; Gupta et
al., 2010; World Bank, 2010a). Shared responsibility can take the form
of ex-ante interventions to reduce vulnerability and poverty, as well as
ex-post disaster response and assistance.
Weather extremes constrain progress toward meeting the MDGs as
expressed in the Millennium Declaration, especially the goal of eradicating
extreme poverty and hunger (UNDP, 2002; Mirza, 2003; Watkins, 2007;
UNISDR, 2009a), which can be interpreted as a direct raison d’être for
international intervention in risk management (UNISDR, 2005b;
Heltberg et al., 2008). Barrett et al. (2007) have shown that ex-ante risk
management strategies on the part of the poor commonly sacrifice
expected gains, such as investing in improved seed, to reduce risk of
suffering catastrophic loss, a situation perpetuating the ‘poverty trap.
The poor can be subject to multiple exposures from climate change
and other stresses like geophysical hazards and changing economic
conditions (e.g., fluctuating exchange rates) leading to vulnerability to
even moderate hazard events (O’Brien and Leichenko, 2000).
Shared responsibility and common human concern have been articulated
most effectively with regard to post-disaster humanitarian assistance,
and the Millennium Declaration gives specific mention to ‘natural’
disasters in this context. Section VI (Protecting the Vulnerable) states: “We
will spare no effort to ensure that children and all civilian populations
that suffer disproportionately the consequences of natural disasters …
are given every assistance and protection so that they can resume normal
life as soon as possible. With growing globalization the principle of
shared responsibility is further enhanced as offers of disaster relief may
provide nations access to new spheres of influence both politically and
in terms of new business opportunities. Governments can piggyback a
humanitarian effort on top of a for-profit operation involving private
companies (Dunfee and Hess, 2000).
Disasters can overwhelm the coping mechanisms of nations, in which
case international relief and assistance, as a form of solidarity, are
required as a matter of saving lives. Humanitarian assistance will
remain essential, but emphasizing disaster response strategies at the
expense of proactive integrated approaches to disaster risk reduction
can have the effect of perpetuating vulnerability (UNDP, 2002; Bhatt,
2007). For this reason, the DRR and CCA communities are placing great
emphasis on pre-disaster investment and planning to redress this balance
and reduce overall costs of disaster management (Kreimer and Arnold,
2000; Linnerooth-Bayer et al., 2005). These efforts include encouraging
the humanitarian community to become a stronger advocate of DRR
and CCA.
Beyond a sense of common human concern, it can be argued that
countries contributing most to climate change have an obligation to
pay to reduce or compensate losses. This is the principle underlying the
‘polluter pays principle.’ In addition, it can be claimed that countries
have a ‘principled’ obligation to support those who are most vulnerable
and who have made a limited contribution to the creation of the climate
change problem. This is the claim underlying the expression of ‘common
but differentiated responsibilities and respective capabilities’ (CBDR),
which has emerged as one principle of international environmental law
(De Lucia, 2007) and has been explicitly formulated in the context of
the 1992 Rio Earth Summit (and subsequently in the Preamble and
Article 3 of the UNFCCC). “In view of the different contributions to global
Chapter 7Managing the Risks: International Level and Integration across Scales
401
environmental degradation, States have common but differentiated
responsibilities. The developed countries acknowledge the responsibility
that they bear in the international pursuit of sustainable development in
view of the pressures their societies place on the global environment and
of the technologies and financial resources they command. (Principle 7,
the Rio Declaration; UNCED, 1992). The CBDR is discussed further in
Section 7.2.5. For purposes here it is important to note that, while
the CBDR principle can apply to climate change in general, including
incremental change, it is relevant to climate-related disasters only if there
is evidence or reason to believe that the disaster would not have occurred
or would have been less severe in the absence of climate change.
Another set of literature (e.g., Adger et al., 2009; Caney, 2010) frames
equity issues around climate change in terms of ‘rights,’ namely the right
‘not to suffer from dangerous climate change’ or ‘to avoid dangerous
climate change’ (Adger, 2004; Caney, 2008). The ‘rights’ argument,
which is highly relevant to international solidarity, can be extended to
suggest that individuals and collectives have the right to be protected
from risk and disaster imposed by others through the processes that
lead to social exclusion, marginality, exposure, and vulnerability.
According to this literature, climate change impacts can jeopardize
fundamental rights to life and livelihood (such as impacts on disease
burden, malnutrition, and food security). Caney (2010, p. 83) also
discusses a potential further undeniable right, ‘not to be forcibly
evicted.This framing, however, raises a number of difficult issues
because of competing fundamental rights (O’Brien et al., 2009).
7.2.4. Subsidiarity
The principle of ‘subsidiarity’ can be invoked to support a case against
international intervention. It is best known as articulated in Article 5 of
the Treaty of Maastricht on European Union (Maastricht Treaty, 1992). It
is based on the concept that centralized governing structures should
only take action if deemed more effective or necessary than action at
lower levels (Jordan, 2000; Craeynest et al., 2010). The intent is to
strengthen accountability and reduce the dangers of making decisions
in places remote from their point of application (Gupta and Grubb,
2000). In Europe, the principle of subsidiarity has been interpreted to
mean, for example, that international- or national-level involvement in
flood protection should only apply to cross-border catchments (Stoiber,
2006). While many regions and river basins are required to develop risk
management flood plans, flood protection is considered predominantly
a national, and in many countries (e.g., Germany and India), primarily a
sub-national (state) responsibility.
The principle also recognizes that multi-level governance requires
cooperation between all levels of government (Begg, 2008). As an
example of this cooperation, in 2004, the African Union developed a
continent-wide African Regional Strategy for Disaster Risk Reduction
(African Union, 2010). Below the continental level, disaster management
strategies are developed at the regional level (e.g., under the
Regional Economic Communities), national level (e.g., National Disaster
Management platforms), district level (e.g., District Disaster Management
Committees), and local levels (e.g., Village Development Committees).
Action at any one level can affect all others in a reflexive fashion.
7.2.5. Legal Obligations
7.2.5.1. Scope of International Law,
Managing Risks, and Adaptation
Contemporary international law concerns the coexistence of states in
times of war and of peace (19th-century conception of international
law, rooted in the Westphalian system), the relationship between a state
and citizens (e.g., human rights law), and the cooperation between
states and other international actors in order to achieve common goals
and address common concerns (e.g., international environmental law).
International law, according to the authoritative Article 38 of the Statute
of the International Court of Justice, emanates from three primary
sources: (1) international conventions, which establish “rules expressly
recognized by the … states,” and result from a deliberate process of
negotiations; (2) international custom, “as evidence of a general practice
accepted as law”; and (3) general principles of law, “recognized by
civilized nations” (see also Birnie et al., 2009). This triumvirate of
conventional and customary international law, and general principles of
law, contains legal norms and obligations that can be used to motivate,
justify, and facilitate international cooperation on climate change
adaptation, such as contained within the UNFCCC, and in anticipation
of and response to natural disasters, such as with the emerging field of
international disaster relief law.
In addition to international sources of ‘hard law,‘soft law’ principles
also exist in the form of non-legally binding resolutions, guidelines,
codes of conduct (Chinkin, 1989; Bodansky, 2010), and other non-legally
binding instruments adopted by states. Collectively, hard law and soft
law provide a framework within which states have obligations (hard
law) or commitments (soft law) of relevance to adapting to climate
change and disaster risk management. These include obligations to
mitigate the effects of drought (United Nations Convention to Combat
Desertification), to formulate and implement measures to facilitate
adaptation (UNFCCC; see Section 7.3.2), to exercise precaution (Rio
Declaration), for international cooperation to protect and promote
human rights (OHCHR, 2009, para. 84 et seq.), and to develop national
legislation to address disaster risk reduction (HFA; see Section 7.3.1).
At the same time as international law appears to provide a normative
framework and to create an obligation to “implement … measures to
facilitate adequate adaptation to climate change” (UNFCCC Article 4.1(b)),
the literature suggests that taken together, international legal instruments
are not equipped to fully facilitate climate adaptation and to reduce
disaster risk. To illustrate, the law of international disaster response,
which aims to establish a legal framework for transborder disaster relief
and recovery, has been characterized as “dispersed, with gaps of scope,
geographic coverage and precision” (Fisher, 2007), with states being
Chapter 7 Managing the Risks: International Level and Integration across Scales
402
“hesitant to negotiate and accept far-reaching treaties that impose
legally binding responsibilities with respect to disaster preparedness,
protection, and response” (Fidler, 2005). A second example, international
refugee law, does not recognize environmental factors as grounds for
granting refugee status to those displaced across borders as a result of
environmental factors (Kibreab, 1997).
7.2.5.2. International Conventions
Few internationally negotiated treaties deal, at the international level,
with managing risk associated with climate extremes or with adaptation
to climate change. As the primary treaty to address climate-related risk
management at the international level, the UNFCCC commits Parties to
facilitate adequate adaptation, to cooperate with planning for extreme
weather, and to consider insurance schemes, though at present it is
unresolved as to whether this implies international insurance schemes.
Specifically, in Article 4.1(b), Parties to the UNFCCC agree to “formulate,
implement, publish and regularly update national and, where appropriate,
regional programmes containing … measures to facilitate adequate
adaptation to climate change. In Article 4.1(e), Parties agree to
“cooperate in preparing for adaptation to the impacts of climate change;
develop and elaborate appropriate and integrated plans for coastal zone
management, water resources and agriculture, and for the protection
and rehabilitation of areas, particularly in Africa, affected by drought and
desertification, as well as floods.” Article 4.8 of the UNFCCC commits
Parties to consider actions “including related to funding, insurance and
the transfer of technology” to meet the specific needs and concerns of
developing countries. In Article 3.14, UNFCCC’s Kyoto Protocol considers
the establishment of funding, insurance, and transfer of technology (see
also Sections 7.4.2, 7.4.3, and 7.4.4).
In addition to the UNFCCC, Parties to the United Nations Convention to
Combat Desertification aim to “combat desertification and mitigate the
effects of drought in countries experiencing serious drought and/or
desertification … through effective action at all levels, supported by
international cooperation and partnership arrangements” (Article 2).
The Tampere Convention on the Provision of Telecommunication Resources
for Disaster Mitigation and Relief Operations is the only contemporary
multilateral treaty on the topic of disaster relief (Fidler, 2005). Aiming to
reduce regulatory barriers for important equipment for disaster response,
and entered into force in 2005, the Tampere Convention’s first application
has been met with limited success, due primarily to limited membership
of many of the most vulnerable states (Fisher, 2007).
7.2.5.3. Customary Law and Soft Law Principles
Customary law and soft law principles, unlike international conventions,
emerge from informal processes and do not exist in canonical form
(Bodansky, 2010, p. 192 et seq.), though such customary law and soft
law principles are often reflected in international treaties. This is the
reality of various customs and principles that justify or mandate
international action on disaster risk reduction and climate change
adaptation. To be established as customary law, two elements are
requisite: evidence of generally uniform and continuous state practice
(regular behavior), and evidence that this practice is motivated by a sense
of legal obligation (opinio juris) (Bodansky, 1995). Soft law principles of
law, by contrast, are not customary norms and do not reflect behavioral
regularities. They are rather an articulation of collective aspiration,
important in shaping the “development of international law and
negotiations to develop more precise norms” (Bodansky, 2010, p. 200).
In practice, the distinction between rules of customary law (reflecting
actual practice of states following a legal obligation) and soft law
principles is frequently blurred. For instance, the principle of common but
differentiated responsibilities and respective capabilities – which would
for example suggest that states have differentiated responsibilities in
addressing disaster risk and financing adaptation – is increasingly
supported by state practice, however opinio juris is lacking as it is unclear
whether most states consider the principle to be a legal obligation. The
principle of common but differentiated responsibilities and respective
capabilities might thus fall closer to a general principle than a customary
norm. Irrespective of this status, the principle of common but differentiated
responsibilities and respective capabilities is nevertheless a principle
that states may apply in their international relations, even if it is not a
norm of customary international law.
The precautionary principle states that scientific uncertainty does not
justify inaction with respect to environmental risks (Trouwborst, 2002),
and is articulated in a number of international instruments including
Principle 15 of the Rio Declaration, and Article 3 of the UNFCCC. That
states have a duty to prevent transboundary harm, provide notice of, and
undertake consultations with respect to such potential harms is a soft
law norm expressed under international environmental law. The more
general duty to cooperate has evolved as a result of the inapplicability
of the law of state responsibility to problems of multilateral concern,
such as global environmental challenges. The Office of the High
Commissioner for Human Rights has noted that “climate change can
only be effectively addressed through cooperation of all members of the
international community” (OHCHR, 2009). From the duty to cooperate
is deduced a duty to notify other states of potential environmental
harm. This is reflected in Principles 18 and 19 of the Rio Declaration (a
non-legal international instrument), that “States shall immediately notify
other States of any natural disasters or other emergencies that are likely
to produce sudden harmful effects on the environment of those States”
(Rio Principle 18) and “States shall provide prior and timely notification
and relevant information to potentially affected States on activities that
may have a significant adverse transboundary environmental effect”
(Rio Principle 19).
7.2.5.4. Non-Legally Binding Instruments
Many international instruments are non-legal in nature (Raustiala,
2005). This is the case with respect to disaster relief where many of the
Chapter 7Managing the Risks: International Level and Integration across Scales
403
most significant international instruments are non-binding. Illustrative are
the Code of Conduct for the International Red Cross and Red Crescent
Movement and Nongovernmental Organizations in Disaster Relief (ICRC,
1995) and the Sphere Project, Humanitarian Charter and Minimum
Standards in Disaster Response (Sphere Project, 2004), which focus on the
quality of relief developed by the international humanitarian community.
These are limited by lack of compliance mechanisms (Fidler, 2005), as
well as in their application, as they are the creation of international
nongovernmental organizations (NGOs) and are rarely recognized in the
policies of national governments. The Guiding Principles on Internal
Displacement (Cohen, 1998) articulate principles of disaster prevention
and of human vulnerability (Fisher, 2007).
International human rights norms as articulated in the International Bill
of Human Rights have also been applied to disaster risk reduction and
adaptation to climate change. Notably, the Report of the Office of the
High Commission for Human Rights observes that climate change and
response measures thereto can have a negative effect on the realization
of human rights including rights to life, adequate food, water, health,
adequate housing, and self-determination (OHCHR, 2009). These rights
could risk being jeopardized when contemplated, for example, in the
context of migration induced by extreme weather events. As discussed
in Section 7.3.1, the HFA further stipulates key tasks for governments
and multi-stakeholder actors; among these is the development of legal
frameworks (UNISDR, 2005a, para. 22). The HFA is an international
framework, a priority area of which is to ensure that disaster risk reduction
is a national priority with an institutional basis for implementation. As
to adaptation, the Bali Action Plan agreed to at the 13th Conference of
the Parties to the UNFCCC recognizes the need to address consideration
of disaster reduction strategies and risk management within adaptation
(UNFCCC, 2007a). Adaptation is further addressed in the Cancun
Agreements (UNFCCC, 2010c).
7.3. Current International Governance
and Institutions
Among the many relevant frameworks and protocols administered by
a host of United Nations and other international agencies, the most
significant for this Special Report are the HFA, to reduce disaster risk,
and the UNFCCC, which includes adaptation to the adverse effects of
climate change. Since both DRR and CCA occur within a broader
development context and are particularly relevant to the challenges
facing developing countries, they are indirectly connected to a third
important international framework: the MDGs.
The UNFCCC was adopted in 1992 following one year of negotiations
and was further complemented by the Kyoto Protocol adopted in 1997.
The Convention came into force in 1994 and the Protocol in 2005. In
parallel, the DRR framework was adopted as a nonbinding instrument
in 2005 following two years of negotiations and is time bound – 2005
to 2015. The HFA recognizes the relevance of addressing climate change
in order to reduce the risk of disasters and, as soon as adopted, the two
processes began to work together, collaborating closely in order to
synchronize frameworks and approaches so as to create added value to
current risk management initiatives. This IPCC Special Report is one
example of the initiatives taken by governments. It is one of the first
official products of the two communities working within different but
related policy frameworks.
This section first introduces the HFA and the UNFCCC, including an
overview of their respective objectives, legal nature, and status of
implementation. It then presents relevant international actors involved
in implementing these two frameworks, as well as a summary of other
relevant international policy frameworks and agencies.
7.3.1. The Hyogo Framework for Action
7.3.1.1. Evolution and Description
The first major collective international attempt to reduce disaster impact,
particularly within hazard-prone developing countries, took place in
1989, when the United Nations (UN) General Assembly designated the
1990s as the International Decade for Natural Disaster Reduction
(IDNDR) (Wisner et al., 2004). About 120 National Committees were
established and in 1994, the first World Conference on Natural Disaster
Reduction was held in Yokohama, Japan. The conference produced the
‘Yokohama Strategy and Plan of Action,’ providing policy guidance with
a strong technical and scientific focus.
In 2000, the IDNDR was followed by the United Nations International
Strategy for Disaster Reduction (UNISDR), which broadened the technical
and policy scope of the IDNDR to include increased social action, public
commitment, and linkages to sustainable development. The UNISDR
system promotes tools and methods to reduce disaster risk while
encouraging collaboration between disaster reduction and climate
change. The UNISDR Secretariat provides information and guidance on
disaster risk reduction and has increasingly widened its focus to embrace
adaptation to climate change. The strategy undertakes global reviews of
disaster risk and promotes national initiatives to reduce disaster risk.
The UNISDR has also promoted the development of National Platforms.
A key function is to assist in the compilation, exchange, analysis, and
dissemination of good practices and lessons learned in disaster risk
reduction (refer to Section 7.4.5).
In January 2005, just three weeks after the Indian Ocean tsunami, the
second World Conference on Disaster Reduction was held in Kobe,
Japan. 168 governments adopted the Hyogo Framework for Action
2005-2015: Building the Resilience of Nations and Communities to
Disasters. The adoption of the framework directly after a devastating
tsunami gave the framework high visibility in many countries. The HFA
was unanimously endorsed by the UN General Assembly (UNISDR, 2005a).
The HFA is not a binding agreement: the governments simply agreed
and adopted the framework as a set of recommendations to be utilized
voluntarily. In international law it can be described as ‘soft law.’ Some
Chapter 7 Managing the Risks: International Level and Integration across Scales
404
regard the voluntary nature of the HFA as a useful flexible commitment,
largely based on self-regulation and trust, while others regard this as its
inherent weakness (Pelling, 2011, p. 44).
The HFA’s Strategic Goals include the integration of DRR into sustainable
development policies and planning; development and strengthening of
institutions, mechanisms, and capacities to build resilience to hazards;
and the systematic incorporation of risk reduction approaches into the
design and implementation of emergency preparedness, response, and
recovery programs (UNISDR, 2005a). The Framework also provides five
Priorities for Action:
1) Ensure that DRR is a national and local priority, with a strong
institutional basis for implementation
2) Identify, assess, and monitor disaster risks, and enhance early
warning
3) Use knowledge, innovation, and education to build a culture of
safety and resilience at all levels
4) Reduce the underlying risk factors
5) Strengthen disaster preparedness for effective response at all levels.
The priorities address all hazards with a multi-hazard approach, hence
the inclusion of climate change risks and adaptation, but they do not
specify the need to factor climate change risks and adaptation into
ongoing action. The HFA does identify ‘critical tasks’ for varied actors,
including states who are to “promote the integration of DRR with climate
variability and climate change into DRR strategies and adaptation to
climate change” (UNISDR 2005a; see also UNISDR, 2009a, 2011a,b;
World Bank, 2011a).
7.3.1.2. Status of Implementation
This section will review the various tools that have been used to measure
the performance of the HFA in fulfilling its Strategic Goals and Priorities
for Action.
The measurement of performance in the implementation of DRR was a
matter of considerable debate when the HFA was drafted. The consensus
was for the final text not to include targets or indicators of progress, but
countries were encouraged to develop their own guidelines to monitor
their own progress in reducing their risks. To assist this process, in 2008,
UNISDR published guidance notes on ‘Indicators of Progress’ (UNISDR,
2008). This provided the template for self-assessment that is used in
national reports. While there is an obvious value in ‘self-assessment’ as
a learning experience, in the absence of external, objective evaluation,
inevitable doubts will always remain concerning such internal reporting
on actual progress with DRR and CCA.
The main instruments to encourage HFA applications are the HFA
Monitoring Service on PreventionWeb acting mainly as a guidance tool
for countries to monitor their own progress in DRR. This is a multi-tier
online tool for regional, national, and local progress review. Core
Indicators are measured for the five HFA Priorities for Action as noted
below, and these are reported with detailed analysis in the Global
Assessment Reports (UNISDR, 2009a, 2011a; refer to Section 7.4.5). In
addition to these biennial reports, the UNISDR has published a mid-term
review of progress in achieving the HFA (UNISDR, 2011b). Further tools
to measure progress include the reports to the biennial sessions of the
Global Platform for DRR and the regional platforms for DRR and other
similar mechanisms. The World Bank and the United Nations
Development Programme (UNDP) also utilize the HFA to guide their
support to national and local programs on DRR and gradually also for
CCA (the HFA is also discussed in Sections 1.3.6 and 6.3.2).
As a result of the adoption of HFA, and the development of performance
indicators, global efforts to address DRR have become more systematic.
In 2009, the first biennial Global Assessment Report (GAR) on Disaster
Risk Reduction was released and in the same year the Global Network
of Civil Society Organisations for Disaster Reduction (GNDR) also
released a report on the performance of the HFA (GNDR, 2009). The GAR
found that since the adoption of the HFA, progress toward decreasing
disaster risk is varied across scales. This variation is based on national
government agencies self-assessment of progress against the indicators
defined by the UNISDR (UNISDR, 2008) and since many of these indicators
require a subjective assessment, progress is not directly comparable
across countries.
Countries have been making improvements toward increasing capacity,
developing institutional systems, and legislation to promote DRR, and
early warning systems have been implemented in many areas. However,
the Global Assessment Reports (UNISDR, 2009a, 2011a) conclude that
progress is still required to mainstream DRR into public investment,
development planning, and governance arrangements. During 2010, at
the mid-point in the HFA, the UN Secretary General echoed this concern
in reporting that “risk reduction is still not hardwired into the ‘business
processes’ of the development sectors, planning ministries and financial
institutions” (UNGA, 2010, p. 5).
Further, both the GARs and the GNDR (2009, 2011) noted that at
national and international levels, policy and institutional frameworks
for climate change adaptation and poverty reduction are not yet
synchronized to those for DRR. For example, the 2011 GAR reports on
weak coordination and separate management between institutional
and program mechanisms (UNISDR, 2011a, p. 150).
The GNDR observed that ecosystem management approaches can provide
multiple benefits, including risk reduction, and thus be a central part of
DRR strategies. But countries have experienced difficulty in addressing
underlying risk drivers (such as food security, social protection, building
codes/standards, poverty alleviation, poor urban and local governance,
vulnerable rural livelihoods, and ecosystem decline) in a way that leads
to a reduction in the risk of damages and economic loss (GNDR, 2009).
This Fourth HFA Priority for Action – ‘Reduce the Underlying Risk Factors’
– remains the greatest challenge to civil society bodies, with all 13
criteria only reaching a rating of 2 on the assessment scale: ‘some activity
but significant scope for improvements’ (GNDR, 2009, pp. 24–26). The
Chapter 7Managing the Risks: International Level and Integration across Scales
405
GARs also note this area of weakness, but note that it is possible for
countries to address underlying risk drivers using an assortment of
mechanisms to increase resilience (e.g., raising awareness, education,
training, risk assessments, early warning systems, building safety, micro-
insurance in macro-financing schemes) (UNISDR, 2009a, 2011a).
It was also acknowledged in the 2009 GAR that weather-related disaster
risk is escalating swiftly, in terms of the regions affected, frequency
of events, and losses reported. This frequency relates to occurrence
patterns as well as improved reporting of all categories of weather-
related hazards. Data was collected from a sample of 12 Asian and Latin
American countries: Argentina, Bolivia, Colombia, Costa Rica, Ecuador,
the Indian states of Orissa and Tamil Nadu, Iran, Mexico, Nepal, Peru,
Sri Lanka, and Venezuela. The report further noted that these increases
will magnify the uneven distribution of risk between wealthier and
poorer countries (UNISDR, 2009a, p. 11). Furthermore, a conclusion is
drawn in the report that climate change is changing the geographical
distribution, intensity, and frequency of these weather-related hazards,
threatening to exceed the capacities of poorer countries and their
communities’ abilities to absorb losses and recover from disaster
impacts (UNISDR, 2009b). However, the 2011 GAR reported significant
progress with a decrease in global mortality risk from tropical cyclones and
flooding, with the only exception being South Asia where vulnerability
is still increasing (UNISDR, 2011a, p. 28).
The 2009 and 2011 GARs, as well as the discussion they generated in
the Global Platforms of 2009, have brought a regional dimension to
performance assessment, in an effort to monitor progress.
When evaluating the progress of HFA on each of its five Priorities for
Action, the GNDR found that the lowest level of progress across all the
five priorities was at the lowest scale in community participation in
decisionmaking on DRR (GNDR, 2009). These findings also indicate the
need for a stronger link between policy formulation at international and
national levels to policy execution at local levels. Rapid progress has
been made in the development of comprehensive seasonal and long-
term early warning systems (EWS) to anticipate droughts, floods, and
tropical storms. These systems have proved to be effective in saving
lives and protecting property. In the 2009 GAR, the status of EWS was
reviewed (UNISDR, 2009a, Box 5.2 on p. 127). This was based on a
detailed progress review of EWS undertaken by WMO (WMO, 2009).
Typical examples of the effectiveness of EWS in reducing the impact of
cyclones and flooding can be found in Mozambique, where their EWS
was first tested in a cyclone in 2007 (Foley, 2007) and in Bangladesh,
where the flood and cyclone EWS has been progressively developed
over three decades (Paul et al., 2010; also see Case Study 9.2.11).
A key finding concerned the importance of education and sharing
knowledge, including indigenous and traditional knowledge, and ensuring
easy and systematic access to best practice tools and international
standards, tailored to specific sectors (see Section 7.4.5). There is some
recognition of the benefits in harmonizing and linking the frameworks
and policies for DRM and CCA as core policy and programmatic objectives
in national development plans and in support of poverty reduction
strategies. DRM policies also need to take account of climate change.
Nevertheless, countries are making significant progress in strengthening
capacities, institutional systems, and legislation to address deficiencies
in disaster preparedness and response (GNDR, 2009; UNISDR, 2009a).
In preparing for the mid-term review of the HFA, the UNISDR secretariat
commissioned a desk review of literature to form “a baseline of the
disaster risk reduction landscape. Forty-seven key documents were
identified, mainly consisting of reports from UNISDR offices and partner
organizations: NGOs and international development banks (UNISDR,
2011b).
The HFA Mid-Term Review 2010-2011 raised two important international
issues. The first need is to develop accountability mechanisms at all
levels to measure the actions taken and progress achieved in DRR. The
second need is for the international community to develop a more
coherent and integrated approach to support the implementation of the
HFA. The review suggests that this will require connected action of the
varied international actors (UNISDR, 2011b).
However, it is important to reflect on the reality that all of these methods
to review international progress in risk reduction – country progress
reports, the 2009 and 2011 GARs, the reports of the GNDR, and the
Mid-Term Review of the HFA – are all internally produced reports by the
participating agencies with external advisory boards and peer review,
but all involving self-assessment. The GNDR’s publications are fully
independent from the UN and governments, but make no claim to be
scientifically accurate assessments. The country HFA reports are online
at www.preventionweb.net/english/hyogo/progress/?pid:73&pih:2.
All the above studies attempted to assess HFA performance and, as
noted above, none were totally separate from the work or institutions
being assessed. Furthermore, none looked specifically at the performance
of the lead organization, UNISDR, in comparison with other multilateral
bodies. This report came in 2011, when the UK Aid Agency, the
Department for International Development (DFID), published a
Multilateral Aid Review. The purpose was to ensure maximum value for
money for UK aid by examining the performance of 43 multilateral
organizations. This peer-reviewed assessment placed the UNISDR in a
43rd-ranked position in an assessment of 43 multilateral organizations
(DFID, 2011).
This independent and comparative assessment included an evaluation
of UNISDR since its foundation and identified its strength as global
coordinator of the three Global Platforms in DRR that have been
successful in advocacy and raising awareness. However, the assessment
also identified a series of shortcomings in UNISDR. They included its poor
performance in international coordination and its focus on national-level
responses rather than its global mandate, which is broad rather than
specific in focus. Further criticisms include inadequate attention to
strategic considerations as well as leadership failures, with the report
stating that there was no clear line of sight from UNISDR’s mandate, to
Chapter 7 Managing the Risks: International Level and Integration across Scales
406
a strategy, to an implementation plan and that there was an absence of
a results-based framework, thus making it difficult to measure results
from input to output (DFID, 2011, p. 211).
UNISDR responded to the assessment by noting that the criticisms were
also reflected in a UN audit as well as in an external evaluation requested
by UNISDR in 2009, and that changes had now been incorporated in a
management-reform work program (UNISDR, 2011c).
Whatever method is adopted to monitor progress with risk reduction
and climate change adaptation (internal or external, self-assessment or
peer review), the implicit problems faced in the measurement of DRR
and CCA before a disaster event must be recognized. It is not easy, even
with detailed objective scientific measurement, to accurately determine
whether a given structural or non-structural measure will actually
provide the necessary level of protection to people and property under
extreme hazard loads. Structural tests can be carried out and simulation
exercises can be usefully conducted to test warning systems or the
effectiveness of preparedness, but at best such performance tests can
only approximate disaster reality. The ultimate test of DRR and CCA
applications will inevitably need to await the impact of the next disaster.
But this limitation does not remove the requirement to monitor and
measure progress in an objective scientific manner to the upper limits
of existing knowledge (Davis, 2004).
7.3.2. The United Nations Framework Convention
on Climate Change
7.3.2.1. Evolution and Description
The UNFCCC is a multilateral treaty aimed at addressing climate change.
Its ultimate objective as stated in Article 2 is (UN, 1992; see also
Oppenheimer and Petsonk, 2005):
“to achieve … stabilization of greenhouse gas concentrations in the
atmosphere at a level that would prevent dangerous anthropogenic
interference with the climate system. Such a level should be achieved
within a time-frame sufficient to allow ecosystems to adapt naturally
to climate change, to ensure that food production is not threatened
and to enable economic development to proceed in a sustainable
manner.”
The UNFCCC was negotiated from February 1991 to May 1992, and
opened for signature at the UN Conference on Environment and
Development in Rio de Janeiro in June 1992. It entered into force on 21
March 1994, and since 1995 the Conference of the Parties (COP) to the
UNFCCC has met in yearly sessions. The rules, institutions, and procedures
of the UNFCCC have been described in detail elsewhere (e.g., Yamin and
Depledge, 2004; Bodansky, 2005). The development of adaptation as a
priority under the UNFCCC has been analyzed by Schipper (2006).
A major thrust of the UNFCCC and subsequent negotiations about its
implementation concerns the mitigation of climate change: all policies
and measures aimed at reducing the emission of greenhouse gases such
as carbon dioxide (CO
2
), or at retaining and capturing them in sinks
such as forests, oceans, and underground reservoirs. As mentioned by
Schipper (2006), adaptation to climate change was initially given little
priority, although it is subject to various commitments in the UNFCCC
(see Box 7-1). When taken together, these commitments acknowledge
the systematic nature of climate change risks and the relevance of the
principles of economic efficiency, solidarity, and subsidiarity in adaptation.
The Kyoto Protocol, agreed at COP3 in 1997 and in force since 2005,
sets binding targets for 37 industrialized countries and the European
Union for reducing greenhouse gas emissions by an average of 5%
compared to 1990 over the five year period 2008-2012. Adaptation is
all but absent in the Kyoto Protocol, with two exceptions. Article 10(b)
specifies that Parties shall formulate, implement, publish, and regularly
update national and, where appropriate, regional programs containing
measures to mitigate climate change and measures to facilitate adequate
adaptation to climate change. Article 12.8, on the Clean Development
Mechanism, provides the basis of what later became the Adaptation
Fund (see Section 7.4.2).
7.3.2.2. Status of Implementation
There is to date no overall assessment of progress on adaptation under
the UNFCCC in the way that the UNISDR has assessed progress under
the HFA in the GARs. However, Parties to the UNFCCC are required to
submit National Communications on their activities toward implementing
the UNFCCC, including adaptation. There is no common reporting
template so reports vary widely in content, making aggregation or
comparison problematic. The annual sessions of the COP also allow
countries to assess their progress toward meeting their commitments
under the UNFCCC, and to negotiate and adopt new decisions for further
implementation. By June 2011, there were 195 Parties to the UNFCCC:
194 countries and one regional economic integration organization (the
European Union).
During the 1990s, adaptation received little attention in the UNFCCC
negotiations, reflecting a similarly low level of attention to adaptation
from the academic community at the time (Burton et al., 2002). The profile
was raised in 2001 with the publication of the IPCC Third Assessment
Report, which contained the chapter Adaptation to Climate Change in
the Context of Sustainable Development and Equity’ (Smit et al., 2001).
Also in 2001, COP7 adopted a decision (5/CP.7) that outlined a range of
activities that would promote adaptation in developing countries,
including the preparation of National Adaptation Programmes of Action
(NAPAs) by least-developed countries. To this end, COP7 established
three funds with which adaptation in developing countries could be
supported, namely the Least Developed Countries Fund (LDCF), the
Special Climate Change Fund (SCCF), and the Strategic Priority ‘Piloting
an Operational Approach to Adaptation’ (SPA) under the Trust Fund of
the Global Environment Facility (GEF). In addition, COP7 took the first
steps toward making operational the Adaptation Fund (Huq, 2002;
Chapter 7Managing the Risks: International Level and Integration across Scales
407
Dessai, 2003; Mace, 2005). Section 7.4.2 provides more information on
the international financing of climate change adaptation.
Since 2001, a number of successive decisions have given increasing
priority to climate change adaptation under the UNFCCC. Decision
1/CP.10 built on Decision 5/CP.7; it reiterated the need for support for
adaptation in developing countries and started a regional consultation
process. Decision 2/CP.11 then established the Nairobi Work Programme
on impacts, vulnerability, and adaptation to climate change, which
originally ran from 2006 to 2010 – a next phase is currently under
consideration, to be decided at COP17 in Durban in 2011. The objective
of the Nairobi Work Programme is to assist all Parties, in particular
developing countries, (i) to improve their understanding and assessment
of impacts, vulnerability, and adaptation to climate change, and (ii) to
make informed decisions on practical adaptation actions and measures
to respond to climate change on a sound scientific, technical, and
socioeconomic basis, taking into account current and future climate
change and variability (Decision 2/CP.11). The Nairobi Work Programme
is implemented by Parties, intergovernmental and nongovernmental
organizations, the private sector, communities, and other stakeholders.
Several of the nine work areas of the Nairobi Work Programme are
relevant to DRR as well as CCA, in particular ‘climate-related risks and
extreme events’ and ‘adaptation planning and practices.
With Decision 1/CP.13 (also known as the Bali Action Plan), agreed in
December 2007, the COP launched “a comprehensive process to enable
the full, effective, and sustained implementation of the Convention
through long-term cooperative action – now, up to, and beyond 2012 –
in order to reach an agreed outcome and adopt a decision at its fifteenth
session” in Copenhagen in December 2009 (COP15). The Bali Action
Plan gave equal priority to mitigation and adaptation, and identified
technology and finance as the key mechanisms for enabling developing
countries to respond to climate change (Clémençon, 2008; Ott et al.,
2008; Persson et al., 2009). It recognized the need for action to enhance
adaptation in five main areas:
1) International cooperation to support urgent implementation of
adaptation actions, including through vulnerability assessments,
prioritization of actions, financial needs assessments, capacity
building, and response strategies, and integration of adaptation
actions into sectoral and national planning […]
2) Risk management and risk reduction strategies, including risk-
sharing and transfer mechanisms such as insurance
3) Disaster reduction strategies and means to address loss and damage
associated with climate change impacts in developing countries
that are particularly vulnerable to the adverse effects of climate
change
4) Economic diversification to build resilience
5) Ways to strengthen the catalytic role of the Convention in
encouraging multilateral bodies, the public and private sectors, and
civil society, building on synergies among activities and processes,
as a means to support adaptation in a coherent and integrated
manner.
Chapter 7 Managing the Risks: International Level and Integration across Scales
Box 7-1 | Commitments on Climate Change Adaptation as Included in the UNFCCC
Article 4.1: All Parties, taking into account their common but differentiated responsibilities and their specific national and regional
development priorities, objectives, and circumstances, shall:
(b) Formulate, implement, publish, and regularly update national and, where appropriate, regional programs containing measures to
mitigate climate change by addressing anthropogenic emissions by sources and removals by sinks of all greenhouse gases not
controlled by the Montreal Protocol, and measures to facilitate adequate adaptation to climate change.
(e) Cooperate in preparing for adaptation to the impacts of climate change; develop and elaborate appropriate and integrated plans
for coastal zone management, water resources, and agriculture, and for the protection and rehabilitation of areas, particularly in
Africa, affected by drought and desertification, as well as floods.
(f) Take climate change considerations into account, to the extent feasible, in their relevant social, economic, and environmental
policies and actions, and employ appropriate methods, for example impact assessments, formulated and determined nationally,
with a view to minimizing adverse effects on the economy, on public health, and on the quality of the environment, of projects or
measures undertaken by them to mitigate or adapt to climate change.
Article 4.4: The developed country Parties and other developed Parties included in Annex II shall also assist the developing country
Parties that are particularly vulnerable to the adverse effects of climate change in meeting costs of adaptation to those adverse effects.
Article 4.8: In the implementation of the commitments in this Article, the Parties shall give full consideration to what actions are
necessary under the Convention, including actions related to funding, insurance, and the transfer of technology, to meet the specific
needs and concerns of developing country Parties […].
Article 4.9: The Parties shall take full account of the specific needs and special situations of the least developed countries in their
actions with regard to funding and transfer of technology.
408
No agreed outcome was reached at COP15, and no comprehensive
decision was adopted that included these five issues. Instead, the COP
decided to take note of the Copenhagen Accord, a nonbinding document
about which there was no consensus among Parties, and which provides
considerably less substance on adaptation than the Bali Action Plan
(Bodansky, 2010; Grubb, 2010; Klein, 2010). As mentioned in Section
7.4.2, however, the Copenhagen Accord was a milestone toward scaled-
up funding for both mitigation and adaptation.
In 2010, Decision 1/CP.16 (part of the Cancun Agreements) established
the Cancun Adaptation Framework (Cozier, 2011). It invites all Parties to
enhance action on adaptation by undertaking nine activities related to
planning, implementation, capacity strengthening, and knowledge
development, including “enhancing climate change related disaster risk
reduction strategies, taking into consideration the Hyogo Framework for
Action where appropriate; early warning systems; risk assessment and
management; and sharing and transfer mechanisms such as insurance,
at local, national, sub-regional, and regional levels, as appropriate.” In
addition, Decision 1/CP.16 established (i) a process to enable least-
developed countries and other developing countries to formulate and
implement national adaptation plans; (ii) an Adaptation Committee that
will, among other things, provide technical support, share relevant
information, promote synergies, and make recommendations on finance,
technology, and capacity building required for further action; and (iii) a
work program in order to consider approaches to address loss and damage
associated with climate change impacts in developing countries that
are particularly vulnerable to the adverse effects of climate change.
Decision 1/CP.16 also established a Technology Mechanism, consisting of
a Technology Executive Committee and a Climate Technology Center and
Network. The Technology Mechanism should accelerate action at different
stages of the technology cycle, including research and development,
demonstration, deployment, diffusion, and transfer of technology in
support of mitigation and adaptation. Finally, Decision 1/CP.16 established
the Green Climate Fund as a new entity operating the financial mechanism
of the UNFCCC under Article 11 (see Section 7.4.2).
The unfolding of international adaptation policy under the UNFCCC shows
the increasing prominence of adaptation in the negotiations, and the
increasing level of detail and concreteness of the relevant COP decisions.
It also shows that adaptation under the UNFCCC is increasingly linked
with disaster risk reduction, with the Hyogo Framework for Action
explicitly mentioned in the Cancun Agreements. Yet, this unfolding, from
Decision 5/CP.7 to Decision 1/CP.16, has taken 10 years.
7.3.3. Current Actors
A wide range of actors play a role in DRM and CCA at the international
level. This section does not attempt a comprehensive review of all of
these, but instead identifies the broad areas in which the international
community is providing support at the interface between DRM and
CCA, describes some of the main actors under each of these categories,
and summarizes, where available, independent assessments of their
strengths and weaknesses in performing these roles.
7.3.3.1. International Coordination in Linking Disaster Risk
Management and Climate Change Adaptation
Given the wide range of actions and actors that are considered necessary
by those involved to carry out DRM and CCA, and to link them to each
other, effective international coordination is essential. Overall, there are
weaknesses in the current systems; the 2009 Global Assessment Report
on Disaster Risk Reduction states that: “Efforts to reduce disaster risk,
reduce poverty and adapt to climate change are poorly coordinated”
(UNISDR, 2009a).
The main coordination mechanism for DRR, contributing to DRM, is the
UNISDR, designed to develop a system of partnerships to support
nations and communities to reduce disaster risk. These partners include
governments, intergovernmental and nongovernmental organizations,
international financial institutions, scientific and technical bodies and
specialized networks as well as civil society and the private sector.
Among the diverse range of stakeholders across scales, the national
governments play the most important roles, including developing
national coordination mechanisms; conducting baseline assessments on
the status of disaster risk reduction; publishing and updating summaries
of national programs; reviewing national progress toward achieving the
objectives and priorities of the Hyogo Framework; working to implement
relevant international legal instruments; and integrating disaster risk
reduction with climate change strategies. Intergovernmental organizations
play a supporting role, including, for example, promotion of DRR programs
and integration into development planning, and capacity building
(UNISDR, 2005b). The fact that the primary roles in planning and
implementation are played by national governments, while the UNISDR
Secretariat and other intergovernmental organizations provide supporting,
monitoring, and information sharing roles at the regional and global
level is consistent with the principle of subsidiarity.
UNISDR has made specific efforts to link DRR and CCA, through advocacy
of the role of DRR in climate change adaptation, and support for scientific
reviews of the linkages (including this report). Two evaluations covering
the effectiveness of UNISDR in linking DRR and CCA have recently been
published. The UN Special Representative of the Secretary-General for
Disaster Risk Reduction and the main donors to UNISDR requested an
independent evaluation of the performance of the secretariat, which
was published in 2010 (Dalberg, 2010). This review endorsed the overall
effectiveness of UNISDR, particularly in advocacy and awareness raising,
and in establishing global and regional platforms, and specifically
highlights its strong contribution to mainstreaming DRR into climate
change policy. However, it also highlights difficulties, including lack of
definition of comparative advantage within CCA implementation, and
the need to balance the focus and resources spent on DRR in climate
change adaptation versus the broader DRR concept. The same review also
illustrates challenges in coordination of implementation, particularly the
Chapter 7Managing the Risks: International Level and Integration across Scales
409
need for effective coordination with UN Country Teams, the World
Bank, and other relevant partners at the country level, and in the full
implementation and sustainable follow-up of new initiatives. The UK
Government also published a review of the performance of the UNISDR
Secretariat, alongside other multilateral agencies, in 2011 (DFID, 2011).
The review is critical of the overall operational and organizational
strengths of the UNISDR, citing a lack of a results-based framework, and
weaknesses in strategic direction, coordination focus, and speed of
reform. The review does, however, highlight the unique coordinating
role of UNISDR, and specifically praises “a good focus on climate
change, especially adaptation.
From the CCA side, the main global mechanism to increase understanding
and share best practice in CCA is the Nairobi Work Programme (NWP),
coordinated by the UNFCCC Secretariat (UNFCCC, 2010a; refer to
Section 7.3.2.2). The NWP functions mainly as a forum for interested
parties and organizations to specify their own contributions to CCA
through ‘action pledges,’ and for sharing, synthesis, and dissemination
of information. Disaster risk reduction is well represented within the
NWP, which identifies DRR as one of its 14 specified adaptation delivery
activities, with an associated ‘call to action’ for strengthened work in
areas such as linking DRR and CCA, risk mapping, and cost-benefit
analysis of adaptation options. Out of the 137 action pledges made by
partners, 59 include a component of DRR. Evaluation of the NWP by Parties
is only now being carried out, so as yet there is no formal assessment
of the degree to which it has supported changes in policy and practice
as well as information exchange.
7.3.3.2. International Technical and Operational Support
DRM and CCA are now beginning to be linked not only in international
coordination activities, but also in mechanisms for international technical
and operational support.
7.3.3.2.1. Climate services for disaster risk reduction and
climate change adaptation
National meteorological and hydrological services (NMHSs) are the
primary source of meteorological observations and forecasts at time
scales relevant to both disaster risk management and climate change
adaptation. These national services also constitute the members of the
WMO, which serves to set international standards and coordinate among
the members, as well as supporting several relevant international
programs, including a Disaster Risk Reduction and Service Delivery
Branch and a Climate Prediction and Adaptation Branch.
In recent years, a number of studies have identified weaknesses in the
way in which the large amount of potentially relevant information that
is available from NMHSs at the national and international level is
incorporated into development decisions, particularly in the most
vulnerable countries. For example a ‘gap analysis’ of this issue in Africa
identified gaps in (i) integrating climate into policy; (ii) integrating climate
into practice; (iii) climate services; and (iv) climate data, concluding that
“the problem is one of ‘market’ atrophy: negli gible demand coupled
with inadequate supply of climate services for development decisions”
(IRI, 2006). Studies on specific sectors (e.g., health: Kuhn et al., 2005),
or at a local level (Vogel and O’Brien, 2006), conclude that the main
deficit is not in generation of data, but in knowledge management. They
conclude that this requires more effective mechanisms for decisionmakers
to identify their information needs, and to work both with providers of
weather and climate information and with institutions working on
other dimensions of human and social vulnerability to address these
needs.
In response to the need for a comprehensive approach to climate
variability and change, and the drive for more demand-driven climate
services the, World Climate Conference-3 agreed in 2009 to begin
development of a Global Framework on Climate Services (GFCS) (WMO,
2010). This has a goal of “the development and provision of relevant
science-based climate information and prediction for climate risk
management and adaptation to climate variability and change,
throughout the world. The framework therefore explicitly links climate
variability (most relevant to DRR), in the context of climate change
(most relevant to CCA), and support for risk management decisions
(common to both). The GFCS has four major components: a User
Interaction Mechanism; a World Climate Services System; Climate
Research; and Observation and Monitoring. The initiative will focus on
improving access and operational use of climate information, especially
in vulnerable, developing countries. The principles and focus of the
initiative therefore correspond closely to the objectives of linking DRM
and CCA in operational planning across international and smaller
scales. In May 2011, the 16th WMO congress committed to “support
and facilitate the implementation of the GFCS as a priority of the
Organization, including the development of an implementation plan
for review and adoption in 2012 (WMO, 2011).
7.3.3.2.2. Technical and operational support from civil society
Some of the largest international civil society organizations involved in
disaster risk management and humanitarian response are now beginning
to integrate climate change adaptation activities into their operational
programs (e.g., CARE International, 2010; Oxfam, 2011). One of the
longest established examples of civil society providing technical support
to CCA and DRM integration is the Red Cross/Red Crescent Climate
Centre. Alongside awareness raising and advocacy, the Centre analyzes
forecast information and integrates knowledge of climate risks into Red
Cross/Red Crescent strategies, plans, and activities, with a particular
focus on implementation at the community level (IFRC, 2011).
The various international civil society organizations working on DRR are
now also beginning to coordinate their operational support, and to
make explicit links to CCA (UNISDR, 2009a). The GNDR was launched in
2007, and constitutes over 300 organizations across 90 countries. It
Chapter 7 Managing the Risks: International Level and Integration across Scales
410
has three objectives of (1) influencing DRR public policy formulation
(development); (2) increasing public accountability for effective policy
administration (implementation); and (3) raising resources and political
will for community-based DRR (mobilization). One of the five core
strategies of the GNDR is to develop synergies between DRR and climate
change to address underlying risk factors (sustainable development),
including adapting local-level DRR monitoring infrastructure for climate
adaptation, and input to the UNFCCC COP negotiations. Given the
recent launch of the initiative there is no evaluation of effectiveness so
far.
7.3.3.3. International Finance Institutions and Donors
7.3.3.3.1. Global Environment Facility
The GEF is an independent financial organization established in 1991
that provides grants to developing countries and countries with
economies in transition for projects related to biodiversity, climate
change, international waters, land degradation, the ozone layer, and
persistent organic pollutants. It has become the largest funder of
projects to address global environmental challenges and it serves as the
financial mechanism for the following conventions:
Convention on Biological Diversity (CBD)
United Nations Framework Convention on Climate Change
(UNFCCC)
Stockholm Convention on Persistent Organic Pollutants (POPs)
UN Convention to Combat Desertification (UNCCD).
The GEF administers the main international funds that have been made
available under the UNFCCC for adaptation: the SCCF, which supports
adaptation alongside development, technology transfer, capacity
building, and sectoral approaches, and the LDCF, which particularly
focuses on the development and implementation of NAPAs in the least-
developed countries (LDCs). Ten international agencies [UNDP, the
United Nations Environment Programme, the World Bank, the Food and
Chapter 7Managing the Risks: International Level and Integration across Scales
Box 7-2 | Disaster Risk Management and Climate Change Adaptation
in the Context of International Development
Vulnerability to extreme weather and to climate change is strongly conditioned by socioeconomic development, including income levels
and distribution, supportive institutional frameworks, and the capacities of specific sectors. Conversely, the effects of climate change,
including through any increase in the frequency of extreme weather events, can also set back economic development (Stern, 2007).
Countries that are relatively poor, isolated, and reliant on a narrow range of economic activities are particularly vulnerable to such
shocks (UNISDR, 2009a). The objectives of climate change adaptation, disaster risk reduction, and sustainable development are therefore
intricately linked, and while the HFA and UNFCCC are the main international frameworks for CCA and DRR, a wider range of other
governance and institutional mechanisms have a major influence. These range, for example, from the agreements of the World Trade
Organization (affecting development and potentially technology transfer for adaptation; WTO, 2011), to the International Health
Regulations (affecting the way that epidemics of climate-sensitive infectious diseases such as cholera are managed across borders; WHO,
2007), to the codes of practice of international humanitarian organizations (such as the Code of Conduct for the International Red Cross
and Red Crescent Movement and NGOs in Disaster Relief; ICRC, 1995).
While approaches such as poverty reduction strategies are important in development planning at the national level, arguably the central
framework for defining global development objectives is the Millennium Declaration and the associated MDGs. These have been agreed
by all members of the United Nations as well as 23 international organizations, with a target date of 2015 (UN, 2011). These are also
supported by international aid agreements, such as the Multilateral Debt Relief Initiative to cancel US$ 40 to 55 million dollars’ worth of
debt (IMF, 2011), and the commitment of economically advanced countries to commit 0.7% of gross national income to overseas
development aid (UN ,1970). The eight MDGs break down into 21 quantifiable targets that are measured by 60 indicators (UN, 2011).
Neither DRM nor CCA are explicitly covered in the MDGs. However, they are strongly linked in practice. First, if disasters occur they can
set back progress across many of the goals. Second, progress toward the MDGs can help to increase resilience to extreme weather
events, and to climate change (Schipper and Pelling, 2006). Linking CCA and DRM with the MDGs is therefore important for the coherence
of international development, and the target date of the Hyogo Framework for Action coincides with the intended completion of the
MDGs (UNISDR, 2005b).
While there are exceptions, the majority of the LDCs, particularly in sub-Saharan Africa, are currently off track to reach most of the MDGs
(UN, 2011). This has been attributed in part to financial, structural, and institutional weaknesses in the affected countries, and also by
failure of most developed countries to reach the 0.7% aid target. Failure or delays in reaching the MDGs are therefore likely to be both a
cause and a consequence of vulnerability to extreme weather and climate change (UNISDR, 2005b).
411
Agriculture Organization (FAO), the Inter-American Development Bank
(IADB), the United Nations Industrial Development Organization, the
International Fund for Agricultural Development, the European Bank for
Reconstruction and Development (EBRD), and the African and Asian
Development Banks] implement GEF projects, usually in partnership
with national or other international agencies. Following a review of the
implementation of the LDCF Fund by the UNFCCC’s Subsidiary Body for
Implementation, parties to the UNFCCC have requested the GEF, inter
alia, to speed up the implementation process, update NAPAs, and work
with its implementing agencies to improve communication with LDCs
(UNFCCC, 2011). The GEF also provides interim secretariat services to
the Adaptation Fund, established under the Kyoto Protocol of the
UNFCCC, funded mainly through a percentage of the proceeds of the
Certified Emission Reductions under the Clean Development Mechanism
(Adaptation Fund, 2011a). The Fund finances climate change adaptation
projects, including DRR projects, in developing countries (Adaptation
Fund, 2011b).
7.3.3.3.2. The World Bank and Regional Development Banks
The major development banks (the African Development Bank, Asian
Development Bank, EBRD, IADB, and World Bank Group) manage much
of the funding for both climate change and disaster reduction. This
includes, for example, the Pilot Program for Climate Resilience, covering
a wide remit, including integration of climate risk and resilience into
development planning (World Bank, 2009; Climate Funds Update,
2011).
Perhaps the clearest example of the strengths and challenges of
international financing for DRM and CCA is provided by the Global
Facility for Disaster Reduction and Recovery (GFDRR), managed by the
World Bank. This is a partnership of the UNISDR system to support the
implementation of the HFA. The GFDRR’s mission is to mainstream
disaster reduction and climate change adaptation into national policies,
plans, and strategies to promote development and achieve the MDGs.
The World Bank provides operational services to the GFDRR, on behalf
of donors and other partnering stakeholders. The GFDRR supports
international collaboration, and provides technical and financial
assistance to low- and middle-income countries that are considered to
be at high risk from disasters (GFDRR, 2010).
Two independent evaluations of the GFDRR have been conducted
(Universalia Management Group, 2010; DFID, 2011). The facility has
mobilized significant funds (over US$ 240 million in contributions and
pledges from 2006 to 2009). The fund is considered relevant and
responsive to stakeholders, and to play a unique role in helping to
bridge knowledge, policy, and practice in DRR services, with good
coverage of climate change adaptation (Universalia Management
Group, 2010). It is also considered to be cost-effective in program
implementation (DFID, 2011). However, the resources that have been
mobilized through the fund remain much lower than those required,
and partnerships, policy integration, and monitoring of results are
considered uneven across countries. Despite these challenges, the
facility is considered to have achieved important progress, and to be
implementing the necessary steps to improve function and to scale up
implementation (Universalia Management Group, 2010; DFID, 2011).
7.4. Options, Constraints, and Opportunities
for Disaster Risk Management and
Climate Change Adaptation at the
International Level
7.4.1. International Law
As demonstrated in Section 7.2.5, existing tools and instruments of
international law can assist with disaster risk reduction and management
and in driving adaptation to climate change, recognizing at the same
time that international law is limited in scope and enforceability when
applied to addressing these challenges.
7.4.1.1. Limits and Constraints of International Law
Structurally, international law is both facilitated and constrained by the
need for explicit or implicit acceptance by nation states, which create
and comprise the system.It follows that the relevance of negotiated
treaties depends on state consent, while customary law only exists if
there is state practice and opinio juris. For instance, in the case of the
Tampere Convention on the Provision of Telecommunication Resources
for Disaster Mitigation and Relief Operations noted in Section 7.2.5,
only four of the 25 most disaster-prone states have signed up, limiting
its relevance to many of the states that would most benefit from its
provisions (Fisher, 2007). The International Bill of Rights, which at face
value is highly relevant to disaster risk response and in supporting an
obligation to assist with adapting to climate change, does not enjoy
universal acceptance. Furthermore, because international law is made
by and applicable to states, the many non-state actors relevant to
disaster risk reduction and climate change adaptation are not subject
to obligations – though as citizens they may benefit from the duty of
states.
Some fields of international law provide tools that seem applicable to
disaster risk management and/or adaptation to climate change, yet are
constrained through inherent limited applicability. International
humanitarian law (IHL) enshrined in the 1949 Geneva Conventions
enjoys wide applicability due to universal adherence (Lavoyer, 2006;
Fisher, 2007), but is limited to situations of armed conflict. In contrast,
‘International Disaster Response Law’ (IDRL) (see Fisher, 2007), sometimes
proposed as a peacetime counterpart to IHL, not only lacks the central
regime and universal adhesion of the Geneva Conventions, but further
experiences challenges in coordination and monitoring (Fisher, 2007).
As a second example, international law has on the one hand been
described as “not yet equipped to respond adequately to the diverse
causes of climate-induced migration” (Von Doussa et al., 2007; generally
Biermann and Boas, 2010), while on the other hand the literature is in
Chapter 7 Managing the Risks: International Level and Integration across Scales
412
disagreement as to whether refugee law should provide the instruments
to deal with the challenge of migration related to climate change. The
application of international refugee law, as codified in the 1951
Convention relating to the Status of Refugees, to those who cross
international borders due to climate-induced migration is indeed
complex and limited (UNHCR, 2009). Reopening the Convention to
expand the term ‘refugee,’ it is argued, would risk a renegotiation of the
Convention and thus potentially result in lower levels of protection for
the displaced (Kolmannskog and Myrstad, 2009).
7.4.1.2. Opportunities for the Application of International Law
The potential expansion of the concepts, definitions, and procedures
known to international law can also be seen as future opportunity for
international law to address the challenges of disaster risk reduction
and adaptation to climate change.
Beyond the current international law obligations to mitigate the effects of
climate change, facilitate disaster response, and mandate international
facilitation of adaptation efforts (see Section 7.2.5), the fact that
international law is shaped by nation states and evolves with state
practice means that international law may also adapt to future realities.
Expanding the interpretation and application of existing international
law, and the introduction of new law for disaster response and climate
change adaptation, are both plausible in the future.
A controversial candidate field for expanded interpretation is international
refugee law. The extant definition of ‘refugee’ per the Refugee Convention
and Protocol is any person who is outside their country of nationality
and who, “owing to a well-founded fear of being persecuted” is unable
or unwilling to return to their country. Some literature proposes the
expansion of ‘persecuted’ to encompass being subject to environmental
disaster or degradation (Warnock, 2007; Kolmannskog and Myrstad,
2009). Comparably, Article 7 of the International Covenant on Civil and
Political Rights prohibits torture and “cruel, inhuman, or degrading
punishment. Some literature notes the potential expansion of the
meaning ‘inhuman treatment’ to include being left without basic levels
of subsistence due to climate change impacts. A step further proposes a
new international agreement to share the “emerging burden of climate-
induced migration flows” and which “upholds the human rights of the
individuals affected” (Von Doussa et al., 2007). The expansion of the
definition of refugee remains highly controversial, with many states
opposing the use of refugee law to address climate-related, transboundary
movement of people.
The emerging legal doctrine of ‘responsibility to protect’ has also been
proposed in application to natural disasters. The emergence of state
practice in observing certain responsibilities “before, during, and after
natural disasters occur” in the absence of obligations to do so supports
an emerging responsibility to protect in the context of natural disaster,
and sources of human rights law are to be used in promoting this doctrine
(Saechao, 2007).
7.4.2. International Finance
The UNFCCC recognizes that in addition to the need to mitigate
emissions of greenhouse gases and adapt to climate change, there is a
responsibility on developed countries to support developing countries in
this process (see Article 4.4 in Box 7-1). A starting point for the delivery
of adaptation finance is the assessment of adaptation finance needs,
which have also been interpreted as a proxy for adaptation costs (see
Section 4.5). The UNFCCC (2007b) estimated the additional investment
and financial flows needed worldwide to be US$ 48 to 171 billion in 2030
(or US$ 60 to 193 billion when also considering current investment needs
for ecosystem adaptation). Some US$ 28 to 67 billion of this amount
would be needed in developing countries (UNFCCC, 2007b). The largest
uncertainty in these estimates is in the cost of adapting infrastructure,
which may require anything between US$ 8 and 130 billion in 2030,
one-third of which would be for developing countries. The UNFCCC
(2007b) also estimated that an additional amount of about US$ 41
billion would be needed for agriculture, water, health, and coastal zone
protection, most of which would be used in developing countries. Other
studies providing estimates of the annual incremental costs of adaptation
in developing countries include those by the World Bank (2006), Stern
(2007), Oxfam International (2007), Watkins (2007), and the World Bank
(2010b). These estimates are shown in Table 7-1, and discussed in more
detail in Parry et al. (2009) and Fankhauser (2010).
While these different estimates highlight the high level of uncertainty,
there appears to be consensus that global adaptation costs will total
tens of billions of US dollars per year in developing countries. A review
by the Organisation for Economic Co-operation and Development (OECD)
of the estimates mentioned above found that there is very little quantified
information on the costs of adaptation in developing countries, and
most studies are constrained to a few sectors within countries (mostly
coastal zones and, to a lesser extent, water, agriculture, and health)
(Agrawala and Fankhauser, 2008). In addition, these studies assume
relatively crude relationships and make strong assumptions, such as
perfect foresight and high levels of autonomous adaptation. Almost
no cross-sector studies have examined cumulative effects within
countries, and only a handful of studies have investigated the wider
macroeconomic consequences of impacts or adaptation. Moreover,
most of the literature only considers adaptation to average changes in
temperature or sea level rise. Little attention has been paid to more
Chapter 7Managing the Risks: International Level and Integration across Scales
Assessment
Year
US$
(Billion)
Time Frame
World Bank 2006 - Present
Stern 2006 Present
Oxfam 2007 Present
UNDP 2007 2015
UNFCCC 2007 2030
World Bank 2010
S
ources: World Bank, 2006, 2010b; Stern, 2007; Oxfam International, 2007; UNFCCC, 2007b; Watkins,
2
007.
9 - 41
4 - 37
28 - 67
> 50
86 - 109
70 - 100
2010 - 2050
Table 7-1 |
Estimated annual adaptation costs and finance needs in developing countries.
413
abrupt changes in mean conditions or to changes in the frequency and
magnitude of extreme events (Agrawala and Fankhauser, 2008).
According to Agrawala and Fankhauser (2008), the consensus on global
adaptation costs, even in order of magnitude terms, may therefore be
premature. In addition, in most cases the estimates are neither attributed
to specific adaptation activities, nor do they articulate the benefits of
adaptation investment. Double counting between sectors and scaling
up to global levels from very limited (and often local) source material
limit utility. At the same time, a point also noted by Parry et al. (2009),
many sectors and adaptations have not been included in the estimates.
In addition to these global estimates, total adaptation finance needs can
also be assessed by aggregating national estimates, although this is
hampered by the absence of a common method to make such estimates,
and the fact that they are not available for all countries. The NAPAs (see
Section 7.3.2 and Chapter 6), which have now been completed by most
LDCs, are the most extensive effort to date to assess adaptation priorities
and finance needs in developing countries. The cumulative cost of
projects prioritized to respond to urgent and immediate adaptation
needs is approximately US$ 1,660 million for the 43 countries that had
completed their NAPAs by September 2009 (UNFCCC, 2010b). The
divergence from the global estimates mentioned above can be
explained by several factors: they cover only 43 LDCs, they include only
prioritized projects, and they consider only urgent and immediate
adaptation needs, not medium- to long-term needs (Persson et al., 2009).
A challenge for the international community is how to meet the adaptation
finance needs that have been identified. The GEF operates the LDCF and
SCCF, to provide funding to eligible developing countries to meet the
‘additional’ or ‘incremental’ costs of adaptation; the baseline costs of a
project or program are borne by the recipient country, by other bilateral
or multilateral donors, or both. The LDCF and SCCF rely on voluntary
contributions from developed countries. As of May 2010, US$ 315 million
had been pledged for adaptation under these two funds (US$ 221 million
to the LDCF and US$ 94 million to the SCCF); of this amount, US$ 220
million has been allocated (US$ 135 million from the LDCF and US$ 85
million from the SCCF) (GEF, 2010a). In addition, the GEF has allocated
all US$ 50 million it had made available to the SPA (GEF, 2008; see also
Klein and Möhner, 2009).
The Adaptation Fund, which became operational in 2009, is operated by
a special Adaptation Fund Board. It is the first financial instrument
under the UNFCCC and its Kyoto Protocol that is not based solely on
voluntary contributions from developed countries. It receives a 2%
share of proceeds from project activities under the Clean Development
Mechanism (CDM), but can also receive funds from other sources to
fund concrete adaptation projects and programs (Persson et al., 2009).
The actual amount of money that will be available from the Adaptation
Fund depends on the extent to which the CDM is used and on the
price of carbon. As of October 2010, the Adaptation Fund had received
US$ 202.09 million, of which US$ 130.55 million was generated through
CDM activities. Estimates of potential resources available for the
Adaptation Fund from 31 October 2010 to 31 December 2012 range
from US$ 288.4 million to US$ 401.5 million (Adaptation Fund, 2010).
While the GEF-managed funds have supported adaptation activities in
some 80 countries (Persson et al., 2009), there has been criticism,
particularly from developing countries, on how the funds are being
managed (e.g., Mitchell et al., 2008; Klein and Möhner, 2009; Ministry
of Foreign Affairs of Denmark and GEF Evaluation Office, 2009). In
addition, concern has been voiced about the predictability and adequacy
of funds, and the perceived equity and fairness of decisionmaking (Mace,
2005; Paavola and Adger, 2006; Müller, 2007; Persson et al., 2009). The
GEF has acknowledged the criticism and indicated in reports to the
COP how it is responding to it (GEF, 2009, 2010b). At the same time,
developed countries have raised concern about fiduciary risks in some
developing countries, which would need to be addressed through
improved accountability and transparency before program-based
adaptation can be supported by international finance (Mitchell et al.,
2008; GEF, 2010b). The Adaptation Fund has not been operational long
enough to allow for such an assessment but the first signals are positive,
particularly regarding its governance structure and the option of direct
access (Czarnecki and Guilanpour, 2009; Brown et al., 2010; Grasso,
2010).
In addition to the funds operating within the context of the UNFCCC,
money for adaptation is provided through several other channels,
including developing countries’ domestic national, sectoral, and local
budgets; bilateral and multilateral development assistance; and private-
sector investments. This makes for an adaptation financing landscape
that is highly fragmented, resulting in a proliferation not only of funds but
also of policies, rules, and procedures (Persson et al., 2009). But despite
the proliferation of funds, the amount of money currently available falls
substantially short of the adaptation finance needs presented above.
In light of this shortfall, the 2009 Copenhagen Accord was a milestone
in international climate finance. It refers to a collective commitment for
developed countries to provide “new and additional resources …
approaching USD 30 billion” in ‘fast start’ money for the 2010-2012
period, balanced between adaptation and mitigation, and sets a longer-
term collective goal of mobilizing US$ 100 billion per year by 2020 from
all sources (public and private, bilateral and multilateral) (Bodansky,
2010). Although the Copenhagen Accord was not adopted by the COP,
the collective commitment and longer-term goal are also part of the
Cancun Agreements, which the COP adopted a year later. Parties agreed
that “scaled-up, new and additional, predictable and adequate funding
shall be provided to developing country Parties, taking into account the
urgent and immediate needs of developing countries that are particularly
vulnerable to the adverse effects of climate change. In the meantime,
the High-level Advisory Group on Climate Change Financing, established
by the UN Secretary-General, had analyzed the feasibility of mobilizing
US$ 100 billion per year by 2020. It concluded that “it is challenging but
feasible to meet that goal. Funding will need to come from a wide
variety of sources, public and private, bilateral and multilateral, including
alternative sources of finance, the scaling up of existing sources, and
Chapter 7 Managing the Risks: International Level and Integration across Scales
414
increased private flows. Grants and highly concessional loans are crucial
for adaptation in the most vulnerable developing countries, such as the
least developed countries, small island developing States and Africa”
(AGF, 2010).
An open question is how climate finance might be linked with other
international finance flows. The Bali Action Plan referred to “means to
incentivize the implementation of adaptation actions on the basis of
sustainable development policies” in its section on the provision of
financial resources. The Copenhagen Accord did not discuss the link
between adaptation and development, even though the issue of
‘mainstreaming’ – integrating adaptation to climate change into
mainstream development planning and decisionmaking – was much
debated in the pre-Copenhagen negotiations on adaptation finance
(Persson et al., 2009; Klein, 2010). From an operational perspective,
mainstreaming adaptation into development makes common sense:
both contribute to enhancing human security, and opportunities to
create synergies between the two are increasingly recognized and
pursued (Gigli and Agrawala, 2007; Klein et al., 2007; Kok et al., 2008;
Gupta and Van de Grijp, 2010). Besides, there is a range of activities
that can be seen as contributing to both adaptation and development
objectives (McGray et al., 2007).
But from a climate policy perspective, mainstreaming creates a dilemma
(Persson and Klein, 2009; Klein, 2010). Financial flows for adaptation
and those for development – for example, official development assistance
(ODA) – are managed separately. One of the arguments in favor of
mainstreamed adaptation is that it makes more efficient use of financial
and human resources than adaptation that is designed, implemented,
and managed as stand-alone activities (i.e., separately from ongoing
development planning and decisionmaking). However, developing
countries have expressed the concern that, as a result of donors seeking
to create synergies between adaptation and development, finance for
adaptation will not be new and additional but in effect will be absorbed
into ODA budgets of a fixed size (Michaelowa and Michaelowa, 2007).
The concern is fueled by the fact that the amount of money currently
available for adaptation falls short of the estimated adaptation finance
needs in developing countries. A second, related concern is that
mainstreaming could divert any new and additional funds for adaptation
into more general development activities, thus limiting the opportunity
to evaluate, at least quantitatively, their benefits with respect to climate
change specifically (Yamin, 2005). Third, there is concern that donors’ use
of ODA to pursue mainstreamed adaptation could impose conditionalities
on what should be a country-driven process (Gupta et al., 2010).
As mentioned in Section 7.3.2, the Cancun Agreements established the
Green Climate Fund as a new entity operating the financial mechanism
under Article 11. The Green Climate Fund is not yet operational and it is
too early to say how it might address the mainstreaming dilemma, or
even how important it will be for climate adaptation in developing
countries. All that can be said at this moment is that in the Cancun
Agreements, Parties decided that “a significant share of new multilateral
funding for adaptation should flow through the Green Climate Fund.
7.4.3. Technology Transfer and Cooperation
7.4.3.1. Technology and Climate Change Adaptation
Technologies receive prominent attention both in adaptation to emerging
and future impacts of climate change as well as in mitigating current
disasters. The sustainability, operation, and maintenance of technologies
can be challenging in many developing countries due to lack of
resources, human capacity, and cultural differences. Moreover, technology
transfer is complex and requires capacity building as well as a client
(technology user) focus as opposed to a developer (technology designer)
focus (O’Brien et al., 2007). Intellectual property rights are rarely an
issue in the availability and use of technologies for adaptation (Murphy,
2011) but when they are, adequate methods are needed that foster
affordable deployment of new technologies but preserve the incentives
for technology developers (Doig, 2008). While the importance of
transferring technologies from developers/owners to would-be users is
widely recognized, the bulk of the literature seems to address the issues
at a rather generic level, without going into the details of what
technologies for adaptation would need to be transferred in different
impact sectors from where to where and via what mechanisms.
Institutional, political, technological, economic, information, financial,
cultural, legal, and participation and consultation obstacles can hinder
the transfer of mitigation and adaptation technologies and concerted
efforts are required to overcome those impediments (IEA, 2001). Private-
public partnership as a policy instrument could well be a mechanism
for transferring the required technologies for adaptation projects
(Agrawala and Fankhauser, 2008). In the adaptation literature,
publications addressing the transfer of technologies important for
reducing vulnerability and increasing the ability to cope with weather-
related disasters are even scarcer. This section reviews literature on
technologies for adaptation and the issues involved in international
technology transfer of such technologies.
The Special Report on Methodological and Technological Issues in
Technology Transfer by the IPCC defines the term ‘technology transfer’
as a “broad set of processes covering the flows of know-how, experience
and equipment for mitigating and adapting to climate change amongst
different stakeholders such as governments, private sector entities,
financial institutions, NGOs and research/education institutions” (IPCC,
2000, p. 3). The report uses a broad and inclusive term ‘transfer’
encompassing diffusion of technologies and technology cooperation
across and within countries. It evaluates international as well as domestic
technology transfer processes, barriers, and policies. This section focuses
on the international aspects.
Adaptation to climate change involves more than merely the application
of a particular technology (Klein et al., 2005). Adaptation measures
include increasing robustness of infrastructural designs and long-term
investments, increasing flexibility of vulnerable managed systems,
enhancing adaptability of natural systems, reversing trends that increase
vulnerability, and improving societal risk awareness and preparedness.
In the case of disasters related to extreme weather events, anticipatory
Chapter 7Managing the Risks: International Level and Integration across Scales
415
adaptation is more effective and less costly than emergency measures
and retrofitting, and immediate benefits can be gained from better
adaptation to climate variability and extreme events. Some factors that
determine adaptive capacity of human systems are the level of economic
wealth, access to technology, information, knowledge and skills, and
existence of institutions, infrastructure, and social capital (Smit et al.,
2001; Christoplos et al., 2009).
An extensive list of ‘soft’ options that are vital to building capacity to
cope with climatic hazards with references to publications that either
describe the technology in detail or provide examples of its application
is available (Klein et al., 2000, 2005). For example, the applications in
coastal system adaptation include various types of geospatial information
technologies such as mapping and surveying, videography, airborne
laser scanning (lidar), satellite and airborne remote sensing, global
positioning systems in addition to tide gauges and historical and
geological methods. These technologies help formulate adaptation
strategies (protection versus retreat), implement the selected strategy
(design, construction, and operation), and provide early warning
(UNFCCC, 2006a). Another set of examples includes technologies to
protect against sea level rise: dikes, levees, floodwalls, seawalls,
revetments, bulkheads, groynes, detached breakwaters, floodgates, tidal
barriers, and saltwater intrusion barriers among the hard structural
options, and periodic beach nourishment, dune restoration and creation,
and wetland restoration and creation as examples of soft structural
options (Klein et al., 2000, 2005). A combination of these technologies
selected on the basis of local conditions constitutes the protection
against extreme events in coastal regions. Structural measures are
localized solutions and there is a need for localized information such as
their environmental and hydrologic impacts. In addition, there are a
series of indigenous options (flood and drought management) that
might be valuable in regions to be affected by similar events (Klein et
al., 2005, p. 19). It is also important to integrate technology transfer
efforts for CCA and DRR needs with sustainable development efforts to
avoid conflicts and foster synergies between them (Hope Sr., 1996; Sanusi,
2005). Adaptation is normally assumed to be benign for development
but Eriksen and Brown (2011) challenge this assumption, arguing that
there is emerging evidence that adaptation measures run counter to
principles of sustainable development, as both social equality and
environmental integrity can be threatened. Placing responses to extreme
events into the larger context of other societal and environmental
changes will be vital for sustainable development (Yohe et al., 2007;
Eriksen et al., 2011).
A report by the UNFCCC (2006a) summarizes the technology needs
identified by Parties not included in Annex I to the Convention.
Curiously, only one country mentioned ‘potential for adaptation’ among
the commonly used criteria for prioritizing technology needs. Among
30 technologies listed in the report, the technology needs relevant for
coping and adapting to weather extremes include, for example,
improved drainage, emergency planning, raising buildings and land, and
protecting against sea level rise. Many of these are good examples of
measures that link DRR and CCA objectives, namely to reduce overall
ecological and social vulnerability. Another UNFCCC report (2006b)
observes that, unlike those for mitigation, the forms of technology for
adaptation are often rather familiar. Many have been used over
generations in coping with floods, for example, by building houses on
stilts or by cultivating floating vegetable plots. Some other types of
technologies draw on new developments in, for example, advanced
materials science and satellite remote sensing (see Box 7-3). The
UNFCCC report (2006b) provides an overview of the old and new
technologies available in adapting to changing environments, including
climate change. The Disaster Reduction Hyperbase in Asia is a web-
based collection of new and traditional indigenous technologies
relevant to DRM that also promotes communication among developing
and industrial countries (Kameda, 2007).
Chapter 7 Managing the Risks: International Level and Integration across Scales
Box 7-3 | Examples of Technologies for Adaptation in Asia
In Asia, adaptation to climate change, variability, and extreme events at the community level are small scale and concentrate mainly on
agriculture, water, and disaster amelioration (Alam et al., 2007). They focus on the livelihood of affected communities, raise awareness to
change practices, diversify agriculture, and promote water conservation. For example, Saudi Arabia has already built 215 dams for water
storage and 30 desalination plants, passed water protection and conservation laws, and initiated leakage detection and control schemes
as well as advanced irrigation water conservation schemes and a system for modified water pumping as part of its climate change
adaptation program (Alam et al., 2007). In India, a combination of traditional and innovative technological approaches is used to manage
drought risk. Technological management of drought (e.g., development and use of drought tolerant cultivars, shifting cropping seasons
in agriculture, flood and drought control techniques in water management) is combined with model-based seasonal and annual to
decadal forecasts. Model results are translated into early warning in order to take appropriate drought protection measures (Alam et al.,
2007). In China, adaptation technologies have been widely used for flood disaster mitigation (Alam et al., 2007). Another example is
related to the Philippines where a typhoon in 1987 completely destroyed over 200,000 homes. The Department of Social Welfare and
Development initiated a program of providing typhoon-resistant housing for the population in the most typhoon-prone areas (Diacon,
1992). The so-called Core Shelter houses have typhoon resistant features and can endure wind speeds up to 180 km hr
-1
. The technology
was proved to be successful by providing the required protection and was adopted recently in regions stricken by a landslide
(Government of the Philippines, 2008) and typhoons (Government of the Philippines, 2010), partly financed by UNDP.
416
7.4.3.2. Technologies for Extreme Events
Approaching the issues of technologies to foster adaptation to extreme
weather events and their impacts from the direction of disaster mitigation,
Sahu (2009) presents an overview of diverse technologies that might be
applied in various stages of disaster management. The list of technologies
for adaptation to weather-related extreme events includes early warning
and disaster preparedness; search and rescue for disaster survivors;
water supply, purification, and treatment; food supply, storage, and safety;
energy and electricity supply; medicine and healthcare for disaster victims;
disease surveillance; sanitation and waste management; and disaster-
resistant housing and construction (Sahu, 2009).
Developing wind-resistant building technologies is crucial for reducing
vulnerability to high-wind conditions like storms, hurricanes, and
tornadoes. A report by the International Hurricane Research Centre
presents hurricane loss reduction devices and techniques (IHRC, 2006).
The Wall of Wind testing apparatus (multi-fan systems that generate up
to 209 km hr
-1
winds and include water-injection and debris-propulsion
systems with sufficient wind field sizes to test the construction of small
single-story buildings) will improve the understanding of the failure
mode of buildings and hence lead to technologies and products to
mitigate hurricane impacts (Fugate and Crist, 2008).
An absolutely crucial aspect of managing weather extremes both under
the present and future climate regime is the ability to forecast and
provide early warning. Downscaling projections from global climate
models could provide useful information about the changing risks. It is
important to note that really useful early warning systems would provide
multi-hazard warning and warnings on vulnerability development to the
extent it is possible. Satellite and aerial monitoring, meteorological
models, and computer tools including geographic information systems
(GIS) as well as local and regional communication systems are the most
essential technical components. (The focus on technology here does not
negate the importance of social and communication aspects of early
warning.) The use of GIS in the support of emergency operations in the
case of both weather and non-weather disasters is becoming increasingly
important in the United States. The benefits of using GIS technologies
include informing the public, enabling officials to make smarter decisions,
and facilitating first-responder efforts to effectively locate and rescue
storm victims (NASCIO, 2006). Lack of locally useable climate change
information about projected changes in extreme weather events
remains an important constraint in managing weather-related disasters,
especially in developing countries. Therefore there is a need to develop
regional mechanisms to support in developing and delivering downscaling
techniques and tools (see Section 3.2.3 for details on downscaling
regional climate models) and transferring them to developing countries.
Space technologies (such as Earth observation, satellite imagery, real-time
application of space sensors, mapping) are important in the reduction of
disasters, including extreme weather and climate events such as drought,
flood, and storms (Rukieh and Koudmani, 2006). These technologies can
be particularly helpful in the risk assessment, mitigation, and preparedness
phases of disaster management by identifying risk-prone areas,
establishing zoning restrictions and escape routes, etc. Space technologies
are important for early warning and in managing the effects of disasters.
For incorporating the routine use of space technology-based solutions
in developing countries, there is a need to increase awareness, build
national capacity, and also develop solutions that are customized and
appropriate to their needs (Rukieh and Koudmani, 2006). A good example
of the application of space technology at international scales and for
early warning is the joint initiative of WMO, the National Oceanic and
Atmospheric Administration, the US Agency for International Development,
and the Hydrologic Research Center on global flash flood guidance. The
system uses global data produced by a global center and downscales the
global information to regional products that are sent to national entities
for further downscaling at the national level and then disseminated to
users and communities (WMO, 2007, 2010). It is also important to note
that there are existing capabilities within some particularly exposed
developing countries (such as India, Bangladesh, China, Philippines)
with well-developed remote-sensing capabilities of their own, or existing
arrangements with other space agency suppliers.
Support for risk reduction and relief agencies and governments depends,
among other factors, on timely availability of information about the scale
and nature of these disasters (Holdaway, 2001). Currently, ground-based
sources provide most of such information. This could be improved by
using input from space-based sensor systems, both for disaster
warning and disaster monitoring where the scale of devastation cannot
adequately be monitored from ground-based information sources alone.
A global space-based monitoring and information system, with the
associated ability to provide advanced warning of many types of hazards,
can be combined with the latest developments in sensor technology
(optical, infrared, radar) including a UK initiative on high-resolution
imaging from a microsatellite (Holdaway, 2001). The literature suggests
that transferring these technologies and the related know-how will be
important for building capacities in CCA and DRR in countries where
they are still missing (medium certainty, limited evidence).
Microsatellites (low weights and small sizes, just under or well below
500 kg) are seen as an important technology for the detection of and
preparation for weather-related hazards in other countries as well.
Shimizu (2008) emphasizes the importance of international cooperation
in this area. He observes that only a few countries are able to develop
large rockets and satellites and launch them from their own territories.
Several Asian countries have been cooperating with OECD countries to
develop small Earth observation satellites, like DAICHI (Advanced Land
Observing Satellite) and WINDS (Wideband Internetworking engineering
test and demonstration satellite) that include both optical and microwave
sensors. DAICHI operated between 2006 and 2011 based on cooperation
of Asian countries with the United States and the European Union and
made an important contribution to emergency observations of regions
hit by major disasters in this period (JAXA, 2011).
Mitigation of adverse cyclone impacts involves reliable tropical cyclone
forecasting and warnings, and efficient ways to convey the information
Chapter 7Managing the Risks: International Level and Integration across Scales
417
to stakeholders, users, and the general public (Lee et al., 2006). It is
important that NMHSs take advantage of the advances in communication
technology such as wireless broadband access, Global Positioning System
(GPS), and GIS to enhance the relevance and effectiveness of warnings,
options, and backup capabilities to disseminate warnings through
multiple and diverse channels (Lee et al., 2006). Natural hazards
management has advanced to address a major challenge: turning real-
time data provided by new technologies (e.g., satellite- and ground-based
sensors and instruments) into information products to help people make
better decisions about their own safety and prosperity (Groat, 2004).
The literature about technology transfer to foster adaptation to changes
in extreme events induced by climate change is very limited. It was
necessary to broaden the scope of the literature review and embrace
climate change adaptation in general in order to gain lessons about the
processes, channels, stakeholders, and barriers of technology transfer. In
addition, useful insights were inferred from the literature on technology
transfer to support climate change mitigation, disaster risk reduction
(prevention, mitigation, and preparedness), and other related areas. The
DRR literature on technology development and transfer documents the
expanding international cooperation in forecasting and monitoring
extreme weather events by collecting and disseminating satellite-based
information and the international transfer of know-how to interpret it.
There is increasing emphasis on the importance of establishing close
linkages across all EWS components ranging from collection of hydro-
meteorological data, forecasting how nature will respond (e.g., weather
or flood forecasting), to communicating information (or warnings) to
decisionmakers (sectoral users or communities) (medium agreement,
limited evidence).
7.4.3.3. Financing Technology Transfer
Climate change mitigation has been the primary focus of the financing
mechanisms and innovative financing in recent years. In contrast, the
transfer of technologies for adaptation is hampered by insufficient
incentive regimes, increased risks, and high transaction costs (Klein et
al., 2005). Yet the lessons from the transfer of mitigation technologies
are relevant for adaptation: results of the penetration of energy and
industrial technologies in the developing countries depend on many
factors ranging from labor skills, market conditions, achieved level of
technological development, the reliability of basic services (electricity
and water), availability of spare parts, etc. A combination of interrelated
socioeconomic, institutional, and governance issues would often
determine the success or failure of technology transfer, rather than the
technologies themselves (Klein et al., 2005, p. 23). These factors are also
important in transferring technologies for adaptation because they
determine the feasibility and efficiency of adopting the transferred
technologies (e.g., regulations to build and install them, skilled labor,
water and electricity to operate them).
UNFCCC (2005) addresses the transfer of environmentally sound
technologies for adaptation to climate change: the needs for and the
identification and evaluation of technologies for adaptation to climate
change, and financing their transfer. Cost is one of the main barriers in
technology transfer; therefore innovative financing for the development
and transfer of technologies is needed. Potential sources of funding for
technology transfer include bilateral activities of Parties, multilateral
activities such as the GEF, the World Bank, or regional banks, the SCCF,
the LDCF, financial flows generated by Joint Implementation and CDM
projects, and the private sector (see also Section 7.3.3.3). The GEF funds
for adaptation activities include the SPA trust fund, the LDCF, and the
SCCF. In addition, the GEF is providing secretariat services to the
Adaptation Fund Board under the Kyoto Protocol (see also Section
7.4.2).
Climate variability is already a major impediment to development and
2% of World Bank funds are devoted to disaster reconstruction and
recovery (World Bank, 2008). In order to use available funds efficiently,
the World Bank (2009) developed the screening tool ADAPT
(Assessment & Design for Adaptation to Climate Change: A Prototype
Tool), a software-based tool for assessing development projects for
potential sensitivities to climate change. The tool combines climate
databases and expert assessments of the threats and opportunities arising
from climate variability and change. As of 2011, the knowledge areas
covered by the tool cover agriculture and irrigation in India and sub-
Saharan Africa and, for all regions, various aspects of biodiversity and
natural resources.
Both conventional and innovative options for financing the transfer of
technologies for adaptation might be explored. As conventional options,
the GEF funds (SPA, LDCF, and SCCF) provide opportunities for accessing
financial resources that could be used for deployment, diffusion, and
transfer of technologies for adaptation, including initiatives on capacity
building, partnerships, and information sharing. Projects identified in
technology needs assessments could also be implemented using these
financial opportunities. Based on these experiences as well as on
special needs of groups of countries such as small island developing
states and LDCs, further guidance could be provided to the GEF on
funding technologies for adaptation. In addition, there is an opportunity
to explore innovative financing mechanisms that can promote, facilitate,
and support increased investment in technologies for adaptation
(UNFCCC, 2005).
Concerning financing of technological development and transfer, a
report by the Expert Group on Technology Transfer (UNFCCC, 2009a)
classifies technologies by stage of maturity, the source of financing
(public or private sector), and whether they are under or outside the
UNFCCC and estimates the financing resources currently available for
technology research, development, deployment, diffusion, and transfer.
The estimates for financing mitigation technologies are between US$ 70
and 165 billion per year. In the adaptation area, the report claims that
research and development is focused on tailoring technologies to specific
sites and applications and thus the related expenditures become part of
the project costs. Current spending on adaptation projects in developing
countries is about US$ 1 billion per year (UNFCCC, 2009a).
Chapter 7 Managing the Risks: International Level and Integration across Scales
418
The literature clearly shows that the transfer of technologies for
adaptation lags behind the transfer of mitigation technologies in terms
of the scales of attention and funding. Funding transfer and funding
mechanisms for technologies that help reduce vulnerability to climate
variability, particularly to weather-related extreme events, appear to be
as important for both CCA and DRR (high confidence).
7.4.4. Risk Sharing and Transfer
This section examines the current and potential role of the international
community – international financial institutions, NGOs, development
organizations, private market actors, and the emerging adaptation
community – in enabling access to insurance and other financial
instruments that share and transfer risks of extreme weather. The
international transfer and sharing of risk is an opportunity for individuals
and governments of all countries that cannot sufficiently diversify their
portfolio of weather risk internally, and especially (as discussed in
Section 6.3.3) for governments of vulnerable countries that do not wish
to rely on ad hoc and often insufficient post-disaster assistance.
Experience shows that the international community can play a role in
enabling individual, national, and international risk-sharing and transfer
strategies (high confidence). The following discussion identifies successful
practices, or value added, as well as constraints on this role.
7.4.4.1. International Risk Sharing and Transfer
Risk transfer (usually with payment) and risk sharing (usually informal
with no payment) are recognized by the international community as an
integral part of DRM and CCA (see Case Study 9.2.13 for definitions).
The 2005 HFA calls on the disaster community “to promote the
development of financial risk-sharing mechanisms, particularly insurance
and reinsurance against disasters” (UNISDR, 2005a, p. 11). Similarly, the
2007 Bali Action Plan calls for consideration of risk-sharing and transfer
mechanisms as a means for enhancing adaptation (UNFCCC, 2007a).
The Plan builds on the mandate to consider insurance as set out by
Article 4.8 of the UNFCCC and Article 3.14 of the Kyoto Protocol.
Often by necessity risk sharing and transfer are international. Local and
national pooling arrangements (discussed in Sections 5.5.2 and 6.3.3)
may not be viable for statistically dependent (co-variant) risks that cannot
be sufficiently diversified. A single event can cause simultaneous losses
to many insured assets, violating the underlying insurance principle of
diversification. For this reason, primary insurers, individuals, and
governments (particularly in small countries) rely on risk-sharing and
transfer instruments that diversify their risks regionally and even globally.
A few examples can serve to illustrate international arrangements:
A government receives international emergency assistance and
loans after a major disaster.
A family locates a relative in a distant country who provides post-
disaster relief through remittances.
After a major disaster, a farm household takes out a loan from an
internationally backed micro-lending institution.
An insurer purchases reinsurance from a private reinsurance
company, which spreads these risks to its international shareholders.
A government issues a catastrophe bond, which transfers risks
directly to the international capital markets.
Many small countries form a catastrophe insurance pool, which
diversifies risks and better enables them to purchase reinsurance.
Not only are these financial arrangements international in character, but
many are supported by the international development and climate
adaptation communities (see, especially, UNISDR, 2005b; UNFCCC,
2009b). At the outset it is important to point out that these instruments
cannot stand alone but must be viewed as part of a risk management
strategy, for which cost-effective risk reduction is a priority.
7.4.4.2. International Risk-Sharing and Transfer Mechanisms
This section reviews international mechanisms for sharing and transferring
risk, including remittances, post-disaster credit, insurance and reinsurance,
alternative insurance mechanisms, and regional pooling arrangements.
7.4.4.2.1. Remittances
Remittances – transfers of money from foreign workers or expatriate
communities to their home countries – make up a large part of informal
risk sharing and transfer, even exceeding official development aid flows.
In 2010, the official worldwide flow of remittances was estimated at
US$ 325 billion, and unrecorded flows may add another 50% or more.
In some cases, remittances can be as large as one-third of the recipient
country’s gross domestic product (World Bank, 2011b).
A number of studies show that remittances increase substantially
following disasters, often exceeding post-disaster donor assistance
(Lucas and Stark, 1985; Miller and Paulson, 2007; Yang and Choi, 2007;
Mohapatra et al., 2009). Payments can be sent through professional
money transfer organizations, but often these channels break down and
remittances are carried by hand (Savage and Harvey, 2007). While simple
in concept, remittances can be complicated by associated transfer fees.
A survey carried out in the United Kingdom found that for an average-
sized transfer, the associated costs could vary between 2.5 and 40%
(DFID, 2005). Information pertinent to the transfer is often obscure or in
an unfamiliar language, and transfers across some borders have been
complicated due to initiatives taken by developed nations to counter
international money laundering and terrorism financing (Fagen and
Bump, 2006). Finally, a major problem is difficulties in communicating
with relatives abroad, as well as the high potential of losing necessary
documents in a disaster.
The international community has been active in reducing the costs and
barriers to post-disaster remittances. DFID, among other development
Chapter 7Managing the Risks: International Level and Integration across Scales
419
organizations, supports financial inclusion policies including mobile
banking and special savings accounts earmarked for disaster recovery
that will greatly reduce transaction costs. High-tech proposals for assuring
security have included biometric identification cards and retina scanners
as forms of identification (DFID, 2005; Pickens et al., 2009).
7.4.4.2.2. Post-disaster credit
One of the most important post-disaster financing mechanisms, credit
provides governments and individuals with resources after a disaster, yet
with an obligation to repay at a later time. Governments and individuals
of highly vulnerable countries, however, can have difficulties borrowing
from commercial lenders in the post-disaster context. Since the early
1980s, the World Bank has thus initiated over 500 loans for recovery and
reconstruction with a total disbursement of more than US$ 40 billion
(World Bank, 2006), and the Asian Development Bank also reports large
loans for this purpose (Arriens and Benson, 1999). With the growing
importance of pre-disaster planning, a recent innovation on the part of
international organizations is to make pre-disaster contingent loan
arrangements – for example, the World Bank’s catastrophe deferred
drawdown option, which disburses quickly after the government
declares an emergency (World Bank, 2008).
For micro-finance institutions (MFIs), post-disaster lending has associated
risks given increased demand that tempts relaxed loan conditions or
even debt pardoning. This risk is particularly acute in vulnerable regions.
Recognizing the need for a risk transfer instrument to help MFIs remain
solvent in the post-disaster period, the Swiss State Secretariat for
Economic Affairs (SECO) and the IADB, as well as private investors,
created the Emergency Liquidity Facility (ELF) (UNFCCC, 2008). Located
in Costa Rica, ELF provides needed and immediate post-disaster liquidity
at low rates to MFIs across the region.
7.4.4.2.3. Insurance and reinsurance
Insurance is an instrument for distributing disaster losses among a pool
of at-risk households, farms, businesses, and/or governments, and is the
most recognized form of international risk transfer. The insured share of
property losses from extreme weather events has risen from a negligible
level in the 1950s to approximately 20% of the total in 2007 (Mills,
2007).
Insurance and reinsurance markets attract capital from international
investors, making insurance an instrument for transferring disaster risks
over the globe. The market is highly international in character, yet
uneven in its cover. In the period 2000 to 2005, for example, US insurers
purchased reinsurance annually from more than 2,000 different non-US
reinsurers (Cummins and Mahul, 2009, p. 115). From 1980 through 2003,
insurance covered 4% of total losses from climate-related disasters
(estimated at about US$ 1 trillion) in developing countries compared to
40% in high-income countries (Munich Re, 2003).
The international community is playing an active role in enabling
insurance in developing countries, particularly by supporting micro- and
sovereign (macro) insurance initiatives. The following four examples
illustrate this role:
The World Bank and World Food Programme provided essential
technical assistance and support for establishing the Malawi pilot
micro-insurance program (see discussion in Section 5.5.2), which
provides index-based drought insurance to smallholder farmers
(Hess and Syroka, 2005; Suarez et al., 2007).
The Mongolian government and World Bank support the
Mongolian Index-Based Livestock Insurance Program (see Section
5.5.2) by absorbing the losses from very infrequent extreme events
(over 30% animal mortality) and providing a contingent debt
arrangement to back this commitment, respectively (Skees and
Enkh-Amgalan, 2002; Skees et al., 2008).
The World Food Programme successfully obtained an insurance
contract through a Paris-based reinsurer to provide insurance to
the Ethiopian government, which assures capital for relief efforts
in the case of extreme drought (Hess, 2007).
The governments of Bermuda, Canada, France, and the United
Kingdom, as well as the Caribbean Development Bank and the
World Bank, have recently pledged substantial contributions to
provide start-up capital for the Caribbean Catastrophe Risk
Insurance Facility (discussed in Section 7.4.4.2.5) (Cummins and
Mahul, 2009).
These early initiatives, especially micro-insurance schemes, are showing
promise in reaching the most vulnerable, but also demonstrate significant
challenges to scaling up current operations. Lack of data, regulation,
trust, and knowledge about insurance, as well as high transaction costs,
are some of the barriers (Hellmuth et al., 2009).
As discussed in Case Study 9.2.13, insurance and other risk transfer
instruments can promote DRR and CCA in multiple ways by providing
the means to finance recovery, thus reducing long-term losses; adding
to knowledge about risks; creating incentives (and imperatives) for
risk reduction; and providing the safety net necessary for farms and
businesses to take on cost-effective, yet risky, investments that reduce
their vulnerability to climate change (Linnerooth-Bayer et al., 2009;
Warner et al., 2009b).
7.4.4.2.4. Alternative insurance instruments
Alternative insurance-like instruments, sometimes referred to as risk-
linked securities, are financing devices that enable risk to be sold in
international capital markets. Given the enormity of these markets, there
is a large potential for alternative or non-traditional risk financing,
including catastrophic risk (CAT) bonds (explained below, and in Section
6.3.3 and Case Study 9.2.13), industry loss warranties, sidecars (a
company purchases a portion or all of an insurance policy to share in the
profits and risks), and catastrophic equity puts, all of which are playing
an increasingly important role in providing risk finance for large-loss
Chapter 7 Managing the Risks: International Level and Integration across Scales
420
events. A discussion of these instruments goes beyond the scope of this
chapter, but it is worth drawing attention to the most prominent risk-
linked security, the CAT bond, which is a fully collateralized instrument
whereby the investor receives an above-market return when a specific
natural hazard event does not occur (e.g., a Category 4 hurricane or
greater), but shares the insurer’s or government’s losses by sacrificing
interest or principal following the event if it does occur.
Over 90% of CAT bonds are issued by insurers and reinsurers in developed
countries. Although it is still an experimental market, CAT bond
placements more than doubled between 2005 and 2006, with a peak at
US$ 4.7 billion in 2006 (Cummins and Mahul, 2009), but declining to
US$ 3.4 billion in 2009 (Munich Re, 2010).
In 2006 and 2009, the first government-issued disaster relief CAT bond
placements were executed by Swiss Re and Deutsche Bank Securities to
provide funds to Mexico to insure its catastrophe fund FONDEN against
earthquake and (in 2009) hurricane risk, and thus to defray costs of
disaster recovery and relief (Cardenas et al., 2007). The World Bank
provided technical assistance for these transactions. Although the
transaction costs of the Mexican CAT bond were large, and basis risk
(the risk that the bond trigger will not be highly correlated with losses)
is a further impediment to their success, it is expected that this form of
risk transfer will become increasingly attractive especially to highly
exposed developing country governments (Lane, 2004). As discussed in
Chapter 6, a large number of government treasuries are vulnerable to
catastrophic risks, and post-disaster financing strategies generally have
high opportunity costs for developing countries.
International and donor organizations have played an important role in
another case of sovereign risk transfer (discussed in Section 9.2.13). In
2006, the World Food Programme purchased an index-based insurance
instrument to support the Ethiopian government-sponsored Productive
Safety Net Programme, which provides immediate cash payments in the
case of food emergencies. While this transaction relied on traditional
reinsurance instruments, there is current interest in issuing a CAT bond
for this same purpose. Tomasini and Van Wassenhove (2009) note the
important role that securitized instruments can play in providing backup
for humanitarian aid when disasters strike.
7.4.4.2.5. Regional risk pools
Regional catastrophe insurance pools are a promising innovation that
can enable highly vulnerable countries, and especially small states, to
more affordably transfer their risks internationally. By amalgamating
risks across individual countries or regions and accumulating reserves
over time, catastrophe insurance pools generate diversification benefits
that can eventually reduce insurance premiums. There is also growing
empirical evidence that catastrophe insurance pools have been able to
diversify intertemporally and thus dampen the volatility of the reinsurance
pricing cycle and offer secure premiums to the insured governments
(Cummins and Mahul, 2009).
As a recent example (discussed in Section 6.3.3 and Case Study 9.2.13),
the Caribbean Catastrophe Risk Insurance Facility (CCRIF) was established
in 2007 to provide Caribbean Community governments with an insurance
instrument at a significantly lower cost (about 50% reduction) than if
they were to purchase insurance separately in the financial markets.
Governments of 16 island states contributed resources commensurate
with their exposure to earthquakes and hurricanes, and claims will be
paid depending on an index for hurricanes (wind speed) and earthquakes
(ground shaking). Early cash payments received after an event will help
to mitigate the typical post-disaster liquidity crunch (Ghesquiere et al.,
2006; World Bank, 2007a,b).
7.4.4.3. Value Added by International Interventions
International financial institutions, donors, and other international actors
have played a strongly catalytic role in the development of catastrophic
risk-financing solutions in vulnerable countries, most notably by:
Exercising convening power, for example, the World Bank coordinated
the development of the CCRIF (Cummins and Mahul, 2009)
Supporting public goods for development of risk market infrastructure,
for example, donors might consider funding the weather stations
necessary for index-based weather derivatives
Providing technical assistance, for example, the World Food
Programme carried out risk assessments and provided other
assistance to support the Ethiopian sovereign risk transfer (Hess,
2007), and the World Bank provided technical assistance for the
Mexican CAT bond (Cardenas et al., 2007)
Enabling markets, for example, DFID is active in creating the legal
and regulatory environment to facilitate access to banking services,
which, in turn, greatly expedite remittances (DFID, 2005; Pickens et
al., 2009)
Financing risk transfer, as examples, the Bill Gates Foundation
subsidizes micro-insurance in Ethiopia (Suarez and Linnerooth-
Bayer, 2010); the World Bank provides low-cost capital backing for
the Mongolian micro-insurance program (Skees et al., 2008); the
Swiss SECO and the IADB provide low-interest credit to the ELF
(UNFCCC, 2008); and many countries have contributed to the
CCRIF reserve fund (Cummins and Mahul, 2009).
Though only a few of many examples of involvement by the international
community in risk-sharing and transfer projects, they show that
international financial institutions and development/donor organizations
can assist and enable risk-sharing and transfer initiatives in diverse
ways, which raises the question of their value added. Largely uncontested
is the value of creating the institutional conditions necessary for
community-based risk sharing and market-based risk transfer, yet, direct
financing, especially of insurance, is controversial. Critics point to the
‘economic efficiency principle’ discussed in Section 7.2.2, and argue
that public and international support, especially in the form of premium
subsidies, can distort the price signal and weaken incentives for taking
preventive measures, thus perpetuating vulnerability. Supporters point
to the ‘solidarity principle’ discussed in Section 7.2.3 and the important
Chapter 7Managing the Risks: International Level and Integration across Scales
421
role that solidarity has played in the social systems of the developed
world (Linnerooth-Bayer and Mechler, 2008). Other types of assistance,
like providing reinsurance to small insurers, can crowd out the (emerging)
role of the private market. Finally, critics point out that it may be more
efficient to provide the poor with cash grants than to subsidize insurance
(Skees, 2001; Gurenko, 2004).
Recognizing these concerns, there may be important and valid reasons
for interfering in catastrophe insurance and other risk-financing markets
in specific contexts (see discussions by Cummins and Mahul, 2009;
Linnerooth-Bayer et al., 2010), especially if:
The private market is non-existent or embryonic, in which case
enabling support (e.g., to improve governance, regulatory
institutions, as well as knowledge creation) may be helpful.
• The private market does not function properly, in particular, if
premiums greatly exceed the actuarially fair market price due, for
example, to limitations on private capital and the uncertainty and
ambiguity about the frequency and severity of future losses
(Kunreuther and Michel-Kerjan, 2009). In this case economically
justified premiums that are lower than those charged by the
imperfect private market may be appropriate (Froot, 1999; Cutler
and Zeckhauser, 2000).
The target population cannot afford sufficient insurance coverage,
in which case financial support that does not appreciably distort
incentives may be called for. The designers of the Mongolian
program, for example, argue that subsidizing the ‘upper layer’ is
less price-distorting than subsidizing lower layers of risk because
the market may fail to provide insurance for this layer (Skees et. al.,
2008).
The alternative is providing ‘free’ aid after the disaster happens.
7.4.5. Knowledge Acquisition, Management,
and Dissemination
A close integration of DRR and CCA and their mainstreaming into
sustainable development agendas for managing risks across scales calls for
multiple ways of knowledge acquisition and development, management,
sharing, and dissemination at all levels. Knowledge on the level of
exposure to hazards and vulnerabilities across temporal and geographical
scales (Louhisuo et al., 2007; Heltberg et al., 2008; Kaklauskas et al.,
2009); the legal aspects of DRM and CCA; financing mechanisms at
different scales; and information on access to appropriate technologies
and risk-sharing and transfer mechanisms for disaster risk reduction
(see Sections 7.4.1-7.4.4) are key to integrated risk management.
Collaboration among scientists of different disciplines, practitioners,
policymakers, and the public is pertinent in knowledge acquisition,
management, and accessibility (Thomalla et al., 2006). The type, level of
detail, and ways of generation and dissemination of knowledge will also
vary across scales, that is, from the local level where participatory
approaches are used to incorporate indigenous knowledge and build
collective ownership of knowledge generated, to national and broader
regional to international levels, thus upholding the principle of subsidiarity
in the organization, sharing, and dissemination of information on disaster
risk management (Marincioni, 2007; Chagutah, 2009).
An internationally agreed mechanism for acquisition, storage and
retrieval, and sharing of integrated climate change risk information,
knowledge, and experiences is yet to be established (Sobel and Leeson,
2007). Where this has been achieved it is fragmented, assumes a top-
down approach, is sometimes carried out by institutions with no clear
international mandate, and the quality of the data and its coverage are
inadequate. In other cases a huge amount of information is collected
but not efficiently used (Zhang et al., 2002; Sobel and Leeson, 2007).
Access to data or information under government institutions is often
constrained by bureaucracy and consolidating shared information can
be hampered by multiple formats and incompatible data sets. The major
challenge in achieving coordinated integrated risk management across
scales is in establishing clear mechanisms for a networked program to
generate and exchange diverse experiences, tools, and information that
can enable various DRR and CCA actors at different levels to use different
options available for reducing climate risks. Such a mechanism will
support efforts to mainstream CCA and DRR into development, for
example, in the case of initiatives by UNDP; development organizations
such as the World Bank, DFID, and the IADB; the Canadian International
Development Agency; the European Commission; and so forth (Benson
and Twigg, 2007). Accounting for climate risks within the development
context will, among other things, be effectively achieved where
appropriate information and knowledge of what is required exist and
are known and shared efficiently (Ogallo, 2010).
7.4.5.1. Knowledge Acquisition
Knowledge acquisition by nature is a complex, continuous, nonlinear,
and life-long process that spans generations. Knowledge acquisition for
DRR and CCA involves acquisition, documentation, and evaluation of
knowledge for its authenticity and applicability over time and beyond its
point of origin (Rautela, 2005). Knowledge acquisition and documentation
has to focus on the shifting emphasis by the HFA from reactive emergency
relief to proactive DRR approaches by aiming at strengthening prevention,
mitigation, and preparedness and linking with changes in CCA that
include greater focus on local scales (refer to Section 7.4.3.2). The
Global Spatial Data Infrastructure (GSDI), which aims to coordinate and
support the development of Spatial Data Infrastructures worldwide,
provides important services for a proactive DRR approach (Köhler and
Wächter, 2006). One of the major breakthroughs facilitating the creation
of the GSDI has been the development of interoperability standards
and technology that form a common foundation for the sharing and
interoperability of, for example, geospatial data. However, global
geospatial data infrastructure is still largely underutilized for site- and/or
application-specific needs (Le Cozannet et al., 2008; Di and Ramapriyan,
2010).
There are huge efforts in DRR- and CCA-related knowledge acquisition,
development, and exchange by universities, government agencies,
Chapter 7 Managing the Risks: International Level and Integration across Scales
422
international organizations, and to some extent the private sector, but
coordination of these efforts internationally is yet to be achieved
(Marincioni, 2007). At the international level, the International Council
for Science (ICSU) is the main international body that facilitates and
funds efforts to generate global environmental change information that
extends into DRR and CCA. ICSU is an NGO with a global membership
of national scientific bodies (121 members) and international scientific
unions (30 members) that maintain a strong focus on natural sciences
(www.icsu.org). However, there have been changes over the years and
ICSU now works closely with the International Social Science Council
(ISSC). There are four major global environmental change (GEC) research
programs facilitated by ICSU: an International Programme of Biodiversity
Science (DIVERSITAS), the International Geosphere Biosphere
Programme, the International Human Dimensions Programme closely
tied to the ISSC, and the World Climate Research Programme. These
programs have been supported by a capacity-building and information
dissemination wing, the System for Analysis, Research and Training. The
four GEC programs have had a significant role in generating the
background science that forms the basis for CCA and DRR (Steffen et
al., 2004). The link between science and policy within the UN system for
CCA is achieved through the IPCC process while for DRR it is through
activities of the UNISDR.
There has been growing concern that GEC programs are not integrated
and provide fragmented information limited to certain disciplines. This
concern led to the establishment of the Earth System Science
Partnership aiming to integrate natural and social sciences from the
regional to the global scale. However, this has proved inadequate to
meet the growing need for integrated information (Leemans et al.,
2009). As a result, a major restructuring of the knowledge generation
process both at the institutional and science level has been launched by
ICSU and the main focus is on increased use of integrated approaches
and co-production of knowledge with potential users to deliver regionally
and locally relevant information to address environmental risks for
sustainable development. These initiatives will influence the process of
integration of DRR and CCA and their linkages to development in the
future (ICSU, 2010; Reid et al., 2010).
An assessment of climate services for DRR and CCA is given in Section
7.3.3.2. But the generation of climate change information has followed
a top-down approach relying on global models to produce broad-scale
information with no clear local context and usually with large
uncertainties and complex for the public to assimilate hence providing
lower incentive for policymakers to act on the risks that are indicated
(Weingart et al., 2000; Schipper and Pelling, 2006). Climate change
information is primarily provided at long temporal ranges, for example,
2050, which is far beyond the usual five-year attention span of most
political governments let alone that of poor people concerned with basic
needs. Climate information at all scales is essential for decisionmaking
although there are various factors other than climate information that
ultimately influence decisionmaking. The IPCC Fifth Assessment Report
will cover near-term climate extending to periods earlier than 2050. Efforts
to enhance delivery of information at interannual to interdecadal scales
will improve assimilation of climate information in risk management
(Goddard et al., 2010; Vera et al., 2010). However, expressing impacts,
vulnerability, and adaption require description of complex interactions
between biophysical characteristics of a risk and socioeconomic
factors and relating to factors that usually span far beyond the area
experiencing the risk. Communicating these linkages has been a challenge
particularly for areas where education levels are low and communication
infrastructure is inadequate (Vogel and O’Brien, 2006).
Knowledge acquisition and documentation requires capacity in terms of
skilled manpower, infrastructure, and appropriate institutions and funding
(Section 7.4.3.1). Long-term research and monitoring with a wide global
coverage of different hazards and vulnerabilities is required (Kinzig, 2001).
For example, forecasting a hazard is a key aspect of disaster prevention
but generating such information comes with a cost. Although weather
forecasting through the meteorological networks of WMO is improving,
the network of meteorological stations is far from spatially adequate
and some have ceased to operate or are not adequately equipped
(Ogallo, 2010). Forecasters are challenged to communicate forecasts
that are often characterized by large uncertainty but which need to be
conveyed in a manner that can be readily understood by policymakers
and the public (Vogel and O’Brien, 2006; Carvalho, 2007).
Interdisciplinary generation of information – that is, bridging the
traditional divide among the social, natural, behavioral, and engineering
sciences – continues to be a great intellectual challenge in climate change
risk reduction. The newly formed Integrated Research on Disaster Risk
(IRDR) program – co-sponsored by ICSU, ISSC, and UNISDR – aims at
applying an integrated approach in understanding natural and human-
induced environmental hazards (ICSU, 2008; McBean, 2010). IRDR is
intended to address these challenges and gradually provide relevant
data, information, and knowledge on vulnerability trends, which are
key information for policy- and decisionmakers to formulate integrated
policies and measures for DRR and CCA.
7.4.5.2. Knowledge Organization, Sharing, and Dissemination
Exchange of disaster information worldwide has increased tremendously
through, for example, mass media and information and communication
technologies (ICT). The role of mass media in addressing the broader
needs of DRR and CCA as opposed to disaster response is still limited,
although various regional initiatives such as the Network of Climate
Journalists of the Greater Horn of Africa (NECJOGHA) that involve climate
and media experts are being established to improve the situation (Ogallo,
2010). NECJOGHA serves to disseminate integrated information based
on, for example, environmental monitoring, climatology, agronomy, public
health, and so forth, to the users to enhance sustainable response to
climate change. Clearly, multiple strategies for disseminating and sharing
knowledge and information are required for different needs at different
scales (Glik, 2007; Maitland and Tapia, 2007; Maibach et al., 2008;
Saab et al., 2008; see also Chapters 5 and 6). In particular, greater efforts
are needed to identify and communicate information on vulnerability
Chapter 7Managing the Risks: International Level and Integration across Scales
423
development, going beyond and adding to the hazards information, to
effectively contribute to reducing risk.
Disaster response and recovery are closely linked to provision of effective
communication prior to and throughout the disaster situation (Zhang et
al., 2002). Mass media, for example, radio, television, and newspapers are
powerful mechanisms for conveying information during and immediately
after disasters although they may over-sensationalize issues, which may
influence perception of risk and subsequent responses (Vasterman et
al., 2005; Glik, 2007). A ‘two-step flow’ approach where the mass media
is combined with interpersonal communication channels has been found
to provide a more effective approach to information dissemination
(Maibach et al., 2008; Chagutah, 2009; Kaklauskas et al., 2009).
Increased use of ICT such as mobile phones, online blogging websites
with interactive functions and links to other web pages and real-time
crowd-sourcing electronic commentary, and other forms of web-based
social-networked communications such as Twitter, Facebook, etc.,
represent current tools for timely information dissemination. They
facilitate rapid exchange of information, for instance, from the disaster
scene to rescuers and/or delivery of vital information to those affected. This
is particularly the case where such information is given in an appropriate
format and language and facilities to deliver information are accessible
(Glik, 2007). There are emerging attempts to develop mobile phone-
based disaster response services, for example, that can translate disaster
information into different languages (Hasegawa et al., 2005); and use
real-time mobile phone-calling data to provide information on location
and movement of victims in a disaster area (Madey et al., 2007). Mobile
phones are now routinely used to disseminate disaster warning information
within industrialized countries and the process is rapidly expanding to
developing countries.
Information sharing and dissemination for disaster relief has improved
through the establishment of the ReliefWeb site (www.reliefweb.int) by
the UN Office for the Coordination of Humanitarian Affairs (OCHA) in
1996. ReliefWeb so far offers the largest Internet-based international
disaster information gathering, sharing, and dissemination mechanism
(Wolz and Park, 2006; Maitland and Tapia, 2007; Saab et al., 2008). The
International Charter (www.disasterscharter.org) provides space data
that serve to augment the ReliefWeb. But the OCHA ReliefWeb does not
cover preparedness and disaster prevention to fully embrace CCA
and DRR compared to the comparatively more recent PreventionWeb
(www.preventionweb.net) where disaster risk reduction is covered.
Despite the growing role of mass media and ICT in disaster response,
significant improvements are still needed to reduce disaster losses. The
full potential of mobile phones and Internet facilities in disaster relief
has yet to be exploited. The OCHA ReliefWeb poorly represents local-
to national-level humanitarian activities; for example, most of this
information is not translated into different languages (Wolz and Park,
2006). There are large sections of the global population who have no
access to Internet and other telecommunication services (Samarajiva,
2005) although evidence shows that improved access by disaster workers
has overall positive effects on disaster relief (Wolz and Park, 2006).
Other initiatives such as RAdio and InterNET (RANET), a satellite broadcast
service that combines radio and Internet to communicate hydro-
meteorological and climate-related information, are examples of
innovative measures being put in place to address the problem of
limited access to the Internet in developing countries (Boulahya et al.,
2005). Sustainable use of ICT for coordination of information for
humanitarian efforts faces challenges of limited resources to mount,
maintain, and upgrade these systems (Saab et al., 2008). ICT is also
limited to explicit knowledge that is comprised of, for example, documents
and data stored in computers but generally lacks tacit knowledge that
is based on experience linked to someone’s expertise, competence,
understanding, professional intuition, and so forth that can be valuable
for disaster relief (Kaklauskas et al., 2009). Increased international
collaboration on disaster management and also the growing use of
interactive web communication facilities provides for the filtering of
tacit knowledge.
7.4.5.2.1. Disaster risk reduction and
climate change adaptation
In addition to disaster management organizations such as UNISDR, the
International Federation of Red Cross and Red Crescent Societies, the
Federal Emergency Management Agency, national Red Cross and Red
Crescent societies, and so forth, a great deal of knowledge dissemination
is accomplished in the academic field. But this knowledge does not
translate automatically to the general public. The use of ICTs such as
computer networks, digital libraries, satellite communications, remote
sensing, grid technology, and GIS for data and information integration
for knowledge acquisition and exchange is growing to be important in
integrating DRR and CCA (UNISDR, 2005b; Louhisuo et al., 2007; see
also Section 7.4.3.2). ICT offers interactive modes of learning that could
be of value in distance education and online data sharing and retrieval.
For example, the Centre for Research on the Epidemiology of Disasters
(CRED) at the Catholic University of Louvain in Belgium maintains the
Emergency Events Database (EM-DAT), which has over 18,000 entries
on disasters in the world dated from 1900 to present (www.cred.be).
The data are recorded on a country-level basis and form a useful
resource for disaster preparedness and vulnerability assessments,
although information on small-scale disasters is difficult to establish
(Tschoegl , 2006). In addition to CRED, a comprehensive database of
global natural catastrophe losses is provided by the Munich Re
NatCatSERVICE, where nearly 800 events are entered in the database
every year; by 2009, the database had more than 25,000 entries with
losses spanning from the 1980s, although records for major events go
up to 2000 years ago (Schmidt et al., 2009; Zschocke and de Leon,
2010). Because of its strong focus on insured losses, the Munich Re
database tends to have less coverage for areas with lower insurance
coverage. At a regional level, the DesInventar database in Latin America
is an example of a regional database that was developed in 1994 by
the Network for Social Studies in Disaster Prevention. The DesInventar
database is an inventory of small-, medium-, and greater-impact disasters
Chapter 7 Managing the Risks: International Level and Integration across Scales
424
(www.desinventar.net) and aims to facilitate dialog for risk management
between actors, institutions, sectors, and provincial and national
governments. This initiative has been extended to the Caribbean, Asia,
and Africa by UNDP, while the UNFCCC provides a more local-scale
database on local coping strategies (maindb.unfccc.int/public/adaptation).
ICT capabilities in disaster risk reduction also lie in enhancing interaction
among individuals and institutions from the national, to regional, to
international level, for example, through e-mail, newsgroups, online
chats, mailing lists, and web forums (Marincioni, 2007). Attempts have
been made, for example, in Japan, to create an integrated disaster risk
reduction system where mobile phone communication operates as part
of a greater information generation and delivery chain that includes
Earth observation data analysis, navigation and web technologies, GIS,
and advanced information technology such as grid (Louhisuo et al.,
2007). When such innovations are transferred to other regions they
contribute to international DRR efforts.
Other initiatives include NetHope International, which combines
development and disaster issues into its ICT-centric mandate (Saab et al.,
2008). RANET (www.oar.noaa.gov/spotlite/archive/spot_ranet.html),
originally developed in Africa for drought and which spread to Asia,
Pacific, Central America, and the Caribbean, has a strong community
engagement and disseminates comprehensive information from global
climate data banks combined with regional and local data and forecasts
resulting in spinoffs to food security, agriculture, and health in rural areas
(Boulahya et al., 2005). A network of extension agents, development
practitioners, and trained members of the community are used in
RANET to translate information into local contexts and languages and
as a result, RANET is being considered for other educational initiatives
such as the Spare Time University to improve access to learning in DRR
with benefits for CCA (Glantz, 2007). RANET has been found to reduce
vulnerability to climate extremes in different areas in Africa, for example,
communication of rainfall forecasts in parts of west Africa assists farmers
with decisions on what crop variety to plant and field to use where a
choice of fields of different soil type existed, and also where to search
for pasture and water for livestock during drought periods. However,
RANET faces challenges of unavailability of technical support, follow-up
training, power supply, and coordination (Boulahya et al., 2005).
The establishment of the PreventionWeb facility by UNISDR demonstrates
the potential of ICT in information sharing for international disaster risk
management across scales. PreventionWeb has been evolving since
2006, and was built on the experience of ReliefWeb with the purpose of
becoming a single entry point to the full range of global disaster risk
reduction information and providing a common platform for institutions
to connect, exchange experiences, and share information on DRR, and
facilitating integration with CCA and the development process. Updated
daily, the PreventionWeb platform contains news, DRR initiatives, event
calendars, online discussions, contact directories, policy and reference
documents, training events, terminology, country profiles, and fact sheets
as well as audio and video content. Hence, while catering primarily to
DRR professionals, it also promotes better understanding of disaster risk
by non-specialists. PreventionWeb is a response to a need for greater
information and knowledge sharing and dissemination advanced in Zhang
et al. (2002), Marincioni (2007), Kaklauskas et al. (2009), and others. The
web site serves a critical role in supporting the implementation of the
HFA where information and knowledge sharing is essential (Zschocke
and de Leon, 2010). But the full potential of PreventionWeb has yet to
be realized and evaluated since it is a relatively new initiative.
In addition to the PreventionWeb with a DRR focus, the number of web-
based resource portals supporting both DRR and CCA has been
increasing. These include, among others, ProVention Consortium,
which had a DRR and climate focus (www.proventionconsortium.org)
but has ceased to operate; the UN Adaptation Learning Mechanism
(www.adaptationlearning.net) with links to related online resources
and documentation of over 140 countries; Linking Climate Adaptation
Network/CBA-X (www.linkingclimateadaptation.org) which has some DRR
focus, had over 1,000 members in 2008, and has continued to provide
current thinking on climate adaptation and resources and publications
for researchers, practitioners, and policy formers; and the WeAdapt/
WikiAdapt, an adaptation focus portal (www.weadapt.org) that goes
beyond networking and dissemination to cover knowledge integration
and other innovative adaptation tools. These portals are relatively new,
remain predominantly used by their respective communities, and have
also been noted by others to be poorly organized (Mitchell and van
Aalst, 2008). Performance of such ICT information resources in disaster
risk management could improve with more coordination and integration
of CCA, DRR, and the development community.
7.4.5.2.2. Constraints in knowledge sharing and dissemination
For all information tools noted, the quality of information transferred and
language used influence their effectiveness. Further, these mechanisms
often collapse during a disaster when most needed (Marincioni, 2007;
Saab et al., 2008). Some of the new technologies are not easily accessible
to the very poor, and even the most innovative tools like RANET show
numerous maintenance constraints particularly in remote areas
(Boulahya et al., 2005).
There are differences in perception on the role of ICT in the exchange of
disaster and hazard risk knowledge as opposed to its role in increased flow
of information, with knowledge here defined simply as understanding of
information while information refers to organized data (Zhang et al.,
2002; Marincioni, 2007). Indications are that, while there is increased
circulation of disaster information, this does not always result in
increased assimilation of new risk reduction approaches, a factor that is
partly attributed to lack of effective sharing although lack of capacity to
use/apply the information could be a major factor (Zhang et al., 2002;
UNISDR, 2005b).The level of assimilation of ICT technology into disaster
risk reduction depends, among other things, on levels of literacy and
the working environment including institutional arrangements, hence
effectiveness may vary with levels of development (Samarajiva, 2005;
Marincioni, 2007; see also Section 7.4.3.2). As a result, the contribution
Chapter 7Managing the Risks: International Level and Integration across Scales
425
of these relatively new facilities such as PreventionWeb will, among
other things, depend on accessibility and assimilation of ICT in the
daily operations of institutions across the globe. Evidence shows that
information alone is not adequate to address disaster risk reduction; rather,
other factors such as availability of resources, effective management
structures, and social networks are critical (Glik, 2007; Lemos et al.,
2007; Maibach et al., 2008; Chagutah, 2009).
A major constraint in climate change risk management results from the
fact that communities working in disaster management, climate change,
and development operate separately and this increases vulnerability to
climate extremes leading to disasters (Schipper and Pelling, 2006;
Lemos et al., 2007). For example, emphasis on humanitarian assistance
has been attributed to development agendas that do not adequately
integrate risk reduction leading to increased vulnerability (Benson and
Twigg, 2007), while development community members are, for example,
better equipped with the use of insurance but fail to link this to climate
risk reduction thus exposing communities to vulnerability to climate
extremes. Similar observations have been made about cities where
urban developers have no link with the climate risk management
community (Wamsler, 2006). But in fact both the development and climate
adaptation communities are concerned with vulnerability to disasters.
This could be a common point of focus facilitating collaboration in
research, information sharing, and practice as part of global security
(Schipper and Pelling, 2006; Lemos et al., 2007).
Communication gaps between professional groups often result from
different language styles and jargons. Heltberg et al. (2008) have
suggested a need for establishing universally shared basic operational
definitions of key terms such as risk, vulnerability, and adaptation
across the different actors as a basis for dissemination of knowledge.
This has also been noted by others, for example, for better coordination
among numerous humanitarian organizations (Saab et al., 2008) and in
the FAO guide for disaster risk management (Baas et al., 2008; also see
Chapter 1). The move toward establishment of national disaster risk
reduction institutions that link to similar regional and international
structures by, for example, UNISDR, provides a framework for bringing
different stakeholders together including the climate change and
development communities at the national level, culminating in greater
integration of risk management at the international level. Other efforts
include international initiatives to integrate, at the national level, disaster
risk reduction with poverty reduction frameworks (Schipper and Pelling,
2006).
In conclusion, there is high agreement in the literature indicating that
efforts are being made internationally to build information and knowledge
bases that support the shift in emphasis by the HFA from reactive
emergency relief to proactive DRR (high confidence). Conventional media
and ICT are major factors in facilitating the required international
exchange and dissemination of information on disaster response, CCA,
and DRR (high confidence). This in turn stimulates generation of new
knowledge and will over time lead to greater integration of DRR and CCA,
which at the present moment is still limited (medium confidence). The
limitation of relying heavily on ICT is that there is still a large part of the
world where the ICT infrastructure is not adequately developed. There is
also high agreement in the literature that an increase in the exchange
of data and information at the international level on its own is not a
complete solution to risk reduction. Resources to generate and supply
information and experience in a usable form for each unique case so as
to translate this to knowledge and action are a critical dimension in risk
reduction (high confidence). Further, more attention is required for the
international community to identify what information is essential for
different stages of climate change risk management, and how it should
be captured and used by different actors under different risk reduction
scenarios. Data gathering, information, and knowledge acquisition and
management for disaster relief has a longer history. The process of
building integrated information resource tools that brings together
experiences from CCA, DRR, and the development community is still
weak, yet these tools hold the promise for reducing vulnerability to
disasters in the future (high confidence).
7.5. Considerations for
Future Policy and Research
How best can experience with disaster risk reduction at the international
level be used to help or strengthen climate change adaptation? The
characteristics of the DRR regime (as exemplified chiefly by the UNISDR
and the Hyogo Framework for Action) and the CCA regime (chiefly the
UNFCCC and the IPCC) have been described in detail and assessed to
the extent that the literature allows. One frequently made assumption
is that the DRR world has much to learn from CCA and vice versa (IPCC,
2009). It is widely proposed in the literature that disaster risk reduction
and climate change adaptation should be ‘integrated’ (Birkmann and von
Teichman, 2010).
The call for integration of disaster risk reduction with climate change
adaptation goes much further, however (UNISDR, 2009a). It is argued
that both disaster risk reduction and climate change adaptation remain
outside the mainstream of development activities (UNISDR, 2009a). The
United Nations Global Assessment Report on Disaster Risk Reduction
calls for “an urgent paradigm shift” in disaster risk reduction to address
the underlying risk drivers such as vulnerable rural livelihoods, poor
urban governance, and declining ecosystems (UNISDR, 2009a). The
report also calls for the harmonization of existing institutional and
governance arrangements for disaster risk reduction and climate
change adaptation (p. 181), and presents a 20-point plan to reduce risk
(pp. 176-177).
These conclusions come from an official UN report (UNISDR, 2009a),
and they are widely supported in the scientific literature (O’Brien et al.,
2006; Schipper, 2009) as well as in other government reports (DFID,
2005; Birkmann et al., 2009; CCD, 2009) and in the advocacy literature
(Venton and La Trobe, 2008). More recently, the widely reviewed ICSU
(2010) report (called the Belmont Challenge) on Regional Environmental
Change: Human Action and Adaptation, which was commissioned by
Chapter 7 Managing the Risks: International Level and Integration across Scales
426
the major global environmental change research funders to assess the
international research capability required to respond to the challenge
of delivering knowledge to support human action and adaptation
to regional environmental change, concluded by calling for a highly
coordinated and collaborative research program to deliver integrated
knowledge required to identify and respond to hazards, risks, and
vulnerability, and develop mitigation and adaptation strategies.
Similarly, ICSU and the ISSC carried out a wide consultative process
to rethink the focus and framework of Earth system research. This
consultation came out with four Grand Challenges that require a balanced
mix of disciplinary and interdisciplinary research to address critical issues
at the intersection of Earth systems science and sustainable development
(Reid et al., 2010):
Improve the usefulness of forecasts of future environmental
conditions and their consequences for people.
Develop, enhance, and integrate observation systems to manage
global and regional environmental change.
Determine how to anticipate, avoid, and manage disruptive global
environmental change.
Determine institutional, economic, and behavioral changes to enable
effective steps toward global sustainability.
Both the Belmont Challenge and the Grand Challenges are setting an
international tone for an integrative approach to challenges such as
DRR, CCA, and development. There is no shortage of policy proposals
designed to integrate disaster risk reduction and climate change
adaptation for their common strengthening and benefit.
Official reports also list many reasons why more movement in this
direction has been slow to develop. One constraint is the difficulty of
integration across scales, which is addressed in Section 7.6. Two other
sets of constraints are described as ‘the normative dimension’ and ‘the
knowledge dimension’ (Birkmann et al., 2009). The extensive list of
challenges and constraints identified includes the following:
Normative Dimensions (adapted from Birkmann et al., 2009)
Absence of uniform methods, standards, and procedures in
vulnerability and capacity assessment and also in the design,
formulation, and implementation of adaptation plans, programs,
and projects. Lack of clear norms when applying vulnerability
and capacity assessment and when designing and implementing
adaptation measures
The desire for stability and the tendency to rapidly restore
normalcy limit the scope to explore and to take advantage of the
opportunity after disaster and recover in an adaptive way by
taking account of future climate change. The notion and desire
for stability may hamper the chance to take advantage of change
and dynamics – after disasters, the chance to use the opportunity
and build back in an adaptive way considering future climate
change is in most cases not taken – more commonly, infrastructure
is rapidly built back to the pre-disaster condition
Knowledge Challenges (adapted from Birkmann et al., 2009)
Differences in the form of terminology used – that is, the different
terms and definitions framed by both DRR and CCA communities
Unavailability of information about the concrete effects of
climate change at the local level (see Section 7.4.5.1)
Limited census-based information on relevant census data
(social and economic parameters) especially in dynamic areas
with, for example, high fluctuations of people and/or economic
instability
Scientific knowledge on climate change acquired by the scientific
community has not been translated or trickled down to
practitioners or it is communicated in a way that is hard to
understand and derive practical knowledge (see Section 7.4.5.2.2)
Absence or lack of appropriate indicators for assessment that
could measure successful adaptation and which could also be
incorporated into funding guidelines as well as monitoring and
evaluation strategies (ICSU, 2010).
For the purposes of this Special Report, the question has been formulated
in terms of what can be learned from the practice of DRR to advance
CCA. It is clear from the literature, however, that cooperation between
the DRR and CCA communities is a two-way process. This has given rise
to questions about how ‘integration’ in practice at local and national
levels might best be facilitated by change at the international level.
7.6. Integration across Scales
7.6.1. The Status of Integration
The literature reflects three different perspectives on the integration of
disaster risk reduction and climate change adaptation. One view
common among the community of experts and practitioners is that
climate change adaptation should be integrated into disaster risk
reduction (CCD, 2008a,b,c; Prabhakar et al., 2009, p. 26). It has even
been suggested that climate change adaptation is a case of ‘reinventing
the wheel’ (Mercer 2010) since disaster risk reduction covers much of the
same ground and is “already well-established within the international
development community” (Lewis, 1999; Wisner et al., 2004).
Practitioners in disaster risk reduction tend to have the view that climate
change is one of a number of factors contributing to vulnerability and
disasters (Mercer, 2010), and that therefore climate change adaptation
needs to be taken on board.
A second view is adopted by many in the climate change adaptation
community. They recognize a diversity of cross-cutting risks that can be
associated with the impacts of climate change and consider disaster risk
to be one of these (Birkmann and von Teichman, 2010). They conclude
that disaster risk reduction should be integrated into climate change
adaptation.
A third and perhaps more widespread view is that both disaster risk
reduction and climate change adaptation should be more effectively
integrated into wider development planning (Glantz, 1999; O’Brien et
al., 2006; Lewis, 2007; CCD, 2009; Christoplos et al., 2009; UNISDR,
2009a).
Chapter 7Managing the Risks: International Level and Integration across Scales
427
At the practical level there are many steps already underway to bring
about such forms of integration (see Chapters 5 and 6). There are
numerous hazards and disasters that are not directly linked to climate
change but their impacts may serve to increase vulnerability to climate
change. Nevertheless, as noted in Section 7.5 there are many obstacles
to integration and it is by no means agreed that full integration
between disaster risk reduction and climate change adaptation is
possible, or desirable.
The potential benefits as well as the obstacles to integration can be
examined in terms of three scales: the spatial, the temporal, and the
functional (Birkmann and von Teichman, 2010).
7.6.2. Integration at a Spatial Scale
The literature reflects a view that DRR and CCA operate at different
spatial scales (Birkmann and von Teichman, 2010) and that therefore their
integration in practice has been problematic or impracticable. Disasters
are often thought of as events occurring at a specific location whereas
climate change is thought of as a global or regional phenomenon. This
view is now being modified as the need for locally based climate
change adaptation becomes evident (Adger et al., 2005), as the impacts
of local disasters are recognized as having more widespread impacts at
a larger spatial scale (see Chapters 4 and 6 and Section 7.2.1).
One commonly cited impediment to integration is that climate change
projections do not provide precise information at a local scale (see
Chapter 3) and that adaptation strategies tend to be designed for entire
countries or regions (German Federal Government, 2008; Red Cross and
Red Crescent Climate Centre, 2007).
7.6.3. Integration at a Temporal Scale
There is also a perceived difference in the temporal scales of CCA
and DRR. The disaster community has traditionally been focused on
humanitarian response including relief and reconstruction in the relatively
short term. (UNISDR, 2009b), whereas climate change has been
recognized as including long-term processes with projections extending
from decades to centuries (Chapter 3), which poses problems for
development communities usually focusing on a shorter time span.
More effective cooperation and integration between the DRR and the
CCA practitioners could help to detect, address, and overcome these
temporal-scale challenges. This essentially requires the stronger
recognition of the risks of climate-related disasters in CCA and the
incorporation of longer-term climate change risk factors into DRR.
7.6.4. Integration at a Functional Scale
The functional separation of CCA and DRR institutions, organizations,
and mechanisms extends across all three levels of management from
local to national to international. At the international level there are
weak links between the climate adaptation ‘regime’ as expressed in the
UNFCCC and the leading DRR ‘regime’ in the form of the UNISDR. The
character of the two ‘regimes’ is radically different, the former having
the task of implementing an international agreement and the latter
being a UN-wide interagency and advocacy program. The history of the
evolution of the two institutional arrangements is markedly different.
The disaster field has long been dominated by humanitarian and
emergency response measures and has only relatively recently been
moving toward a stronger DRR approach (Burton, 2003). Similarly,
climate change was initially conceived as an atmospheric pollution
issue with greater emphasis on the need to reduce greenhouse gas
emissions and has slowly been repositioned, as in the UNFCCC
negotiations, as also being a development issue. One consequence of
the different evolution has been that the emerging international climate
‘regime’ (UNFCCC) is linked at the national level to environment
ministries, whereas the disaster ‘regime’ (UNISDR) is linked to emergency
planning and preparedness agencies or, in other cases, to the office of
President. Neither DRR nor CAA are well linked to economic planning
and development agencies (UNISDR, 2009b).
There is also a ‘top-down’ versus ‘bottom-up’ distinction (Rayner, 2010).
Natural hazards and associated disasters have a long history, and DRR
has moved slowly from local to national to international levels in
response to the rationale described in Section 7.2. Climate change, on
the other hand, came to attention as a result of the work of atmospheric
scientists and was first recognized primarily as a global problem, and
has subsequently moved down scale as the need for CCA became
more apparent and pressing. This shows that the opportunity exists
for the two to complement each other, at the international level where
DRR has progressed, and at the national and local level to which CCA
is moving.
7.6.5. Toward More Integration
The mandate of this Special Report is in part to consider how CCA could
be enhanced by learning from the experience of the DRR community,
and vice versa. The literature shows a widespread view that the two
could both benefit from closer integration with each other and that
both would benefit society better if there was more integration into
sustainable development (UNISDR, 2009a). Integration in this sense is
meant as symbiosis or synthesis rather than formal integration at the
institutional level. Integration across scales can be facilitated if
integration between DRR and CCA were also to take place at local,
national, and international levels. Integration at the international level
might help to facilitate integration at national and local levels although
the opposite is also possible. This Special Report is itself a prime example
of emerging cooperation. It is in line with a wider evolution in the global
environmental change science research community whose products
serve both disaster risk reduction and climate change adaptation at the
international level of management.
Chapter 7 Managing the Risks: International Level and Integration across Scales
428
References
A digital library of non-journal-based literature cited in this chapter that
may not be readily available to the public has been compiled as part of
the IPCC review and drafting process, and can be accessed via either the
IPCC Secretariat or IPCC Working Group II web sites.
Adaptation Fund, 2010: Financial Status of the Adaptation Fund Trust Fund.
AFB/EFC.3/7Rev.1, Adaptation Fund, GEF, Washington, DC.
Adaptation Fund, 2011a: About the Adaptation Fund. Adaptation Fund, GEF,
Washington, DC, adaptation-fund.org/about.
AdaptationFund, 2011b: Funded Projects. Adaptation Fund, GEF, Washington, DC,
adaptation-fund.org/funded_projects.
Adger, W.N., 2004: The right to keep cold. Environment and Planning, A, 36, 1711-1715.
Adger, W.N., N.W. Arnell, and E.L. Tompkins, 2005: Successful adaptation to climate
change across scales. Global Environmental Change, 15, 77-86.
Adger, W.N., I. Lorenzoni, and K. O’Brien (eds.), 2009: Adapting to Climate Change:
Thresholds, Values, Governance. Cambridge University Press, Cambridge, UK.
African Union, 2010: Extended Programme of Action for the Implementation of the
Africa Regional Strategy for Disaster Risk Reduction (2006-2015). In: Extended
Programme of Action for the Implementation of the Africa Regional Strategy for
Disaster Risk Reduction (2006 - 2015) and Declaration of the 2nd African
Ministerial Conference on Disaster Risk Reduction 2010. African Union
Commission and United Nations International Strategy for Disaster Reduction,
Addis Ababa, Ethiopia, p. 25-54.
AGF, 2010: Report of the Secretary-General’s High-Level Advisory Group on Climate
Change Financing. United Nations, New York, NY, www.un.org/wcm/webdav/
site/climatechange/shared/Documents/AGF_reports/AGF%20Report.pdf.
Agrawala, S. and S. Fankhauser (eds.), 2008: Economic Aspects of Adaptation to
Climate Change: Costs, Benefits and Policy Instruments. OECD, Paris, France.
Agrawala, S. and M. van Aalst, 2008: Adapting development cooperation to adapt
to climate change. Climate Policy, 8, 183-193.
Alam, M., A. Rahman, M. Rashid, G. Rabbani, P. Bhandary, S. Bhadwal, M. Lal, and
M.H. Soejachmoen, 2007: Impacts, Vulnerability and Adaptation to Climate
Change In Asia. Background Paper for the UNFCCC, UNFCCC, Bonn, Germany,
unfccc.int/files/adaptation/methodologies_for/vulnerability_and_adaptation/
application/pdf/unfccc_asian_workshop_background_paper.pdf.
Alexander, D., 2006: Globalization of disaster: trends, problems and dilemmas.
Journal of International Affairs, 59, 1-22.
Arriens, W.T.L. and C. Benson, 1999: Post disaster rehabilitation: The experience of
the Asian Development Bank. Paper presented to the IDNR-ESCAP Regional
Meeting for Asia: Risk Reduction and Society in the 21st Century, Bangkok,
Thailand, 23-26 February 1999.
Baas, S., S. Ramasamy, J.D. DePryck, and F. Battista, 2008: Disaster Risk Management
Systems Analysis. Environment, Climate Change and Bioenergy Division, FAO,
Rome, Italy, 68 pp.
Bamberger, R.L. and L. Kumins, 2005: Oil and Gas: Supply Issues after Katrina.
Congressional Research Service, Library of Congress, Washington, DC.
Barnett, B.J., C.B. Barrett, and J.R. Skees, 2008: Poverty traps and index-based risk
transfer products. World Development, 36(10), 1766-1785.
Barnett, J. and W.N. Adger, 2007: Climate change, human security and violent conflict.
Political Geography, 26, 639-655.
Barnett, J. and M. Webber, 2009: Accommodating Migration to Promote Adaptation
to Climate Change: A Policy Brief
. World Bank and SCCCD, Washington, DC and
Stockholm, Sweden.
Barrett, C.B., B.J. Barnett, M.R. Carter, S. Chantarat, J.W. Hansen, A.G. Mude, D.E
Osgood, J.R. Skees, C.G. Turvey, and M.N. Ward, 2007: Poverty Traps and Climate
Risk: Limitations and Opportunities of Index-Based Risk Financing. IRI Technical
Report 07-02, International Research Institute for Climate and Society,
Palisades, NY, 53 pp.
Basher, R., 2006: Global early warning systems for natural hazards: systematic and
people-centred. Philosophical Transactions of the Royal Society, 364, 2167-2182.
Begg, I., 2008: Subsidiarity in regional policy. In: Subsidiarity and economic reform
in Europe [Gelauff, G., I. Grilo, and A. Lejour (eds.)]. Springer, Berlin, Germany,
pp. 291-310.
Benson, C. and J. Twigg, 2007: Tools for Mainstreaming Disaster Risk Reduction:
Guidance Notes for Development Organisations. International Federation of
Red Cross and Red Crescent Societies/the ProVention Consortium, ProVention
Consortium Secretariat, Geneva, Switzerland, 178 pp.
Berkhout, F. 2005: Rationales for adaptation in EU climate change policies. Climate
Policy, 5, 377-391.
Bhatt, M.R., 2007: Good practice in local approaches to climate change adaptation
and disaster risk management in South Asia: Lessons from Tsunami Evaluation
Coalition. In: Climate Change, Humanitarian Disasters and International
Development: Linking vulnerability, risk reduction and response capacity.
Research and Policy Workshop, 27 April 2007. Oslo Center for Interdisciplinary
Environmental and Social Research Forskningsparken, Oslo, Norway.
Biermann, F. and I. Boas, 2010: Global adaptation governance: the case of protecting
climate refugees. In: Global Climate Governance Beyond 2012: Architecture,
Agency and Adaptation [Bierman, F., P. Pattberg, and f. Zelli (eds.)]. Cambridge
University Press, Cambridge, UK, pp. 255-269.
Birkmann, J. and K. von Teichman, 2010: Integrating disaster risk reduction and
climate change adaptation: key challenges – scales, knowledge, and norms.
Sustainability Science, 5, 171-184
Birkmann, J., K. von Teichman, P. Aldunce, C. Bach, N. T. Binh, M. Garschagen, S.
Kanwar, N. Setiadi, and L.N. Thach, 2009: Addressing the challenge:
Recommendations and quality criteria for linking disaster risk reduction and
adaptation to climate change. In: DKKV Publications Series 38 [Birkmann, J., G.
Tetzlaff, and K.-O. Zentel (eds).]. German Committee for Disaster Reduction,
Bonn, Germany.
Birnie, P., A. Boyle, and C. Redgwell, 2009: International Law and the Environment.
3rd ed. Oxford University Press, Oxford, UK.
Bodansky, D., 1995: Customary (and not so customary) international environmental
law. Indiana Journal of Global Legal Studies, 3(1), 105-119.
Bodansky, D., 2005: The international climate change regime. In: Perspectives on
Climate Change: Science, Economics, Politics, Ethics [Sinnott-Armstrong, W. and
R.B. Howarth (eds.)]. Elsevier, Amsterdam, The Netherlands, pp. 147-180.
Bodansky, D., 2010: The Art and Craft of International Environmental Law. Harvard
University Press, Cambridge, MA.
Boulahya, M., M.S. Cerda, M. Pratt, and K. Sponberg, 2005: Climate, communications,
and innovative technologies: Potential impacts and sustainability of new radio
and internet linkages in rural African communities. Climatic Change, 70,
299-310.
Brown, J., N. Bird, and L. Schalatek, 2010: Direct Access to the Adaptation Fund:
Realising the Potential of National Implementing Entities. Overseas
Development Institute, London, UK, 10 pp.
Brown, O., 2008: Migration and Climate Change
. Research Series No. 31,
International Organization for Migration, Geneva, Switzerland.
Burton, I., 2003. Do we have the adaptive capacity to develop and use the adaptive
capacity to adapt? In: Climate Change, Adaptive Capacity and Development
[Smith, J.B., R.J.T. Klein, and S. Huq (eds.)]. Imperial College Press. London, UK,
pp 137-161.
Burton, I., S. Huq, B. Lim, O. Pilifosova, and E.L. Schipper, 2002: From impacts
assessment to adaptation priorities: the shaping of adaptation policy. Climate
Policy, 2(2), 145-159.
Caney, S., 2008: Human rights, climate change and discounting. Environmental
Politics, 17, 536-555.
Caney, S., 2010: Climate change, human rights and moral thresholds. In: Human
Rights and Climate Change [Humphreys, S. (ed.)]. Cambridge University Press,
Cambridge, UK, pp. 69-90.
Cardenas, V., S. Hochrainer, R. Mechler, G. Pflug, and J. Linnerooth-Bayer, 2007:
Sovereign financial disaster risk management: the case of Mexico.
Environmental Hazards, 7, 40-53.
Chapter 7Managing the Risks: International Level and Integration across Scales
429
CARE International, 2010: Community-Based Adaptation Toolkit. Cooperative for
Assistance and Relief Everywhere (CARE), Chatelaine, Switzerland,
www.careclimatechange.org/files/toolkit/CARE_CBA_Toolkit.pdf.
Carvalho, A., 2007: Ideological cultures and media discourses on scientific
knowledge: re-reading news on climate change. Public Understanding of
Science, 16, 223.
Castles, S., 2002: Environmental Change and Forced Migration: Making Sense of the
Debate. New Issues in Refugee Research, Working Paper No. 70, United Nations
High Commissioner for Refugees, Geneva, Switzerland.
CCD, 2008a: Incentives and Constraints to Climate Change Adaptation and Disaster
Risk Reduction – A Local Perspective. Commission on Climate Change and
Development, Stockholm, Sweden.
CCD, 2008b: Overview of Adaptation Mainstreaming Initiative. Commission on
Climate Change and Development, Stockholm, Sweden.
CCD, 2008c: Links between Disaster Risk Reduction, Development and Climate
Change. Commission on Climate Change and Development, Stockholm, Sweden.
CCD, 2009: Closing the Gaps: Disaster Risk Reduction and Adaptation to Climate
Change in Developing Countries. Commission on Climate Change and
Development, Stockholm, Sweden.
Chagutah, T., 2009: Towards improved public awareness for climate related disaster
risk reduction in South Africa: A Participatory Development Communication
perspective. Jàmbá: Journal of Disaster Risk Studies, 2(2), 113-126.
Chinkin, C.M., 1989: The challenge of soft law: Development and change in
international law. International and Comparative Law Quarterly, 38(4),
850-866.
Christian Aid, 2007: Human Tide: The Real Migration Crisis. Christian Aid, London,
UK, www.christianaid.org.uk/Images/human-tide.pdf.
Christoplos, I., S. Anderson, M. Arnold, V. Galaz, M. Hedger, R.J.T. Klein, and K. Le
Goulven, 2009: The Human Dimension of Climate Adaptation: The Importance
of Local and Institutional Issues. Commission on Climate Change and
Development, Stockholm, Sweden.
CIENS, 2007: Climate Change, Humanitarian Disasters and International
Development: Linking Vulnerability, Risk Reduction and Response Capacity.
Research and Policy Workshop. 27 April 2007. Oslo Center for Interdisciplinary
Environmental and Social Research (CIENS), Forskningsparken, Oslo, Norway.
Clémençon, R., 2008: The Bali Road Map: A first step on the difficult journey to a
post-Kyoto Protocol agreement. Journal of Environment and Development,
17(1), 70-94.
Climate Funds Update, 2011: Pilot Program for Climate Resilience. Climate Funds
Update, www.climatefundsupdate.org/listing/pilot-program-for-climate-
resilience.
Cohen, R., 1998: The Guiding Principles on Internal Displacement: a New Instrument
for International Organizations and NGOs. Office for the Coordination of
Humanitarian Affairs, DC 1-1568, 1 UN Plaza, 10017 NY, New York, USA. Also
available on http://www.fmreview.org/FMRpdfs/FMR02/fmr209.pdf.
Cozier, M., 2011: Restoring confidence at the Cancun Climate Change Conference.
Greenhouse Gases: Science and Technology, 1(1), 8-10.
Craeynest, L., L. Gallagher, and C. Sharkey, 2010: Business as Unusual. Direct Access:
Giving power back to the poor? Caritas Internationalis and CIDSE, Vatican City,
Italy and Brussels, Belgium.
Cummins, J.D. and O. Mahul, 2009:
Catastrophe Risk Financing in Developing
Countries: Principles for Public Intervention. World Bank, Washington, DC.
Cutler, D.M. and R. Zeckhauser, 2000: The anatomy of health insurance. In:
Handbook of Health Economics. Vol. 1 [Culyer, A.J. and J.P. Newhouse (eds.)].
Elsevier Science, Amsterdam, The Netherlands, pp. 563-643.
Czarnecki, R. and K. Guilanpour, 2009: The Adaptation Fund after Poznan. Carbon
and Climate Law Review, 3(1), 79-87.
Dalberg, 2010: United Nations International Strategy for Disaster Reduction
(UNISDR) Secretariat Evaluation: Final Report.
Davis, I., 2004: The application of performance targets to promote effective
earthquake risk reduction strategies. Engineering Paper no 2726 presented at
the Thirteenth World Conference on Earthquakes, Vancouver, Canada 1-6
August 2004.
De Lucia, V., 2007: Common but differentiated responsibility. In: The Encyclopedia of
Earth [C.J. Cleveland (ed.)]. Environmental Information Coalition, National
Council for Science and the Environment, Washington, DC, www.eoearth.org/
article/Common_but_differentiated_responsibility.
Dessai, S., 2003: The Special Climate Change Fund: origins and prioritisation
assessment. Climate Policy, 3(3), 295-302.
DFID, 2005: Sending Money Home? A Survey of Remittance Products and Services in
the United Kingdom. Department for International Development, London, UK.
DFID, 2011: The Multilateral Aid Review. Ensuring maximum value for money for UK
aid through multilateral organisations, Annex 6, Assessment Summary ISDR.
Criteria for Assessment Table 2, Department for International Development,
London, UK, pp 217-218.
Di, L. and H.K. Ramapriyan, 2010: Standards-based data and information systems for
Earth observations – An introduction. In: Standard-Based Data and Information
Systems for Earth Observation [Di, L. and H.K. Ramapriyan (eds.)]. Springer-
Verlag, Berlin, Germany, pp. 1-6.
Diacon, D., 1992: Typhoon resistant housing in the Philippines: The Core Shelter
Project. Disasters, 16(3), 266-271.
Doig, A. 2008: Setting the Bar High at Poznan. Christian Aid, London, UK,
www.christianaid.org.uk/images/poznan-report.pdf.
Dun, O. and F. Gemenne, 2008: Defining environmental migration. Forced Migration
Review, 31, 10-11.
Dunfee, T.W. and D. Hess, 2000: The legitimacy of direct corporate humanitarian
investment. Business Ethics Quarterly, 10(1), 95-109.
Eriksen, S. and K. Brown, 2011: Sustainable adaptation to climate change: Prioritising
social equity and environmental integrity. Climate and Development, 3(1), 3-6.
Eriksen, S., P. Aldunce, C.S. Bahinipati, R.D.’A. Martins, J.I. Molefe, C. Nhemachena, K.
O’Brien, F. Olorunfemi, J. Park, L. Sygna, and K. Ulstrud, 2011: When not every
response to climate change is a good one: identifying principles for sustainable
adaptation. Climate and Development, 3(1), 7-20.
Fagen, P. and M. Bump (ed.), 2006: Remittances in Conflict and Crises: How
Remittances Sustain Livelihoods in War, Crises, and Transitions to Peace.
International Peace Academy, New York, NY.
Fankhauser, S., 2010: The costs of adaptation. Wiley Interdisciplinary Reviews:
Climate Change, 1(1), 23-30.
Fidler, D., 2005: Disaster relief and governance after the Indian tsunami: What role
for international law? Melbourne Journal of International Law, 6(2), 458.
Fisher, D., 2007: The law of international disaster response. In: Global Legal
Challenges: Command of the Commons, Strategic Communications and Natural
Disasters [M.D. Carsten (ed.)]. International Law Studies Vol. 83, US Naval War
College, Newport, Rhode Island, pp. 293-320.
Foley, C., 2007: Mozambique: A Case Study in the Role of the Affected State in
Humanitarian Action. HPG Working Paper, Humanitarian Policy Group and
Overseas Development Institute, London, UK.
Froot, K.A. (ed.), 1999: The Financing of Catastrophe Risk. The University of Chicago
Press, Chicago, IL.
Fugate, W.C. and C. Crist, 2008: Florida Hurricane Loss Mitigation Program. Report
to the Florida Legislature, Florida Division of Emergency Management,
Tallahassee, FL, www.floridadisaster.org/Mitigation/Documents/RCMPAnnualRpt
SFY07-08-final-secured.pdf.
Gassebner, M., A. Keck, and R. Teh, 2010: Shaken, not stirred: The impact of disasters
on international trade. Review of International Economics, 18(2), 351-368.
GEF, 2008: Report on the Completion of the Strategic Priority on Adaptation.
GEF/C.34/8, Global Environment Facility, Washington, DC, www.thegef.org/gef/
sites/thegef.org/files/documents/C.34.8%20Report%20on%20the%20Complet
ion%20of%20the%20SPA.pdf.
GEF, 2009: Report of the GEF to the Fifteenth Session of the Conference of the Parties to
the United Nations Framework Convention on Climate Change. FCCC/CP/2009/9,
UNFCCC, Bonn, Germany, unfccc.int/resource/docs/2009/cop15/eng/09.pdf.
GEF, 2010a: Status Report on Least Developed Countries Fund and Special Climate
Change Fund. GEF/LDCF.SCCF.9/Inf.2/Rev.2, GEF, Washington, DC, www.thegef.org/
gef/sites/thegef.org/files/documents/Status%20Report%20on%20the%20Clim
ate%20Change%20Funds%20-%20Oct%208,%202010-Rev2.pdf.
Chapter 7 Managing the Risks: International Level and Integration across Scales
430
GEF, 2010b: Report of the GEF to the Sixteenth Session of the Conference of the
Parties to the United Nations Framework Convention on Climate Change.
FCCC/CP/2010/5, UNFCCC, Bonn, Germany, http://unfccc.int/resource/
docs/2010/cop16/eng/05.pdf.
German Federal Government, 2008: German Strategy for Adaptation to Climate
Change. Adopted by the German Federal Cabinet on 17 December 2008, The
Federal Government, Berlin, Germany, www.bmu.de/files/english/pdf/application/
pdf/das_gesamt_en_bf.pdf.
GFDRR, 2010: Partnership Charter. Global Facility for Disaster Reduction and
Recovery, Washington, DC, gfdrr.org/gfdrr/sites/gfdrr.org/files/publication/
GFDRR_Partnership_Charter_2010.pdf.
Ghesquiere, F., O. Mahul, M. Forni, and R. Gartley, 2006: Caribbean Catastrophe Risk
Insurance Facility: A solution to the short-term liquidity needs of small island
states in the aftermath of natural disasters. IAT03-13/3, World Bank,
Washington, DC.
Gigli, S. and S. Agrawala, 2007: Stocktaking on Progress on Integrating Adaptation
to Climate Change into Development Co-operation Activities.
COM/ENV/EPOC/DCD/DAC(2007)1/FINAL, Environment Directorate and
Development Co-operation Directorate, OECD, Paris, France, 74 pp.
Glantz, M.H., 1999: El Niño as a hazard-spawner. In: The Nature of Hazards:
Commemorative Volume to Celebrate Achievements of International Decade for
Natural Disaster Reduction [J. Ingleton (ed.)]. Tudor Rose Publishers, Leicester,
UK, pp. 78-79.
Glantz, M.H., 2007: How about a spare-time university? WMO Bulletin, 56(2), 1-6.
Glik, D.C., 2007: Risk communication for public health emergencies. Annual Review
of Public Health, 28, 33-54.
GNDR, 2009 : Clouds but Little Rain: Views from the Frontline - A local perspective
of progress towards implementation of the Hyogo Framework for Action. Global
Network of Civil Society Organisations for Disaster Reduction, London, UK.
GNDR, 2011: If We do not Join Hands Summary Report, Views from the Frontline,
Local reports of progress on implementing the Hyogo Framework for Action,
Global Network of Civil Society Organisations for Disaster Reduction, London,
UK.
Goddard, L., Y. Ait Chellouche, W. Baethgen, M. Dettinger, R. Graham, P. Hayman, M.
Kadi, R. Martínez, and H. Meinke, with additional contributions by E. Conrada,
2010: Providing seasonal-to-interannual climate information for risk management
and decision-making. Procedia Environmental Sciences, 1, 81-101
Government of the Philippines, 2008: DSWD leads turnover of core shelter units
to landslide victims in Southern Leyte. Government of the Philippines, Manila,
Philippines, www.reliefweb.int/rw/rwb.nsf/db900SID/FBUO-7JSDJY?OpenDocument.
Government of the Philippines, 2010: Philippines: DSWD, UNDP turn over core
shelter project to Sorsogon beneficiaries. Government of the Philippines,
Manila, Philippines, www.reliefweb.int/rw/rwb.nsf/db900sid/MYAI-
82W3GG?OpenDocument.
Grasso, M., 2010: An ethical approach to climate adaptation finance. Global
Environmental Change, 20(1), 74-81.
Groat, Ch.G., 2004: Seismographs, sensors, and satellites: better technology for safer
communities. Technology in Society, 26(2-3), 169-179.
Grubb, M., 2010: Copenhagen: back to the future. Climate Policy, 10(2), 127-130.
Gupta, J. and M. Grubb, 2000: Competence and subsidiarity. In: Climate change and
European leadership: a sustainable role for Europe? [Gupta, J. and M. Grubb
(eds.)]. Kluwer Academic Publishers, Dordrecht, The Netherlands, 372 pp.
Gupta, J. and N. van de Grijp (eds.), 2010: Mainstreaming Climate Change in
Development Cooperation: Theory, Practice and Implications for the European
Union. Cambridge University Press, Cambridge, UK, 347 pp.
Gupta, J., Å. Persson, L. Olsson, J. Linnerooth-Bayer, N. van der Grijp, A. Jerneck, R.J.T.
Klein, M. Thompson, and A.G. Patt, 2010: Mainstreaming climate change in
development co-operation policy: conditions for success. In: Making Climate
Change Work for Us: European Perspectives on Adaptation and Mitigation
Strategies [Hulme, M. and H. Neufeldt (eds.)]. Cambridge University Press,
Cambridge, UK, pp. 319-339.
Gurenko, E., 2004: Catastrophe Risk and Reinsurance: A Country Risk Management
Perspective. Risk Books, London, UK.
Guzmán, J.M., 2009: The use of population census data for environmental and
climate change analysis. In: Population Dynamics and Climate Change
[Guzmán, J.M., G. Martine, G. McGranahan, D. Schensul, and C. Tacoli (eds.)].
UNFPA and IIED, New York, NY and London, UK, pp. 192-205.
Haines, A., R. Kovats, D. Campbell-Lendrum, and C. Corvalan, 2006: Climate change
and human health: impacts, vulnerability, and mitigation. The Lancet,
367(9528), 2101-2109.
Harrington, K. and J. O’Connor, 2009: How Cisco succeeds at global risk management.
Supply Chain Management Review, July/August 2009.
Hasegawa, S., K. Sato, S. Matsunuma, M. Miyao, and K. Okamoto, 2005: Multilingual
disaster information system: information delivery using graphic text for mobile
phones. AI & Society, 19, 265-278.
Hellmuth, M., D. Osgood, U. Hess, A. Moorhead, and H. Bhojwani (eds.), 2009: Index
Insurance and Climate Risk: Prospects for Development and Disaster
Management. Climate and Society No. 2, International Research Institute for
Climate and Society, Columbia University, New York, NY, 112 pp.
Heltberg, R., P.B. Siegel, and S.L. Jorgensen, 2008: Climate Change, Human
Vulnerability and Social Risk Management. World Bank, Washington, DC.
Hess, U., 2007: Risk management framework- the big LEAP in Ethiopia, presentation,
Insurance in catastrophe risk.
Hess, U. and J. Syroka, 2005: Weather-based Insurance in Southern Africa: The Case
of Malawi. Agriculture and Rural Development Discussion Paper 13, World
Bank, Washington, DC.
Holdaway, R., 2001: Is space global disaster warning and monitoring now nearing
reality? Space Policy, 17(2), 127-132.
Hope Sr., K.H., 1996: Promoting sustainable community development in developing
countries: The role of technology transfer. Community Development Journal,
31(3), 193-200.
Huq, S., 2002: The Bonn–Marrakech agreements on funding. Climate Policy, 2(2),
243-246.
ICRC, 1995: The Code of Conduct for the International Red Cross and Red Crescent
Movement and NGOs in Disaster Relief. International Committee of the Red
Cross, Geneva, Switzerland, www.icrc.org/eng/resources/documents/misc/code-
of-conduct-290296.htm.
ICSU, 2008: A Science Plan for Integrated Research on Disaster Risk: addressing the
challenge of natural and human induced environmental hazards. International
Council for Science, Paris, France.
ICSU, 2010: Regional Environmental Change: Human Action and Adaptation.
International Council for Science, Paris, France.
IEA, 2001: Technology without Borders. Case Studies of Successful Technology
Transfer. OECD/International Energy Agency, Paris, France.
IFRC, 2011: Red Cross/Red Crescent Climate Centre. International Federation of Red
Cross and Red Crescent Societies, Geneva, Switzerland, www.climatecentre.org.
IHRC, 2006: News Updates on the Release of Public Hurricane Loss Model
(Winter/Spring 2006). International Hurricane Research Centre, Florida
International University, Miami, FL.
IMF, 2011: International Monetary Fund Factsheet: The Multilateral Debt Relief
Initiative. International Monetary Fund, Washington, DC, www.imf.org/external/
np/exr/facts/pdf/mdri.pdf.
IPCC, 1990: Climate Change: The IPCC Scientific Assessment. The First Assessment
Report of the Intergovernmental Panel on Climate Change. Cambridge
University Press, Cambridge, United Kingdom and New York, NY, USA.
IPCC, 1996: Climate Change 1995. The Second Assessment Report of the
Intergovernmental Panel on Climate Change. Cambridge University Press,
Cambridge, United Kingdom and New York, NY, USA.
IPCC, 2000: Special Report on Emissions Scenarios. Cambridge University Press,
Cambridge, UK.
IPCC, 2007: Climate Change 2007. The Fourth Assessment Report of the
Intergovernmental Panel on Climate Change. Cambridge University Press,
Cambridge, United Kingdom and New York, NY, USA.
IPCC, 2009. Scoping Meeting for an IPCC Special Report on Extreme Events and
Disasters: Managing the Risks. Proceedings [Barros, V., et al. (eds.)], Oslo,
Norway, 23-26 March 2009.
Chapter 7Managing the Risks: International Level and Integration across Scales
431
IRI, 2006. A Gap Analysis for the Implementation of the Global Climate Observing
System Programme in Africa. IRI Technical Report IRI-TR/06/1, International
Research Institute for Climate and Society, Palisades, NY, portal.iri.columbia.edu/
portal/server.pt/gateway/PTARGS_0_2_2806_0_0_18/GapAnalysis.pdf.
JAXA, 2011: DAICHI (ALOS) Operation Completion. Japan Aerospace Exploration
Agency, Tokyo, Japan, www.jaxa.jp/press/2011/05/20110512_daichi_e.html.
Jordan, A., 2000: The politics of multilevel environmental governance: Subsidiarity
and environmental policy in the European Union. Environment and Planning A,
32(7), 1307-1324.
Kaklauskas A., D. Amaratunga, and R. Haigh, 2009: Knowledge model for post-
disaster management. International Journal of Strategic Property Management,
13, 117-128.
Kameda, H., 2007: Networking disaster risk reduction technology and knowledge
through Disaster Reduction Hyperbase (DRH). In: Proceedings of the Disaster
Reduction Hyperbase (DRH) Contents Meeting, Kobe, Japan, 12-13 March 2007,
drh.edm.bosai.go.jp/Project/Phase2/1Documents/9_EXr.pdf.
Kibreab, G., 1997: Environmental causes and impact of refugee movements: a
critique of the current debate. Disasters, 21(1), 20-38.
Kinzig, A.P., 2001: Bridging disciplinary divides to address environmental and
intellectual challenges. Ecosystems, 4, 709-715.
Klein, R.J.T., 2010: Linking adaptation and development finance: a policy dilemma
not addressed in Copenhagen. Climate and Development, 2(3), 203-206.
Klein, R.J.T. and A. Möhner, 2009: Governance limits to effective global financial
support for adaptation. In: Adapting to Climate Change: Thresholds, Values,
Governance [Adger, W.N., I. Lorenzoni, and K.L. O’Brien (eds.)]. Cambridge
University Press, Cambridge, UK, pp. 465-475.
Klein, R.J.T. and A. Persson, 2008: Financing Adaptation to Climate Change: Issues
and Priorities. European Climate Platform (ECP) Report No. 8. An Initiative of
Mistra’s Climate Policy Research Programme (Clipore) and the Centre for
European Policy Studies (CEPS).
Klein, R.J.T., J. Aston, E.N. Buckley, M. Capobianco, N. Mizutani, R.J. Nicholls, P.D.
Nunn, and S. Ragoonaden, 2000: Coastal adaptation. In: Methodological and
Technological Issues in Technology Transfer [Metz, B., O.R. Davidson, J.-W.
Martens, S.N.M. van Rooijen, and L. Van Wie McGrory (eds.)]. Cambridge
University Press, Cambridge, UK, pp. 349-372.
Klein, R.J.T., W.W. Dougherty, M. Alam, and A. Rahman, 2005: Technology to understand
and manage climate risks. Background Paper for the UNFCCC Seminar on the
Development and Transfer of Environmentally Sound Technologies for
Adaptation to Climate Change, Tobago, 14-16 Jun 2005, UNFCCC, Bonn,
Germany, unfccc.int/ttclear/pdf/Workshops/Tobago/backgroundPaper_old.pdf.
Klein, R.J.T., S.E.H. Eriksen, L.O. Næss, A. Hammill, T.M. Tanner, C. Robledo, and K.L.
O’Brien, 2007: Portfolio screening to support the mainstreaming of adaptation
to climate change into development assistance. Climatic Change, 84(1), 23-44.
Kleindorfer, P.R., 2009: Climate Change and Insurance: Integrative Principles and
Regulatory Risk. INSEAD Working Paper No. 2009/43/TOM/INSEAD, INSEAD,
Fontainebleau, France, ssrn.com/abstract=1456862.
Köhler, P. and J. Wächter, 2006: Towards an open information infrastructure for
disaster research and management: Data management and information systems
inside DFNK. Natural Hazards, 38, 141-157.
Kok, M., B. Metz, J. Verhagen, and S. van Rooijen, 2008: Integrating development and
climate policies: national and international benefits. Climate Policy, 8(2), 103-118.
Kolmannskog, V., 2008: Future floods of refugees. Norwegian Refugee Council,
Oslo, Norway.
Kolmannskog, V. and F. Myrstad, 2009: Environmental displacement in European
asylum law. European Journal of Migration and Law, 11, 313-326.
Kreimer, A. and M. Arnold, 2000: Managing Disaster Risk in Emerging Economies.
Disaster Risk Management Series No 2, World Bank, Washington, DC.
Kuhn, K., D. Campbell-Lendrum, A. Haines, and J. Cox, 2005. Using Climate to Predict
Infectious Disease Epidemics. World Health Organization, Geneva, Switzerland,
whqlibdoc.who.int/publications/2005/9241593865.pdf.
Kunreuther, H. and E. Michel-Kerjan, 2009: At War with the Weather: Managing
Large-Scale Risks in a New Era of Catastrophes. 1st ed. The MIT Press,
Cambridge, MA.
Lamb, H., 1995: Climate, History and a Modern World. 2nd ed. Routledge, New York,
NY.
Lane, M., 2004: The viability and likely pricing of “cat bonds” for developing countries.
In: Catastrophe Risk and Reinsurance: A Country Risk Management Perspective
[Gurenko, E. (ed.)]. Risk Books, London, UK, pp. 239-268.
Lavoyer, J.-P., 2006: International humanitarian law: Should it be reaffirmed,
clarified or developed? In: Issues in International Law and Military Operations
[Jaques, R.B. (ed.)]. International Law Studies Vol. 80, US Naval War College,
Newport, Rhode Island, pp. 287-310.
Le Cozannet, G., S. Hosford, J. Douglas, J.-J. Serrano, D. Coraboeuf, and J. Comte,
2008: Connecting hazard analysts and risk managers to sensor information.
Sensors, 8, 3932-3937.
Lee, W.-J., R.A. Pielke Jr., and L. Anderson-Berry, 2006: Disaster mitigation, warning
systems and societal impact. Rapporteur Report. In: Sixth International
Workshop on Tropical Cyclones, San José, Costa Rica, 21-30 Nov 2006, WMO,
Geneva, Switzerland, severe.worldweather.org/iwtc/document/
Topic_5_M_C_Wong.pdf.
Leemans, R., G. Asrar, A. Busalacchi, J. Canadell, J. Ingram, A. Larigauderie, H.
Mooney, C. Nobre, A. Patwardhan, M. Rice, F. Schmidt, S. Seitzinger, H. Virji, C.
Vorosmarty, and O. Young, 2009: Developing a common strategy for integrative
global environmental change research and outreach: the Earth System Science
Partnership (ESSP) Strategy paper. Current Opinion in Environmental
Sustainability, 1, 4-13.
Lemos, M.C., E. Boyd, E.L. Tompkins, H. Osbahr, and D. Liverman, 2007. Developing
adaptation and adapting development. Ecology and Society, 12(2), 26,
www.ecologyandsociety.org/vol12/iss2/art26/.
Lewis, J., 1999: Development in Disaster-prone Places: Studies of Vulnerability.
Intermediate Technology Publications, London, UK.
Lewis, J., 2007: Climate and disaster reduction. Tiempo Climate Newswatch,
www.tiempocyberclimate.org/newswatch/comment070217.htm.
Linnerooth-Bayer, J. and R. Mechler, 2008: Insurance against Losses from Natural
Disasters in Developing Countries. Background paper for United Nations World
Economic and Social Survey, United Nations Department of Economic and
Social Affairs, New York, NY.
Linnerooth-Bayer, J., R. Loefstedt, and G. Sjostedt (eds.), 2001: Transboundary Risk
Management. Earthscan Publications, London, UK.
Linnerooth-Bayer, J., R. Mechler, and G. Pflug, 2005: Refocusing disaster aid.
Science, 309, 1044-1046.
Linnerooth-Bayer, J., K. Warner, C. Bals, P. Höppe, I. Burton, T. Loster, and A. Haas,
2009: Insurance mechanisms to help developing countries respond to climate
change. The Geneva Papers on Risk and Insurance - Issues and Practice, 34,
381-400.
Linnerooth-Bayer, J., C. Bals, and R. Mechler, 2010: Insurance as part of a climate
adaptation strategy. In: Making Climate Change Work for Us: European
Perspectives on Adaptation and Mitigation Strategies [Hulme, M. and H.
Neufeldt (eds.)].Cambridge University Press, Cambridge, UK.
Louhisuo, M., T. Veijonen, J. Ahola, and T. Morohoshi, 2007: A disaster information
and monitoring system utilizing earth observation. Management of
Environmental Quality, 18(3), 246-262.
Lucas, R. and O. Stark, 1985: Motivations to remit: evidence from Botswana. Journal
of Political Economy, 93(5), 901-918.
Maastricht Treaty, 1992 : Provisions Amending the Treaty Establishing the
European Economic Community with a view to Establishing the European
Community, www.eurotreaties.com/maastrichtec.pdf.
Mace, M.J., 2005: Funding for adaptation to climate change: UNFCCC and GEF
developments since COP-7. Review of European Community & International
Environmental Law, 14(3), 225-246.
Madey, G.R., A. Barabási, N.V. Chawla, M. Gonzalez, D. Hachen, B. Lantz, A. Pawling,
T. Schoenhar, G. Szabó, P. Wang, and P. Yan, 2007: Enhanced situational awareness:
Application of DDDAS concepts to emergency and disaster management. In:
Computational Science – ICCS 2007: 7th International Conference, Beijing,
China, 27-30 May 2007, Proceedings, Part I, LNCS 4487 [Shi, Y. and G.D. Van
Albada (eds.)]. Springer-Verlag, Berlin, Germany, pp. 1090-1097.
Chapter 7 Managing the Risks: International Level and Integration across Scales
432
Maibach, E.W., C. Roser-Renouf, and A. Leiserowitz, 2008: Communication and
marketing as climate change–intervention assets. A public health perspective.
American Journal of Preventive Medicine, 35(5), 488-500.
Maitland, C. and A. Tapia, 2007: Outcomes from the UN OCHA 2002 Symposium &
HIN Workshops on Best Practices in Humanitarian Information Management
and Exchange. College of Information Sciences and Technology, Pennsylvania
State University, University Park, PA.
Marincioni, F., 2007: Information technologies and the sharing of disaster knowledge:
the critical role of professional culture. Disasters, 31(4), pp. 459-476.
McBean, G.A., 2010: Introduction of a New International Research Program:
Integrated Research on Disaster Risk - The Challenge of Natural and Human-
Induced Environmental Hazards. In: Geophysical Hazards: Minimizing Risk,
Maximizing Awareness. Part II [Beer, T. (ed.)]. Springer, Berlin, Germany, pp. 59-
69, www.cprm.gov.br/33IGC/1340016.html.
McGray, H., A. Hammill, R. Bradley, E.L. Schipper, and J.-E. Parry, 2007: Weathering
the Storm: Options for Framing Adaptation and Development. World Resources
Institute, Washington, DC, 57 pp.
Mercer, J., 2010. Disaster risk reduction or climate change adaptation: Are we
reinventing the wheel? Journal of International Development, 22, 247-264.
Michaelowa, A. and K. Michaelowa, 2007: Climate or development: is ODA diverted
from its original purpose? Climatic Change, 84(1), 5-21.
Miller, D. and A. Paulson, 2007: Risk Taking and the Quality of Informal Insurance:
Gambling and Remittances in Thailand. Working Paper No. 2007-01, Federal
Reserve Bank of Chicago, Chicago, IL, ssrn.com/abstract=956401.
Mills, E., 2007: From Risk to Opportunity: Insurer Responses to Climate Change. A
Ceres Report, Ceres, Boston, MA.
Ministry of Foreign Affairs of Denmark and GEF Evaluation Office, 2009:
Joint External Evaluation: Operation of the Least Developed Countries Fund for
Adaptation to Climate Change. MFA, Copenhagen, Denmark, 112 pp.
Mirza, M.M.Q., 2003: Climate change and extreme weather events: Can developing
countries adapt? Climate Policy, 3, 233-248.
Mitchell, T. and M. van Aalst, 2008: Convergence of Disaster Risk Reduction and
Climate Change Adaptation. A Review for DFID 31st October 2008. Department
for International Development (DFID), London, UK, www.preventionweb.net/
files/7853_ConvergenceofDRRandCCA1.pdf.
Mitchell, T., S. Anderson, and S. Huq, 2008: Principles for Delivering Adaptation Finance.
Institute of Development Studies, University of Sussex, Brighton, UK, 6 pp.
Mohapatra, S., J. George, and D. Ratha, 2009: Remittances and Natural Disasters:
Ex-post Response and Contribution to Ex-ante Preparedness. Policy Research
Working Paper 4972, World Bank, Washington, DC.
Morrissey, J., 2009: Environmental Change and Forced Migration - A State of the
Art Review. Background Paper, Refugee Studies Centre, Oxford Department
of International Development, Queen Elizabeth House, University of Oxford,
48 pp., www.rsc.ox.ac.uk/events/environmental-change-and-migration/
EnvChangeandFmReviewWS.pdf.
Müller, B., 2007: Nairobi 2006: Trust and the Future of Adaptation Funding. Oxford
Institute for Energy Studies, Oxford, UK, 26 pp.
Munich Re, 2003: TOPICSgeo: Annual Review – Natural Catastrophes 2002
. Munich
Reinsurance Group, Geoscience Research, Munich, Germany.
Munich Re, 2010: Insurance-linked Securities (ILS). Market Update Q1 2010, Munich
Re, Munich, Germany.
Murphy, B., 2011: Briefing Paper: Technology for Adapting to Climate Change.
weADAPT, the Collaborative Platform on Climate Adaptation, weadapt.org/
knowledge-base/wikiadapt/technology-and-adaptation.
Myers, N., 2001: Environmental refugees: A growing phenomenon of the 21st
century. Philosophical Transactions of the Royal Society B, 357, 609-613.
Myers, N., 2005: Environmental refugees: An emergent security issue. In: Session III
– Environment and Migration, 13th Meeting of the OSCE Economic Forum,
Prague, Czech Republic, 23-27 May 2005, pp. 23-27.
NASCIO, 2006: State of Louisiana GIS Support for Emergency Operations Before,
During, and After the Hurricanes of 2005. National Association of State Chief
Information Officers, Lexington, KY, www.nascio.org/awards/nominations/
2006Louisiana9.pdf.
O’Brien, G., P. O’Keefe, J. Rose, and B. Wisner, 2006: Climate change and disaster
management. Disasters, 30(1), 64-80.
O’Brien, G., P. O’Keefe, and J. Rose, 2007: Energy, poverty and governance.
International Journal of Environmental Studies, 64(5), 607-618.
O’Brien, K. and R. Leichenko, 2000: Double exposure: assessing the impacts of
climate change within the context of economic globalization. Global
Environmental Change, 10, 221-232.
O’Brien, K., L. Sygna, R. Leichenko, W.N. Adger, J. Barnett, T. Mitchell, L. Schipper, T.
Tanner, C. Vogel, and C. Mortreux, 2008: Disaster Risk Reduction, Climate
Change Adaptation and Human Security. GECHS Report 2008:3, prepared for
the Royal Norwegian Ministry of Foreign Affairs by the Global Environmental
Change and Human Security (GECHS) Project.
O’Brien, K., B. Hayward, and F. Berkes, 2009: Rethinking social contracts: building
resilience in a changing climate. Ecology and Society, 14(2), 12,
www.ecologyandsociety.org/vol14/iss2/art12/.
OCHA/IDMC, 2009: Monitoring Disaster Displacement in the Context of Climate
Change. UN Office for the Coordination of Humanitarian Affairs (OCHA) and
Internal Displacement Monitoring Center (IDMC), Geneva, Switzerland.
Ogallo, L., 2010: The mainstreaming of climate change and variability information
into planning and policy development for Africa. Procedia Environmental
Sciences, 1, 405-410.
OHCHR, 2009: Report of the Office of the United Nations High Commissioner for
Human Rights on the Relationship between Climate Change and Human Rights.
UN Doc. A/HRC/10/61, Office of the United Nations High Commissioner for
Human Rights, Geneva, Switzerland.
Oppenheimer, M. and A. Petsonk, 2005: Article 2 of the UNFCCC: historical origins,
recent interpretations. Climatic Change, 73(3), 195-226.
Ott
, H.E., W. Sterk, and R. Watanabe, 2008: The Bali roadmap: new horizons for glob-
al climate policy. Climate Policy, 8(1), 91-95.
Oxfam, 2011: Adaptation and Risk Reduction. Oxfam, Oxford, UK, www.oxfam.org.uk/
resources/issues/climatechange/introduction.html#adaptation.
Oxfam International, 2007: Adapting to Climate Change: What’s Needed in Poor
Countries, and Who Should Pay. Oxfam Briefing Paper 104, Oxfam International
Secretariat, Oxford, UK, 47 pp.
Paavola, J. and W.N. Adger, 2006: Fair adaptation to climate change. Ecological
Economics, 56(4), 594-609.
Parry, M., N. Arnell, P. Berry, D. Dodman, S. Fankhauser, C. Hope, S. Kovats, R. Nicholls,
D. Satterthwaite, R. Tiffin, and T. Wheeler, 2009: Assessing the Costs of
Adaptation to Climate Change: A Review of the UNFCCC and Other Recent
Estimates. International Institute for Environment and Development and
Grantham Institute for Climate Change, London, UK, 111 pp.
Paul, B.K., H. Rashid, M. Shahidul Islam, and L. Hunt, 2010: Cyclone evacuation in
Bangladesh: Tropical cyclones Gorky (1991) vs. Sidr (2007). Environmental
Hazards, 9, 89-101.
Pelling, M., 2011: Adaptation to Climate Change: From Resilience to Transformation.
Routledge, Abingdon, UK.
Persson, Å. and R.J.T. Klein, 2009: Mainstreaming adaptation to climate change into
official development assistance: challenges to foreign policy integration. In:
Climate Change and Foreign Policy: Case Studies from East to West [Harris, P.
(ed.)]. Routledge, London, UK, pp. 162-177.
Persson, Å., R.J.T. Klein, C. Kehler Siebert, A. Atteridge, B. Müller, J. Hoffmaister, M.
Lazarus, and T. Takama, 2009: Adaptation Finance under a Copenhagen Agreed
Outcome. Stockholm Environment Institute, Stockholm, Sweden, 187 pp.
Pickens, M., D. Porteous, and S. Rotman, 2009: Banking the Poor via G2P Payments.
Focus Note 58, Consultative Group to Assist the Poor, Washington, DC.
Prabhakar, S.V.R.K., S. Anch, and R. Shaw, 2009: Climate change and local level
disaster risk reduction planning: need, opportunities and challenges. Mitigation
and Adaptation Strategies for Global Change, 14, 7-33.
Raustiala, K., 2005: Form and substance of international agreements. American
Journal of International Law, 99, 581.
Rautela, P., 2005: Indigenous technical knowledge inputs for effective disaster
management in the fragile Himalayan ecosystem. Disaster Prevention and
Management, 14(2), 233-241.
Chapter 7Managing the Risks: International Level and Integration across Scales
433
Rayner, S., 2010. How to eat an elephant: a bottom-up approach to climate policy.
Climate Policy, 10, 615-621.
Red Cross and Red Crescent Climate Centre, 2007: Red Cross, Red Crescent
Climate Guide. Red Cross, The Hague, The Netherlands.
Reid, W.V., D. Chen, L. Goldfarb, H. Hackmann, Y.T. Lee, K. Mokhele, E. Ostrom, K.
Raivio, J. Rockström, H.J. Schellnhuber, and A. Whyte, 2010: Earth system science
for global sustainability: Grand challenges. Science, 330, 916-917.
Rodriguez, J., F. Vos, R. Below, and D. Guha-Sapir, 2009: Annual Disaster Statistical
Review 2008 –The numbers and trends. Centre for Research on the
Epidemiology of Disasters, Université Catholique de Louvain, Louvain, Belgium,
www.emdat.be/publications.
Rukieh, M. and M. Koudmani, 2006: Use of Space Technology for Natural Disaster
Detection and Prevention. General Organisation of Remote Sensing, Damascus,
Syria, www.iemss.org/iemss2006/papers/s11/288_RUKIEH_0.pdf.
Saab, D., E. Maldonado, R. Orendovici, L. Ngamassi, A. Gorp, K. Zhao, C. Maitland, and
A. Tapia, 2008: Building global bridges: Coordination bodies for improved
information sharing among humanitarian relief agencies. In: Proceedings of the
5th International ISCRAM Conference, Washington, DC, May 2008 [Fiedrich, F.
and B. Van de Walle (eds.)].
Saechao, T.R., 2007: Natural disasters and the responsibility to protect: From chaos
to clarity. Brooklyn Journal of International Law, 32, 663.
Sahu, S., 2009: Guidebook on Technologies for Disaster Preparedness and Mitigation.
Prepared for the Asian and Pacific Centre for Transfer of Technology, New Delhi,
India, technology4sme.net/docs/Guidebook%20on%20Technologies%20for%
20Disaster%20Preparedness%20&%20Mitigation.pdf.
Samarajiva, R., 2005: Policy Commentary: Mobilizing information and communications
technologies for effective disaster warning: lessons from the 2004 tsunami.
New Media & Society, 7(6), 731-747.
Sanusi, Z.A., 2005: Technology Transfer under Multilateral Environmental
Agreements: Analyzing the Synergies. UNU-IAS Working Paper No. 134,
Institute of Advanced Studies, United Nations University, Yokohama, Japan.
Savage, K. and P. Harvey (eds.), 2007: Remittances during Crises: Implications for
Humanitarian Response. Overseas Development Institute, London, UK.
Schipper, E.L.F., 2006: Conceptual history of adaptation in the UNFCCC process.
Review of European Community & International Environmental Law, 15(1),
82-92.
Schipper, E.L.F., 2009: Meeting at the crossroads?: Exploring the linkages between
climate change adaptation and disaster risk reduction. Climate and
Development, 1, 16-30.
Schipper, L. and Pelling, M., 2006: Disaster risk, climate change, and international
development, scope for, and challenges to integration. Disasters, 30 (1) 19-38
Schmidt, S., C. Kemfert, and P. Höppe, 2009: Tropical cyclone losses in the USA and
the impact of climate change – A trend analysis based on data from a new
approach to adjusting storm losses. Environmental Impact Assessment Review,
29, 359-369.
Sheffi, Y., 2005: The Resilient Enterprise. MIT Press, Cambridge, MA.
Shimizu, T., 2008: Disaster Management Satellite System Development and
International Cooperation Promotion in Asia. Science and Technology Trends
Quarterly Review, No. 27/April 2008, 93-108.
Skees, J., 2001: The bad harvest: More crop insurance reform: A good idea gone
awry. Regulation: The CATO Review of Business and Government, 24, 16-21.
Skees, J.R. and A. Enkh-Amgalan, 2002: Examining the feasibility of livestock insurance
in Mongolia. World Bank Working Paper 2886, World Bank, Washington, DC.
Skees, J., B. Barnett, and A. Murphy, 2008: Creating insurance markets for natural
disaster risk in lower income countries: the potential role for securitization.
Agricultural Finance Review, 68, 151-157.
Smit, B., O. Pilifosova, I. Burton, B. Challenger, S. Huq, R.J.T. Klein, and G. Yohe, 2001:
Adaptation to climate change in the context of sustainable development and
equity. In: Climate Change 2007. Impacts, Adaptation and Vulnerability.
Contribution of Working Group II to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P.
Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 877-912.
Sobel, R.S. and P.T. Leeson, 2007: The use of knowledge in natural-disaster relief
management. The Independent Review, 11(4), 519- 532.
Sphere Project, 2004: Humanitarian Charter and Minimum Standards in Disaster
Response. The Sphere Project, Geneva, Switzerland.
Steffen, W., A. Sanderson, P.D. Tyson, J. Jäger, P.A. Matson, B. Moore III, F. Oldfield,
K. Richardson, H.J. Schellnhuber, B.L. Turner, and R.J. Wasson, 2004: Global
Change and the Earth System: A Planet Under Pressure. Springer-Verlag, Berlin,
Germany.
Stern, N., 2007: The Economics of Climate Change: The Stern Review. Cambridge
University Press, Cambridge, UK.
Stoiber, E., 2006. Why Europe needs a subsidiarity early-warning mechanism.
Europe’s World, Summer 2006.
Suarez, P. and J. Linnerooth-Bayer, 2010: Micro-insurance for local adaptation. Wiley
Interdisciplinary Reviews: Climate Change, 1(2), 271-278.
Suarez, P., J. Linnerooth-Bayer, and R. Mechler, 2007: The Feasibility of Risk Financing
Schemes for Climate Adaptation: The Case of Malawi. DEC-Research Group,
Infrastructure and Environment Unit, World Bank, Washington, DC.
Tacoli, C., 2009: Crisis or adaptation? Migration and climate change in a context of
high mobility. Environment and Urbanization, 21, 513-525.
Thomalla, F., T. Downing, E. Spanger-Siegfried, G. Han, and J. Rockström, 2006:
Reducing hazard vulnerability: towards a common approach between disaster
risk reduction and climate adaptation. Disasters, 30(1), 39-48.
Tomasini, R. and L. Van Wassenhove, 2009: Humanitarian Logistics. Palgrave
Macmillan, London, UK.
Tompkins, E.L. and W.N. Adger, 2005: Defining a response capacity for climate
change. Environmental Science and Policy, 8, 562-571.
Trouwborst, A., 2002: Evolution and Status of the Precautionary Principle in
International Law. Kluwer, Dordrecht, The Netherlands.
Tschoegl, L. (with R. Below and D. Guha-Sapir), 2006: An analytical review of selected
data sets on natural disasters and impacts. In: UNDP/CRED Workshop on
Improving Compilation of Reliable Data on Disaster Occurrence and Impact,
CRED, Brussels, Belgium, pp. 2-21.
UN, 1970: UN General Assembly resolution 2626, The International Development
Strategy for the Second United Nations Development Decade (25th session).
Earthscan, London, UK.
UN, 1992: United Nations Framework Convention on Climate Change. United
Nations, New York, NY, unfccc.int/resource/docs/convkp/conveng.pdf.
UN, 2011: Millennium Goals Report 2011. United Nations, New York, NY,
www.un.org/millenniumgoals/pdf/%282011_E%29%20MDG%20Report%202
011_Book%20LR.pdf.
UNCED, 1992: Rio Declaration 1992. United Nations Conference on Environment and
Development, Rio de Janeiro, Brazil.
UNDP, 2002: A Climate Risk Management Approach to Disaster Reduction and
Adaptation to Climate Change. United Nations Development Programme
Expert Group Meeting, Integrating Disaster Reduction with Adaptation to
Climate Change, Havana, Cuba, 19-21 June 2002, 24 pp.
UNFCCC, 2005: Report on the seminar on the development and transfer of technologies
for adaptation to climate change. Note by the secretariat. United Nations
Framework Convention on Climate Change, Bonn, Germany, unfccc.int/
resource/docs/2005/sbsta/eng/08.pdf.
UNFCCC, 2006a: Synthesis report on technology needs identified by Parties not
included in Annex I to the Convention. Note by the secretariat. United Nations
Framework Convention on Climate Change, Bonn, Germany, unfccc.int/
resource/docs/2006/sbsta/eng/inf01.pdf.
UNFCCC, 2006b: Technologies for adaptation to climate change. United Nations
Framework Convention on Climate Change, Bonn, Germany, unfccc.int/
resource/docs/publications/tech_for_adaptation_06.pdf.
UNFCCC, 2007a: Bali Action Plan. Report of the Conference of the Parties (COP 13),
Bali, adopted by Decision 1/CP.13 of the COP-13, United Nations Framework
Convention on Climate Change, Bonn, Germany. unfccc.int/files/meetings/
cop_13/application/pdf/cp_bali_action.pdf.
UNFCCC, 2007b: Investment and Financial Flows to Address Climate Change. United
Nations Framework Convention on Climate Change, Bonn, Germany, 272 pp.
Chapter 7 Managing the Risks: International Level and Integration across Scales
434
UNFCCC, 2008: Mechanisms to manage financial risks from direct impacts of climate
change in developing countries. Technical paper, FCCC/TP/2008/9, United
Nations Framework Convention on Climate Change, Bonn, Germany.
UNFCCC, 2009a: Recommendations on future financing options for enhancing the
development, deployment, diffusion and transfer of technologies under the
Convention. FCCC/SB/2009/2, United Nations Framework Convention on Climate
Change, Bonn, Germany, unfccc.int/resource/docs/2009/sb/eng/02sum.pdf.
UNFCCC, 2009b: Draft decision CP.15 Copenhagen Accord. FCCC/CP/2009/L.7,
United Nations Framework Convention on Climate Change, Bonn, Germany.
UNFCCC, 2010a: Nairobi Work Programme. United Nations Framework Convention
on Climate Change, Bonn, Germany, unfccc.int/adaptation/nairobi_work_
programme/items/3633.php.
UNFCCC, 2010b: Potential Costs and Benefits of Adaptation Options: A Review of
Existing Literature. FCCC/TP/2009/2/Rev.1, United Nations Framework
Convention on Climate Change, Bonn, Germany, unfccc.int/resource/docs/2009/
tp/02r01.pdf.
UNFCCC, 2010c: Report of the Conference of the Parties on its Sixteenth Session,
held in Cancun from 29 November to 10 December 2010. United Nations
Framework Convention on Climate Change, UNFCCC/CP/2010/7/Add.1. Also
available on unfccc.int/resource/docs/2010/cop16/eng/07a01.pdf#page=2.
UNFCCC, 2011: Draft decision -/CP.16: Further guidance for the operation of the
Least Developed Countries Fund. United Nations Framework Convention on
Climate Change, Bonn, Germany, unfccc.int/files/meetings/cop_16/conference_
documents/application/pdf/20101204_cop16_ldcf.pdf.
UNGA, 2010: Implementation of the International Strategy for Disaster Reduction.
Report of the Secretary General to the Sixty- fifth session, Ref A/65/ 388, United
Nations General Assembly, New York, NY.
UNGA, 2000: United Nations Millennium Declaration. Resolution adopted by the
General Assembly, United Nations General Assembly (A/55/L.2), New York, NY.
UNHCR, 2009: Climate Change, Natural Disasters and Human Displacement: A
UNHCR Perspective. United Nations High Commissioner for Refugees, Geneva,
Switzerland, www.unhcr.org/refworld/docid/4a8e4f8b2.html.
UNISDR, 2005a: Hyogo Framework for Action 2005-2015: Building the resilience of
nations and communities to disasters. In: Report of the World Conference on
Disaster Risk Reduction, Jan. 2005, Kobe, Japan, United Nations International
Strategy for Disaster Reduction, Geneva, Switzerland, pp. 40-62.
UNISDR, 2005b: The Link between Millennium Development Goals (MDGs) and
Disaster Risk Reduction. United Nations International Strategy for Disaster
Reduction, Geneva, Switzerland, www.unisdr.org/eng/mdgs-drr/link-mdg-drr.htm.
UNISDR, 2005c: World Conference on Disaster Reduction (WCDR): Proceedings of
the Conference held at Kobe Japan. United Nations International Strategy for
Disaster Reduction, Geneva, Switzerland, www.unisdr.org/wcdr/thematic-ses-
sions/WCDR-proceedings-of-the-Conference.pdf.
UNISDR, 2008: Indicators of Progress: Guidance on Measuring the Reduction of
Disaster Risks and the Implementation of the Hyogo Framework of Action. United
Nations International Strategy for Disaster Reduction, Geneva, Switzerland.
UNISDR, 2009a: Global Assessment Report on Disaster Risk Reduction. United Nations
International Strategy for Disaster Reduction, Geneva, Switzerland, 207 pp.,
www.preventionweb.net/english/hyogo/gar/report/index.php?id=1130&pid:34
&pih.
UNISDR, 2009b: Second Global Platform on Disaster Risk Reduction, Geneva, June
2009: Concluding Summary by the Platform Chair. United Nations International
Strategy for Disaster Reduction, Geneva, Switzerland.
UNISDR, 2011a: Revealing Risk, Redefining Development. Global Assessment Report
on Disaster Risk Reduction. United Nations International Strategy for Disaster
Reduction, Geneva, Switzerland.
UNISDR, 2011b: Hyogo Framework for Action 2005-2015: Building the Resilience of
Nations and Communities to Disasters, Mid-Term Review 2010-2011. United
Nations International Strategy for Disaster Reduction, Geneva, Switzerland.
UNISDR, 2011c: Response to the United Kingdom’s Department for International
Development : Multilateral Aid Review of the UN Secretariat of the
International Strategy for Disaster Reduction. United Nations International
Strategy for Disaster Reduction, Geneva, Switzerland.
Universalia Management Group, 2010: Evaluation of the World Bank Global
Facility for Disaster Reduction and Recovery (GFDRR). Universalia Management
Group, Montreal, Canada, gfdrr.org/docs/GFDRR_EvaluationReportVol-I.pdf.
Vasterman, P., C.J. Yzermans, and A.J.E. Dirkzwager, 2005: The role of the media
and media hypes in the aftermath of disasters. Epidemiologic Reviews, 27,
107-114.
Venton, P. and S. La Trobe, 2008: Linking Climate Change Adaptation and Disaster
Risk Reduction. Tearfund, Teddington, UK.
Vera, C., M. Barange, O.P. Dube, L. Goddard, D. Griggs, N. Kobysheva, E. Odada, S.
Parey, J. Polovina, G. Poveda, B. Seguin and K. Trenberth, 2010: Needs assessment
for climate information on decadal timescales and longer. Procedia
Environmental Sciences, 1, 275-286.
Vernon, T., 2008: The economic case for pro-poor adaptation: what do we know? IDS
Bulletin, 39(4), 32-41.
Vogel, C. and K. O’Brien, 2006: Who can eat information? Examining the effectiveness
of seasonal climate forecasts and regional climate-risk management strategies.
Climate Research, 33, 111-122.
Von Doussa, J., A. Corkery, and R. Chartres, 2007: Human rights and climate change.
Australian International Law Journal, 14, 161.
Wamsler, C., 2006: Mainstreaming risk reduction in urban planning and housing: a
challenge for international aid organizations. Disasters, 30(2), 151-177.
Warner, K., M. Hamza, A. Oliver-Smith, F. Renaud, and A. Julca, 2009a: Climate
change, environmental degradation and migration. Natural Hazards,
55(3),689-715.
Warner, K., N. Ranger, S. Surminski, M. Arnold, J. Linnerooth-Bayer, E. Michel-Derjan,
P. Kovacs, and C. Herweijer, 2009b: Adaptation to Climate Change: Linking
Disaster Risk Reduction and Insurance. United Nations International Strategy
for Disaster Reduction, Geneva, Switzerland.
Warnock, A., 2007. Small island developing states of the Pacific and climate change:
Adaptation and alternatives. New Zealand Yearbook of International Law, 4,
247-286.
Watkins, K. (ed.), 2007: Fighting Climate Change: Human Solidarity in a Divided
World. Human Development Report 2007/2008 of the United Nations
Development Programme, Palgrave Macmillan, Basingstoke, UK and New York,
NY, USA, 384 pp.
Weingart, P., A. Engels, and P. Pansegrau, 2000: Risks of communication: discourses
on climate change in science, politics, and the mass media.
Public
Understanding of Science, 9(3), 261-283.
WHO, 2007: International Health Regulations. World Health Organization, Geneva,
Switzerland, www.who.int/gb/ebwha/pdf_files/WHA58/WHA58_3-en.pdf.
Wisner, B., P. Blaikie, T. Cannon, and I. Davis, 2004: At Risk: Natural Hazards, People’s
Vulnerability and Disasters. 2nd Ed. Routledge, London, UK.
WMO, 1995: Resolution 40 (Cg-XII), WMO Policy and Practice for the Exchange of
Meteorological and Related Data and Products Including Guidelines on
Relationships in Commercial Meteorological Activities. Twelfth WMO Congress,
World Meteorological Organization, Geneva, Switzerland.
WMO, 2007: Commission for Basic Systems Management Group Seventh Session.
Final Report. World Meteorological Organization, Geneva, Switzerland.
WMO, 2009: Thematic progress review on Early Warning Systems. Background paper
for Global Assessment Report on Disaster Risk Reduction, World Meteorological
Organization, Geneva, Switzerland.
WMO, 2010: High-level Taskforce towards the Global Framework for Climate Services
(GFCS). World Meteorological Organization, Geneva, Switzerland, www.wmo.int/
hlt-gfcs/index_en.html.
WMO, 2011: Resolution 48 (Cg-XVI) - Response to the Report of the High-Level Task
Force on the Global Framework for Climate Services. World Meteorological
Organization, Geneva, Switzerland.
Wolz, C., and N. Park, 2006: Evaluation of ReliefWeb. Prepared for Office for the
Coordination of Humanitarian Affairs, United Nations by Forum One
Communications, www.reliefweb.int/rw/lib.nsf/db900SID/LTIO-6VLQJP?
OpenDocument.
World Bank, 2006: Clean Energy and Development: Towards an Investment
Framework. DC2006- 0002, World Bank, Washington, DC, 146 pp.
Chapter 7Managing the Risks: International Level and Integration across Scales
435
World Bank, 2007a: A Partnership for Mainstreaming Disaster Mitigation in Poverty
Reduction Strategies. Global Facility for Disaster Reduction and Recovery
(GFDRR), World Bank, Washington, DC.
World Bank, 2007b: The Caribbean catastrophe risk insurance initiative: results of
preparation work on the design of a Caribbean Catastrophe Risk Insurance
Facility. World Bank, Washington, DC.
World Bank, 2008: Catastrophe risk deferred drawdown option (DDO), or CAT DDO.
Background Note, World Bank, Washington, DC.
World Bank, 2009: Climate Investment Funds: PPCR Programming and Financing
Modalities. World Bank, Washington, DC, siteresources.worldbank.org/INTCC/
Resources/ppcrprogrammingdraftvers2april23.pdf.
World Bank, 2010a: World Development Report 2010: Development and Climate
Change. World Bank, Washington, DC.
World Bank, 2010b: Economics of Adaptation to Climate Change: Synthesis Report.
World Bank, Washington, DC, 101 pp.
World Bank, 2011a: Disaster Risk Management Programs for Priority Countries.
World Bank, Global Facility for Disaster Reduction and Recovery (GFDRR),
Washington, DC.
World Bank, 2011b: Migration and Remittances Factbook. World Bank, Washington,
DC.
WTO, 2011: Legal Texts of the World Trade Organization. World Trade Organization,
Geneva, Switzerland, docsonline.wto.org/gen_browseDetail.asp?preprog=3.
Yamin, F., 2005: The European Union and future climate policy: Is mainstreaming
adaptation a distraction or part of the solution? Climate Policy, 5(3), 349-361.
Yamin, F. and J. Depledge, 2004: The International Climate Change Regime: A Guide
to Rules, Institutions and Procedures. Cambridge University Press, Cambridge,
UK, 699 pp.
Yang, D. and H. Choi, 2007: Are remittances insurance? Evidence from rainfall shocks
in the Philippines. The World Bank Economic Review, 21(2), 219-248.
Yohe, G.W., R.D. Lasco, Q.K. Ahmad, N.W. Arnell, S.J. Cohen, C. Hope, A.C. Janetos, and
R.T. Perez, 2007: Perspectives on climate change and sustainable development.
In: Climate Change 2007. Impacts, Adaptation and Vulnerability. Contribution of
Working Group II to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der
Linde, and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK, pp.
811-841.
Zhang, D., L. Zhou, and J.F. Nunamaker Jr., 2002: A knowledge management
framework for the support of decision making in humanitarian assistance/
disaster relief. Knowledge and Information Systems, 4, 370-385.
Zschocke, T. and J.C.V. de Leon, 2010: Towards an ontology for the description of
learning resources on disaster risk reduction. In: Knowledge Management,
Information Systems, E-Learning, and Sustainability Research [Lytras, M.D., P.
Ordóñez De Pablos, A. Ziderman, A. Roulstone, H. Maurer, and J.B. Imber (eds.)].
Proceedings of the Third World Summit of the Knowledge Society, WSKS 2010,
Corfu, Greece, 22-24 Sep. 2010, Springer-Verlag, Berlin, Germany, pp. 60-74.
Chapter 7 Managing the Risks: International Level and Integration across Scales
436
Chapter 7Managing the Risks: International Level and Integration across Scales
437
Coordinating Lead Authors:
Karen O’Brien (Norway), Mark Pelling (UK), Anand Patwardhan (India)
Lead Authors:
Stephane Hallegatte (France), Andrew Maskrey (Switzerland), Taikan Oki (Japan), Úrsula Oswald-Spring
(Mexico), Thomas Wilbanks (USA), Pius Zebhe Yanda (Tanzania)
Review Editors:
Carlo Giupponi (Italy), Nobuo Mimura (Japan)
Contributing Authors:
Frans Berkhout (Netherlands), Reinette Biggs (South Africa), Hans Günter Brauch (Germany),
Katrina Brown (UK), Carl Folke (Sweden), Lisa Harrington (USA), Howard Kunreuther (USA),
Carmen Lacambra (Colombia), Robin Leichenko (USA), Reinhard Mechler (Germany),
Claudia Pahl-Wostl (Germany), Valentin Przyluski (France), David Satterthwaite (UK), Frank Sperling
(Germany), Linda Sygna (Norway), Thomas Tanner (UK), Petra Tschakert (Austria), Kirsten Ulsrud
(Norway), Vincent Viguié (France)
This chapter should be cited as:
O’Brien, K., M. Pelling, A. Patwardhan, S. Hallegatte, A. Maskrey, T. Oki, U. Oswald-Spring, T. Wilbanks, and P.Z. Yanda, 2012:
Toward a sustainable and resilient future. In: Managing the Risks of Extreme Events and Disasters to Advance Climate
Change Adaptation [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach,
G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups I and II of the
Intergovernmental Panel on Climate Change (IPCC). Cambridge University Press, Cambridge, UK, and New York, NY, USA,
pp. 437-486.
8
Toward a Sustainable
and Resilient Future
Toward a Sustainable and Resilient Future
438
Executive Summary .................................................................................................................................439
8.1. Introduction..............................................................................................................................441
8.2. Disaster Risk Management as Adaptation:
Relationship to Sustainable Development Planning ...............................................................443
8.2.1. Concepts of Adaptation, Disaster Risk Reduction, and Sustainable Development and how they are Related..............443
8.2.2. Sustainability of Ecosystem Services in the Context of Disaster Risk Management and Climate Change Adaptation......445
8.2.3. The Role of Values and Perceptions in Shaping Response ..............................................................................................446
8.2.4. Technology Choices, Availability, and Access...................................................................................................................447
8.2.5. Tradeoffs in Decisionmaking: Addressing Multiple Scales and Stressors........................................................................448
8.3. Integration of Short- and Long-Term Responses to Extremes.................................................450
8.3.1. Implications of Present-Day Responses for Future Well-Being .......................................................................................450
8.3.2. Barriers to Reconciling Short- and Long-Term Goals.......................................................................................................451
8.3.3. Connecting Short- and Long-Term Actions to Promote Resilience..................................................................................453
8.4. Implications for Access to Resources, Equity, and Sustainable Development ........................454
8.4.1. Capacities and Resources: Availability and Limitations...................................................................................................454
8.4.2. Local, National, and International Winners and Losers....................................................................................................456
8.4.3. Potential Implications for Human Security ......................................................................................................................457
8.4.4. Implications for Achieving Relevant International Goals................................................................................................458
8.5. Interactions among Disaster Risk Management, Adaptation to
Climate Change Extremes, and Mitigation of Greenhouse Gas Emissions..............................458
8.5.1. Thresholds and Tipping Points as Limits to Resilience ....................................................................................................458
8.5.2. Adaptation, Mitigation, and Disaster Risk Management Interactions............................................................................459
8.5.2.1. Urban....................................................
................................................................................................
............................................460
8.5.2.2. Rural .................................................................................................................................................................................................461
8.6. Options for Proactive, Long-Term Resilience to Future Climate Extremes..............................462
8.6.1. Planning for the Future....................................................................................................................................................462
8.6.2. Approaches, Tools, and Integrating Practices ..................................................................................................................463
8.6.2.1. Improving Analysis and Modeling Tools.....................................................
.......................................................................................464
8.6.2.2. Institutional Approaches...................................................................................................................................................................464
8.6.2.3. Transformational Strategies and Actions for Achieving Multiple Objectives.....................................................................................465
8.6.3. Facilitating Transformational Change ...
....
.......................................................................................................................466
8.6.3.1. Adaptive Management ................................................................................................................................................
.....................467
8.6.3.2. Learning............................................................................................................................................................................................467
8.6.3.3. Innovation ........................................................................................................................................................................................468
8.6.3.4. Leadership ........................................................................................................................................................................................469
8.7. Synergies between Disaster Risk Management and
Climate Change Adaptation for a Resilient and Sustainable Future.......................................469
References ...............................................................................................................................................471
Chapter 8
Table of Contents
439
Actions that range from incremental steps to transformational changes are essential for reducing risk from
weather and climate extremes (high agreement, robust evidence). [8.6, 8.7] Incremental steps aim to improve
efficiency within existing technological, governance, and value systems, whereas transformation may involve alterations
of fundamental attributes of those systems. The balance between incremental and transformational approaches
depends on evolving risk profiles and underlying social and ecological conditions. Disaster risk, climate change
impacts, and capacity to cope and adapt are unevenly distributed. Vulnerability is often concentrated in poorer countries
or groups, although the wealthy can also be vulnerable to extreme events. Where vulnerability is high and adaptive
capacity relatively low, changes in extreme climate and weather events can make it difficult for systems to adapt
sustainably without transformational changes. Such transformations, where they are required, are facilitated through
increased emphasis on adaptive management, learning, innovation, and leadership.
Evidence indicates that disaster risk management and adaptation policy can be integrated, reinforcing,
and supportive – but this requires careful coordination that reaches across domains of policy and practice
(high agreement, medium evidence). [8.2, 8.3, 8.5, 8.7] Including disaster risk management in resilient and
sustainable development pathways is facilitated through integrated, systemic approaches that enhance capacity to
cope with, adapt to, and shape unfolding processes of change, while taking into consideration multiple stressors,
different prioritized values, and competing policy goals.
Development planning and post-disaster recovery have often prioritized strategic economic sectors and
infrastructure over livelihoods and well-being in poor and marginalized communities. This can generate
missed opportunities for building local capacity and integrating local development visions into longer-term
strategies for disaster risk reduction and adaptation to climate change (high agreement, robust evidence).
[8.4.1, 8.5.2] A key constraint that limits pathways to post-disaster resilience is the time-bound nature of reconstruction
funding. The degradation of ecosystems providing essential services also limits options for future risk management
and adaptation actions locally.
Learning processes are central in shaping the capacities and outcomes of resilience in disaster risk
management, climate change adaptation, and sustainable development (high agreement, robust evidence).
[8.6.3, 8.7] An iterative process of monitoring, research, evaluation, learning, and innovation can reduce disaster risks
and promote adaptive management in the context of extremes. Technological innovation and access may help achieve
resilience, especially when combined with capacity development anchored in local contexts.
Progress toward resilient and sustainable development in the context of changing climate extremes can
benefit from questioning assumptions and paradigms, and stimulating innovation to encourage new
patterns of response (medium agreement, robust evidence). [8.2.5, 8.6.3, 8.7] Successfully addressing disaster
risk, climate change, and other stressors often involves embracing broad participation in strategy development, the
capacity to combine multiple perspectives, and contrasting ways of organizing social relations.
Multi-hazard risk management approaches provide opportunities to reduce complex and compound hazards
in rural and urban contexts (high agreement, robust evidence). [8.2.5, 8.5.2, 8.7] Considering multiple types of
hazards reduces the likelihood that risk reduction efforts targeted at one type of hazard will increase exposure and
vulnerability from other hazards, both in the present and future. Building adaptation into multi-hazard risk management
involves consideration of current climate variability and projected changes in climate extremes, which pose different
challenges to affected human and natural systems than changes in the means. Where changes in extremes cause
greater stresses on human and natural systems, direct impacts may be more unpredictable, increasing associated
adaptation challenges.
The most effective adaptation and disaster risk reduction actions are those that offer development benefits
in the relative near term, as well as reductions in vulnerability over the longer term (high agreement,
Chapter 8 Toward a Sustainable and Resilient Future
Executive Summary
440
medium evidence). [8.2.1, 8.3.1, 8.3.2, 8.5.1, 8.6.1] There are tradeoffs between current decisions and long-term
goals linked to diverse values, interests, and priorities for the future. Short-term and long-term perspectives on both
disaster risk management and adaptation to climate change thus can be difficult to reconcile. Such reconciliation
involves overcoming the disconnect between local risk management practices and national institutional and legal
frameworks, policy, and planning. Resilience thinking offers some tools for reconciling short- and long-term responses,
including integrating different types of knowledge, an emphasis on inclusive governance, and principles of adaptive
management. However, limits to resilience are faced when thresholds or tipping points associated with social and/or
natural systems are exceeded.
Building a strong foundation for integrating disaster risk management and adaptation to climate change
includes making transparent the values and interests that underpin development, including who wins and
loses from current policies and practices, and the implications for human security (high agreement, medium
evidence). [8.2.3, 8.2.4, 8.4.2, 8.4.3, 8.6.1.2] Both disaster risk management and adaptation to climate change share
challenges related to (1) reassessing and potentially transforming the goals, functions, and structure of institutions
and governance arrangements; (2) creating synergies across temporal and spatial scales; and (3) increasing access to
information, technology, resources, and capacity. These challenges are particularly demanding in countries and localities
with the highest climate-related risks and weak capacities to manage those risks. Countries with significant capacity
and strong risk management records also benefit from addressing these challenges.
Social, economic, and environmental sustainability can be enhanced by disaster risk management and
adaptation approaches. A prerequisite for sustainability is addressing the underlying causes of vulnerability,
including the structural inequalities that create and sustain poverty and constrain access to resources
(medium agreement, robust evidence). [8.6.2, 8.7] This involves integrating disaster risk management in other
social and economic policy domains, as well as a long-term commitment to managing risk.
The interactions among climate change mitigation, adaptation, and disaster risk management will have a
major influence on resilient and sustainable pathways (high agreement, low evidence). [8.2.5, 8.5.2, 8.7]
Interactions between the goals of mitigation and adaptation in particular will play out locally, but have global
consequences.
There are many approaches and pathways to a sustainable and resilient future. Multiple approaches and
development pathways can increase resilience to climate extremes (medium agreement, medium evidence).
[8.2.3, 8.4.1, 8.6.1, 8.7] Choices and outcomes for adaptive actions to climate extremes must reflect divergent
capacities and resources and multiple interacting processes. Actions are framed by tradeoffs between competing
prioritized values and objectives, and different visions of development that can change over time. Iterative, reflexive
approaches allow development pathways to integrate risk management so that diverse policy solutions can be
considered, as risk and its measurement, perception, and understanding evolve over time. Choices made today can
reduce or exacerbate current or future vulnerability, and facilitate or constrain future responses.
Chapter 8Toward a Sustainable and Resilient Future
441
8.1. Introduction
This chapter focuses on the implications of changing climate extremes
for development, and considers how disaster risk management and
climate change adaptation together can contribute to a sustainable and
resilient future. Changes in the frequency, timing, magnitude, and
characteristics of extreme events pose challenges to the goals of reducing
disaster risk and vulnerability, both in the present and in the future (see
Chapter 3). Enhancing the capacity of social-ecological systems to cope
with, adapt to, and shape change is central to building sustainable and
resilient development pathways in the face of climate change. The
concept for social-ecological systems recognizes the interdependence of
social and ecological factors in the generation and management of risk,
as well as in the pursuit of sustainable development. Despite 20 years
on the policy agenda, sustainable development remains contested and
elusive (Hopwood et al., 2005). However, within the context of climate
change, it is becoming increasingly clear that the sustainability of humans
on the Earth is closely linked to resilient social-ecological systems, which
is influenced by social institutions, human agency, and human capabilities
(Pelling, 2003; Bohle et al., 2009; Adger et al., 2011).
Extremes are translated into impacts by the underlying conditions of
exposure and vulnerability associated with development contexts. For
example, there is robust evidence that institutional arrangements and
governance weaknesses can transform extreme events into disasters
(Hewitt, 1997; Pelling, 2003; Wisner et al., 2004; Ahrens and Rudolph,
2006). The potential for concatenated global impacts of extreme events
continues to grow as the world’s economy becomes more interconnected,
but in relative terms most impacts will occur in contexts with severe
environmental, economic, technological, cultural, and cognitive limits to
adaptation (see Section 5.5.3). In relation to extreme events, global risk
assessments show that social losses – as well as economic losses as a
proportion of livelihood or GDP – are disproportionately concentrated in
developing countries, and within these countries in poorer communities
and households (UNDP, 2004; UNISDR, 2009, 2011; World Bank, 2010a).
This chapter recognizes that outcomes of changing extreme events depend
on responses and approaches to disaster risk reduction and climate
change adaptation, both of which are closely linked to development
processes. The assessment of literature presented in this chapter shows
that changes in extreme events call for greater alignment and integration
of climate change responses and sustainable development strategies,
and that this alignment depends on greater coherence between short-
and long-term objectives. Yet there are different interpretations of
development, different preferences and prioritized values and motivations,
different visions for the future, and many tradeoffs involved. Research
on the resilience of social-ecological systems provides some lessons for
addressing the gaps among these objectives. Transformative social,
economic, and environmental responses can facilitate disaster risk
reduction and adaptation (see Box 8-1). Transformations often include
questioning of social values, institutions, and technical practices (Loorbach
et al., 2008; Hedrén and Linnér, 2009; Pelling 2010a). A resilient and
sustainable future is a choice that involves proactive measures that
promote transformations, including adaptive management, learning,
innovation, and leadership capacity to manage risks and uncertainty.
In this chapter, we assess a broad literature presenting insights on how
diverse understandings and perspectives on disaster risk reduction and
climate change adaptation can help to promote a more sustainable and
resilient future. Drawing on many of the key messages from earlier
chapters, the objective is to assess scientific knowledge on the
incremental and transformative changes needed, particularly in relation
to integrating disaster risk reduction and climate change adaptation
into development policies and pathways. Bringing together experience
from a range of disciplines, this chapter identifies proven pathways that
can help move from an incremental to an integrative approach that also
Chapter 8 Toward a Sustainable and Resilient Future
Box 8-1 | Transformation in Response to Changing Climate Extremes
Transformation involves fundamental changes in the attributes of a system, including value systems; regulatory, legislative, or bureaucratic
regimes; financial institutions; and technological or biophysical systems (see Glossary). This chapter focuses on the transformation of
disaster risk management systems in the context of climate extremes, through integration with climate change adaptation strategies and
wider systems of human development. This is similar to, yet distinct from, other types of transformation associated with climate change.
For example, there have been attempts to understand climate change and development failures by identifying the scope for political
(Harvey, 2010), social (Kovats et al., 2005), economic (Jackson, 2009), and value (Leiserowitz et al., 2006) transformation, and so too for
disaster risk management (Klein, 2007). Across these cases, observed processes of stasis and change are analogous (often using common
language), but actors and objectives are distinct. That said, transformation in wider political, economic, social, and ethical systems can
open or close policy space for a more resilient and sustainable form of disaster risk management (Birkland, 2006), just as acts aimed at
transformation in managing climate extremes can have implications for wider systems. This is particularly true where contemporary
development goals, paths, and hierarchies are identified and addressed as part of the root or proximate causes of vulnerability and risk,
that is, when they are seen as part of the solution for building resilient and sustainable futures (Wisner, 2003; Pelling, 2010a). Although
there has been some research on how and why social lock-in makes it difficult to move away from established development priorities
and trajectories (Pelling and Manuel-Navarrete, 2011), there has been only limited academic work to date on the ways in which wider
transformations impact on disaster risk management, and vice versa.
442
embraces transformation – as illustrated in Figure 8-1, which depicts
resilience as a moving target that is positioned somewhere between the
acceptability of residual risk and the costs of risk management. The target
moves as the relationship between risk and uncertainty changes (driven
by climate extremes, as well as development trends such as urbanization)
in relation to the capacity for risk management (which integrates climate
change adaptation, disaster risk management, and development). As
risk and uncertainty increase, incremental adjustments in practices may
no longer be sufficient to achieve resilience, and at some point the
growing resilience gap will provoke a search for transformative solutions.
Through enhanced experimentation and learning approaches, climate
change adaptation, disaster risk management, and development may
provide a pathway for keeping pace with the dynamic drivers and
expressions of risk.
After this introduction, this chapter discusses the relationship between
disaster risk management, climate change adaptation, and sustainable
and resilient development (Section 8.2), highlighting the synergies and
conflicts between these objectives and the common obstacles to reaching
them (Section 8.2.1) and the specific role of ecosystems and biodiversity
(Section 8.2.2). In particular, it emphasizes the importance of values and
perceptions (Section 8.2.3) and the role of technologies (Section 8.2.4) in
designing sustainability policies. Finally, it highlights the importance of
tradeoffs between temporal scales, spatial scales, and multiple stressors
(Section 8.2.5).
Focusing on time perspectives, Section 8.3 then discusses options to
integrate short- and long-term objectives, by looking at the long-term
consequences of present-day responses to disasters (Section 8.3.1),
investigating the barriers to integrating short- and long-term responses
(Section 8.3.2), and proposing options to overcome these barriers and
promote resilience (Section 8.3.3).
Section 8.4 assesses the implications of disaster risk reduction and
climate change adaptation for equity and access to resources, and in
particular the importance of capacities and resource availability to
implement policies for adaptation and disaster risk reduction (Section
8.4.1). It also highlights the existence of losers and winners from
disasters and disaster risk reduction and adaptation policies (Section
8.4.2), and the consequences of these distributive effects for human
security (Section 8.4.3) and for the possibility to achieve international
goals such as the Millennium Development Goals (Section 8.4.4).
Section 8.5 focuses on the specific issue of combining disaster risk
management and adaptation with climate change mitigation policies. It
starts by stressing the role of thresholds and tipping points as limits
to what can be achieved in terms of disaster risk management and
adaptation, and thus the importance of considering the three policies
together (Section 8.5.1). It then discusses synergies and conflicts
between mitigation, adaptation, and disaster risk management in urban
and rural areas (Section 8.5.2).
Section 8.6 identifies the tools and options to promote resilience to
climate extremes and combine adaptation, disaster risk management,
and other policy goals. It first discusses various approaches to planning
for the future, including the use of scenarios (Section 8.6.1). It then
highlights the existence of a continuum of options to make progress
over the short and long term, from incremental to transformational
Chapter 8Toward a Sustainable and Resilient Future
Risk
Management-
Development
Synergy and
Capacity
Risk and Uncertainty
Innovation
Adaptive
Management
Learning Leadership
Point at which a lack of resilience becomes
unacceptable, provoking a shift from incremental
to transformative risk management approaches
Unacceptable Risk
Management Burden
Unacceptable Risk
Adjusting
Pathway
Resilience
Trajectory
Point at which the costs
of risk management are
perceived as too high,
such that measures are
no longer taken to keep
up with changing risk
and uncertainty
Incremental
Adjustments
Transformative
Changes
en
en
en
enc
enc
nc
nc
nc
nc
e
e
e
e
e
e
e
e
e
a
a
a
a
jjecjecjecjec
jecjec
jec
jec
t
t
t
t
t
to
to
to
to
Figure 8-1 | Incremental and transformative pathways to resilience.
443
changes (Section 8.6.2). These increasingly ambitious changes include the
use of analysis and modeling tools to improve disaster risk management
and adaptation (Section 8.6.2.1), the implementation of new institutional
tools (Section 8.6.2.2), and transformational strategies to reach multiple
objectives (Section 8.6.2.3). Such transformational changes can be
facilitated using a combination of approaches (Section 8.6.3), including
adaptive management (Section 8.6.3.1), learning (Section 8.6.3.2),
innovation (Section 8.6.3.3), and leadership (Section 8.6.3.4). The chapter
concludes (Section 8.7) by discussing synergies between disaster risk
reduction and climate change adaptation to achieve a resilient and
sustainable future.
8.2. Disaster Risk Management as Adaptation:
Relationship to Sustainable Development
Planning
Earlier chapters discussed the concepts of and relationship between
disaster risk management (including disaster risk reduction) and climate
change adaptation. The two concepts and practices overlap considerably
and are strongly complementary. Disaster risk management considers
hazards other than those that are climate-derived, such as earthquakes
and volcanoes, while climate change adaptation considers and addresses
vulnerabilities related to phenomena that would not normally be
classified as discrete disasters, such as gradual changes in precipitation,
temperature, or sea level. Examples of hazards that are addressed by
both communities include flooding, droughts, and heat waves.
Disaster risk management is increasingly considered as one of the
‘frontlines’ of adaptation, and perhaps one of the most promising
arenas for mainstreaming or integrating climate change adaptation into
sustainable development planning (Sperling and Szekely, 2005; G. O’Brien
et al., 2006; Schipper and Pelling, 2006; Schipper, 2009). However, it
requires modifying development policies, mechanisms, and tools, and
identifying and responding to those who gain and lose from living with
and creating risk. Contested notions of development and hence differing
perspectives on sustainable development planning lead to different
conclusions about how disaster risk reduction can contribute to adaptation.
This section reviews the definitions of some of the key concepts used in
this chapter, and considers the roles that ecosystems services, values
and perceptions, technologies, and tradeoffs in decisionmaking can play
in influencing sustainable development planning and outcomes. It also
considers the tradeoffs that are involved in decisionmaking.
8.2.1. Concepts of Adaptation, Disaster Risk Reduction,
and Sustainable Development and how they are
Related
Adaptation can be defined as the process of adjustment to actual or
expected climate and its effects in order to moderate harm or exploit
beneficial opportunities (see Section 1.1.2). Adaptation actions may be
undertaken by public or private actors, and can be anticipatory or
reactive, and incremental or transformative (Adger et al., 2007; Stafford
Smith et al., 2011). In both principle and practice, adaptation is more
than a set of discrete measures designed to address climate change; it
is an ongoing process that encompasses responses to many factors,
including evolving experiences with both vulnerabilities and vulnerability
reduction planning and actions, as well as risk perception (Tschakert
and Dietrich, 2010; Weber, 2010; Wolf, 2011).
Adaptive capacity underlies action and is defined in this report as the
combination of strengths, attributes, and resources available to an
individual, community, society, or organization that can be used to
prepare for and undertake adaptation. Adaptive capacity can also be
described as the capability for innovation and anticipation (Armitage,
2005), the ability to learn from mistakes (Adger, 2003), and the capacity
to generate experience in dealing with change (Berkes et al., 2003).
Enhancing adaptive capacity under climate change entails paying
attention to learning about past, present, and future climate threats,
accumulated memory of adaptive strategies, and anticipatory action to
prepare for surprises and discontinuities in the climate system (Nelson
et al., 2007).
Adaptive capacity is uneven across and within sectors, regions, and
countries (K. O’Brien et al., 2006). Although wealthy countries and
regions have more resources to direct to adaptation, the availability of
financial resources is only one factor determining adaptive capacity (Moss
et al., 2010; Ford and Ford, 2011). Other factors include the ability to
recognize the importance of the problem in the context of multiple
stresses, to identify vulnerable sectors and communities, to translate
scientific knowledge into action, and to implement projects and programs
(Moser and Ekstrom, 2010). The capacity to adapt is in fact dynamic
and influenced by economic and natural resources, social networks,
entitlements, institutions and governance, human resources, and
technology (Parry et al., 2007). It is particularly important to understand
that places with greater wealth are not necessarily less vulnerable to
climate impacts and that a socioeconomic system might be as vulnerable
as its weakest link (K. O’Brien et al., 2006; Tol and Yohe, 2007). Therefore,
even wealthy locations can be severely impacted by extreme events,
socially as well as economically, as Europeans experienced during the
2003 heat wave (Salagnac, 2007; see also Case Study 9.2.1).
Current adaptation planning in many countries, regions, and localities
involves identification of a wide range of options, although the available
knowledge of their costs, benefits, wider consequences, potentials, and
limitations is still incomplete (NRC, 2010; see Section 4.5). In many cases,
the most attractive adaptation actions are those that offer development
benefits in the relatively near term, as well as reductions of vulnerabilities
in the longer term (Agrawala, 2005; Klein et al., 2007; McGray et al., 2007;
Hallegatte, 2008a; NRC, 2010). This is a lesson already noted, though not
always practiced, in disaster preparedness and risk reduction (IFRC,
2002; Pelling, 2010b). An emerging literature discusses adaptation
through the lens of sustainability, recognizing that not all adaptation
responses are necessarily benign; there are tradeoffs, potentials for
negative outcomes, competing interests, different types of knowledge,
Chapter 8 Toward a Sustainable and Resilient Future
444
and winners and losers inherent in adaptation responses (Eriksen and
O’Brien, 2007; Ulsrud et al., 2008; Barnett and O’Neill, 2010; Beckman,
2011; Brown, 2011; Eriksen et al., 2011; Gachathi and Eriksen, 2011;
Owuor et al., 2011). Sustainable adaptation is defined as a process
that addresses the underlying causes of vulnerability and poverty,
including ecological fragility; it is considered a way of generating social
transformation, or changes in the fundamental attributes of society that
contribute to vulnerability (Eriksen and O’Brien, 2007; Eriksen and Brown,
2011).
Disaster risk can be defined in many ways (see Section 1.1.2). In general,
however, it is closely associated with the concepts of hazards, exposure,
and vulnerability. Hazards are defined in this report as the potential
occurrence of a natural or human-induced physical event that may
cause negative consequences. Exposure is defined as the presence of
people, livelihoods, environmental services and resources, infrastructure,
and economic, social, and cultural assets in places that could be
adversely affected by climate extremes. Hazards and exposure are
changing, not only as the result of climate change, but also due to
human activities. For example, hazards associated with floods, landslides,
storm surges, and fires can be influenced by declines in ecosystem
services that regulate runoff, erosion, etc. The drainage of wetlands,
deforestation, the destruction of mangroves, and the changes associated
with urban development (such as the impermeability of surfaces and
overexploitation of groundwater) are all factors that can modify
hazard patterns (Nobre et al., 1991, 2005; MEA, 2005; Nicholls et
al., 2008). Consequently, most weather-related hazards now have an
anthropogenic element (Cardona, 1999; Lavell, 1999).
Vulnerability has many different (and often conflicting) definitions and
interpretations, both across and within the disaster risk and climate
communities (see Sections 1.1.2 and 2.2). Vulnerability can increase or
decrease over time as a result of both environmental and socioeconomic
changes (Blaikie et al., 1994; Leichenko and O’Brien, 2008). In general,
improvements in a country’s development indicators have been associated
with reduced mortality risk, yet an increase in economic loss and insurance
claims (UNDP, 2004; Pielke Jr. et al., 2008; Schumacher and Strobl, 2008;
ECA, 2009; UNISDR, 2009; World Bank, 2010a). Indeed, recent evidence
confirms that, despite increasing exposure, mortality risk from tropical
cyclones and floods is now decreasing globally, as well as in heavily
exposed regions like Asia (UNISDR, 2011). In contrast, the risk of
economic loss is increasing globally because reductions in vulnerability
are not compensating for rapid increases in the exposure of economic
assets. In the Organisation for Economic Co-operation and Development
(OECD) countries, for example, economic losses are increasing at a faster
rate than GDP per capita. In other words, the risk of losing wealth in
disasters is increasing faster than that wealth is being created (UNISDR,
2011). However, some types of development may increase vulnerability
or transfer it between social groups, particularly if development is
unequal or degrades ecosystem services (Guojie, 2003). Even where
growth is more equitable, vulnerabilities can be generated (e.g., when
modern buildings are not constructed to prescribed safety standards)
(Hewitt, 1997; Satterthwaite, 2007).
Climate change can magnify many preexisting risks through changes in
the frequency, severity, and spatial distribution of weather-related
hazards, as well as through increases in vulnerability due to climate
impacts (e.g., decreased water availability, decreased agricultural
production and food availability, or increased heat stress) (see Section
4.3). Like adaptation, disaster risk reduction may be anticipatory
(ensuring that new development does not increase risk) or corrective
(reducing existing risk levels) (Lavell, 2009). Given expected population
increases in hazard-prone areas, anticipatory disaster risk reduction is
fundamental to addressing the risk associated with future climate
extremes. At the same time, investments in corrective disaster risk
reduction are required to address the accumulation of exposure and
susceptibility to existing climate risks, for example, those inherited from
past urban planning or rural infrastructure decisions.
Climate change adaptation and disaster risk management (especially
disaster risk reduction) are critical elements of long-term sustainability
for economies, societies, and environments at all scales (Wilbanks and
Kates, 2010). The generally accepted and most widespread definition of
sustainable development comes from the Brundtland Commission
Report, which defined sustainable development as “development that
meets the needs of the present without compromising the ability of
future generations to meet their own needs” (WCED, 1987). A number
of principles of sustainable development have emerged, including the
achievement of a standard of human well-being that meets human
needs and provides opportunities for social and economic development;
that sustains the life support systems of the planet; that broadens
participation in development processes and decisions; and that accelerates
the movement of knowledge into action in order to provide a wider
range of options for resolving issues (WCED, 1987; Meadowcroft, 1997;
NRC, 1999; Swart et al., 2003; MEA, 2005). Because sustainable
development means finding pathways that achieve socioeconomic and
environmental goals without sacrificing either, it is a concept that is
fundamentally political (Wilbanks, 1994).
Discussions of the relationships between sustainable development and
climate change have increased over the past decades (Cohen et al., 1998;
Yohe et al., 2007; Bizikova et al., 2010). The literature on development
has considered how development paths relate to vulnerabilities both to
climate change and to climate change policies (e.g., Davis, 2001; Garg
et al., 2009), as well as to other hazards. Clearly, some climate change-
related environmental shifts are potentially threatening to sustainable
development, but they can also help move toward sustainability,
especially if the trends or events are severe enough to require significant
adjustment of unsustainable development practices or development
paths (e.g., the relocation of population or economic activities to less
vulnerable areas). In such cases, both disaster risk reduction and climate
change adaptation can be important – even essential – contributors to
sustainable development.
There are some examples of successful decreases in vulnerability
through disaster risk management, but less evidence in relation to climate
change adaptation, in part because the ability to attribute observed
Chapter 8Toward a Sustainable and Resilient Future
445
environmental stresses from and responses to climate change is still
limited (Fankhauser et al., 1999; Adger et al., 2007; Repetto, 2008). In
terms of disaster risk reduction, a large number of lives have been saved
over the last decade due to improved disaster early warning systems
(IFRC, 2005), and to increased development and human welfare
(UNISDR, 2011). There remains, however, much more that can be done
to reduce mortality and counteract growth in the number of people
affected by disasters and climate extremes. For example, recent self-
assessments of progress by over 100 countries on the objectives of the
Hyogo Framework of Action (UNISDR, 2009, 2011) indicate that few
developing countries have conducted comprehensive, accurate, and
accessible risk assessments, which are a prerequisite for both
anticipatory and corrective disaster risk reduction. Furthermore, the
assessment shows that few countries are able to quantify their investment
in disaster risk reduction. There are numerous ways to evaluate success of
disaster risk management or climate adaptation, including gauging the
extent to which the goals of a given action (determined in anticipation
of a given environmental stressor) are achieved, independent of whether
the environmental stressor materializes. Both climate adaptation and
disaster risk management can contribute to responses to changes in
extreme events due to climate change, yet neither approach alone is
sufficient.
Econometric analyses at the national scale have reached different
conclusions about the impact of disasters on economic growth, but the
balance of evidence suggests a negative impact. Whereas Noy and
Nualsri (2007), Noy (2009), Hochrainer (2009), Jaramillo (2009), and
Raddatz (2007) suggest that the overall impact on growth is negative,
Albala-Bertrand (1993) and Skidmore and Toya (2002) argue that natural
disasters have a positive influence on long-term economic growth, often
due to both the stimulus effect of reconstruction and the productivity
effect. As suggested by Cavallo and Noy (2009) and Loayza et al. (2009),
this difference may arise from the different impacts of small and large
disasters, the latter having a negative impact on growth and the former
enhancing growth. In any case, whether or not disaster losses translate
into other social and economic impacts depends on how each individual
disaster is managed (Moreno and Cardona, 2011) which in turn is related
to capacities and political priorities. At the local scale, Strobl (2011)
investigates the impact of hurricane landfall on county-level economic
growth in the United States. This analysis shows that a county that is
struck by at least one hurricane in a year sees its economic growth
reduced on average by 0.79%, and increased by only 0.22% the following
year. Noy and Vu (2010) investigate the impact of disasters on economic
growth at the province level in Vietnam, and find that lethal disasters
decrease economic production while costly disasters increase short-
term growth. Rodriguez-Oreggia et al. (2009) focus instead on poverty
and the World Bank’s Human Development Index at the municipality
level in Mexico. They show that municipalities affected by disasters
experienced an increase in poverty by 1.5 to 3.6%. Considering these
important links between disasters and development, there is a need
to consider disaster risk reduction, climate change adaptation, and
sustainable development in a consistent and integrated framework (G.
O’Brien et al., 2006; Schipper and Pelling, 2006).
8.2.2. Sustainability of Ecosystem Services
in the Context of Disaster Risk Management
and Climate Change Adaptation
Reducing human pressures on ecosystems and managing natural
resources more sustainably can facilitate efforts to mitigate climate
change and to reduce vulnerabilities to extreme climate and weather
events. The degradation of ecosystems is undermining their capacity to
provide ecosystem goods and services upon which human livelihoods and
societies depend (MEA, 2005; WWF, 2010), and to withstand disturbances,
including climate change. There is evidence that the likelihood of collapse
and subsequent regime shifts in ecological and coupled social-ecological
systems may be increasing in response to the magnitude, frequency,
and duration of climate change and other disturbance events (Folke et al.,
2004; MEA, 2005; Woodward, 2010). Large, persistent shifts in ecosystem
services not only affect the total level of welfare that people in a
community can enjoy, they also impact the welfare distribution between
people within and between generations and hence may give rise to new
conflicts over resource use and questions on inter-generational equity as
a component of sustainable development (Thomas and Twyman, 2005).
They could result in domino effects of increased pressure on successive
resource systems, as has been suggested in the case of depletion of
successive fish stocks (Berkes et al., 2006). However, the thresholds at
which ecosystems undergo regime shifts and the points at which these
may catalyze social stress remain largely unknown, partly due to
variability over space and time (Biggs et al., 2009; Scheffer, 2009).
Ecosystems can act as natural barriers against climate-related hazardous
extremes, reducing disaster risk (Conde, 2001; Scholze et al., 2005). For
example, mangrove forests are a highly effective natural flood control
mechanism that will become increasingly important with sea level rise,
and are already used as a coastal defense against extreme climatic and
non-climatic events (Adger et al., 2005). The benefits of such ecosystem
services are determined by ecosystem health, hazard characteristics,
local geomorphology, and the geography and location of the system
with respect to the hazard (Lacambra and Zahedi, 2011). In assessing
the ecological limits of adaptation to climate change, Peterson (2009)
emphasizes that ecosystem regime shifts can occur as the result of
extreme climate shocks, but that such shifts depend upon the resilience
of the ecosystem, and are influenced by processes operating at multiple
scales. In particular, there is evidence that the loss of regulating services
(e.g., flood regulation, regulation of soil erosion) erodes ecological
resilience (MEA, 2005).
Ecosystems and ecosystem approaches can also facilitate adaptation to
changing climatic conditions (Conde, 2001; Scholze et al., 2005).
Conservation of water resources and wetlands that provide hydrological
sustainability can further aid adaptation by reducing the pressures and
impacts on human water supply, while forest conservation for carbon
sinks and alternative sources of energy such as biofuels can reduce
carbon emissions and have multiple benefits (Reid, 2006), as can coastal
defenses and avalanche protection (Silvestri and Kershaw, 2010). In
New York, for example, untreated storm water and sewage regularly
Chapter 8 Toward a Sustainable and Resilient Future
446
flood the streets because the aging sewage system is no longer
adequate. After heavy rains, overflowing water flows directly into rivers
and streams instead of reaching water treatment plants. The US
Environmental Protection Agency has estimated that around US$ 300
billion over 20 years would be needed to upgrade sewage infrastructure
across the country (UNISDR, 2011). In response, New York City will
invest US$ 5.3 billion in green infrastructure on roofs, streets, and
sidewalks. This promises multiple benefits: the new green spaces may
absorb more rainwater and reduce the burden on the city’s sewage
system, improve air quality, andreduce water and energy costs. Such
changes in the constituents of an ecosystem can be used as levers to
enhance the resilience of coupled social-ecological systems (Biggs et al.,
2009).
Biodiversity is also important to adaptation. Functionally diverse systems
have more scope to adapt to climate change and climate variability than
functionally impoverished systems (Elmqvist et al., 2003; Hughes et al.,
2003; Lacambra and Zahedi, 2011). A larger gene pool will facilitate the
emergence of genotypes that are better adapted to changed climatic
conditions. Conservation of biodiversity and maintenance of ecosystem
integrity may therefore be a key objective in improving the adaptive
capacity of society to cope with climate change extremes (Peterson et
al., 1997; Elmqvist et al., 2003; SCBD, 2010).
Strategies that are adopted to reduce climate change through greenhouse
gas mitigation can affect biodiversity both negatively and positively
(Edenhofer et al., 2011), which in turn influences the capacity to adapt
to climate extremes. For example, some bioenergy plantations replace
sites with high biodiversity, introduce alien species, and use damaging
agrochemicals, which in turn may reduce ecosystem resilience and
hence their capacity to respond to extreme events (Foley et al., 2005;
Fargione et al., 2009). Large hydropower schemes can cause loss of
terrestrial and aquatic biodiversity, inhibit fish migration, and lead to
mercury contamination (Montgomery et al., 2000), as well as change
watershed sediment dynamics, leading to sediment starvation in coastal
areas, which in turn could lead to coastal erosion and make coasts more
vulnerable to sea level rise and storm surges (Silvestri and Kershaw,
2010).
The increasing international attention and support for efforts focused
on Reducing Emissions from Deforestation and Forest Degradation,
maintaining/enhancing carbon stocks, and promoting sustainable forest
management (REDD+) is an example of where incentives for the
protection and sustainable management of natural resources driven by
mitigation concerns also have the potential of generating co-benefits
for adaptation. By mediating runoff and reducing flood risk, protecting
soil from water and wind erosion, providing climate regulation, and
providing migration corridors for species, ecosystem services supplied
by forests can increase the resilience to some climatic changes (Locatelli
et al., 2010). Primary forests tend to be more resilient to disturbance
and environmental changes, such as climate change, than secondary
forests and plantations (Thompson et al., 2009). However, forests are also
vulnerable to climatic extremes (Nepstad et al., 2007) and the modeled
effects of global warming (Vergara and Scholz, 2011). Hence, the role of
forest ecosystems in climate change mitigation and adaptation will
itself depend on the rate and magnitude of climate change and whether
the crossing of ecological tipping points can be avoided.
8.2.3. The Role of Values and Perceptions
in Shaping Response
Values and perceptions are important in influencing action on climate
change extremes, and they can have significant implications for
sustainable development. The disaster risk community has used several
points of view for resolving decisions about where to invest limited
resources, including considerations of economic rationality and moral
obligation (Sen, 2000). Value judgments are embedded in problem
framing, solutions, development decisions, and evaluation of outcomes,
thus it is important to make them explicit and visible. Values describe
what is desirable or preferable, and they can be used to represent the
subjective, intangible dimensions of the material and nonmaterial world
(O’Brien and Wolf, 2010). They are closely linked to worldviews and
beliefs, including perceptions of change and causality (Rohan, 2000;
Leiserowitz, 2006; Weber, 2010). Values both inform and are shaped by
action, judgment, choice, attitude, evaluation, argument, exhortation,
rationalization, and attribution of causality (Rokeach, 1979). However,
values do not always clearly translate to particular behaviors
(Leiserowitz et al., 2005). Recognizing and reconciling conflicting values
increases the need for inclusiveness in decisionmaking and for finding
ways to communicate across social and professional boundaries
(Rosenberg, 2007; Vogel et al., 2007; Oswald Spring and Brauch, 2011).
Losses from extreme events can have implications beyond objective,
measurable impacts such as loss of lives, damage to infrastructure, or
economic costs. They can lead to a loss of what matters to individuals,
communities, and groups, including the loss of elements of social capital,
such as sense of place or of community, identity, or culture. This has long
been observed within the disaster risk community (Hewitt, 1997; Mustafa,
2005) and in more recent work in the climate change community (O’Brien,
2009; Adger et al., 2010; Pelling, 2010a). A values-based approach
recognizes that socioeconomic systems are continually evolving, driven
by innovations, aspirations, and changing values and preferences of the
constituents (Simmie and Martin, 2010; Hedlund-de Witt, 2011). This
approach raises not only the ethical question of ‘whose values count?’,
but also the important political question of ‘who decides?’. These
questions have been asked in relation to both disaster risk (Blaikie et al.,
1994; Wisner, 2003; Wisner et al., 2004) and climate change (Adger,
2004; Hunt and Taylor, 2009; Adger et al., 2010; O’Brien and Wolf, 2010),
and are significant when considering the interaction of climate change
and disaster risk, including the complexity of the temporal consequences
of policies and decisions (Pelling, 2003).
The probabilistic risk assessments that form the basis for current
models of cost-benefit analysis (CBA) rarely take into account the wider
consequences that account for a substantial proportion of disaster
Chapter 8Toward a Sustainable and Resilient Future
447
damage for poorer households and communities (UNISDR, 2004, 2009;
Marulanda et al., 2010). These include outcomes such as increased
poverty and inequality (Hallegatte, 2006; de la Fuente et al., 2009),
health effects (Murray and Lopez, 1996; Grubb et al., 1999; Viscusi and
Aldy, 2003), cultural assets and historical building losses (ICOMOS,
1998), and environmental impacts, which are often very difficult to
measure in monetary terms. Specific approaches allow accounting for
distributive effects in CBA (e.g., distributional-weight CBA, see
Harberger, 1978; basic-needs CBA, see Harberger, 1984; or social welfare
function built as a sum of individual welfare function that increases
nonlinearly with income), but none of them are consensual. Other types of
valuation emphasize institutional elements such as the ‘moral economy’
associated with the collective memory and identities of people living in
non-western cultures in many parts of the world (Rist, 2000; Hughes,
2001; Trawick, 2001; Scott, 2003).
Two important philosophical value frameworks have dominated
attempts to establish priorities for risk management: human rights and
utilitarian approaches. Human rights-based approaches (Wisner, 2003;
Gardiner, 2010) emphasize moral obligation to reduce avoidable risk
and contain loss, which was recognized in the UN Universal Declaration
of Human Rights in 1948: Article 3 provides for the right to “life, liberty
and security of person, while Article 25 protects “a standard of living
adequate for the health and well-being … in the event of unemployment,
sickness, disability, widowhood, or old age or other lack of livelihood in
circumstances beyond his [sic] control.
The humanitarian community, and civil society more broadly, has made
considerable progress in addressing these aspirations (Kent, 2001),
perhaps best exemplified by the Sphere standards. These are a set of
self-imposed guidelines for good humanitarian practices that require
impartiality in post-disaster actions including shelter management and
access to and distribution of relief and reconstruction aid (Sphere,
2004). The ethics and equity dimensions of risk management have also
been explored in adaptation through the application of Rawls’ theory of
justice (Rawls, 1971; Paavola, 2005; Paavola and Adger 2006; Paavola
et al., 2006; Grasso, 2009, 2010). From this perspective, priority is given
to reducing risk for the most vulnerable, even if this limits the absolute
numbers who benefit.
In contrast to focusing on the most excluded or economically poor,
utilitarian approaches assume that interpersonal welfare comparisons
are possible, and that a social welfare function that summarizes the
welfare of a population can be built (Pigou, 1920). Assuming its existence,
maximizing this social welfare function reveals where economic benefits
of public investments exceed costs. The calculated economic benefits of
investing in risk reduction vary, but are often considered significant (see
Ghesquiere et al., 2006; World Bank 2010a; UNISDR, 2011). There are,
however, extreme difficulties in accounting for the complexity of disaster
costs and risk reduction investment benefits (Pelling et al., 2002;
Hallegatte and Przyluski, 2010). A key point here is that value frameworks
can significantly influence the types of responses to climate and weather
extremes.
8.2.4. Technology Choices, Availability, and Access
Technology choices can contribute to both risk reduction and risk
enhancement, relative to extreme climate and weather events. As
discussed in Section 7.4.3, technologies receive prominent attention in
both climate change adaptation and disaster risk reduction. Continuing
transitions from one socio-technological state to another frame many
aspects of responses to climate change risks. Assessments of roles of
technology choices, availability, and access in responding to climate
extremes are enmeshed in a wide range of technologies that must be
considered within a broad range of development contexts. However, in
nearly every case, issues are raised about the balance between risk
reduction and risk creation. Technology is a broad concept that
embraces a range of areas, including information and communication
technologies, roads and infrastructure, food and production technologies,
energy systems, and so on. Technology choices can alleviate disaster
risk, but they can also significantly increase risks and add to adaptation
challenges (Jonkman et al., 2010). For example, some modern energy
systems and centralized communication systems are dependent on
physical structures that can be vulnerable to storm damage. It has been
suggested that relatively centralized high-technology systems are ‘brittle,
offering efficiencies under normal conditions but subject to cascading
effects in the event of emergencies (Lovins and Lovins, 1982).
In many cases, technologies are considered to be an important part of
responses to climate extremes and disaster risk. This includes, for example,
attention to physical infrastructure, including how to ‘harden’ built
infrastructure such as bridges or buildings, or natural systems such as
hillsides or river channels, such that they are able to withstand higher
levels of stress (UNFCCC, 2006; Larsen et al., 2007; CCSP, 2008). Another
focus is on technologies that assist with information collection and
diffusion, including technologies to monitor possible stresses and
vulnerabilities, technologies to communicate with populations and
responders in the event of emergencies, and technology applications to
disseminate information about possible threats and contingencies –
although access to such technologies may be limited in some developing
regions. Seasonal climate forecasts based on the results from numerical
climate models have been developed in recent decades to provide
multi-month forecasts, which can be used to prepare for floods and
droughts (Stern and Easterling, 1999). Modern technological development
is exploring a wide variety of innovative concepts that may eventually
hold promise for disaster risk reduction, for example, through new food
production technologies, although ecological, ethical, and human health
implications are often as yet unresolved (Altieri and Rosset, 1999).
Attention to technology alternatives and their benefits, costs, potentials,
and limitations in cases where disaster risk is created and when risk
reduction takes place involve two different time horizons. In the near
term, technologies to be considered are those that currently exist or that
can be modified relatively quickly. In the longer term, it is possible to
consider potentials for new technology development, given identified
needs (Wilbanks, 2010). In some circumstances, technologies put in
place to reduce short-term risk and vulnerability can increase future
Chapter 8 Toward a Sustainable and Resilient Future
448
vulnerability to extreme events or ongoing trends. For example, the
use of irrigation has reduced farmer vulnerabilities to low and variable
precipitation patterns. However, when the irrigation water is from a
nonrenewable source (e.g., the Ogallala-High Plains aquifer system of the
United States), the foreseeable reduction in future irrigation opportunities
would mean an increase in vulnerability and the risk of increasing crop
failures (AAG, 2003; Harrington, 2005).
Similarly, while large dams could mitigate drought and generate
electricity, well known costs of social and ecological displacement may
be unacceptable (Baghel and Nusser, 2010). Furthermore, unless dams
are constructed to accommodate future climate change, they may present
new risks to society by encouraging a sense of security that ignores
departures from historical experience (Wilbanks and Kates, 2010). In the
Mekong region, dikes, dams, drains, and diversions established for flood
protection have unexpected consequences for risk over the longer term,
because they influence risk-taking behavior (Lebel et al., 2009). In the
United States, past building in floodplain areas downstream from dams
that have now exceeded their design life has become a major concern;
tens of thousands of dams are now considered as having high hazard
potential (McCool, 2005; FEMA, 2009; ASCE, 2010).
Investments in physical infrastructure cast long shadows through time,
because they tend to assume lifetimes of three to four decades or
longer. The gradual modernization of a city’s housing stock, transport, or
water and sanitation infrastructure takes many decades without targeted
planning. If they are maladaptive rather than adaptive, the consequences
can be serious. This suggests a reappraisal of technology that might
promote more distributed solutions, for example, multiple, smaller dams
that can resolve local as well as more distant needs, or widely spread,
local energy production (perhaps utilizing micro-solar, wind and water,
or geothermal power) that can reduce exposure to secondary impacts
from natural disasters when large power generators or power transmission
lines are lost during a natural disaster, or when power plants generate
secondary disasters after being impacted by a natural hazard, as has
happened recently in Japan. The goal of a more distributed and less
maximizing development vision has been expressed in Thailand’s
‘Sufficiency Economy’ approach, where local development is judged
against its contribution to local, national, and international wealth
generation (UNDP, 2007a).
Technology choices, availability, and access depend on more than
technology development alone. Unless the technologies, the skills
required to use them, and the institutional approaches appropriate to
deploy them are effectively transferred from providers to users
(‘technology transfer’), the effects of technology options, however
promising, are minimized (see Section 7.4.3). Challenges in putting
science and technology to use for sustainable development have
received considerable attention (e.g., Nelson and Winter, 1982; Patel
and Pavit, 1995; NRC, 1999; ICSU, 2002; Kristjanson et al., 2009),
emphasizing the wide range of contexts that shape both barriers and
potentials. If obstacles related to intellectual property rights can be
overcome, however, the growing power of the information technology
revolution could accelerate technology transfer (linked with local
knowledge) in ways that would be very promising (Wilbanks and
Wilbanks, 2010).
8.2.5. Tradeoffs in Decisionmaking:
Addressing Multiple Scales and Stressors
Sustainable development involves finding pathways that achieve a
variety of socioeconomic and environmental goals, without sacrificing
any one for the sake of the others. As a result, the relationships between
adaptation, disaster risk management, and sustainability are highly
political. Successful reconciliation of multiple goals “lies in answers to
such questions as who is in control, who sets agendas, who allocates
resources, who mediates disputes, and who sets rules of the game”
Chapter 8Toward a Sustainable and Resilient Future
FAQ 8.1 | Why is there not a greater emphasis on technology as the solution to climate extremes?
Technology is an essential part of responses to climate extremes, at least partly because technology choices and uses are so often a part
of the problem. Enhancing early warning systems is one example where technology can play an important role in disaster risk management.
This example also flags the importance of considering ‘hard’ (engineering) and ‘soft’ (social and administrative) technology. Great
advances have been made in hard technology around hazard identification, and this has saved many lives. Communicating warnings
through the ‘soft’ technology of institutional reform and communication networks has been less well developed. Both hard and soft
technology systems must be responsive to different cultures, environments, and types of governance. Most fundamentally, it is clear that
technologies are the product of research and development choices, which reflect particular values, interests, and priorities. The successful
transfer of technology is sensitive to local needs, capacities, and development goals. Technologies can have unintended consequences
that contribute to maladaptations. For example, some modern agricultural technologies may reduce local biodiversity and constrain
future adaptation. Technologies only matter if they are both appropriate and accessible. Technology development and use are necessary
for reducing vulnerabilities to climate extremes, both through mitigation and adaptation, but they need to be the right technologies that
are deployed in the right ways. This calls for greater reflection on the social, economic, and environmental consequences of technology
across both space and time. In many cases, responses to climate extremes can be improved by addressing social vulnerability, rather
than focusing exclusively on technological responses.
449
(Wilbanks, 1994, p. 544). This means that conflicts of interest must be
acknowledged and addressed, whether they are between government
departments, sectors, or policy arenas, and suggests that simple
panaceas are unlikely without tradeoffs in decisionmaking (Brock and
Carpenter, 2007).
There is no single or optimal way of adapting to climate change or
managing risks, because contexts for risk management vary so widely.
For example, risk management decisions can be oriented toward
incremental responses to frequent events that are disruptive but
perhaps not ‘extreme.’ Often, tradeoffs between multiple objectives are
ambiguous. For example, focusing on and taking actions to protect
against frequent events may lead to greater vulnerability to larger and
rarer extreme events (Burby, 2006). This is a particular challenge for
investing in fixed physical infrastructure. Social investments and risk
awareness, including early warning systems, can be strengthened by
more frequent low-impact events that maintain risk visibility and allow
preparedness for larger, less frequent events (see Case Studies 9.2.11
and 9.2.14). Pielke Jr. et al. (2007) also warn that locating adaptation
policy in a narrow risk framework by concentrating only on identifiable
anthropogenic risks can distort public policy because vulnerabilities are
created through multiple stresses.
As one salient example, during disaster reconstruction, tensions frequently
arise between demands for speed of delivery and sustainability of outcome.
Response and reconstruction funds tend to be time-limited, often requiring
expenditure within 12 months or less from the time of disbursement.
This pressure is compounded by multiple agencies working with often
limited coordination. Time pressure and competition between agencies
tends to promote centralized decisionmaking and the subcontracting of
purchasing and project management to non-local commercial actors.
Both outcomes save time but miss opportunities to include local people
in decisionmaking and learning from the event, with the resulting
reconstruction in danger of failing to support local cultural and economic
priorities (Berke et al., 1993; Pearce, 2003). At the same time it is important
not to romanticize local actors or their viewpoints, which might at times
be unsustainable or point to maladaptation, or to accept local voices
as representative of all local actors. When successful, participatory
reconstruction planning has been shown to build local capacity and
leadership, bind communities, and provide mechanisms for information
exchange with scientific and external actors (Lyons et al., 2010). As part
of any participatory or community-based reconstruction, the importance
of a clear conflict resolution strategy has been recognized.
Tradeoffs may also arise through conflicts between economic development
and risk management (Kahl, 2003, 2006). The current trend of development
in risk-prone areas (e.g., coastal areas in Asia) is driven by socioeconomic
benefits yielded by these locations, with many benefits accruing to
private investors or governments through tax revenue. For example,
export-driven economic growth in Asia favors production close to large
ports to reduce transportation time and costs. Consequently, the
increase in risk has to be balanced against the socioeconomic gains of
development in at-risk areas. Additional construction in at-risk areas is
not unacceptable a priori, but has to be justified by other benefits, and
sometimes complemented by other risk-reducing actions (e.g., early
warning and evacuation, improved building norms, specific flood
protection). This introduces the possibility for those benefiting financially
to offset produced risk through risk reduction mechanisms ranging from
fair wages and disaster-resistant housing (to enhance worker resilience)
to support for early warning, preparedness, and reconstruction. Such
approaches have been considered in some businesses through corporate
social responsibility agendas (Twigg, 2001).
One climate change/development tradeoff linked both to timeframes
and the magnitude of climate extremes is the future need for risk
reduction infrastructure that would require changes in ecologically or
historically important areas. For example, when considering additional
protection (e.g., dikes and seawalls) in historical centers, aesthetic and
cultural elements as well as building costs will be taken into account.
Existing planning and design standards to protect cultural heritage or
ecological integrity may need to be balanced with the needs of adaptation
(Hallegatte et al., 2011a). Difficulties in attributing value to cultural and
ecological assets mean that CBAs are not the best tool to approach
these types of problems. Multi-criteria decisionmaking tools (Birkmann,
2006) that incorporate a participatory element and can recognize the
political, ethical, and philosophical aspects of such decisions can also be
useful (Mercer et al., 2008). But the magnitude of emerging climate
extremes is an important issue. If climate change is relatively severe,
rather than moderate, then the focus on preserving iconic areas is likely
to increase, as will the costs.
Another contextual complication that introduces tradeoffs is the fact
that impacts of climate change extremes extend across multiple scales.
The challenge is to find ways to combine the strengths of addressing
multiple scales, rather than having them work against each other
(Wilbanks, 2007, 2009). Local scales offer potentials for bottom-up
actions that ensure participation, flexibility, and innovation. At the same
time, efforts to develop initiatives from the bottom up are often limited
by a lack of information, limited resources, and limited awareness of
larger-scale driving forces (AAG, 2003). Larger scales offer potentials for
top-down actions that assure resource mobilization and cost sharing.
Integrating these kinds of assets across scales is often essential for
resilience to extremes, but in fact, integration is profoundly impeded by
differences in who decides, who pays, and who benefits, and perceptions
of scalar effects that often reflect striking ignorance and misunderstanding
(Wilbanks, 2007). In recent years, there have been a number of calls for
innovative co-management structures that cross scales in order to
promote sustainable development (e.g., Bressers and Rosenbaum, 2003;
Cash et al., 2006; Campbell et al., 2010).
What might be done to realize potentials for integrating actions at
different scales to make them more complementary and reinforcing?
Many top-down interventions (from international donor development
and disaster response and reconstruction funding to new adaptation
fund mechanisms and national programming) may unintentionally
discourage local action by imposing bureaucratic conditions for access
Chapter 8 Toward a Sustainable and Resilient Future
450
to financial and other resources (Christoplos et al., 2009). Top-down
sustainability initiatives are often preoccupied with input metrics, such as
criteria for partner selection and justifications (often based on relatively
detailed quantitative analyses of such attributes as ‘additionality’),
rather than on outcome metrics, such as whether the results make a
demonstrable contribution to sustainability (regarding metrics, see
NRC, 2005).
To manage tradeoffs and conflicts in an open, efficient, and transparent
way, institutional and legal arrangements are extremely important. The
existing literature on legislation for adaptation at the state level is not
comprehensive, but those countries studied lack many of the institutional
mechanisms and legal frameworks that are important for coordination
at the state level (Richardson et al., 2009). This has been found to be the
case for Vietnam, Laos, and China (Lin, 2009). In the South Pacific, high
exposure to climate change risk has yet to translate into legislative
frameworks to support adaptation – with only Fiji, Papua New Guinea,
and Western Samoa formulating national climate change regulatory
frameworks (Kwa, 2009). Without a supporting and implemented
national legislative structure, achieving local disaster reduction and
climate change adaptation planning can be complicated (La Trobe and
Davis, 2005; Pelling and Holloway, 2006; see also Section 6.4). Still,
where local leadership is determined, skillful planning is possible, even
without legislation. This has been the experience of Ethekwini
Municipality (the local government responsible for the city of Durban,
South Africa), which has developed a Municipal Climate Protection
Programme with a strong and early focus on adaptation without national-
level policy or legal frameworks to guide adaptation planning at the
local level (Roberts, 2008, 2010).
One way around the challenges of tradeoffs is to ‘bundle’ multiple
objectives through broader participation in strategy development and
action planning, both to identify multiple objectives and to encourage
attention to mutual co-benefits. In this sense, both the pathway and
outcomes of development planning have scope to shape future social
capacity and disaster risk management. Policies and actions to achieve
multiple objectives include stakeholder participation, participatory
governance (IRGC, 2009), capacity building, and adaptive organizations,
including both private and public institutions where there is a considerable
knowledge base reflecting both research and practice to use as a starting
point (e.g., NRC, 2008). Multi-hazard risk management approaches
provide opportunities to reduce complex and compound hazards, both
in rural and urban contexts.
8.3. Integration of Short- and Long-Term
Responses to Extremes
When considering the linkages between disaster management, climate
change adaptation, and development, time scales play an important role.
Disaster management increasingly emphasizes vulnerability reduction in
addition to the more traditional emergency response and relief measures.
This requires addressing underlying exposure and sensitivity in the
context of hazards with different frequencies and return periods. As
discussed in Chapter 2, there is now a converging focus on vulnerability
reduction in the context of disaster risk management and adaptation to
climate change (Sperling and Szekely, 2005).
Cross-scale (spatial and temporal) interactions between responses
focusing on the short term and those required for long-term adjustment
can potentially create both synergies and contradictions among disaster
risk reduction, climate change adaptation, and development. This section
assesses the literature regarding synergies and tradeoffs between short-
and long-term adjustments. First, we consider the implications of present-
day responses for future well-being. The barriers to reconciling short-
and long-term goals are then assessed. Insights from research on the
resilience of social-ecological systems are then considered as a potential
means of addressing integration in a long-term perspective.
8.3.1. Implications of Present-Day Responses
for Future Well-Being
The implications of present-day responses to both disaster risk and
climate change can be either positive or negative for human security
and well-being in the long term. Positive implications can include
increased resilience, capacity building, broad social benefits from
extensive participation in risk management and resilience planning, and
the value of multi-hazard planning (see Sections 5.4 and 6.5). Negative
implications can include threats to sustainability if the well-being of
future generations is not considered; issues related to the economic
discounting of future benefits; ‘silo effects’ of optimizing responses for
one system or sector without considering interaction effects with others
(see Burby et al., 2001); equity issues regarding who benefits and who
pays; and the ‘levee effect,’ where the adaptive solution to a current risk
management problem builds confidence that the problem has been
solved, blinding populations to the possibility that conditions may
change and make the present adaptation inadequate (Burby, 2006;
Burby et al., 2006).
The terms ‘coping’ and ‘adaptation’ reflect strategies for adjustments to
changing climatic and environmental conditions. In the case of a set of
policy choices, both coping and adaptation denote forms of conduct
that aim and indeed may achieve modifications in the ways in which
society relates to nature, and nature to society (Stehr and von Storch,
2005). As discussed in Section 2.4, coping actions are those that take
place when trying to alleviate the impacts or to live with the costs of a
specific event. They are usually found during the unfolding of disaster
impacts, which can continue for some time after an event – for example,
if somebody loses their job or is traumatized. Coping strategies can help
to alleviate the immediate impact of a hazard, but may also increase
vulnerabilities over the medium to longer term (Swift, 1989; Davies,
1993; Sperling et al., 2008). The different time frames for coping and
adaptation can present challenges for risk management. Focusing on
short-term responses and coping strategies can limit the scope for
adaptation in the long term. For example, drought can force agriculturalists
Chapter 8Toward a Sustainable and Resilient Future
451
to remove their children from school or delay medical treatment, which
may have immediate survival benefits, yet in aggregate undermines the
human resources available for long-term adaptation (Norris, 2005;
Alderman et al., 2006; Santos, 2007; Sperling et al., 2008).
In both developed and developing countries, a focus on coping with the
present is often fueled by the perception that climate change is a long-
term issue and that other challenges, including economic growth, food
security, water supply (Bradley et al., 2006), sanitation, education, and
health care, require more immediate attention (Klein et al., 2005; Adly
and Ahmed, 2009; Kameri-Mbote and Kindiki, 2009). Particularly in poor
rural contexts, short-term coping may be a tradeoff that increases
longer-term risks (UNISDR, 2009; Brauch and Oswald Spring, 2011).
Adaptation, on the other hand, is often focused on minimizing potential
risk to future losses (Oliver-Smith, 2007). This ‘long-term’ framing of
adaptation can constrain both short-term coping and adaptive capacity,
for example, when relocation of settlements to avoid coastal hazards
undermines social capital and local livelihoods, limiting household coping
and adaptive capacity (Hunter, 2005). There is a large literature and
much experience related to slum relocation that is of direct relevance to
urban coping and adaptation (Gilbert and Ward, 1984; Davidson et al.,
1993; Viratkapan and Perera, 2006). Context is important in discussing
tradeoffs between addressing short- and long-term risks, and even in well-
governed systems, political expediency will often distort the regulatory
process in a way that favors the short term (Platt, 1999).
Disasters can destroy assets and wipe out savings, and can push
households into ‘poverty traps,’ that is, situations where productivity is
reduced, making it impossible for households to rebuild their savings
and assets (Zimmerman and Carter, 2003; Carter et al., 2007; Dercon
and Outes, 2009; López, 2009; van den Berg, 2010). The process by which
a series of events generates a vicious spiral of impacts, vulnerability,
and risk was first recognized by Chambers (2006), who described it as
the ratchet effect of disaster, risk, and vulnerability. These micro-level
poverty traps can also be created by health and social impacts of natural
disasters: it has been shown that disasters can have long-lasting
consequences for psychological health (Norris, 2005), and for child
development from reduction in schooling and diminished cognitive
abilities (see Alderman et al., 2006; Santos, 2007; Bartlett, 2008).
Where disaster loss is widespread, micro-level poverty traps can aggregate
to the regional level. Here, poor regions impacted by disaster are unable
to fully recover so that capacity is reduced and vulnerability heightened,
making future disasters more likely. Without enough time to rebuild
between events, such regions may end up in a state of permanent
reconstruction, with resources devoted to repairing and replacing rather
than accumulating infrastructure and equipment. This obstacle to capital
accumulation and infrastructure development can lead to a permanent
disaster-related underdevelopment (Hallegatte et al., 2007; Hallegatte
and Dumas, 2008). This can be amplified by other long-term mechanisms,
such as changes in risk perception that reduce investments in the affected
regions or reduced services that make qualified workers leave the
region. These effects have been discussed by Benson and Clay (2004),
and investigated by Noy (2009) and Hochrainer (2009), who found that
natural disasters have a negative impact on economic growth and
development, especially when direct losses are large. This negative
impact is found to be larger when the disaster affects a smaller country,
with lower GDP per capita, weaker institutions, lower openness to
trade, lower literacy rates, and lower levels of government spending,
and when foreign aid and remittances are lower. Such effects have been
modeled by Hallegatte et al. (2007) and Hallegatte and Dumas (2008)
using a reduced-form economic model that shows that the average GDP
impact of natural disasters can be either close to zero if reconstruction
capacity is large enough, or very large if reconstruction capacity is too
limited, which may be the case in less-developed countries. There are,
however, many uncertainties in the ways in which people’s spontaneous
and organized responses to increasing climate-related hazards feed
back to influence long-term adaptive capacity and options. Migration,
which can be traumatic for those involved, might lead to enhanced life
chances for the children of migrants, building long-term capacities and
potentially also contributing to the movement of populations away
from places exposed to risk (IOM, 2007, 2009a,b; Ahmed, 2009; Oswald
Spring, 2009b; UNDP, 2009).
A broad literature on experiences of community-based and local-level
disaster risk reduction indicates options for transiting from short- to
longer-term responses, at least in the context of frequently occurring
risk manifestations (Lavell, 2009; UNISDR, 2009; Maskrey, 2011). Such
approaches, many of which are based on community participation, have
progressively moved from addressing disaster preparedness and capacities
for emergency management toward addressing the vulnerability of
livelihoods, the decline of ecosystems, the lack of social protection,
unsafe housing, the improvement of governance, and other underlying
risk factors (Bohle, 2009). While managing existing risk will contain loss,
addressing underlying risk drivers will contribute to a reduction in
future risk to climate extremes.
8.3.2. Barriers to Reconciling Short- and Long-Term Goals
Although there is robust evidence in the literature to support disaster
risk reduction as a strategy for long-term climate change adaptation,
there are numerous barriers to reconciling short- and long-term goals.
Many poor countries are very vulnerable to natural hazards but cannot
implement the measures that could reduce this vulnerability for financial
reasons, or due to a lack of governance capacity or technology. The recent
national self-assessments of progress toward achieving the UNISDR
Hyogo Framework for Action indicated that some least-developed
countries lack the human, institutional, technical, and financial capacities
to address even emergency management concerns (UNISDR, 2009). A
recently developed index that measures capacities and conditions for
risk reduction shows that low- and lower-middle-income countries with
weak governance have, with some exceptions, great difficulty addressing
underlying drivers of vulnerability. Those at the bottom of the index,
such as Haiti, Chad, or Afghanistan, are also experiencing conflict or
political instability (UNISDR, 2011). Another obstacle to reconciling
Chapter 8 Toward a Sustainable and Resilient Future
452
short- and long-term goals is access to technology and maintenance of
infrastructure. An example is the introduction of water reuse technologies,
which have been developed in a few countries, which could bring a
great improvement in the management of droughts if they could be
disseminated in many developing countries (Metcalf & Eddy, 2005).
Money and technology are not enough to implement efficient disaster
risk reduction and adaptation strategies. Indeed, differences in
resources cannot explain the differences in exposure and vulnerability
among regions (Nicholls et al., 2008). Governance capacities and the
inadequacy of and lack of synergy between institutional and legislative
arrangements for disaster risk reduction, climate change adaptation,
and poverty reduction are also as much a part of the problem as the
shortage of resources. Institutional and legal environments and political
will are important, as illustrated by the difference in risk management
in various regions of the world (Pelling and Holloway, 2006). In many
countries disaster risk management and adaptation to climate change
measures are overseen by different institutional structures (see Section
1.1.3). This is explained by the historical evolution of both approaches.
Disaster risk management originated from humanitarian assistance
efforts, evolving from localized, specific response measures to preventive
measures, which seek to address the broader environmental and
socioeconomic aspects of vulnerability that are responsible for turning
a hazard into a disaster in terms of human and/or economic losses.
Within countries, disaster risk management efforts are often coordinated
by civil defense agencies, while measures to adapt to climate change
are usually developed by environment ministries. Responding to climate
change was originally more of a top-down process, where advances in
scientific research led to international policy discussions and frameworks.
While the different institutional structures may represent an initial
coordination challenge, the converging focus on vulnerability reduction
represents an opportunity for managing disaster and climate risks more
comprehensively within the development context (Sperling and Szekely,
2005). A change in the culture of public administration toward creative
partnerships between national and local governments and empowered
communities has been found in some cases to dramatically reduce costs
(Dodman et al., 2008).
In addition to the barriers described above, there is also a tendency for
individuals and groups to focus on the short-run and to ignore low-
probability, high-impact events. The following studies discuss some of
the psychological and economic barriers shaping how people make
decisions under uncertainty:
Underestimation of the risk: Even when individuals are aware of
the risks, they often underestimate the likelihood of the event
occurring (Smith and McCarty, 2006). This bias can be amplified by
natural variability (Pielke Jr. et al., 2008), where there is expert
disagreement, and where there is uncertainty. Magat et al. (1987),
Camerer and Kunreuther (1989), and Hogarth and Kunreuther
(1995), for example, provide considerable empirical evidence that
individuals do not seek information on probabilities in making
their decisions.
Budget constraints: If there is a high upfront cost associated with
investing in adaptation measures, individuals will often focus on
short-run financial goals rather than on the potential long-term
benefits in the form of reduced risks (Kunreuther et al., 1978;
Thaler, 1999).
Difficulties in making tradeoffs: Individuals are also not skilled in
making tradeoffs between costs and benefits of these measures,
which requires comparing the upfront costs of the measure with
the expected discounted benefits in the form of loss reduction over
time (Slovic, 1987).
Procrastination: Individuals are observed to defer choosing
between ambiguous choices (Tversky and Shafir, 1992; Trope and
Liberman, 2003).
Samaritan’s dilemma: Anticipated availability of post-disaster
support can undermine self-reliance when there are no incentives
for risk reduction (Burby et al., 1991).
Politician’s dilemma: Time delays between public investment in
risk reduction and benefits when hazards are infrequent, and the
political invisibility of successful risk reduction can be pressures for
a ‘not in my term of office’ attitude that leads to inaction (Michel-
Kerjan, 2008).
Work in West and East Africa has shown that rural communities tend to
underestimate external forces that influence their region while
overestimating their own response capacity (Enfors et al., 2008;
Tschakert et al., 2010). Misjudging external drivers may be explained by
the low degree of control people feel they have over these drivers,
resulting in reactions that range from powerlessness to denial. Another
issue that makes it difficult to reconcile short- and long-term goals
arises from the challenges in projecting the long-term climate and
corresponding risks (see Section 3.2.3). Examples of this challenge are
reflected in the demographic growth of Florida in the 1970s and 1980s,
which unfolded during a period of low hurricane activity but may
expose larger populations to the risks associated with extreme climate
and weather events. Major engineering projects with long lead times
from planning to implementation have difficulty factoring in climate
change futures, and have instead been planned according to historic
hazard risk (Pielke Jr. et al., 2008). Managing natural risks and adapting
to climate change requires anticipating how natural hazards will
change over the next decades, but uncertainty about climate change
and natural variability is a significant obstacle to such anticipation
(Reeder et al., 2009).
Climate change is typically viewed as a slow-onset, multigenerational
problem. Consequently, individuals, governments, and businesses have
been slow to invest in adaptation measures. Research in South Asia shows
that in those regions where past development had prioritized short-term
gains over long-term resilience, agricultural productivity is in decline
because of drought and groundwater depletion, rural indebtedness is
increasing, and households are sliding into poverty with particularly
insidious consequences for women, who face the brunt of nutritional
deprivation as a result (Moench et al., 2003; Moench and Dixit, 2007).
Connecting short- and long-term perspectives is thus seen as critical to
Chapter 8Toward a Sustainable and Resilient Future
453
realizing the synergies between disaster risk management and climate
change adaptation.
8.3.3. Connecting Short- and Long-Term Actions
to Promote Resilience
The previous section has highlighted the importance of linking short-
and long-term responses so that disaster risk reduction and climate
change adaptation mutually support each other. A systems approach
that emphasizes cross-scale interactions can provide important insights on
how to realize synergies between disaster risk management and climate
change adaptation. Resilience, a concept fundamentally concerned with
how a system, community, or individual can deal with disturbance and
surprise, increasingly frames contemporary thinking about sustainable
futures in the context of climate change and disasters (Folke, 2006;
Walker and Salt, 2006; Brand and Jax, 2007; Bahadur et al., 2010). It has
developed as a fusion of ideas from several bodies of literature: ecosystem
stability (e.g., Holling, 1973; Gunderson, 2009), engineering robust
infrastructures (e.g., Tierney and Bruneau, 2007), the behavioral sciences
(Norris, 2010), psychology (e.g., Lee et al., 2009), disaster risk reduction
(e.g., Cutter et al., 2008), vulnerabilities to hazards (Moser, 2009), and
urban and regional development (e.g., Simmie and Martin, 2010). In the
context of this report, resilience refers to a system’s capacity to anticipate
and reduce, cope with, and respond to and recover from external
disruptions (see Sections 1.1.2.1 and 1.3.2). Resilience perspectives can
be used as an approach for understanding the dynamics of human-
environmental systems and how they respond to a range of different
perturbations (Carpenter et al., 2001; Walker et al., 2004).
‘Resilience thinking’ (Walker and Salt, 2006) may provide a useful
framework to understand the interactions between climate change and
other challenges, and in reconciling and evaluating tradeoffs between
short- and longer-term goals in devising response strategies.
Approaches that focus on resilience emphasize the need to manage for
change, to see change as an intrinsic part of any system, social or
otherwise, and to ‘expect the unexpected.’ Resilience thinking goes
beyond the conventional engineering systems’ emphasis on capacity to
control and absorb external shocks in systems assumed to be stable (Folke,
2006). For social-ecological systems (examined as a set of interactions
between people and the ecosystems they depend on), resilience
involves three properties: the amount of change a system can undergo
and retain the same structure and functions; the degree to which it can
reorganize; and the degree to which it can build capacity to learn and
adapt (Folke, 2006). Resilience can also be considered a dynamic
process linked to human agency, as expressed in the ability to deal with
hazards or disturbance, to engage with uncertainty and future changes,
to adapt, cope, learn, and innovate, and to develop leadership capacity
(Bohle et al., 2009; Obrist et al., 2010).
Resilience approaches offer four key contributions for living with
extremes: first, in providing a holistic framework to evaluate hazards in
coupled social-ecological systems; second, in putting emphasis on the
capacities to deal with hazard or disturbance; third, in helping to explore
options for dealing with uncertainty and future changes; and fourth, in
identifying enabling factors to create proactive responses (Berkes, 2007;
Obrist et al., 2010).The concept of resilience is already being applied as
a guiding principle to disaster risk reduction and adaptation issues, as
well as to examine specific responses to climate change in different
developed and developing country contexts (e.g., Cutter et al., 2008).
Eakin and Webbe (2008) use a resilience framework to show that the
interplay between individual and collective adaptation can be related to
wider system sustainability. Goldstein (2009) uses resilience concepts to
strengthen communicative planning approaches to dealing with surprise.
Linnenluecke and Griffiths (2010) use a resilience framework to explore
organizational adaptation to climate change and weather extremes, and
suggest that organizations may need to develop multiple capabilities
and response approaches in response to changing extremes. Nelson et
al. (2007) have shown how resilience thinking can enhance analyses of
adaptation to climate change: as adaptive actions affect not only the
intended beneficiaries but have repercussions for other regions and
times, adaptation is part of a path-dependent trajectory of change.
Resilience also considers a distinction between incremental adjustments
and system transformation, which may broaden the expanse of adaptation
and also provide space for agency (Nelson et al., 2007). Resilience
approaches can be seen as complementary to agent-based analyses of
climate change responses that emphasize processes of negotiation and
decisionmaking, as they can provide insights into the systems-wide
implications. Adger et al. (2011) show that dealing with specific risks
without taking into account the nature of system resilience can lead to
responses that potentially undermine long-term resilience.
Recent work on resilience and governance has focused on communication
of science between actors and depth of inclusiveness in decisionmaking
as key determinants of the character of resilience. In support of these
approaches it is argued that inclusive governance facilitates better
flexibility and provides additional benefit from the decentralization of
power. On the down side, greater participation can lead to loose
institutional arrangements that may be captured and distorted by
existing vested interests (Adger et al., 2005; Plummer and Armitage,
2007). Still, the balance of argument (and existing centrality of
institutional arrangements) calls for a greater emphasis to be placed on
the inclusion of local and lay voices and of diverse stakeholders in shaping
agendas for resilience through adaptation and adaptive management
(Nelson et al., 2007). Striking the right balance between top-down
command-and-control approaches, which offer stability over the short
term but reduced long-term resilience, and more flexible, adaptive forms
of risk management is the core practical challenge that disaster risk
management brings to climate change adaptation under conditions of
climatic extremes and projected increases in disaster risk and impacts
(Sperling and Szekely, 2005).
Resilience thinking is not without its critiques (Nelson, 2009; Pelling,
2010a). Shortcomings include the downplaying of human agency in
systems approaches and difficulty in including analysis of power in
explanations of change, which combine to effectively promote stability
Chapter 8 Toward a Sustainable and Resilient Future
454
rather than flexibility, that is, maintaining the status quo and thus serving
particular interests rather than supporting adaptive management, social
learning, or inclusive decisionmaking. One challenge to enhancing
resilience of desired system states is to identify how responses to any
single stressor influence the larger, interconnected social-ecological
system, including the system’s ability to absorb perturbations or shocks,
its ability to adapt to current and future changes, and its ability to learn
and create new types or directions of change. Responses to one stressor
alone may inadvertently undermine the capacity to address other
stressors, both in the present and future. For example, coastal towns in
eastern England, experiencing worsening coastal erosion exacerbated
by sea level rise, are taking their own action against immediate erosion
in order to protect livelihoods and homes, affecting sediments and erosion
rates down the coast (Milligan et al., 2009). While such actions to protect
the coast are effective in the short term, in the long term, investing to
‘hold the line’ may diminish capital resources for other adaptations and
hence reduce adaptive capacity to future sea level rise. Thus, dealing
with specific risks without a full accounting of the nature of system
resilience can lead to responses that can potentially undermine long-
term resilience. Despite an increasing emphasis on managing for
resilience (Walker et al., 2002; Lebel et al., 2006), the resilience lens
alone may not sufficiently illuminate how to enhance agency and move
from the understanding of complex dynamics to transformational
action.
8.4. Implications for Access to Resources,
Equity, and Sustainable Development
The previous section assessed the links between short- and long-term
responses to climate extremes. This section takes the idea of links further.
It explores the relationships between climate change adaptation, disaster
risk management, and mitigation, and larger issues related to equity,
access to resources, environmental and ecosystem protection, and related
development processes. This draws out the importance of governance in
determining the relationship between disaster risk and underlying
processes of unequal socioeconomic development and environmental
injustices (Maskrey, 1994; Sacoby et al., 2010). The section discusses
issues related to capacity and equity, the existence of winners and losers
from disaster and disaster management policy, and opportunities for
contributing to wider development goals including the enhancing of
human security.
8.4.1. Capacities and Resources:
Availability and Limitations
The capacity to manage risks and adapt to change is unevenly distributed
within and across nations, regions, communities, and households
(Hewitt, 1983; Wisner et al., 2004; Beck, 2007). The literature on how
these capacities contribute to disaster risk management and climate
change adaptation emphasizes the role of economic, financial, social,
cultural, human, and natural capital, and of institutional context (see
Sections 1.4 and 2.4). When the poor are impacted by disasters, limited
resources are quickly expended in coping actions that can further
undermine household sustainability in the long run, reducing capital
and increasing hazard exposure or vulnerability. In these vicious cycles
of decline, households tend first to expend savings and then, if pressures
continue, to withdraw members from non-productive activities such as
school, and finally to sell productive assets. As households begin to
collapse, individuals may be forced to migrate or in some cases enter
into culturally inappropriate, dangerous, or illegal livelihoods such as
the sex industry (Mgbako and Smith, 2010; Ferris, 2011). This poverty
and vulnerability trap means that recovery to pre-disaster levels of well-
being becomes increasingly difficult (Burton et al., 1993; Adger, 1996,
Wisner et al., 2004; Chambers, 2006).
Children, the elderly, and women stand out as more vulnerable to
extreme climate and weather events. The vulnerability of children and
their capacity to respond to climate change and disasters is discussed in
Box 8-2 (see also Section 5.5.1 and Case Study 9.2.14). Among the
elderly, increasing numbers will become exposed to climate change
impacts in the coming decades, particularly in OECD countries where
populations are aging most rapidly. By 2050, it is estimated that one in
three people will be older than 60 years in OECD countries, as well as one
in five at the global scale (UN, 2002). The elderly are made additionally
vulnerable to climate change-related hazards by characteristics that also
increase vulnerability to other social and environmental hazards (thus
compounding overall vulnerability): deterioration of health, personal
lifestyles, social isolation, poverty, and inadequate access to health and
social infrastructures (OECD, 2006). Gender impacts vulnerability in
many ways. In the 1991 cyclone in Bangladesh, the death toll among
women was reportedly five times higher than among men (UNDP,
2007b). Cultural as well as physiological factors are widely cited for the
over-representation of female deaths from flooding. Gender inequality
extends into female-headed households to compound the vulnerability
of dependent children or elderly (Cannon, 2002; UNISDR, 2008; Oxfam,
2010). Inequality has many other important faces: race, caste, religious
affiliation, and physical disability, all of which help determine individual
and household vulnerability, and they cross-cut gender and age effects.
Importantly, the social construction of vulnerability through these
characteristics highlights the ways in which vulnerability changes over
time – in this case with changes in family structure and access to
services in response to economic cycles and political and cultural trends
evolving as the climate changes with potentially compounding effects
(Leichenko and O’Brien, 2008).
Studies also show that female-headed households more often borrow
food and cash than rich and male-headed households during difficult
times. This coping strategy is considered to be a dangerous one as the
households concerned will have to return the food or cash soon after
harvests, leaving them more vulnerable as they have less food or cash
to last them the season and to be prepared if disaster strikes (Young and
Jaspars, 1995). This may leave households in a cycle of poverty from one
season to the next. Literature shows that this outcome is linked to
unequal access by women to resources, land, and public and privately
Chapter 8Toward a Sustainable and Resilient Future
455
Chapter 8 Toward a Sustainable and Resilient Future
Box 8-2 | Children, Extremes, and Equity in a Changing Climate
The linkages between children and extreme events have been addressed through two principal lenses.
1. Differentiated Impacts and Vulnerability
The literature estimates that 66.5 million children are affected annually by disasters (Penrose and Takaki, 2006). Research on disaster
impacts among children focuses on short- and long-term physical and psychological health impacts (Norris et al., 2002; Bunyavanich et
al., 2003; del Ninno and Lindberg, 2005; Balaban, 2006; Waterson, 2006; Bartlett, 2008). Vulnerability to these impacts in part is due to
the less-developed physical and mental state of children, and therefore differential capacities to cope with deprivation and stress in
times of disaster (Cutter, 1995; Bartlett, 2008; Peek, 2008).
Most literature points toward higher mortality and morbidity rates among children due to climate stresses and extreme events (Cutter,
1995; Bunyavanich et al., 2003; Telford et al., 2006; Waterson, 2006; Bartlett, 2008; Costello et al., 2009). This is especially acute in
developing countries, where climate-sensitive health outcomes such as malnutrition, diarrhea, and malaria are already common and
coping capacities are lowest (Haines et al., 2006), although research in the United States found relatively low child mortality from
disasters and considerable differences across age groups for different types of hazard (Zahran et al., 2008).
Recent studies conducted in Bolivia, Indonesia, Mexico, Mozambique, Nepal, the Philippines, and Vietnam provide evidence of how
extensive (low impact/high frequency) disasters negatively affect children’s education, health, and access to services such as water and
sanitation, an issue of critical importance given the importance of primary education for human and long-term economic development.
In areas in Bolivia that experienced the greatest incidence of extensive disasters, the gender gap in primary education achievement
widened, preschool enrollment rates decreased, and dropout rates increased. Equivalent areas in Nepal and Vietnam saw, respectively,
reduced primary enrollment rates and a drop in the total number of children in primary education. Extensive disasters also led to an
increased incidence of diarrhea in children under five years of age in Bolivia, an increased proportion of malnourished children under
three in Nepal, an increased infant mortality rate in Vietnam, and an increase in the incidence of babies born with low birth weight in
Mozambique. This study also found evidence of negative impacts in terms of access to water and sanitation in Mexico and Vietnam
(UNISDR, 2011).
These studies underpin the need for resources for child protection during and after disaster events (Last, 1994; Jabry, 2003; Bartlett,
2008; Lauten and Lietz, 2008; Weissbecker et al., 2008). These include protection from abuse, especially during displacement, social safety
nets to guard against withdrawal from school due to domestic or livelihood duties, and dealing with psychological and physical health
issues (Norris et al., 2002; Keenan et al., 2004; Evans and Oehler-Stinnett, 2006; Waterson, 2006; Bartlett, 2008; Davies et al., 2008;
Lauten and Lietz, 2008; Peek, 2008).
2. Children’s Agency and Resource Access
Rather than just vulnerable victims requiring protection, children also have a critical role to play in tackling extreme events in the context
of climate change (Tanner, 2010). There is also increasing attention on child-centered approaches to preventing, preparing for, coping
with, and adapting to extreme events (Peek, 2008; Tanner, 2010).
While often centered on disaster preparedness and climate change programs in education and schools (Wisner, 2006; Bangay and Blum,
2010), more recent work emphasizes the latent capacity of children to participate directly in disaster risk reduction or adaptation supported
through child-centered programs (Back et al., 2009; Tanner et al., 2009). This emphasis acknowledges the unique risk perceptions and
risk communication processes of children, and their capacity to act as agents of change before, during, and after disaster events (see
collections of case studies in Peek, 2008; Back et al., 2009; and Tanner, 2010). Examples demonstrate the ability to reduce risk behavior
at household and community scales, but also to mobilize adults and external policy actors to change wider determinants of risk and
vulnerability (Mitchell et al., 2008; Tanner et al., 2009).
456
provided services (Agarwal, 1991; Thomas-Slayter et al., 1995;
Nemarundwe, 2003; Njuki et al., 2008). But women are also often the
majority holders of social capital and the mainstay of social movements
and local collective action, providing important mechanisms for household
and local risk reduction and potentially transformative resilience.
Important here are local saving groups and microcredit/micro-finance
groups, some of which extend to micro-insurance. In a review of micro-
finance for disaster risk reduction and response in South Asia,
Chakrabarti and Bhatt (2006) identify numerous initiatives, including
those that build on extensive social networks and connect to the formal
financial services sector.
Demographic and sociological diversity is also difficult to capture
in decisionmaking, where non-scientific knowledge is less easily
incorporated into formal decisionmaking. The importance of culture,
including traditional knowledge, in shaping strategies for adaptation is
recognized (Heyd and Brooks, 2009; ISET, 2010). There is a long tradition
of seeking to identify and bring such knowledge into planned disaster
risk management in urban and rural contexts through participatory and
community-based disaster risk management (Bruner et al., 2001;
Fearnside, 2001; Pelling, 2007; Mercer et al., 2008) and tools such as
participatory geographic information systems that explicitly seek to
bring scientific and local knowledge together (Tran et al., 2009). Both
development planning and post-disaster recovery have tended to
prioritize strategic economic sectors and infrastructure over local
livelihoods and poor communities (Maskrey, 1989, 1996). However, this
represents a missed opportunity for building local capacity and including
local visions for the future in planning the transition from reconstruction
to development – opportunities that could increase long-term
sustainability (Christoplos, 2006).
8.4.2. Local, National, and International
Winners and Losers
While climate-related disasters cannot always be prevented, the scale of
loss and its social and geographical distribution do differ significantly,
determined by the characteristics of those at risk and overarching
structures of governance including the legacy of preceding development
paths for social institutions, economies, and physical assets (Oliver-
Smith, 1994). But some people also benefit from disasters. These may be
organizations or individuals who benefit economically from reconstruction
or response (West and Lenze, 1994; Hallegatte, 2008b), through supplying
materials, equipment, and services – often at a premium price generated
by local scarcity and inflationary pressures (Benson and Clay, 2004) or
as a result of poorly managed tendering processes (Klein, 2007). Areas
not impacted by disaster can also experience economic benefits, for
example, in the Caribbean where hurricanes have caused international
tourist flows to be redirected (Pelling and Uitto, 2001). Political actors
can also benefit by demonstrating strong post-disaster leadership, at
times even when past political decisions have contributed to generating
disaster risk (Olson and Gawronski, 2003; Le Billon and Waizenegger,
2007; Gaillard et al., 2008). The same can be said for climate change,
with very unequal consequences in various regions of the world and
various economic sectors and social categories (Adger et al., 2003;
O’Brien et al., 2004; Tol et al., 2004). Less directly, those who have
benefited from policies and processes, such as expansion of commercial
agriculture or logging, can also be described as benefiting from decisions
that have generated vulnerability and prefigured disaster for others.
Such costs and benefits are often separated geographically and
temporally, making any efforts at distributional equity challenging. For
example, in the case of Hurricane Mitch, which killed more than 10,000
people and caused as much as US$ 8.5 billion in damages, deforestation
and rapid urban growth are often cited among the key causes of the
disaster-related losses (Alves, 2002; Pielke Jr. et al., 2003), with those
benefiting from such development including distant speculators.
Analyses of winners and losers associated with climate change and
discrete hazards need differentiation. In almost every circumstance,
what one part of society views as a win can be viewed by another part
as a loss. In examining possible responses to risks of climate extremes,
it is essential to recognize that possible impacts interact with vested
interests of different locations, sectors, and population groups in very
different ways. In virtually every case, the question is: benefits for
whom? Who says that this course of action is best for society as a
whole? What compensation is offered for those who are losers? In
particular, who is listening to views of those parts of society that have
less political power and influence?
While individual events can be assessed as a snapshot of winners and
losers, climate change as an ongoing process has no final state. Over
time, it may produce different distributions of winners and losers, for
example, as areas experience positive and then negative consequences
of changes in temperature or precipitation. Whether or not a particular
place produces winners or losers from an extreme event or a combination
of climate extremes and other driving forces also depends on perceptions.
These may be shaped by the recovery process, but are strongly influenced
by prioritized values (Quarantelli, 1984, 1995; O’Brien, 2009; O’Brien
and Wolf, 2010). In considering winners and losers from extreme climate
and weather events, and also from the outcomes of policies directed at
reducing disaster risk or responding to climate change, it is thus vital to
recognize the subjective understanding of winners and losers.
Much depends upon an individual, group, or society’s dominant values,
perspectives, and access to information. While some regard winners and
losers as a natural and inevitable outcome of ecological changes and/or
economic development, others suggest that winners and losers are
deliberately generated by unequal political and social conditions
(O’Brien and Leichenko, 2003; Wisner, 2003). Lurking behind discourses
about winners and losers are issues of liability and compensation for
losses: that is, if a population or an area experiences severe losses due to
an extreme event (at least partly) attributed to climate change, can fault
be prescribed? Does responsibility lie with those who have generated
local environmental change through settling a hazard-exposed area,
those who have promoted or permitted such settlement, or those who
have failed to mitigate local hazard or global environmental change?
Chapter 8Toward a Sustainable and Resilient Future
457
Issues of equity, justice, and compensation are emerging in climate
change adaptation, but few have begun to deal with questions of liability
for disaster risk production beyond the local scale (Kent, 2001; Mitchell,
2001; Wisner, 2001; O’Brien et al., 2010b). It seems that efforts to assign
responsibility will emerge as an issue for both governments and courts
at a range of scales (Farber, 2007).
8.4.3. Potential Implications for Human Security
Changes in extreme climate and weather events threaten human security,
and both disaster risk management and climate change adaptation
represent strategies that can improve human security while also avoiding
disasters. Human security addresses the combined but related challenges
of upholding human rights, meeting basic human needs, and reducing
social and environmental vulnerability (UNDP, 1994; Sen, 2003; Bogardi
and Brauch, 2005; Brauch 2005a,b, 2009; Fuentes and Brauch, 2009). It
also emphasizes equity, ethics, and reflexivity in decisionmaking and a
critical questioning and contestation of the drivers of climate change
(O’Brien et al., 2010b) and local impacts (Pelling, 2010a). Human security
is realized through the capacity of individuals and communities to respond
to threats to their environmental, social, and human rights (GECHS, 2000;
Barnett et al., 2010). A number of studies have assessed the relationship
between climate change and human security, demonstrating that the
linkages are often both complex and context-dependent (Barnett, 2003;
Barnett and Adger, 2007; Brauch et al., 2008, 2009, 2011; Buhaug et al.,
2008; O’Brien et al., 2010a). Among the most likely human security
threats are impacts felt through damage to health, food, water, or soil
conditions (Oswald Spring, 2009a, 2011b).
Among the most widely discussed humanitarian and human security
issues related to climate change are the possibilities of increased migration
and/or violent conflict resulting from the biophysical or ecological
disruptions associated with climate change (Reuveny, 2007; O’Brien et
al., 2008; Raleigh et al., 2008; Warner et al., 2010). There are indications
that migration conditions followed disasters in the distant past, as well
as in current situations (see, e.g., Le Roy Ladurie, 1971; Kinzig et al.,
2006; Peeples et al., 2006). Migration is a key coping mechanism for
poor rural households, not only in extreme circumstances (e.g., during a
prolonged drought, as with the 20th-century US Dustbowl period and
Sahelian droughts) but also as a means of diversifying and increasing
income (Harrington et al., 2009; Oswald Spring, 2011a; Scheffran, 2011).
The opportunities that population movement opens for risk reduction are
seen in international remittance flows from richer to poorer countries.
These are estimated to have exceeded US$ 318 billion in 2007, of which
developing countries received US$ 240 billion (World Bank, 2008).
Disasters linked to extreme events often lead to forced displacement of
people, as well as provoke voluntary migration among the less poor. The
relationship between climate risk and displacement is a complex one
and there are numerous factors that affect migration (UNDP, 2009).
Nonetheless, recent research suggests that adverse environmental
impacts associated with climate change have the potential to trigger
displacement of an increased number of people (Kolmannskog, 2008;
Feng et al., 2010). Studies suggest that most migration will take place
internally within individual countries; that in most cases when climatic
extremes occur in developing countries they will not lead to net out-
migration because people tend to return to re-establish their lives after
a disaster; and that while long-term environmental changes may cause
more permanent migration this will also tend to be internal (Piguet,
2008; UNDP, 2009). More negatively, forced land abandonment is
stressful for migrants whose culture and sense of identity are affected
(Mortreux and Barnett, 2009; Brauch and Oswald Spring, 2011). The
social dislocation provoked by migration can lead to a breakdown in
traditional rural institutions and associated coping mechanisms, for
example, in the erosion of traditional community-based water management
committees in central and west Asia (Birkenholtz, 2008). Local collective
coping and adaptive capacity can also be limited by increases in the
number of female-headed households as men migrate (Oswald Spring,
1991, 2009a).
Attention has been mainly focused on population displacement associated
with large disasters. Pakistan’s 2010 floods have to date left an estimated
6 million people in need of shelter; India’s 2008 floods also uprooted
roughly 6 million people; Hurricane Katrina displaced more than half a
million people in the United States; and Cyclone Nargis uprooted
800,000 people in Myanmar and South Asia. However, the compound
effect of smaller, more frequent events can also contribute to displacement.
Hazards such as floods, although often causing relatively low mortality,
destroy many houses and hence cause considerable displacement.
Between 1970 and 2009 in Colombia, for example, 24 disaster loss
reports detailed floods that killed fewer than 10 people but destroyed
more than 500 houses. In total, around 26,500 houses were destroyed,
potentially displacing more than 130,000 people. In the Indian state of
Orissa, 265 floods with similar low mortality rates destroyed more than
half a million houses. It is estimated that such extensive disasters
account for an additional 19% displacement of people who are typically
less visible than those displaced in larger events that attract international
media and humanitarian assistance (UNISDR, 2011).
Despite the opportunities to enhance development, disaster response is
often better at meeting basic needs than securing or extending human
rights. Indeed, the political neutrality that underpins the humanitarian
imperative makes any overt actions to promote human rights by
humanitarian actors difficult. In this way, disaster response and
reconstruction can to only a partial extent claim to enhance human
security (Pelling and Dill, 2009). Work at the boundaries between
humanitarian and development actors, new partnerships, the involvement
of government, and meaningful local participation are all emerging as ways
to resolve this challenge. One successful case has been the reconstruction
process in Aceh, Indonesia, following the Indian Ocean tsunami, where
collaboration between government and local political interests facilitated
by international humanitarian actions on the ground and through political-
level peace-building efforts have increased rights locally, contained
armed conflict, and provided an economic recovery plan (Le Billon and
Waizenegger, 2007; Gaillard et al., 2008; Törnquist et al., 2010).
Chapter 8 Toward a Sustainable and Resilient Future
458
Coping with the new and unprecedented threats to human security
posed by climate change has raised questions about whether existing
geopolitics and geo-strategies have become obsolete (Dalby, 2009). The
concepts, strategies, policies, and measures of the geopolitical and strategic
toolkits of the past as well as the short-term interests dominating
responses to climate change have been increasingly questioned, while
the potential for unprecedented disasters has led to a consideration of
the security implications of climate change (UNSC, 2007; EU, 2008;
UNGA, 2009; UNSG, 2009). Concerns range from increased needs for
humanitarian assistance to concerns over environmental migration,
emergent diseases for humans or in food chains, potentials for conflict
between nations or localities over resources, and potential for political/
governmental destabilization due to climate-related stresses in
combination with other stresses, along with efforts to assign blame
(Ahmed 2009; Brauch and Oswald Spring, 2011).
Climate change is generally regarded to act as a threat multiplier for
instability in some of the most volatile regions of the world (CNA
Corporation, 2007). Even in stable polities, adaptation planning that
seeks long-term resilience is confronted by political instability directly
after disasters (Drury and Olson, 1998; Olson, 2000; UNDP, 2004; Pelling
and Dill, 2009). When disasters strike across national boundaries or
within areas of conflict, they can provide a space for rapprochement,
but effects are usually short-lived unless the underlying political and
social conditions are addressed (Kelman and Koukis, 2000; Kelman,
2003). New interest in disaster and climate change as a security concern
has brought in lessons from international law (Ammer et al., 2010) and
security policy (Campbell et al., 2007) on planning for relatively low-
probability, high-consequence futures. Although during times of stress it
is easy for polities to drift toward militarization and authoritarianism for
managing disaster risk (Albala-Bertrand, 1993), there are alternatives,
such as inclusive governance, that can meet the goals of sustainable
development and human security over the long term (Olson and
Gawronski, 2003; Brauch, 2009; Pelling and Dill, 2009; Bauer, 2011).
8.4.4. Implications for Achieving
Relevant International Goals
Addressing or failing to address disaster risk reduction and climate
change adaptation can influence the success of international goals,
particularly those linked to development. Successive reports have noted
the potential for climate change to derail the Millennium Development
Goals (MDGs). In 2003, the Asian Development Bank and nine other
development organizations first highlighted that climate change may
impact on progress toward the MDGs and in particular constrain
progress beyond 2015, underlying the importance of managing climate
risks within and across development sectors. The UK Department for
International Development (DFID, 2004, 2006) and UNDP (2004) show
how each of the MDGs is dependent on some aspect of disaster risk for
success. Disaster impacts on the MDG targets are both direct and
indirect. For example, MDG 1 (to eradicate extreme poverty and hunger)
is impacted directly by damage to productive and reproductive assets of
the poor and less poor (who may remain in poverty or slip into
poverty as a result of disaster loss), and indirectly affected by negative
macroeconomic impacts. The 2007 UN Human Development Report noted
that enhanced adaptation is required to protect the poor, with climate
change potentially acting as a brake on development beyond 2015.
The UNISDR Hyogo Framework for Action 2005-2015: Building the
Resilience of Nations and Communities to Disasters (HFA) explicitly
recognizes that climate variability and change are important contributors
to disaster risk and includes strong support for better linking disaster
management and climate change adaptation efforts (Sperling and
Szekely, 2005; see also Section 7.3.1). The HFA priorities for action have
proven foresightful in including resilience explicitly as a component.
Priority Three calls for “knowledge, innovation and education to build a
culture of safety and resilience at all levels.” This provides a strong
justification for international actors to support investment in institutional
and human capacity for national and local resilience building, and one
that does not require the addition of new international agreements to
start work. Frameworks for such action exist, for example, in Common
Country Assessments, United Nations Development Assistance
Frameworks, National Adaptation Plans of Action, and Poverty Reduction
Strategies, but limited progress has been made on this to date (DFID,
2004). Some have proposed integrating climate change and disaster
risk management into any equivalent of the MDGs post-2015 (Tribe and
Lafon, 2009; Gostin et al., 2011).
8.5. Interactions among Disaster Risk
Management, Adaptation to Climate
Change Extremes, and Mitigation of
Greenhouse Gas Emissions
8.5.1. Thresholds and Tipping Points
as Limits to Resilience
Recent literature suggests that climate change could trigger large-scale,
system-level regime shifts that could significantly alter climatic and
socioeconomic conditions (MEA, 2005; Lenton et al., 2008; Hallegatte et
al., 2010; see Section 3.1.7). Examples of potential system changes include
dieback of the Amazon rainforest, decay of the Greenland ice sheet, and
changes in the Indian summer monsoon (Lenton et al., 2008). At smaller
scales, also, climate change is exacerbating well-established examples
of environmental regime shifts, such as freshwater eutrophication
(Carpenter, 2003), shifts to algae-dominated coral reefs (Hughes et al.,
2003), and woody encroachment of savannas (Midgley and Bond,
2001). The abruptness and persistence of such changes in social and
ecological systems, coupled with the fact that they can be difficult and
sometimes impossible to reverse, means that they can have substantial
impacts on human well-being (Scheffer et al., 2001; MEA, 2005; Scheffer,
2009). The notion of regimes shifting once thresholds or tipping points
are crossed contrasts with discussions of climate thresholds (see Section
3.1.1) and traditional thinking about gradual, linear, and more predictable
changes in ecological and social-ecological systems, and emphasizes
Chapter 8Toward a Sustainable and Resilient Future
459
the possibility of multiple futures determined by the crossing of critical
thresholds (Levin, 1998; Gunderson and Holling, 2002). Similar discussion
on socioeconomic systems has, for example, identified profitability limits
in economic activities as critical thresholds that can bring about sudden
collapse and regime change (Schlenker and Roberts, 2006; OECD, 2007).
The metaphor of tipping points, or points at which a system shifts from
one state to another, can also be applied to disaster events. Disasters
themselves are threshold-breaching events, where coping capacities of
communities are overwhelmed (e.g., Blaikie et al., 1994; Sperling et al.,
2008). Disasters may lead to secondary hazards, for example, when the
impacts from one disaster breach the coping capacities of related systems,
as when hurricane impacts trigger landslides or when different disasters
produce concatenated impacts over time (Biggs et al., 2011). For example,
losses associated with droughts and fires during the 1997/1998 El Niño-
Southern Oscillation event in Central America increased landslide and
flood hazard during Hurricane Mitch in 1998 (Villagrán de León, 2011).
Critical social thresholds may be crossed as disaster impacts spread
across society. Disaster response is as much about containing such losses
as assisting those hurt by the initial disaster impact.
For the poor, life and health are immediately at risk; for those living in
societies that take measures to protect infrastructure and economic and
physical assets, the lives and health of the population are less at risk.
However, a threshold can be crossed when hazards exceed anticipated
limits, are novel or unexpected in a specific risk management domain
(Beniston, 2004; Schär et al., 2004; Salagnac, 2007), or when vulnerability
has increased or resilience decreased due to spillover from market and
other shocks (Wisner, 2003).
In 2010, for example, western Russia experienced the hottest summer
since the beginning of systematic weather data recording 130 years
ago. Lack of rainfall in early 2010 and July temperatures almost 8°C
above the long-term average led to parched fields, forests, and peatlands
that posed a high wildfire risk. During and after the wildfires, Russia’s
mortality rate increased by 18%. In August alone, 41,300 more people
died as compared to August 2009, due to both the extreme heat and
smoke pollution. Social and economic change had greatly increased the
risk posed by wildfires. Traditional agricultural livelihoods have declined,
accompanied by out-migration and reduced management of surrounding
forests, arguably exacerbated by the decentralization of national
management and increased exploitation by the private sector (UNISDR,
2011).
The recognition of nonlinearity and the importance of thresholds as
limiting points for existing systems has led climate scientists to increase
their attention to the ‘tails’ of impact probability density functions
(Weitzman, 2009). This is in contrast to the disasters research community,
which, after focusing on major extremes, is now recognizing the
importance of small or local disasters and the secondary disasters that
make up concatenated events (UNDP, 2004; UNISDR, 2009). Both lenses
are valuable for a comprehensive understanding of the interaction of
disaster impact with development.
Tipping points in natural and human systems are more likely to arise
with relatively severe and/or rapid climate change than with moderate
levels and rates (Wilbanks et al., 2007). Because of this, less success with
climate change mitigation implies greater challenges for adaptation and
disaster risk management. Not only does adaptation need to consider
incremental change in hazard and vulnerability, but the possibility of
threshold breaching, systems-wide changes. The nonlinear changes
associated with breaching thresholds may exceed adaptation capacity
to avoid serious disruptions. Examples of ecological system changes of this
kind and social impacts include the disappearance of glaciers currently
feeding urban and agricultural water supply (Orlove, 2009), effects of
climate change on traditional livelihoods for the sustainability of
indigenous cultures (Turner and Clifton, 2009), widespread loss of corals
in acidifying oceans and fisher livelihoods (Reaser et al., 2000), and
profitability limits for important economic activities like agriculture,
fisheries, and tourism. When socioeconomic systems are already under
stress (e.g., fisheries in many countries), sustainability thresholds may be
more easily passed. Responses to potential thresholds or tipping points
include efforts to improve the information available to decisionmakers,
for example, through monitoring systems to provide early warning of an
impending system collapse (Biggs et al., 2009; Scheffer et al., 2009),
but also initiatives researching the balance of risks associated with
geoengineering (Royal Society, 2009) aiming to avoid such tipping
points (Virgoe, 2002; Kiehl, 2006).
8.5.2. Adaptation, Mitigation, and
Disaster Risk Management Interactions
As indicated above, the extent to which future adaptation and disaster
risk reduction action will be required is likely to be dependent on the
extent and rapidity with which climate change mitigation actions may
be taken and resulting risk unfolds for any given development context.
This section assesses the ways in which mitigation, disaster risk reduction,
and adaptation interact with development in urban and rural contexts.
In many instances, climate change mitigation and adaptation may be
synergistic, such as land use planning to reduce transport-related energy
consumption and limit exposure to floods, or building codes to reduce
heating energy consumption and enhance robustness to heat waves
(McEvoy et al., 2006). There is an emerging literature exploring the
linkages between climate change mitigation and adaptation, and the
possibility of approaches that address both objectives simultaneously
(Wilbanks and Sathaye, 2007; Hallegatte, 2009; Bizikova et al., 2010;
Wilbanks, 2010; Yohe and Leichenko, 2010). In this section we enlarge
the scope of the interactions to include disaster risk management. This
builds on experience within the disaster management community that
has recently sought to integrate risk modeling (UNDP, 2004; UNISDR,
2009, 2011) and planning to consider multi-hazard contexts. An important
lesson from this is that avoiding superficial integration means seeking
out and addressing shared root causes of exposure and vulnerability to
hazards, and not just addressing the surface expressions of risk (Wisner,
2011). The extent of adaptation required will depend on the climate
Chapter 8 Toward a Sustainable and Resilient Future
460
change mitigation efforts undertaken, and it is possible that these
requirements could increase drastically if levels of climate change exceed
systemic thresholds – whether in geophysical or socioeconomic systems.
Practical integration of climate change mitigation and adaption into a
development context is complicated because of a differential distribution
in costs and benefits (e.g., mitigation benefits are distributed and
accrue globally; adaptation benefits, like disaster risk management, are
often easier to measure locally). In addition, the research and policy
discourses of these three policy domains are quite separated and in
areas technically unrelated, and the constituencies and decisionmakers
are often different (Wilbanks et al., 2007). In many cases, the challenge
of bringing the entire range of issues and options into focus – seeking
synergies and avoiding conflicts – is most likely to occur in discussions
of climate change responses and development objectives in particular
places: localities and small regions where compliance with national or
international mitigation agendas provides a logic for local action
(Wilbanks, 2003). The following subsections present the urban and rural
contexts as examples.
8.5.2.1. Urban
In an increasingly urbanized world, global sustainability in the context
of a changing climate will depend on achieving sustainable and climate-
resilient cities. Urban spatial form is critical for energy consumption,
emission patterns, and disaster risk management (Desplat et al., 2009),
and it influences where and how residents live and the modes of transport
that they use. Urban planning is a tool that can be used to pursue climate
change mitigation, adaptation, and disaster risk reduction as part of the
everyday development process (Newman and Kenworthy, 1989; Bento
et al., 2005; Handy et al., 2005; Ewing and Rong, 2008; Grazi et al.,
2008; Brownstone and Golob, 2009; Glaeser and Kahn, 2010). Urban
form also influences the spatial and social inequalities that largely shape
vulnerability, coping, and adaptive capacity (Pelling, 2003; Gusdorf
et al., 2008; Leichenko and Solecki, 2008). The historical failure of
urban planning in most developing country cities has had tremendous
environmental and social consequences (UN-HABITAT, 2009; World Bank,
2010a). Also in richer countries, where planning is not comprehensive,
maladaptation can take place rather than synergistic risk reduction, for
example, where urban heat wave risk management results in increased
private air conditioning without decarbonized energy available (Lindley et
al., 2006). Similarly, a denser city may reduce greenhouse gas emissions
but increase heat wave vulnerability (Hamin and Gurran, 2009).
However, since urban forms influence both greenhouse gas emissions
and vulnerability (McEvoy et al., 2006), scope for synergistic planning
and action can also be found. For example, managing car use may
contribute to decreased greenhouse gas emissions, but also lower local
particulate pollution and reduce the health impacts of urban heat waves
(Dennekamp and Carey, 2010).
As yet there is only limited evidence that opportunities for synergistic
planning offered by urbanization are being realized, especially for those
most marginalized and vulnerable. More typically, urbanization
compounds environmental problems. As countries urbanize, the risks
associated with economic asset loss tend to increase through rapid
growth in infrastructure and productive and social assets, while mortality
risk tends to decrease (Birkmann, 2006). As cities grow, they also modify
the surrounding rural environment, and consequently may generate a
significant proportion of the hazard to which they are also exposed. For
example, as areas of hinterland are paved over, runoff increases during
storms, greatly magnifying flood hazards (Mitchell, 1999; Pelling, 1999).
As mangroves are destroyed in coastal cities, storm-surge hazard can
increase (Hardoy et al., 2001). Likewise, within urban areas (though
often beyond the reach of urban planning), the expansion of informal
settlements can lead to increased local population exposure to landslide
and flood hazards (Satterthwaite, 1997; UNDP, 2004). Global risk
models indicate that expansion of urban risk is primarily due to rapidly
increasing exposure, which outpaces improvements in capacity to
reduce vulnerability (such as through improvements legislating and
applying building standards and land use planning), at least in rapidly
growing low- and middle-income nations (UNISDR, 2009, 2011). As
a consequence, risk is becoming increasingly urbanized (Mitchell,
1999; Pelling, 2003; Leichenko and O’Brien, 2008). There are dramatic
differences, nonetheless, between developed and developing countries.
In most developed countries (and increasingly in a number of cities in
middle-income countries, e.g., Bogota, Mexico City), risk-reducing
capacities exist that can manage increases in exposure. In contrast, in
much of the developing world (and particularly in the poorest least-
developed countries where large proportions of the urban population live
in unplanned settlements) such capacities are greatly restricted, while
population growth drives exposure. Financial and technical constraints
matter for risk management, but differences in wealth alone do not
explain differences in risk reduction investments, which also depend on
risk perceptions and political choice (e.g., Satterthwaite, 1998; Hardoy
et al., 2001; Hanson et al., 2011).
Urban planning can be a vehicle for synergy, but it takes time to produce
significant effects. Synergy in planning requires anticipation of future
climate change, taking into account how climate will change over many
decades, the uncertainty of this information, the vulnerability of urban
systems, and the capacity of social agents. The Asian Cities Climate
Change Resilience Network found that catalyzing city-level actors to
assess these plans are essential, rather than depending on external
experts or national agencies to prepare urban plans (Tyler et al., 2010).
Built forms are difficult to change because they exhibit strong inertia
and irreversibility: when a low-density city is created, transforming it
into a high-density city is a long, expensive, and difficult process
(Gusdorf et al., 2008). This point is crucial in the world’s most rapidly
growing cities, where urban forms of the future are being decided based
on actions taken in the present, and where current trends indicate that
low-density, automobile-dependent forms of suburban settlement are
rapidly expanding (Solecki and Leichenko, 2006). Some work has started
to investigate these aspects of climate change adaptation and mitigation
(Newman et al., 1996). At the same time, there are specific opportunities
when cities enter periods of large-scale transformation. This is happening
Chapter 8Toward a Sustainable and Resilient Future
461
in Delhi, Mumbai, and other cities in India as private capital redevelops
low-income city neighborhoods into commercial districts and middle-
and high-income housing areas with associated low-income housing.
There is rare scope here to promote disaster risk reduction, climate change
adaptation, and mitigation alongside existing demands for market
profitability and social justice in urban and building design. There are
also growing numbers of large-scale slum/informal settlement upgrading
programs that aim to improve housing and living conditions for low-
income households (Boonyabancha, 2005; Satterthwaite, 2010).
Disaster reconstruction also creates opportunities for synergistic
development planning. For example, reconstruction after the 2005
Hurricane Katrina disaster in New Orleans, Louisiana, included rebuilding
to Green Building Council ‘Leadership in Energy and Environmental
Design’ (LEED) standards (USGBC, 2010). Similarly, in May 2007,
Greensburg, Kansas, was virtually destroyed by a tornado and LEED
standards have been applied (Harrington, 2010). Echoing the tradeoffs
between speed and sustainability presented in Section 8.2.5, the actions
in Greensburg have also slowed rebuilding of the town, leading in this
instance to an erosion in community and associated aspects of
resilience in the short run, while attempting to create a model ‘green’
community in the long run.
In short, despite the many opportunities for building synergy into urban
development planning and practice, examples of success are not plentiful.
Lack of synergy is more the norm, to take just one example of urbanization
in central Dhaka, Bangladesh. These flood-prone areas had until recently
been occupied by natural water bodies and drains, vital to the regulation
of floods. The Dhaka Metropolitan Development Plan restricts development
in many of these areas, but despite the Plan, infilling continues with
both private- and public-sector projects. Destruction of retention ponds
and drains increases risks of flooding and building in the drained
wetlands generates new risks of liquefaction following earthquakes
(UNISDR, 2011).
8.5.2.2. Rural
Rural areas are the primary site for climate change mitigation. Rural areas
have considerable experience in disaster risk management and more
recently in climate change adaptation (UNDP, 2007b). Nonetheless, as
for urban areas, the evidence base is limited for consciously synergistic
development projects and policies that consider climate change
mitigation, adaptation, and disaster management together. There are,
however, several important opportunities where climate change mitigation
and adaptation or risk management have shown scope for integration
and opportunities are being explored, for example in agroforestry
(Verchot et al., 2007).
Any scope for synergy needs to be seen within the context of contemporary
development pressures (Goklany, 2007). For small farms in particular,
pressures are strong for diversification into non-farm activities, where such
opportunities exist, but strong support is needed to enable transitions
in economic activity (Roshetko et al., 2007). Climate change affects the
range of choices available, for example, in low-lying coastal zones
where saltwater intrusion and coastal flooding are already making
traditional agriculture marginal and leading to the adoption of saltwater
tolerant crops or a shift from agriculture to aquaculture (Adger, 2000).
While urban areas have expanded in size and influence, the majority of
the poor continue to reside in rural areas in many countries, particularly
in Africa, and are among the most resource-scare and capacity-limited
population groups (UNDP, 2009). For populations that may also be
isolated from markets and communication networks, even small increases
in the frequency or severity of hazard can cause local livelihoods to
collapse, though recent developments in communication technology may
bridge this gap (Aker and Mbiti, 2010). Where political and economic
systems disrupt food distribution and market functioning, vulnerability
to food insecurity escalates (Misselhorn, 2005).
Hard choices also have to be made between expanding rural populations
or economies and natural capital. Too often, local natural assets are
exploited not by local actors to build local capacities but by external
agents, such that resources are extracted with little benefit accruing
locally. The balance and implementation of controls on natural resource
exploitation is both a potential damper on current capacity building and
a critical mechanism for ensuring long-term sustainability of rural
livelihoods and ecosystem services (Chouvy and Laniel, 2007). Non-farm
income now represents a substantial proportion of total income for
many rural households and can, in turn, increase resilience to weather-
and climate-related shocks (Brklacich et al., 1997; Smithers and Smit,
1997; Wandel and Smit, 2000). The implications of these transitions for
local rural risk, and how far they may provide scope for mitigation, has
not been fully explored in the literature.
While urban sites offer opportunities for mitigation through diversified
(household) production and energy conservation, rural areas are a focus
for concentrated low- or no-carbon energy production ranging from
hydroelectric power (HEP) to solar and wind farms, biofuel crops, and
carbon sink functions associated with forestry in particular and REDD+
projects. These investments can have significant local impacts on disaster
risk through changes in land use and land cover that may influence
hydrology, or through economic effects and consequences for livelihoods.
There is scope for synergy, for example, through small HEP/flood or
water conservation dams, and some have gone as far to say that this
joined-up approach is part of a transformed development policy for
meeting combined energy and water demands in vulnerable rural
communities, most particularly in sub-Saharan Africa (Foster and Briceño-
Garmendia, 2011). Some impacts can even go beyond local places.
Recent impacts of biofuel production on rural livelihoods and global
food security indicate the interdependence of vulnerability in rural and
urban systems, and the care required in transformations of this kind
where impacts can quickly spread and be amplified through global
markets (Dufey, 2006; de Fraiture et al., 2008).
Flows of investment, remittances, migration, and material transfers
through trade and also in the movement of resources (water, food,
Chapter 8 Toward a Sustainable and Resilient Future
462
waste, and energy) intimately connect rural and urban economies and
societies, and the local with the global, such that the sustainability of
one will influence the other. The existence of multiple, intersecting
stressors in rural and urban contexts draws attention to the importance
of addressing the underlying drivers of risk as a means of both disaster
risk management and adaptation, and of promoting climate change
mitigation.
8.6. Options for Proactive, Long-Term
Resilience to Future Climate Extremes
Considering the broad challenges described previously, it is important
to assess the range of existing planning tools and the ways they are
used, who uses them, and how they interact or change over time.
Pursuing sustainable and resilient development pathways requires
integrated and ambitious policy that is science-based and knowledge-
driven, and that is capable of addressing issues of heterogeneity and
scale. The latter issues are particularly vexing, as the consequences of,
and responses to, extreme climate and weather events are local, but
these responses need to be supported and enabled by actions at regional,
national, and global scales.
This section first considers the challenges of planning for the future,
then assesses the literature pertaining to tools and practices that can
help address these issues. As the preceding sections in this chapter and
other chapters in the report have argued, achieving a sustainable and
resilient future draws attention to the need for both incremental and
transformational changes. Based on an assessment of the literature, the
final section discusses why such changes may involve a combination of
adaptive management, learning, innovation, and leadership.
8.6.1. Planning for the Future
Disaster risk management and climate change adaptation are
fundamentally about planning for an uncertain future, a process that
involves combining one’s own aspirations (individual and collective)
with perspectives on what is to come (Stevenson, 2008). Planning for
the future is challenging when the stakes are high, values disputed,
and decisions urgent, and these factors often create tensions among
different visions of development. Typically, decisionmakers (representing
households, local or national governments, international institutions,
etc.) look to the future partly by remembering the past (e.g., projections
of the near future are often derived from recent experiences with extreme
events) and partly by projecting how the future might be different
(using forecasts, scenarios, visioning processes, or story lines – either
formal or informal) (Miller, 2007). Projections further into the future are
necessarily shrouded in larger uncertainties. The most common
approach for addressing these uncertainties is to develop multiple
visions of the future (quantitative scenarios or narrative storylines), that
in early years can be compared with actual directions of change
(Boulanger et al., 2006a,b; Moss et al., 2010).
Scenario development has become an established research tool both in
the natural sciences (e.g., Nakicenovic et al., 2000; Lobell et al., 2008)
and in the social sciences (e.g., Wack, 1985; Davis, 1998; Robinson,
2003; Galer, 2004; Kahane, 2004; Rosegrant et al., 2011). Scenarios can
be based on different spatial (e.g., global, national, and local) and
temporal scales (e.g., from a few years to several decades or centuries).
The challenges for integrated disaster risk management and climate
change adaptation scenarios are to generate climate data that can be
downscaled to at least regional and sub-national scales, while extending
disaster risk projections to longer time scales (see Gaffin et al., 2004;
Theobald, 2005; Bengtsson et al., 2006; van Vuuren et al., 2006; Grübler
et al., 2007; Moss et al., 2010; Hallegatte et al., 2011a).
Scenario development has traditionally been carried out in a sequential
manner (Moss et al., 2010). For example, a first step in developing climate
change scenarios has typically involved structural projections of key
determinants of greenhouse gas emissions (e.g., population changes,
urbanization, etc.). These have been used to estimate concentrations
and radiative forcing from emissions, leading to climate projections that
can be used in impacts research. One difficulty in using climate scenarios
for disaster risk management and climate change adaptation has been
the uncertainties associated with extreme climate and weather events,
including the behavior of local climates (see Section 3.2.3). Future
socioeconomic changes (e.g., demography, population preferences,
technologies) are also highly uncertain, thus scenarios must consider a
wide range of possible futures to design adaptation strategies and
analyze tradeoffs (e.g., Hall, 2007; Lempert, 2007; Lempert and Collins,
2007; WGBU, 2008; Dessai et al., 2009a,b; Hallegatte, 2009). Alternative
approaches have focused first on scenarios of radiative forcing, followed
by an analysis of the combinations of economic, technological,
demographic, policy, and institutional factors that can influence such
trajectories (Moss et al., 2010). Other approaches are based on robust
decisionmaking (e.g., Groves and Lempert, 2007; Lempert and Collins,
2007; Groves et al., 2008); information gap analysis (Hine and Hall,
2010); or on the search for co-benefits, no regrets strategies, flexibility,
and reversibility (e.g., Fankhauser et al., 1999; Goodess et al., 2007;
Hallegatte, 2009).
Scenario development requires substantial climate, social, environmental,
and economic data, which are not equally available or accessible for all
parts of the world. Qualitative scenarios can also be produced based on
expert judgment (e.g., Delphi exercises) or on storylines designed through
consultative processes. Such scenarios often reflect different mindsets
or worldviews that represent contrasting visions of the future.
To adapt to changing climate and weather extremes, difficult choices
may become increasingly necessary. In many locations, for example,
adapting to scenarios of reduced water availability may involve
increased investments in water infrastructure to provide enough
irrigation to maintain existing agricultural production, or a shift from
current production to less water-consuming crops (see Rosenzweig et
al., 2004; ONERC, 2009; Gao and Hu, 2011). In considering adaptation
to future flood risk in the Thames Estuary, the UK Environment Agency
Chapter 8Toward a Sustainable and Resilient Future
463
(2009) applied four scenarios over three time periods to flood
management. Through a wide consultation process, it was determined
that improving the current infrastructure continues to be the preferred
strategy until 2070, when construction of an outer barrage may become
justifiable, especially as economic and climate change conditions
change over time.
Evaluating choices among different options depends on how the
stakeholders view the region in coming decades, and on adaptation
decisions that are informed by political processes. One scenario
approach that explicitly acknowledges both social and environmental
uncertainties entails identification of flexible adaptation pathways for
managing the future risks associated with climate change (Yohe and
Leichenko, 2010). Based on principles of risk management (which
emphasize the importance of diversification and risk-spreading
mechanisms in order to improve social and/or private welfare in situations
of profound uncertainty), this approach can be used to identify a
sequence of adaptation strategies that are designed to keep society at
or below acceptable levels of risk. These strategies, which policymakers,
stakeholders, and experts develop and implement, are expected to
evolve over time as knowledge of climate change and associated climate
hazards progresses. The flexible adaptation or adaptive management
approach that underpins this also stresses the connections between
adaptation and mitigation of climate change, recognizing that climate
change mitigation will be needed in order to sustain society at or below
an acceptable level of risk (Yohe and Leichenko, 2010).
In contrast to predictive scenarios and risk management approaches,
exploratory and normative approaches can be used to develop scenarios
that represent desirable alternative futures. This is particularly important
in the case of sustainability, where the most likely futures may not be
the most desirable (Robinson, 2003), and where poverty, inequity, and
injustice are recognized by many as incompatible with sustainable
development (Redclift 1987, 1992; St. Clair, 2010). Pathways that
require considerable transformation to reach sustainable futures of this
kind can be supported by backcasting techniques. The process of
backcasting involves developing normative scenarios that explore the
feasibility and implications of achieving certain desired outcomes
(Robinson, 2003; Carlsson-Kanyama et al., 2008). It is concerned with
how desirable futures can be attained, focusing on policy measures that
would be required to reach such conditions. Participatory backcasting,
which involves local stakeholders in visionary activities related to
sustainable development, can also open deliberative opportunities and
inclusiveness in decision framing and making. Where visioning is
repeated it can also open possibilities for tracking development and
learning processes that make up adaptive strategies for disaster risk
management, based on the explicit acknowledgement of the beliefs,
values, and preferences of citizens (Robinson, 2003). Changing attitudes
and core beliefs, including those on climate change, its causes, and
consequences, is a slow process (Volkery and Ribeiro, 2009).
Adding an anticipatory dimension to planning for the future is critical
for striving toward transformational actions in the face of multiple and
dynamic uncertainties. The literature on anticipatory action learning
provides some experience on what this might look like (Stevenson,
2002; Kelleher, 2005). The framing and negotiation of decisionmaking
and policy is made inclusive and reflexive through multiple rounds of
stakeholder engagement to explore meanings of what different futures
may involve, reflect upon unavoidable tradeoffs and the winners and
losers, and establish confidence to creatively adapt to new challenges
(Inayatullah, 2006). This type of learning stresses the skills, knowledge,
and visions of those at risk and aims to support leadership from even
the most vulnerable. A combination of local- and global-scale scenarios
that link storylines developed at several organizational levels (Biggs et
al., 2007), personalizing narratives to create a sense of ownership
(Frittaion et al., 2010), and providing safe and repeated learning spaces
(Kesby, 2005) can enhance learning.
While scenarios, projections, and forecasts are all useful and important
inputs for planning, actual planning and decisionmaking is a complex
socio-political process involving different stakeholders and interacting
agents. Although much progress has been made by employing scenario
building and narrative creation to explore uncertainties, surprises,
extreme events, and tipping points, the transition from envisioning to
planning, policymaking, and implementation remains poorly understood
(Lempert, 2007). Similarly, more widespread uptake of even scientifically
highly robust scenarios may be hampered by conflicting understandings
of and practical approaches to uncertainty, different scalar needs, and
lack of training among users (Gawith et al., 2009). Experiences in
scenario building emphasize their usefulness for raising awareness on
climate change (Gawith et al., 2009). However, to move from framing
public debates to policymaking and implementation, useful scenario
building requires procedural stability, permanent yet flexible institutional
and governance structures that build trust, and experience to take
advantage of new insights for effective and fair risk management
(Volkery and Ribeiro, 2009).
Developing the capacity for adaptive learning to accommodate
complexity and uncertainty requires exploratory and imaginative visions
for the future that support choices and can accommodate multiple values
and aspirations (Miller, 2007). Disaster risk management and climate
change adaptation, and synergies between the two, can contribute
toward planning for a sustainable and resilient future, but this involves
expanding the diversity of futures that are considered and identifying
those that are desirable, as well as the short- and long-term values and
actions that are consistent with them (Lempert, 2007).
8.6.2. Approaches, Tools, and Integrating Practices
As discussed above, scenarios, narrative storylines (Tschakert and Dietrich,
2010) and simulations (Nicholls et al., 2007) can help to project and
facilitate discussion of possible futures. This section considers the tools
that are available for helping decisionmakers and planners think about
and plan for the future in the context of extreme climate and weather
events. Past experiences with enhancing resilience to climate extremes
Chapter 8 Toward a Sustainable and Resilient Future
464
include examples of both specific decision support tools and the
governance and institutional contexts in which these tools are used and
subsequent decisions are made (OECD, 2009b; Burch et al., 2010;
Whitehead et al., 2010). Tools include those that enable information
gathering, monitoring, analysis, and assessment; simulate threats;
develop projections of possible impacts; and explore implications for
response. Effective approaches combine understandings of potential
stresses from climate extremes, along with possible tipping points for
affected social and physical systems, with monitoring systems for tracking
changes and identifying emerging threats in time for adaptive responses.
This is, of course, challenging, requiring methodologies that can be open
to both quantitative and qualitative data and their analysis, including
participatory deliberation (NRC, 2010).
Institutional innovations aimed at improving the availability of disaster
information to decisionmakers include the creation of national or
regional institutions to manage and distribute disaster risk information
(von Hesse et al., 2008; Corfee-Morlot et al., 2011), bringing together
previously fragmented efforts centered in national meteorological,
geological, oceanographic, and other agencies. The World Meteorological
Organization and partner organizations have proposed the creation of
a Global Framework for Climate Services, a collaborative effort to help
the global community to better adapt to climate variability and change
by developing and incorporating science-based climate information and
prediction into planning, policy, and practice (WCC-3, 2009). New open-
source tools for comprehensive probabilistic risk assessment are beginning
to offer ways of compiling information at different scales and from
different institutions (OECD, 2009b). A growing number of countries are
also systematically recording disaster loss and impacts at the local level
(DesInventar, 2010) and developing mechanisms to use such information
to inform and guide public investment decisions (Comunidad Andina,
2007, 2009; von Hesse et al., 2008) and national planning. Unfortunately,
there is as yet only limited experience with the integrated deployment
of such tools and institutional approaches, especially in ways that cross
scales of risk management strategy development and decisionmaking,
and very limited evaluations of such deployment.
8.6.2.1. Improving Analysis and Modeling Tools
Various tools can be used to design environmental and climate policies.
Among them, integrated environment-energy-economy models produce
long-term projections taking into account demographic, technological,
and economic trends (e.g., Edenhofer et al., 2006; Clarke and Weyant,
2009). These models can be used to assess the consequences of various
policies. However, most such models are at spatial and temporal scales
that do not resolve specific climate extremes or disasters (Hallegatte et
al., 2007). At higher spatial resolution, numerical models (e.g., input-
output models, calculable general equilibrium models) can help to
assess disaster consequences and, therefore, balance the cost of disaster
risk management actions and their benefits (Rose et al., 1997; Gordon
et al., 1998; Okuyama, 2004; Rose and Liao, 2005; Tsuchiya et al., 2007;
Hallegatte, 2008b). In particular, they can compare the cost of responding
to disasters with the cost of preventing disasters. Since disasters have
intangible consequences (e.g., loss of lives, ecosystem losses, cultural
heritage losses, distributional consequences) that are difficult to measure
in economic terms, the quantitative models are necessary but not sufficient
to determine desirable policies and disaster risk management actions.
Whether incorporated in models or used in other forms of analysis, CBA
is useful to compare costs and benefits; but when intangibles play a
large role and when no consensus can be reached on how to value these
intangibles, other decisionmaking tools and approaches are needed.
Multi-criteria decisionmaking (Birkmann, 2006), robust decisionmaking
(e.g., Lempert 2007; Lempert and Collins, 2007), transition management
approaches (e.g., Kemp et al., 2007; Loorbach, 2010), and group-process
analytic-deliberative approaches (Mercer et al., 2008) are examples of
such alternative decisionmaking methodologies.
Also necessary are indicators to measure the successes and failures of
policies. For example, climate change adaptation policies often target
the enhancement of adaptive capacity. The effects and outcomes of
policies are often measured using classical economic indicators such as
GDP. The limits of such indicators are well known, and have been
summarized in several recent reports (e.g., CMEPSP, 2009; OECD,
2009a). To measure progress toward a resilient and sustainable future,
one needs to include additional components, such as measures of
stocks, other capital types (natural capital, human capital, social capital),
distribution issues, and welfare factors (health, education, etc.). Many
alternative indicators have been proposed in the literature, but no
consensus exists. Examples of these alternative indicators include the
Human Development Index, the Genuine Progress Indicator, the Index of
Sustainable Economic Welfare, the Ecological Footprint, the normalized
GDP, and various indicators of vulnerability and adaptive capacity
(Costanza, 2000; Yohe and Tol, 2002; Lawn, 2003; Costanza et al., 2004;
Eriksen and Kelly, 2006; Jones and Klenow, 2010).
8.6.2.2. Institutional Approaches
Among the most successful disaster risk management and adaptation
efforts have been those that have facilitated the development of
partnerships between local leaders and other stakeholders, including
extra-local governments (Bicknell et al, 2009; Pelling and Wisner, 2009;
Gero et al, 2011). This allows local strength and priorities to surface in
disaster risk management, while acknowledging also that communities
(including local government) have limited resources and strategic scope
and alone cannot always address the underlying drivers of risk
(Bhattamishra and Barrett, 2010). Local programs are now increasingly
moving from a focus on strengthening disaster preparedness and
response to reducing both local hazard levels and vulnerability (e.g.,
through slope stabilization, flood control measures, improvements in
drainage, etc.) (Lavell, 2009; UNISDR, 2009; Reyos, 2010). Most of the
cases where sustainable local processes have emerged are where
national governments have decentralized both responsibilities and
resources to the local level, and where local governments have become
more accountable to their citizens, as for example in cities in Colombia
Chapter 8Toward a Sustainable and Resilient Future
465
such as Manizales (Velásquez, 1998, 2005). In Bangladesh and Cuba,
successes in disaster preparedness and response leading to drastic
reduction in mortality due to tropical cyclones, built on solid local
organization, have relied on sustained support from the national level
(Haque and Blair, 1992; Bern et al., 1993; Ahmed et al., 1999;
Chowdhury, 2002; Kossin et al., 2007; Elsner et al., 2008; Karim and
Mimura, 2008; Knutson et al., 2010; World Bank, 2010b). A growing
number of examples now exist of community-driven approaches that are
supported by local and national governments as well as by international
agencies, through mechanisms such as social funds (Bhattamishra and
Barrett, 2010).
Risk transfer instruments, such as insurance, reinsurance, insurance
pools, catastrophe bonds, micro-insurance, and other mechanisms, shift
economic risk from one party to another and thus provide compensation
in exchange for a payment, often a premium (ex-post effect) (see
Sections 5.6.3, 6.5.3, and 7.4, and Case Study 9.2.13). In addition, these
mechanisms can also help to anticipate and reduce (economic) risk as
they reduce volatility and increase economic resilience at the household,
national, and regional levels (Linnerooth-Bayer et al., 2005). As one
example, with such insurance, drought-exposed farmers in Malawi have
been able to access improved seeds for higher yielding and higher risk
crops, thus helping them to make a leap ahead in terms of generating
higher incomes and the adoption of higher return technologies (World
Bank, 2005; Hazell and Hess, 2010). However, many obstacles to such
schemes still exist, particularly in low-income and many middle-income
countries, including the absence of comprehensive risk assessments and
required data, legal frameworks, and the necessary infrastructure, and
probably more experience is required to determine the contexts in
which they can be effective (Linnerooth-Bayer and Mechler, 2007;
Cummins and Mahul, 2009; Mahul and Stutley, 2010).
Disaster risk management and adaptation can also be addressed
through the enhancement of generic adaptive capacity alongside hazard-
specific response strategies (IFRC, 2010). This capacity includes access
to information, the skills and resources needed to reflect upon and apply
new knowledge, and institutions to support inclusive decisionmaking.
These are cornerstones of both sustainability and resilience. While
uncertainty may make it difficult for decisionmakers to commit funds for
hazard-specific risk reduction actions, these barriers do not prevent
investment in generic foundations of resilient and sustainable societies
(Pelling, 2010a). Importantly, from such foundations, local actors may
be able to make better-informed choices on how to manage risk in their
own lives, certainly over the short and medium terms. For instance,
federations formed by slum dwellers have become active in identifying and
acting on disaster risk within their settlements and seeking partnerships
with local governments to make this more effective and larger scale
(IFRC, 2010).
Changes in systems and structures may call for new ways of thinking
about social contracts, which describe the balance of rights and
responsibilities between different parties. Social contracts that are
suitable for technical problems can be limiting and insufficient for
addressing adaptive challenges (Heifetz, 2010). Pelling and Dill (2009)
describe the ways that current social contracts are tested when disasters
occur, and how disasters may open up a space for social transformation,
or catalyze transformative pathways building on pre-disaster trajectories.
O’Brien et al. (2009) consider how resilience thinking can contribute to
new debates about social contracts in a changing climate, drawing
attention to tradeoffs among social groups and ecosystems, and to the
rights of and responsibilities toward distant others and future generations.
8.6.2.3. Transformational Strategies and Actions
for Achieving Multiple Objectives
If extreme climate and weather events increase significantly in coming
decades, climate change adaptation and disaster risk management are
likely to require not only incremental changes, but also transformative
changes in systems and institutions. Transformation can be defined as a
fundamental qualitative change, or a change in composition or structure
that is often associated with changes in perspectives or initial conditions
(see Box 8-1). It often involves a change in paradigm and may include
shifts in perception and meaning, changes in underlying norms and
values, reconfiguration of social networks and patterns of interaction,
changes in power structures, and the introduction of new institutional
arrangements and regulatory frameworks (Folke et al., 2009, 2010;
Pahl-Wostl, 2009; Smith and Stirling, 2010).
Although transformational policies and measures may be deliberately
invoked as a strategy to reduce disaster risk and adapt to climate change,
in many cases such strategies are precipitated by an extreme event,
sometimes referred to as a ‘focusing event’ (Birkland, 1996). However,
whether an extreme event leads to any change at all is unclear, as
processes of policy change are often subtle and complex and linked to
learning processes (Birkland, 2006). Exploring the relationship between
systematic learning processes and small disasters, Voss and Wagner
(2010) find that a failure to learn is the most common prerequisite for
future disasters. There are, however, many dimensions to learning (e.g.,
cognitive, normative, and relational; see Huitema et al., 2010), and
learning may be a necessary but insufficient condition for initiating
transformational change.
Understanding processes of deliberate change and change management
can provide insights on societal responses to extreme climate and
weather events. Traditional approaches to managing change successfully
in businesses and organizations focus on a series of defined steps
(Harvard Business Essentials, 2003). Kotter (1996), for example, identifies
an eight-step process for promoting change: (1) create a sense of urgency;
(2) pull together the guiding team; (3) develop the change vision and
strategy; (4) communicate for understanding and buy in; (5) empower
others to act; (6) produce short-term wins; (7) don’t let up; and (8)
create a new culture. Kotter (1995) also identifies eight errors that are
often made when leading change, including, for example, allowing too
much complacency, failing to create a sufficiently powerful guiding
coalition, and underestimating the power of a sound vision. It is also
Chapter 8 Toward a Sustainable and Resilient Future
466
important to recognize that many change initiatives create uncertainty
and disequilibria, and are considered disruptive or disorienting (Heifetz
et al., 2009). Furthermore, vested interests seldom choose transformation,
particularly when there is much to lose from change (Christensen, 1997).
As discussed in Section 8.5.2, there are winners and losers not only from
extreme climate and weather events, but also from responses.
Consequently, fundamental change is often resisted by the people that it
affects the most (Kotter, 1996; Kegan and Lahey, 2009). Helping people,
groups, organizations, and governments to manage the resulting
disequilibria is seen as essential to successful transformation.
Many of the recent approaches to change and transformation focus on
learning organizations, and the importance of changing individual and
collective mindsets or mental models (Senge, 1990; Heifetz et al., 2009;
Kegan and Lahey, 2009; Scharmer, 2009). This transformational change
literature distinguishes between technical problems that can be
addressed through management based on existing organizational and
institutional structures and cultural norms, and adaptive challenges that
require a change in mindsets, including changes in assumptions, beliefs,
priorities, and loyalties (Heifetz et al., 2009; Kegan and Lahey, 2009).
Treating disaster risk management and climate change adaptation as
technical problems may focus attention only on improving technologies,
reforming institutions, or managing displaced populations, whereas
viewing them as an adaptive challenge shifts attention toward gaps
between values and behaviors (e.g., values that promote human security
versus policies or behaviors that undermine health and livelihoods),
beliefs (e.g., a belief that disasters are inevitable or that adaptation will
occur autonomously), and competing commitments (e.g., a commitment
to maintaining aid dependency or preserving social hierarchies).
Although most problems have both technical and adaptive elements,
treating an adaptive challenge only as a technical problem limits
successful outcomes (Heifetz et al., 2009).
Transformative changes that move society towards the path of
openness and adaptability depend not only on changes in mindsets, but
also on changes in systems and structures. Case studies of social-
ecological systems suggest that there are three phases involved in systems
transformations. The first phase includes being prepared, or preparing
the system, for change. The second phase calls for navigating the
transition by making use of a sudden crisis as an opportunity for
change, whether the crisis is real or perceived. The third phase involves
building resilience of the new system (Olsson et al., 2004; Chapin et al.,
2010). Traditional management approaches emphasize the reduction of
uncertainties, with the expectation that this will lead to systems that
can be predicted and controlled. However, in the case of climate change,
future projections of climate variables and extremes will contain
uncertainty (see Section 3.2.3). Consequently, there is a need for
management approaches that are adaptive and robust in the presence
of large and irreducible uncertainties.
8.6.3. Facilitating Transformational Change
Adapting to climate and weather extremes associated with rapid and
severe climate change, such as a warming beyond 4°C within this
century, without transformational policy and social change will be
difficult: if not chosen through proactive policies, forced transformations
and crises are likely to result (New et al., 2011). Adaptation that is
Chapter 8Toward a Sustainable and Resilient Future
FAQ 8.2 | Are transformational changes desirable and even possible, and if so, who will lead them?
Transformation in and of itself is not always desirable. It is a complex process that involves changes at the personal, cultural, institutional,
and systems levels. Transformation can imply the loss of the familiar, which can create a sense of disequilibrium and uncertainty. In
some cases, notable changes in the nature, form, or appearance of a system or process may be inconsistent with the values and preferences
of some groups. Transformation can thus be perceived as threatening by some and instrumental by others, as the potential for real or
perceived winners and losers at different scales stimulates social unease or tension. Desirable or not, it is important to recognize that
transformations are now occurring at an unprecedented rate and scale, influenced by globalization, social and technological development,
and environmental change. Climate change itself represents a system-scale transformation that will have widespread consequences for
ecology and society, including through changes in climate extremes. Responses to climate change and changes in disaster risk can be
both incremental and transformational. Transformational responses are not always radical or monumental – sometimes they simply
involve a questioning of assumptions or viewing a problem from a new perspective. Transformational responses are not only possible,
but they can be facilitated through learning processes, especially reflexive learning that explores blind spots in current thinking and
approaches to disaster risk management and climate change adaptation. However, because there are risks and barriers, transformation
also calls for leadership – not only from authority figures who hold positions and power, but from individuals and groups who are able
to connect present-day actions with their values, and with a collective vision for a sustainable and resilient future. Considering the
balance between incremental and transformative adjustments flags the importance of scale: first, because of the opportunities for
enhancing leadership capacity that come from greater involvement of those locally at risk or undertaking adaptive experimentation for
risk management; and second, because of the potential for transformation, incremental change, or stability at one systems level or sector
(e.g., administrative, social, technical) to provoke or restrict adjustments in other systems and scales. Inter-scale and inter-sectoral
communication therefore become important tools for managing adaptive disaster risk management.
467
transformative marks a shift from emphasizing finite projects with linear
trajectories and readily identifiable, discrete strategies and outcomes
(Schipper, 2007) toward an approach that includes adaptive management,
learning, innovation, and leadership, among other elements. These
aspects of adjustment are increasingly seen as being embedded in
ongoing socio-cultural and institutional learning processes. This can be
observed in the many adaptation projects that emphasize learning about
risks, evaluating response options, experimenting with and rectifying
options, exchanging information, and making tradeoffs based on public
values using reversible and adjustable strategies (McGray et al., 2007;
Leary et al., 2008; Hallegatte, 2009; Hallegatte et al., 2011b).
Transformational adaptations are likely to be enabled by a number of
factors. Some of the factors arise from external drivers such as focal
events that catalyze attention to vulnerabilities or the presence of other
sources of stress that also encourage considerations of major changes.
Supportive social contexts such as the availability of understandable
and socially acceptable options, access to resources for action, and the
presence of incentives may also be important. Other factors are related
to effective institutions and organizations, including those described in
the following subsections.
8.6.3.1. Adaptive Management
In general terms, adaptive management can be defined as a structured
process for improving management policies and practices by systemic
learning from the outcomes of implemented strategies, and by taking
into account changes in external factors in a proactive manner (Pahl-
Wostl et al., 2007; Pahl-Wostl, 2009; Section 6.6.2). Principles of adaptive
management can contribute to a more process-oriented approach to
disaster risk management, and have already shown some success in
promoting sustainable natural resource management under conditions
of uncertainty (Medema et al., 2008). Adaptive management is often
associated with ‘adaptive’ organizations that are not locked into rigid
agendas and practices, such that they can consider new information, new
challenges, and new ways of operating (Berkhout et al., 2006; Pelling et
al., 2007). Organizations that can monitor environmental, economic,
and social conditions and changes, respond to shifting policies and
leadership changes, and take advantage of opportunities for innovative
interventions are a key to resilience, especially with respect to conceivable
but long-term and/or relatively low-probability events. Those social
systems that appear most adept at adapting are able to integrate formal
organizational roles with cross-cutting informal social spaces for learning,
experimentation, communication, and for trust-based and speedy disaster
response that is nonetheless accountable to beneficiaries (Pelling et al.,
2007).
Adaptive management is a challenge for those organizations that
perceive reputational risk from experimentation and the knowledge
that some local experiments may fail (Fernandez-Gimenez et al., 2008).
Where this approach works best, outcomes have gone beyond specific
management goals to include trust-building among stakeholders a
resource that is fundamental to any policy environment facing an
uncertain future, and which also has benefits for quality of life and
market competitiveness (Fernandez-Gimenez et al., 2008). It requires
revisiting the relationship between the state and local actors concerning
facilitation of innovation, particularly when experiments go wrong.
Investing in experimentation and innovation necessarily requires some
tolerance for projects that may not be productive or cost effective, or at
least not in the short term or under existing risk conditions. However, it
is exactly the existence of this diversity of outcomes that makes societies
fit to adapt once risk conditions change, particularly in unexpected and
nonlinear directions.
8.6.3.2. Learning
The dynamic notion of adaptation calls for learning as an iterative
process in order to build resilience and enhance adaptive capacity now,
rather than targeting adaptation in the distant future (see Section 1.4.2).
Social and collective learning includes support for joint problem solving,
power sharing, and iterative reflection (Berkes, 2009). The need to take
into account the arrival of new information in the design of response
strategies has also been mentioned for mitigation policies (Ha Duong et
al., 1997; Ambrosi et al., 2003). Adaptive management is an incremental
and iterative learning-by-doing process, whereby participants make sense
of system changes, engage in actions, and finally reflect on changes and
actions. Lessons from learning theories, including experiential learning
(Kolb, 1984) and transformative learning (Mezirow, 1995), stress the
importance of learning-by-doing in concrete learning cycles, problem-
solving actions, and the reinterpretation of meanings and values
associated with learning activities.
Learning is a key component for living with uncertainty and extreme
events, and is nurtured by building the right kind of social/institutional
space for learning and experimentation that allows for competing
worldviews, knowledge systems, and values, and facilitates innovative
and creative adaptation (Thomas and Twyman, 2005; Armitage et al.,
2008; Moser, 2009; Pettengell, 2010). Examples include promoting
shared platforms for dialogs and participatory vulnerability assessments
that include a wide range of stakeholders (see ISET, 2010). It is equally
important to acknowledge that abrupt and surprising changes may
surpass existing skills and memory (Batterbury, 2008). Adaptation projects
have demonstrated that fostering adaptive capacity and managing
uncertainty in real-time, by adjusting as new information, techniques, or
conditions emerge, especially among populations exposed to multiple
risks and stressors, is more effective than more narrowly designed
planning approaches that target a given impact and are dependent on
particular future climate information (McGray et al., 2007; Pettengell,
2010). In the humanitarian sector, institutionalized processes of learning
have contributed to leadership innovation (see Box 8-3).
Action research and learning provide a powerful complement to
resilience thinking, as they focus explicitly on iterative or cyclical
learning through reflection of successes and failures in experimental
Chapter 8 Toward a Sustainable and Resilient Future
468
action, transfer of knowledge between learning cycles, and the next
learning loop that will lead to new types of action (List, 2006; Ramos,
2006). Referring to the learning processes described in Section 1.4.2,
critical reflection is paramount to triple-loop learning; it also constitutes
the key pillar of double-loop learning, or the questioning of what works
and why that is fundamental to shifts in reasoning and behavior (Kolb
and Fry, 1975; Argyris and Schön, 1978; Keen et al., 2005). Allowing time
for reflection in this iterative learning process is important because it
provides the necessary space to develop and test theories and strategies
under ever-changing conditions. It is through such learning processes
that individual and collective empowerment can emerge and potentially
be scaled up to trigger transformation (Kesby, 2005).
8.6.3.3. Innovation
The transformation of society toward sustainability and resilience
involves both social innovations and technological innovations –
incremental as well as radical. Innovation can refer to non-material
changes related to knowledge, cognition, communication, or intelligence,
or it can refer to any kind of material resources. In some cases, small
adjustments in practices or technologies may represent innovative steps
toward sustainability, while in other cases there is a strong need for more
radical transformations. Some of the literature on innovation focuses on
ensuring economic competitiveness for firms in an increasingly
globalized economy (Fløysand and Jakobsen, 2010), and some
concentrates on the relationship between environment on the one hand
and the competitiveness of firms on the other (Mol and Sonnenfeld,
2000). In addition, there is a body of social science literature on innovation
that has emerged during the last 15 years, motivated by the need for
transforming society as a whole in more sustainable directions. Recent
literature has brought out new ideas and frameworks for understanding
and managing technology and innovation-driven transitions, such as
the Multi Level Perspective (MLP) (Rip and Kemp, 1998; Geels, 2002;
Geels and Schot, 2007; Markard and Truffer, 2008). Combining insights
from evolutionary theory and sociology of technology, MLP conceptualizes
major transformative change as the product of interrelated processes
occurring at the three levels of niches, regimes, and landscape. The model
emphasizes the incremental nature of innovation in socio-technical
regimes. Transitions – that is, shifts from one stable socio-technical
regime to another – occur when regimes are destabilized through
landscape pressures, which provide breakthrough opportunities for
niche innovations.
In this field of research, there is a strong focus on systems innovation
and transformation of socio-technical systems, with the potential of
facilitating transitions from established systems for transport, energy
supply, agriculture, housing, etc., to alternative, sustainable systems
(Geels, 2002; Hoogma et al., 2002; Smith et al., 2005; Raven et al., 2010).
The systems innovation literature analyzes the emergence and dynamics
of large-scale, long-term socio-technical transformations (Kemp et al.,
1998).
Though not directly dependent on changes in technology, technological
and social innovations are often closely interrelated, not the least in
that they involve changes in social practices, institutions, cultural values,
knowledge systems, and technologies (Rohracher, 2008). Box 8-4
describes such innovation in water management. A central, basic insight
established within this research is that social and technological change
is an interactive process of co-development between technology and
society (Kemp, 1994; Hoogma et al., 2002; Rohracher, 2008). Throughout
history, new socio-technical systems have emerged and replaced old ones
in so-called technological revolutions, and an important characteristic of
such transitions is the interactions and conflicts between new, emerging
systems and established and dominating socio-technical regimes, with
strong actors defending business as usual (Kemp, 1994; Perez, 2002).
Chapter 8Toward a Sustainable and Resilient Future
Box 8-3 | Institutionalized Research and Learning in the Humanitarian Sector
An important attribute of the humanitarian sector is its readiness to learn. Research and learning unfolds at multiple levels, including
sector-wide reviews of performance and practice such as those undertaken by the Active Learning Network for Accountability and
Performance in Humanitarian Action. Research and learning is also structured around the internal needs of organizations (e.g., Red Cross
and Red Crescent Societies) or the outcomes of individual events (e.g., the landmark report on humanitarian sector practice following
the Indian Ocean tsunami (Telford et al., 2006). Organizations have different methodologies, target audiences, and frames of reference,
making cross-sector learning difficult (Amin and Goldstein, 2008), but they all have led to practical and procedural changes. Less well-
developed is active experimentation in the field of practice, with a view of proactive learning (Corbacioglu and Kapucu, 2006). This is
difficult in the humanitarian sector, where stakes are high and rapid action has typically made it difficult to implement learning-while-doing
experiments. Where experimentation may be more observable, for example, in disaster prevention and risk reduction or reconstruction
activities, there are significant gaps in documentation that have slowed down the transferring of learning outcomes between organizations.
Hierarchical models of governance have fostered a lack of cooperation and generated competition between agencies within the
humanitarian and development sectors, partly explaining why there is more learning based on the sharing of experience inside
organizations than across sectors (Kapucu, 2009). But the increasing scale and diversity of risk associated with climate change, and
compounded by other development trends such as growing global inequality and urbanization, puts more pressure on donors to promote
cross-sector communication of productive innovations and of the research and experimentation such innovation builds upon.
469
8.6.3.4. Leadership
Leadership can be critical for disaster risk management and climate
change adaptation, particularly in initiating processes and sustaining
them over time (Moser and Ekstrom, 2010). Change processes are shaped
both by the action of individual champions (as well as by those resisting
change) and their interactions with organizations, institutional structures,
and systems. Leadership can be a driver of change, providing direction and
motivating others to follow, thus the promotion of leaders by institutions
is considered an important component of adaptive capacity (Gupta et al.,
2010), although knowledge about how to create and enable leadership
remains elusive. Leadership and leaders often do not develop independent
of the institutional context, which includes institutional rules, resources,
and organizational culture (Kingdon, 1995).
Leaders who facilitate transformation have the capacity to understand
and communicate a wide set of technical, social, and political perspectives
related to a particular issue or problem. They are also able to reframe
meanings, overcome contradictions, synthesize information, and create
new alliances that transform knowledge into action (Folke et al., 2009).
Leadership also involves diagnosing the kinds of losses that some
people, groups, organizations, or governments may experience through
transformative change, such as the loss of status, wealth, security,
loyalty, or competency, not to mention loved ones (Heifetz, 2010).
Leaders help individuals and groups to take action to mobilize ‘adaptive
work’ in their communities, such that they and others can thrive in a
changing world by managing risk and creating alternative development
pathways or engaging and directing people during times of choice and
change (Heifetz, 2010).
8.7. Synergies between Disaster Risk
Management and Climate Change
Adaptation for a Resilient and
Sustainable Future
Drawing on the assessment presented in this chapter, it becomes clear that
there are many potential synergies between disaster risk management
and climate change adaptation that can contribute to social, economic,
and environmental sustainability and a resilient future. There is, however,
no single approach, framework, or pathway to achieve this. Responding
to a diversity of extremes in the present and under varying social and
environmental conditions can contribute to future resilience in situations
of uncertainty. Nonetheless, some important contributing factors have
been identified and discussed in this chapter, and are confirmed by the
wider literature (e.g., Lemos et al., 2007; Tompkins et al., 2008; Pelling,
2010a; Wisner, 2011). Eight critical factors stand out as important:
1) A capacity to reconcile short- and long-term goals
2) A willingness to reconcile diverse expressions of risk in multi-hazard
and multi-stressor contexts
3) The integration of disaster risk reduction and climate change
adaptation into other social and economic policy processes
4) Innovative, reflexive, and transformative leadership (at all levels)
5) Adaptive, responsive, and accountable governance
6) Support for flexibility, innovation, and learning, locally and across
sectors
7) The ability to identify and address the root causes of vulnerability
8) A long-term commitment to managing risk and uncertainty and
promoting risk-based thinking.
Lessons learned in climate change adaptation and disaster risk
management illustrate that managing uncertainty through adaptive
management, anticipatory learning, and innovation can lead to more
flexible, dynamic, and efficient information flows and adaptation
plans, while creating openings for transformational action. Reducing
vulnerability has been identified in many contemporary disaster studies
as the most important prerequisite for a resilient and sustainable future.
Research has consistently found that for long-term sustainability, disaster
risk management is most impactful when combined with structural
reforms that address underlying causes of vulnerability and the structural
Chapter 8 Toward a Sustainable and Resilient Future
Box 8-4 | Innovation and Transformation
in Water Management
The impacts of climate change in many regions are
predominantly linked to the water system, in particular
through increased exposure to floods and droughts (Lehner
et al., 2006; Smith and Barchiesi, 2009; see Section 2.5).
Considering water as a key structuring element or guiding
principle for landscape management and land use planning
requires technology, integrated systems thinking, and the art
of thinking in terms of attractiveness and mutual influence, or
even mutual consent, between different authorities, experts,
interest groups, and the public. One of the most pronounced
changes can be observed in The Netherlands, where the
government has requested a radical rethinking of water
management in general and flood management in particular.
The resulting policy stream, initiated through the ‘Room for the
River’ (Ruimte voor de Rivier) policy, has strongly influenced
other areas of government policy. Greater emphasis is now
given to the integration of water management and spatial
planning, with the regulating services provided by landscapes
with natural flooding regimes being highly valued. This
requires a revision of land use practices and reflects a gradual
movement toward integrated landscape planning, whereby
water is recognized as a natural, structural element. The societal
debate about the plans to build in deep-lying polders and
other hydrologically unfavorable spots, and new ideas on
floating cities, indicate a considerable social engagement of
both public and private parties with the issue of sustainable
landscapes and water management. However, although such
innovative ideas have been adopted in policy, they take time
to implement, as there is considerable social resistance
(Wolsink, 2006).
470
inequalities that create and sustain poverty and constrain access to
resources (Hewitt, 1983; Wisner et al., 2004; Lemos et al., 2007; Collins,
2009; Pelling, 2010b).
Engaging possible and desirable futures and options for decisionmaking
fosters knowledge generation essential for adaptive risk management
as well as iterative change processes. Zooming in on uncertain elements
and their potential impacts (e.g., changes in rainfall and variability)
and identifying factors that currently limit adaptive capacity (e.g.,
marginalization, lack of access to resources, or information gaps) allows
for more robust decisionmaking that also integrates local contexts (asset
portfolios, spreading and managing risks) with the climate context
(current trends, likely futures, and uncertainties) to identify the most
feasible, appropriate, and equitable response strategies, policies, and
external interventions (Pettengell, 2010). Creating space and recognizing
a diversity of voices often means reframing what counts as knowledge,
engaging uncertainties, nourishing the capacity for narrative imagination,
and articulating agency and strategic adaptive responses in the face of
already experienced changes and to anticipate and prepare for future
disturbances and shocks (Tschakert and Dietrich, 2010).
Challenges remain with respect to anticipating low-probability, high-
impact events and potentially catastrophic tipping points that represent
futures too undesirable to imagine, especially under circumstances
where exposure and vulnerability are high and adaptive capacity low
(Volkery and Ribeiro, 2009). At a practical level, there are many gaps and
barriers to realizing synergies for integration to foster a sustainable and
resilient future. For example, overcoming the current disconnect
between local risk management practices and national institutional and
legal frameworks, policy, and planning can be considered key to
reconciling short- and long-term goals for vulnerability reduction. Even
where capacity is present, it can take effort to shift into more critical,
learning modes of governance (Corbacioglu and Kapucu, 2006).
Moreover, anticipating vulnerabilities as well as feasible and fair actions
may also reveal limits of adaptation and risk management, and thus
raise the potential need for transformation. Because transformative
changes open up questions about the values and priorities shaping
development and risk futures, who wins and loses, and the balance of
tradeoffs, decisions about when and where to facilitate transformative
change and to whose benefit are inherently normative and political.
Transformation cannot be approached without understanding related
ethical and governance dimensions. At the same time, incremental
changes, in supporting many aspects of business-as-usual, also possess
implicit ethical and normative aspects. At heart, it is perhaps in failing
to fully reveal and question these normative positions that current
disaster risk management practice and policy has remained outside
of development planning and policy processes, inhabiting a long
acknowledged, but still present ‘disasters archipelago’ in the policy
world (Hewitt, 1983, p. 12).
Disasters often require urgent action and represent a time when everyday
processes for decisionmaking are disrupted. Although it is a useful
approach in responding to emergency events and disaster relief, such
top-down command and control frameworks work less well in disaster
risk reduction and this is likely to be the case too in integrated adaptive
risk management. In such systems, it is often the most vulnerable to
hazards that are left out of decisionmaking processes (Pelling, 2003,
2007; Cutter, 2006; Mercer et al., 2008), whether it is within households
Chapter 8Toward a Sustainable and Resilient Future
FAQ 8.3 | What practical steps can we take to move toward a sustainable and resilient future?
The disruptions caused by disaster events often reveal development failures. They also provide an opportunity for reconsidering
development through reconstruction and disaster risk reduction. Practical steps can address both the root causes of risk found in
development relations, including enhancing human rights, gender equity, and environmental integrity, and more proximate causes
expressed most commonly through a need for extending land and property rights, access to critical services and basic needs, including
social safety nets and insurance mechanisms, and transparent decisionmaking, especially at the local level. Identifying the drivers of
hazard and vulnerability in ways that empower both those at risk and risk managers to take action is key. This can be done best where
local and scientific knowledge is combined in the generation of risk maps or risk management plans. Greater use of local knowledge
when coupled with local capacity can initiate enhanced accountability in integrated risk decisionmaking that helps to break unsustainable
development relations.
The uncertainty that comes with climatic variability and extremes reinforces arguments for better coordination and accountability within
governance hierarchies and across sectors, as well as between generations and for non-human species in development. Local, national,
and international actors bring different strengths and tools to questions of environmental change and its relationship to trends in
human development. While offering a range of specific practical measures, both local and national approaches to risk management can
better meet the flexibility demands of adaptation and resilience when they have strong, accountable leadership, and are enhanced by
systematic experimentation and support for innovation in the development of tools as part of planned adaptive risk management
approaches. International actors can help by providing an institutional framework to support experimentation, innovation, and flexibility.
This can be part of national and local strategies to move development away from incentives that promote short-term gain and toward
those that promote longer-term sustainability and flexibility.
471
(where the knowledge of women, children, or the elderly may not be
recognized), within communities (where divisions among social groups
may hinder learning), or within nations (where marginalized groups may
not be heard, and where social division and political power influence
the development and adaptation agenda). Disaster periods are frequently
the times when the development visions and aspirations for the future
of those most affected are not recognized. This reflects a widespread
limitation on the quality and comprehensiveness of local participation
in disaster risk reduction and its integration into everyday development
planning. Instead, the humanitarian imperative, limited-term reconstruction
budgets, and an understandable desire for rapid action over deliberation
means that too often international social movements and humanitarian
nongovernmental organizations, government agencies, and local relief
organizations impose their own values and visions, often with the best
of intentions. It is also important to recognize the potential for some
people or groups to prevent sustainable decisions by employing their
veto power or lobbying against reforms or regulations based on short-
term political or economic interests (Klein, 2007). The distribution of
power in society and who has the responsibility or right to shape the
future through decisionmaking today is thus significant, and includes
the role of international as well as national and local actors. Within the
international humanitarian community, efforts such as the Sphere
Standards and the Humanitarian Accountability Partnership are steps
toward addressing this challenge.
Actions to reduce disaster risk and responses to climate change
invariably involve tradeoffs with other societal goals, and conflicts
related to different values and visions for the future. Innovative and
successful solutions that combine multiple perspectives, differing
worldviews, and contrasting ways of organizing social relations have been
described by Verweij et al. (2006) as ‘clumsy solutions.’ Such solutions,
they argue, depend on institutions in which all perspectives are heard
and responded to, and where the quality of interactions among competing
viewpoints foster creative alternatives. Drawing on the development
ethics literature, St. Clair (2010) notes that when conflict and broad-
based debate arise, alternatives often flourish and many potential
spaces for action can be created, tapping into people’s innovation and
capacity to cope, adapt, and build resilience. Pelling (2010a) stresses
the importance of social learning for transitional or transformational
adaptation, and points out that it requires a high level of trust, a
willingness to experiment and accept the possibility of failure in
processes of learning and innovation, transparency of values, and active
engagement of civil society. Committing to such a learning process is,
as Tschakert and Dietrich (2010, p. 17) argue, preferable to alternatives
because “learning by shock is neither an empowering nor an ethically
defensible pathway.
The conjuncture of hazard and vulnerability, realized through disasters,
forces coping and adaptation on individuals and society. Climate change
and ongoing development place more people and assets at risk.
Noteworthy progress in disaster risk management has been made,
especially through the action of early warning on reducing mortality,
but underlying vulnerability remains high (as indicated by increasing
numbers of people affected and economic losses from disaster) and
demographic and economic development trends continue to raise the
stakes and present a choice: risk can be denied or faced, and adaptation
can be forced or chosen. A reduction in the disaster risks associated
with climate and weather extremes is therefore a question of political
choice that involves addressing issues of equity, rights, and participation
at all levels.
Chapter 8 Toward a Sustainable and Resilient Future
References
A digital library of non-journal-based literature cited in this chapter that
may not be readily available to the public has been compiled as part of
the IPCC review and drafting process, and can be accessed via either the
IPCC Secretariat or IPCC Working Group II web sites.
AAG, 2003: Global Change and Local Places: Estimating, Understanding, and
Reducing Greenhouse Gases. Association of American Geographers, Cambridge
University Press, Cambridge, UK, 290 pp.
Adger, W.N., 1996: Approaches to Vulnerability to Climate Change. CSERGE Working
Paper #GEC 96-05, University of East Anglia, Norwich, UK, 66 pp.
Adger, W.N., 2000: Social and ecological resilience: are they related? Progress in
Human Geography, 24(3), 347-364.
Adger, W.N., 2003: Social capital, collective action, and adaptation to climate
change. Economic Geography, 79(4), 387-404.
Adger, W.N., 2004: The right to keep cold. Environment and Planning A, 36(10),
1711-1715.
Adger, W.N., S. Huq, K. Brown, D. Conway, and M. Hulme, 2003: Adaptation to
climate change in the developing world. Progress in Development Studies, 3(3),
179-195.
Adger, W.N., T.P. Hughes, C. Folke, S.R. Carpenter, and J. Rockstrom, 2005: Social-
ecological resilience to coastal disasters. Science, 309, 1036-1039.
Adger,W.N., S. Agrawala, M.M.Q. Mirza, C. Conde, K. O’Brien, J. Pulhin, R. Pulwarty,
B. Smit, and K. Takahashi, 2007: Assessment of adaptation practices, options,
constraints and capacity. In: Climate Change 2007: Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani,
J.P. Palutikof, P.J. van der Linden and C.E. Hanson (eds.)], Cambridge University
Press, Cambridge, UK, pp. 717-743.
Adger, W.N., J. Barnett, and H. Ellemor, 2010: Unique and valued places. In: Climate
Change Science and Policy [Schneider, S. H., A. Rosencranz, M. Mastrandrea,
and K. Kuntz-Duriseti (eds.)]. Island Press, Washington, DC, pp. 131-138.
Adger, W.N., K. Brown, D. Nelson, F. Berkes, H. Eakin, C. Folke, K. Galvin, L.
Gunderson, M. Goulden, K. O’Brien, J. Ruitenbeek, and E. Tompkins, 2011:
Resilience implications of policy responses to climate change. WIRES Climate
Change, doi:10.1002/WCC.133.
Adly, E. and T. Ahmed, 2009: Water and food security in the river Nile basin:
Perspectives of the government and NGOs in Egypt. In: Facing Global
Environmental Change: Environmental, Human, Energy, Food, Health and Water
Security Concepts [Brauch, H.G., N.C. Behera, P. Kameri-Mbote, J. Grin, Ú.
Oswald Spring, B. Chourou, C. Mesjasz, and H. Krummenacher (eds.)]. Springer,
Berlin, Germany, pp. 645-654.
Agarwal, B., 1991: Social security and the family: Coping with seasonality and
calamity in rural India. In: Social Security in Developing Countries [Ahmad, E., J.
Dreze, J. Hills, and A. Sen (eds.)]. Clarendon Press, Oxford, UK, pp. 171-244.
Agrawala, S. (ed.), 2005: Bridge over Troubled Waters: Linking Climate Change and
Development. Organisation for Economic Co-operation and Development, Paris,
France, 153 pp.
472
Ahmed, A.U., M. Alam, and A.A. Rahman, 1999: Adaptation to climate change in
Bangladesh: future outlook. In: Vulnerability and Adaptation to Climate Change
for Bangladesh [Huq, S., M. Asaduzzaman, Z. Karim, and F. Mahtab (eds.)]. J.
Kluwer Academic Publishers, Dordrecht, The Netherlands, pp. 125-143.
Ahmed, I., 2009: Environmental refugees and environmental distressed migration as
a security challenge for India and Bangladesh. In: Facing Global Environmental
Change: Environmental, Human, Energy, Food, Health and Water Security
Concepts [Brauch, H.G., Ú. Oswald Spring, J. Grin, C. Mesjasz, P. Kameri-Mbote,
N. Chadha Behera, B. Chourou, and H. Krummenacher (eds.)]. Springer, Berlin,
Germany, pp. 295-308.
Ahrens, J. and P.M. Rudolph, 2006: The importance of governance in risk reduction
and disaster management. Journal of Contingencies and Crisis Management,
14(4), 207-220.
Aker, J.C. and I.M.Mbiti, 2010: Mobile phones and economic development in Africa.
The Journal of Economic Perspectives, 24(3), 207-232.
Albala-Bertrand, J.M., 1993: The Political Economy of Large Natural Disasters
with Special Reference to Developing Countries. Clarendon Press, Oxford, UK,
259 pp.
Alderman, H., J. Hodditnott, and B. Kinsey, 2006: Long-term consequences of early
childhood malnutrition. Oxford Economic Papers, 58(3), 450-474.
Altieri, M.A. and P. Rosset, 1999: Ten reasons why biotechnology will not ensure
food security, protect the environment, and reduce poverty in the developing
world. AgBioForum, 2(3/4), 155-162.
Alves, D., 2002: An analysis of the geographical patterns of deforestation in Brazilian
Amazonia in the 1991–1996 period. In: Patterns and Processes of Land Use and
Forest Change in the Amazon [Wood, C. and R. Porro (eds.)]. University of
Florida Press, Gainesville, Florida, pp. 95-105.
Ambrosi, P., J.-C. Hourcade, S. Hallegatte, P. Lecocq, P. Dumas, and M. Ha Duong,
2003: Optimal control models and elicitation of attitudes towards climate
damages. Environmental Modeling and Assessment, 8(3), 133-147.
Amin, S. and M. Goldstein (eds.), 2008: Data against Natural Disasters: Establishing
Effective Systems for Relief, Recovery and Reconstruction. World Bank,
Washington, DC, 342 pp.
Ammer, M., M. Nowak, L. Stadlmayr, and G. Hafner, 2010: Legal Status and Legal
Treatment of Environmental Refugees. Federal Environment Ministry, Dessau-
Roßlau, Germany, 16 pp.
Argyris, C. and D. Schön, 1978: Organizational Learning: A Theory of Action
Perspective. Addison Wesley, Reading, MA.
Armitage, D., 2005: Adaptive capacity and community-based natural resource
management. Environmental Management, 35(6), 703-715.
Armitage, D., M. Marschke, and R. Plummer, 2008: Adaptive co-management and
the paradox of learning. Global Environmental Change, 18(1), 86-98.
ASCE, 2010: Report Card for America’s Infrastructure: Dams. American Society of Civil
Engineers, Reston, VA.
Back, E., C. Cameron, and T.M. Tanner, 2009:
Children and Disaster Risk Reduction:
Taking Stock and Moving Forward. Children in a changing climate report,
Institute of Development Studies, Brighton, UK, 44 pp.
Baghel, R. and M. Nusser, 2010: Discussing large dams in Asia after the World
Commission on Dams: is a political ecology approach the way forward. Water
Alternatives, 3(2), 231-248.
Bahadur, A.V., M. Ibrahim, and T. Tanner, 2010: The Resilience Renaissance?
Unpacking of Resilience for Tackling Climate Change and Disasters.
Strengthening Climate Resilience Discussion Paper, 1, Institute of Development
Studies, Brighton, UK ,43 pp.
Balaban, V., 2006: Psychological assessment of children in disasters and emergencies.
Disasters, 30(2), 178-198.
Bangay, C. and N. Blum, 2010: Education responses to climate change and quality:
two parts of the same agenda? International Journal of Educational
Development, 30(4), 359-368.
Barnett, J., 2003: Security and climate change. Global Environmental Change, 13(1),
7-17.
Barnett, J. and W.N. Adger, 2007: Climate change, human security and violent
conflict. Political Geography, 26(6), 639-655.
Barnett, J. and S. O’Neill, 2010: Maladaptation. Global Environmental Change, 20,
211-213.
Barnett, J., R. Matthew, and K. O’Brien, 2010: Global environmental change and
human security: An introduction. In: Global Environmental Change and Human
Security [Matthew, R.A., B. McDonald, J. Barnett, and K. O’Brien (eds.)]. MIT
Press, Cambridge, MA, pp. 3-32.
Bartlett, S., 2008: After the tsunami in Cooks Nagar: The challenges of participatory
rebuilding. Children, Youth and Environments, 18(1), 470-484.
Batterbury, S., 2008: Anthropology and global warming: the need for environmental
engagement. Australian Journal of Anthropology, 19(1), 62-68.
Bauer, S., 2011: Stormy weather: International security in the shadow of climate change.
In: Coping with Global Environmental Change, Disaster and Security [Brauch,
H.G., Ú. Oswald Spring, C. Mesjasz, J. Grin, P. Kameri-Mbote, B. Chourou, P.
Dunay, and J. Birkmann (eds.)]. Springer, Berlin, Germany, pp. 719-734.
Beck, U., 2007: World at Risk. Polity, Cambridge, UK, 269 pp.
Beckman, M., 2011: Converging and conflicting interests in adaptation to env.
change in Central Vietnam. Climate and Development, 3(1), 32-41.
Bengtsson, M., Y. Shen, and T. Oki, 2006: A SRES-based gridded global population
dataset for 1990–2100. Population and Environment, 28(2), 113-131.
Beniston, M., 2004: The 2003 heat wave in Europe: a shape of things to come? An
analysis based on Swiss climatological data and model simulations.
Geophysical Research Letters, 31(2), L02202.
Benson, C. and E. Clay, 2004: Understanding the Economic and Financial Impact of
Natural Disasters. International Bank for Reconstruction and Development,
Disaster Risk Management Series 4, World Bank, Washington, DC, 134 pp.
Bento, A.M., M.L. Cropper, A.M. Mobarak, and K. Vinha, 2005: The effects of urban
spatial structure on travel demand in the United States. Review of Economics
and Statistics, 87(3), 466-478.
Berke, P.R., J. Kartez, and D. Wenger, 1993: Recovery after disaster: achieving
sustainable development, mitigation and equity. Disasters, 17(2), 93-109.
Berkes, F., 2007: Understanding uncertainty and reducing vulnerability: Lessons from
resilience thinking. Natural Hazards, 41(2), 283-295.
Berkes, F., 2009: Evolution of co-management: role of knowledge generation, bridging
organizations and social learning. Journal of Environmental Management,
90(5), 1692-1702.
Berkes, F., J. Colding, and C. Folke (eds.), 2003: Navigating Social-ecological Systems:
Building Resilience for Complexity and Change. Cambridge University Press,
Cambridge, UK, 398 pp.
Berkes, F., T.P. Hughes, R.S. Stenech, J.A. Wilson, D.R. Bellwood, B. Crona, C. Folke, L.H.
Gunderson, H.M. Leslie, J. Norberg, M. Nyström, P. Olsson, H. Österblom, M.
Scheffer, and B. Worm, 2006: Globalization, roving bandits, and marine
resources. Science, 311(5767), 1557-1558.
Berkhout, F., J. Hertin, and D.M. Gann, 2006: Learning to adapt: Organisational
adaptation to climate change impacts. Climatic Change, 78(1), 135-156.
Bern, C., J. Sniezek, G.M. Mathbor, M.S. Siddiqi, C. Ronsmans, A.M. Chowdhury, A.E.
Chowdhury, K. Islam, M. Bennish, E. Noji, and R.I. Glass, 1993: Risk factors for
mortality in the Bangladesh Cyclone 1991. Bulletin of the World Health
Organization, 71(1), 73-78.
Bhattamishra, R. and C.B. Barrett, 2010: Community-based risk management
arrangements: A review. World Development, 38(7), 923-932.
Bicknell, J., D. Dodman, and D. Satterthwaite (eds.), 2009: Adapting Cities to Climate
Change: Understanding and Addressing the Development Challenges.
Earthscan Publications, London, p. 384.
Biggs, D., R. Biggs, V. Dakos, R.J. Scholes, and M.L. Schoon, 2011: One crisis after
another: Are we entering an era of concatenated global shocks? Ecology and
Society, 16(2), 27.
Biggs, R., C. Raudsepp-Hearne, C. Atkinson-Palombo, E. Bohensky, E. Boyd, G.
Cundill, H. Fox, S. Ingram, K. Kok, S. Spehar, M. Tengö, D. Timmer, and M. Zurek,
2007: Linking futures across scales: a dialog on multiscale scenarios. Ecology
and Society, 12(1), 17.
Biggs, R., S.R. Carpenter, and W.A. Brock, 2009: Turning back from the brink: detecting
an impending regime shift in time to avert it. Proceedings of the National
Academy of Sciences, 106(3), 826-831.
Chapter 8Toward a Sustainable and Resilient Future
473
Birkenholtz, T., 2008: Irrigated landscapes, produced scarcity and adaptive social
institutions in Rajasthan, India. Annals of the Association of American
Geographers, 99(1), 118-137.
Birkland, T.A., 1996: Natural disasters as focusing events: Policy communities and
political response. International Journal of Mass Emergencies and Disasters,
14(2), 221-243.
Birkland, T.A., 2006: Lessons of Disaster: Policy Change after Catastrophic Events.
Georgetown University Press, Washington, DC, 224 pp.
Birkmann, J., (ed.), 2006: Measuring Vulnerability to Natural Hazard: Towards
Disaster Resilient Societies. UNU Press, Tokyo, Japan, 400 pp.
Bizikova, L., S. Burch, S. Cohen, and J. Robinson, 2010: Linking sustainable development
with climate change adaptation and mitigation. In: Climate Change, Ethics and
Human Security [O’Brien, K., A. St. Clair, and B. Kristoffersen (eds.)]. Cambridge
University Press, Cambridge, UK, pp. 157-179.
Blaikie, P., T. Cannon, I. Davis and B. Wisner, 1994: At Risk: Natural Hazards, People’s
Vulnerability and Disaster. 1st ed. Routledge, London, UK, 286 pp.
Bogardi, J. and H.G. Brauch, 2005: Global environmental change: A chal lenge for
human security – Defining and conceptualizing the environmental dimension of
human security. In: UNEO – Towards an International Environmental
Organisation - Approaches to a Sustainable Reform of Global Environmental
Governance [Rechkemmer, A. (ed.)]. Nomos, Baden-Baden, Germany, pp. 85-109.
Bohle, H.G., 2009: Sustainable livelihood security. Evolution and application. In:
Facing Global Environ mental Change: Environmental, Human, Energy, Food,
Health and Water Security Con cepts [Brauch, H.G., Ú. Oswald Spring, J. Grin, C.
Mesjasz, P. Kameri-Mbote, N. Chadha Behera, B. Chourou, and H.
Krummenacher (eds.)]. Springer, Berlin, Germany, pp. 521-528.
Bohle, H.G., B. Etzold, and M. Keck, 2009: Resilience as agency. IHDP Update, 2,
8-13.
Boonyabancha, S., 2005: Baan Mankong; going to scale with ‘slum’ and squatter
upgrading in Thailand. Environment and Urbanization, 2(2), 309-329.
Boulanger, J.-P., F. Martinez, and E.C. Segura, 2006a: Projection of future climate
change conditions using IPCC simulations, neural networks and Bayesian
statistics. Part 1. Temperature mean state and seasonal cycle in South America.
Climate Dynamics, 27, 233-259.
Boulanger, J.-P., F. Martinez, and E.C. Segura, 2006b: Projection of future climate
change conditions using IPCC simulations, neural networks and Bayesian
statistics. Part 2. Precipitation mean state and seasonal cycle in South America.
Climate Dynamics, 28, 255-271.
Bradley, R.S., M. Vuille, H. Diaz, and W. Vergara, 2006: Threats to water supplies in
the tropical Andes. Science, 312, 1755-1756.
Brand, F.S. and K. Jax, 2007: Focusing the meaning(s) of resilience: resilience as a
descriptive concept and a boundary object. Ecology and Society, 12(1), 23.
Brauch, H.G., 2005a: Environment and Human Security. Towards Freedom from
Hazard Impact. InterSecTions No. 2, United Nations University–Institute for
Environment and Human Security, Bonn, Germany, 60 pp.
Brauch, H.G., 2005b: Threats, Challenges, Vulnerabilities and Risks in Environmental
and Human Security. Source No. 1, United Nations University–Institute for
Environment and Human Security, Bonn, Germany, 100 pp.
Brauch
, H.G., 2009: Introduction: facing global environmental change and
sectorialization of security. In: Facing Global Environ mental Change:
Environmental, Human, Energy, Food, Health and Water Security Concepts
[Brauch, H.G., Ú. Oswald Spring, J. Grin, C. Mesjasz, P. Kameri-Mbote, N. Chadha
Behera, B. Chourou, and H. Krummenacher (eds.)]. Springer, Berlin, Germany,
pp. 21-42.
Brauch, H.G. and Ú. Oswald Spring, 2011: Introduction: coping with global
environmental change in the anthropocene. In: Coping with Global
Environmental Change, Disaster and Security [Brauch, H.G., Ú. Oswald Spring,
C. Mesjasz, J. Grin, P. Kameri-Mbote, B. Chourou, P. Dunay, and J. Birkmann
(eds.)]. Berlin, Springer, Germany, pp. 31-60.
Brauch, H.G., Ú. Oswald Spring, C. Mesjasz, J. Grin, P. Dunay, N. Chadha Behera, B.
Chourou, P. Kameri-Mbote, and P.H. Liotta (eds.), 2008: Globalization and
Environmental Challenges: Reconceptualizing Security in the 21 st Century.
Springer, Berlin, Germany, 1148 pp.
Brauch, H.G., Ú. Oswald Spring, J. Grin, C. Mesjasz, P. Kameri-Mbote, N. Chadha
Behera, B. Chourou, and H.Krummenacher (eds.), 2009: Facing Global
Environmental Change: Environmental, Human, Energy, Food, Health and Water
Security Concepts. Springer, Berlin, Germany, 1588 pp.
Brauch, H.G., Ú. Oswald Spring, C. Mesjasz, J. Grin, P. Kameri-Mbote, B. Chourou, P.
Dunay, and J. Birkmann (eds.), 2011: Coping with Global Environmental
Change, Disasters and Security – Threats, Challenges, Vulnerabilities and Risks.
Springer, Berlin, Germany, 1815 pp.
Bressers, H.T.A. and W.A. Rosenbaum (eds), 2003: Achieving Sustainable
Development: the Challenge of Governance Across Social Scales. Praeger
Publishers, Westport, CT.
Brklacich, M., D. McNabb, C. Bryant, and J. Dumanski, 1997: Adaptability of agriculture
systems to global climate change: A Renfrew county, Ontario, Canada pilot study.
In: Agricultural Restructuring and Sustainability: A Geographical Perspective
[Ilbery, B., Q. Chiotti, and T. Rickard (eds.)]. CABI, Wallingford, UK, pp. 351-364.
Brock, W.A. and S.R. Carpenter, 2007: Panaceas and diversification of environmental
policy. Proceedings of the National Academy of Sciences, 104(39), 15206-15211.
Brown, K., 2011: Sustainable adaptation: an oxymoron? Climate and Development,
3(1), 21-31.
Brownstone, D. and T.F. Golob, 2009: The impact of residential density on vehicle
usage and energy consumption. Journal of Urban Economics, 65(1), 91-98.
Bruner, A.G., R.E. Gullison, R.E. Rice, and G.A.B. da Fonseca, 2001: Effectiveness of
parks in protecting tropical biodiversity. Science, 291, 125-128.
Buhaug, H., N.P. Gleditsch, and O.M. Theisen, 2008: Implications of Climate Change
for Armed Conflict. World Bank, Washington DC, 52 pp.
Bunyavanich, S., C.P. Landrigan, A.J. McMichael, and P.R. Epstein, 2003: The impact
of climate change on child health. Ambulatory Pediatrics, 3(1),44-52.
Burby, R.J., 2006: Hurricane Katrina and the paradoxes of government disaster policy:
bringing about wise governmental decisions for hazardous areas. The Annals of
the American Academy of Political and Social Science, 604(1), 171-191.
Burby, R.J., B.A. Cigler, S.P. French, E.J. Kaiser, J. Kartez, D. Roenigk, D. Weist, and D.
Whittington, 1991: Sharing Environmental Risks: How to Control Governments’
Losses in Natural Disasters
. Westview, Boulder, CO, 280 pp.
Burby, R.J., A.C. Nelson, D. Parker, and J. Handmer, 2001: Urban containment policy and
exposure to natural hazards: is there a connection? Journal of Environmental
Planning and Management, 44(4), 475-490.
Burby, R.J., A.C. Nelson, and T.W. Sanchez, 2006: The problem of containment and
the promise of planning. In: Rebuilding Urban Places After Disaster, Lessons
from Hurricane Katrina [Birch, E.L. and S.M. Wachter (eds.)]. University of
Pennsylvania Press, Philadelphia, PA, pp. 47-65.
Burch, S., S.R.J. Sheppard, A. Shaw, and O. Flanders, 2010: Planning for climate
change in a flood-prone community: municipal barriers to policy action and the
use of visualizations as decision-support tools. Journal of Flood Risk
Management, 3(2), 126-139.
Burton, I., R.W. Kates, and G.F. White: 1993: The Environment as Hazard. 2nd ed.
Guildford Press, London, UK, 290 pp.
Camerer, C. and H. Kunreuther, 1989: Decision processes for low probability events:
policy implications. Journal of Policy Analysis and Management, 8(4), 565-592.
Campbell, B.M., J.A. Sayer, and B. Walker, 2010: Navigating trade-offs: working for
conservation and development outcomes. Ecology and Society, 15(2), 16.
Campbell, K.M., J. Gulledge, J.R. McNeill, J. Podesta, P. Ogden, L. Fuerth, R.J. Woolsey,
A.T.J. Lennon, J. Smith, R. Weitz, and D. Mix, 2007: The Age of Consequences:
The Foreign Policy and National Security Implications of Global Climate Change.
Center for Strategic and International Studies, Washington, DC, 124 pp.
Cannon, T., 2002: Gender and climate hazards in Bangladesh. Gender and
Development, 10(2), 45-50.
Cardona, O., 1999: Environmental management and disaster prevention: Two
related topics. In: Cities at Risk: Environmental Degradation, Urban Risks and
Disaster in Latin America [Fernandez, M.A. (ed.)]. La Red/USAID, Quito, Ecuador,
pp. 79-102.
Carlsson-Kanyama, A., K.H. Dreborg, H.C. Moll, and D. Padovan, 2008: Participative
backcasting: a tool for involving stakeholders in local sustainability planning.
Futures, 40(1), 34-46.
Chapter 8 Toward a Sustainable and Resilient Future
474
Carpenter, S.R., 2003: Regime Shifts in Lake Ecosystems: Pattern and Variation.
Excellence in Ecology Series, 15, Ecology Institute, Oldendorf/Luhe, Germany.
Carpenter, S.R., B.H. Walker, J.M. Anderies, and N. Abel, 2001: From metaphor to
measurement: resilience of what to what? Ecosystems, 4(8), 765-781.
Carter, M., P.D. Little, T. Mogues, and W. Negatu, 2007: Poverty traps and natural
disasters in Ethiopia and Honduras. World Development, 35(5), 835-856.
Cash, D.W., W. Adger, F. Berkes, P. Garden, L. Lebel, P. Olsson, L. Pritchard, and O.
Young, 2006: Scale and cross-scale dynamics: governance and information in a
multilevel world. Ecology and Society, 11(2), 8.
Cavallo, E. and I. Noy, 2009: The Economics of Natural Disasters: A Survey. IDB
Working Paper Series 124, Inter-American Development Bank, Washington, DC,
50 pp.
CCSP, 2008: Weather and Climate Extremes in a Changing Climate. Regions of Focus:
North America, Hawaii, Caribbean, and U.S. Pacific Islands [Karl, T.R., G.A. Meehl,
C.D. Miller, S.J. Hassol, A.M. Waple, and W.L. Murray (eds.)]. A report by the U.S.
Climate Change Science Program and the Subcommittee on Global Change
Research, Department of Commerce, NOAA National Climatic Data Center,
Washington, DC, 164 pp.
Chakrabarti, P.G.D. and M.R. Bhatt (eds.), 2006: Micro-finance and Disaster Risk
Reduction. Knowledge World, New Delhi, India.
Chambers, R., 2006: Vulnerability, coping and policy. IDS Bulletin, 37(4), 33-40.
Chapin III, F.S., S.R. Carpenter, G.P. Kofinas, C. Folke, N. Abel, W.C. Clark, P. Olsson,
D.M. Stafford Smith, B.H. Walker, O.R. Young, F. Berkes, R. Biggs, J.M. Grove, R.L.
Naylor, E. Pinkerton, W. Steffen, and F.J. Swanson, 2010: Ecosystem stewardship:
sustainability strategies for a rapidly changing planet. Trends in Ecology and
Evolution, 25(4), 241-249.
Chouvy, P.-A. and L.R. Laniel, 2007: Agricultural drug economies: cause or alternative
to intra-state conflicts? Crime, Law, Social Change, 48, 133-150.
Chowdhury, K.M.M.H., 2002: Cyclone preparedness and management in
Bangladesh. In: Improvement of Early Warning System and Responses in
Bangladesh Towards Total Disaster Risk Management Approach. Bangladesh
Public Administration Training Centre, Dhaka, Bangladesh, pp. 97-106.
Christensen, C. M., 1997: The Innovator’s Dilemma: When New Technologies Cause
Great Firms to Fail. Harvard Business School Press, Boston, MA, 256 pp.
Christoplos, I., 2006: Links between Relief, Rehabilitation and Development in the
Tsunami Response. Tsunami Evaluation Coalition, London, UK, 102 pp.
Christoplos, I., S. Anderson, M. Arnold, V. Galaz, M. Hedger, R.J.T. Klein, and K. Le
Goulven, 2009: The Human Dimension of Climate Adaptation: The Importance
of Local and Institutional Issues. Commission on Climate Change and
Development, Edita Sverige AB, Stockholm, Sweden, 38 pp.
Clarke, L. and J. Weyant, 2009: Introduction to the EMF 22 special issue on climate
change control scenarios. Energy Economics, 31(2), 63.
CMEPSP, 2009:
Report of the Commission on the Measurement of Economic
Performance and Social Progress. Commission on the Measurement of
Economic Performance and Social Progress, France, 292 pp.
CNA Corporation, 2007: National Security and the Threat of Climate Change. CNA
Corporation, Alexandria, VA, 63 pp.
Cohen, S., D. Demeritt, J. Robinson, and D. Rothman, 1998: Climate change and
sustainable development: towards dialogue. Global Environmental Change,
8(4), 341-371.
Collins, A.E., 2009: Disaster and Development. Routledge Perspectives on
Development, Routledge, London, UK.
Comunidad Andina, 2007: Agenda Ambiental Andina 2006–2010 (Andean
Environmental Agenda 2006–2010). Secretary General, San Isidro, Perú, 12 pp.
Comunidad Andina, 2009: Articulando la gestión del riesgo y la adaptación al
cambio climático en el sector agropecuario (Articulating risk assessment and
adaptation to climate change in agriculture and livestock). Secretary General,
San Isidro, Perú, 128 pp.
Conde, J.E., 2001: The Orinoco River Delta Venezuela. In: Coastal Marine Ecosystems
of Latin America [Seeliger, U. and B. Kjerfve (eds.)]. Springer, Berlin, Germany,
pp. 61-70.
Corbacioglu, S. and N. Kapucu, 2006: Organisational learning and self adaptation
in dynamic disaster environments. Disasters, 30(2), 212-233.
Corfee-Morlot, J., I. Cochran, P.-J. Teasdale, and S. Hallegatte, 2011: Multilevel
governance and deliberative practice to support local climate action. Climatic
Change, 104(1), 169-197.
Costanza, R., 2000: The dynamics of the ecological footprint concept. Ecological
Economics, 32, 341-345.
Costanza, R., J. Erickson, K. Fligger, A. Adams, C. Adams, G. Altschuler, S. Balter, B.
Fisher, J. Hike, J. Kelly, T. Kerr, M. McCauley, K. Montone, M. Rauch, K.
Schmiedeskamp, D. Saxton, L. Sparacino, W. Tusinski, and L. Williams, 2004:
Estimates of the Genuine Progress Indicator (GPI) for Vermont, Chittenden
County and Burlington, from 1950 to 2000. Ecological Economics, 51(1-2),
139-155.
Costello, A., M. Abbas, A. Allen, S. Ball, S. Bell, R. Bellamy, S. Friel, N. Groce, A.
Johnson, M. Kett, M. Lee, C. Levy, M. Maslin, D. McCoy, B. McGuire, H.
Montgomery, D. Napier, C. Pagel, J. Patel, J. Antonio, P. de Oliveira, N. Redclift, H.
Rees, D. Rogger, J. Scott, J. Stephenson, J. Twigg, J. Wolff, and C. Patterson, 2009:
Managing the health effects of climate change. The Lancet, 373, 1693-1733.
Cummins, J.D. and O. Mahul, 2009: Catastrophe Risk Financing in Developing
Countries: Principles for Public Intervention. World Bank, Washington, DC, 268 pp.
Cutter, S.L., 1995: The forgotten casualties: women, children, and environmental
change. Global Environmental Change, 5(3), 181-194.
Cutter, S.L., 2006: Hazards, Vulnerability and Environmental Justice. Earthscan,
London, UK, 418 pp.
Cutter, S.L., L. Barnes, M. Berry, C. Burton, E. Evans, E. Tate, and J. Webb, 2008: A
place-based model for understanding community resilience to natural disasters.
Global Environmental Change,
18(4), 598-606.
Dalby, S., 2009: Security and Environmental Change. Polity, Cambridge, UK, 199 pp.
Davidson, F., M. Zaaijer, M. Peltenburg, and M. Rodell, 1993. Relocation and
Resettlement Manual: A Guide to Managing and Planning Relocation. Institute
for Housing and Urban Development Studies, Rotterdam, The Netherlands, 69 pp.
Davies, M., B. Guenther, J. Leavy, T. Mitchell, and T.M. Tanner, 2008: Adaptive social
protection: synergies for poverty reduction. IDS Bulletin, 39(4), 105-112.
Davies, S., 1993: Are coping strategies a cop-out? IDS Bulletin, 24(4), 60-72.
Davis, G., 1998: Creating Scenarios for Your Company’s Future. The 1998 Conference
on Corporate Environmental, Health, and Safety Excellence, Bringing
Sustainable Development Down to Earth, New York, USA, 5 pp.
Davis, M., 2001: Late Victorian Holocausts. El Niño Famines and the Making of the
Third World. Verso, London, UK, 464 pp.
de Fraiture, C., M. Giordano, and Y. Liao, 2008: Biofuels and implications for
agricultural water use: blue impacts of green energy. Water Policy, 10(S1), 67-81.
de la Fuente, A., A. Revi, F. Lopez-Calva, J. Serje, F. Ramirez, C. Rosales, A. Velasquez,
and S. Dercan, 2009: Deconstructing disaster: Risk patterns and poverty trends
at the local level. In: Global Assessment Report on Disaster Risk Reduction.
United Nations International Strategy for Disaster Reduction, Geneva,
Switzerland, pp. 59-85.
del Ninno, C. and M. Lindberg, 2005: Treading water: the long-term impact of the
1998 flood on nutrition in Bangladesh. Economics and Human Biology, 3(1),
67-96.
Dennekamp, M. and M. Carey, 2010: Air quality and chronic disease: why action on
climate change is also good for health. New South Wales Public Health Bulletin,
21(6), 115-121.
Dercon, S. and I. Outes, 2009: Income Dynamics in Rural India: Testing for Poverty
Traps and Multiple Equilibria. Background paper for the U.N.-World Bank
Assessment on the Economics of Disaster Risk Reduction, Washington, DC,
40 pp.
DesInventar, 2010: online.desinventar.org/desinventar/.
Desplat, J., J-L. Salagnac, R. Kounkou, A. Lemonsu, M. Colombert, M. Lauffenburger,
and V. Masson, 2009: Multidisciplinary study of the impacts of climate change
on the scale of Paris. In: Proceedings of the Seventh International Conference
on Urban Climate, Yokohama, Japan, June 2009, 5 pp.
Dessai, S., M. Hulme, R. Lempert, and R. Pielke, Jr., 2009a: Climate prediction: a limit
to adaptation? In: Adapting to Climate Change: Thresholds, Values, Governance
[Adger, W.N., I. Lorenzoni, and K.L. O’Brien (eds.)]. Cambridge University Press,
Cambridge, UK, pp. 64-78.
Chapter 8Toward a Sustainable and Resilient Future
475
Dessai, S., M. Hulme, R. Lempert, and R. Pielke, Jr., 2009b: Do we need better
predictions to adapt to a changing climate? Eos, 90(13), 111-112.
DFID, 2004: Disaster Risk Reduction: a Development Concern. Department for
International Development, London, UK, 74 pp.
DFID, 2006: Eliminating World Poverty: Making Governance Work for the Poor. A
White Paper on International Development, Department for International
Development, London, UK, 132 pp.
Dodman, D., J. Hardoy, and D. Satterthwaite, 2008: Urban Development and
Intensive and Extensive Risk. Contribution to the Global Assessment Report on
Disaster Risk Reduction (2009), International Institute for Environment and
Development, London, UK, 67 pp.
Drury, A.C. and R.S. Olson, 1998: Disasters and political unrest: an empirical
investigation. Journal of Contingencies and Crisis Management, 6(3), 153-161.
Dufey, A., 2006: Biofuels Production, Trade and Sustainable Development: Emerging
Issues. Sustainable Markets Discussion Paper 2, International Institute for
Environment and Development, London, UK, 62 pp.
Eakin, H. and M. Webbe, 2008: Linking local vulnerability to system sustainability in
a resilience framework: Two cases from Latin America. Climatic Change, 93,
355-377.
ECA, 2009: Shaping Climate-Resilient Development: A Framework for Decision-
Making. A report of the Economics of Climate Adaptation Working Group.
ClimateWorks Foundation, Global Environment Facility, European Commission,
McKinsey and Company, The Rockefeller Foundation, Standard Chartered Bank,
and Swiss Re, 164 pp.
Edenhofer, O., K. Lessmann, C. Kemfert, M. Grubb, and J. Köhler, 2006: Induced
technological change: exploring its implications for the economics of atmospheric
stabilization: synthesis report from innovation modeling comparison project.
The Energy Journal, 27(S1), 57-107.
Edenhofer, O., R. Pichs-Madruga, Y. Sokona, K. Seyboth, P. Matschoss, S. Kadner, T.
Zwickel, P. Eickemeier, G. Hansen, S. Schlömer, and C. von Stechow (eds.), 2011:
IPCC Special Report on Renewable Energy Sources and Climate Change
Mitigation. Cambridge University Press, Cambridge, UK, 1544 pp.
Elmqvist, T., C. Folker, M. Nyström, G. Peterson, J. Bengtsson, B. Walker, and J.
Norberg, 2003: Response diversity, ecosystem change, and resilience. Frontiers
in Ecology and the Environment, 1(9), 488-494.
Elsner, J.B., J.P. Kossin, and T.H. Jagger, 2008: The increasing intensity of the strongest
tropical cyclones. Nature, 455, 92-95.
Enfors, E. I., L.J. Gordon, G.D. Peterson, and D. Bossio, 2008: Making investments in
dryland development work: participatory scenario planning in the Makanya
catchment, Tanzania. Ecology and Society, 13(2), 42.
Eriksen, S. and K. Brown, 2011: Sustainable adaptation to climate change. Climate
and Development, 3(1), 3-6.
Eriksen, S.H. and P.M. Kelly, 2006: Developing credible vulnerability indicators for
climate adaptation policy assessment. Mitigation and Adaptation Strategies for
Global Change, 12(4), 495-524.
Eriksen, S. and K. O’Brien, 2007: Vulnerability, poverty and the need for sustainable
adaptation measures. Climate Policy, 7(4), 337-352.
Eriksen, S., P. Aldunce, C.S. Bahinipati, R. Martins, J.I. Molefe, C. Nhemachena, K.
O’Brien, F. Olorunfemi, J. Park, L. Sygna, and K. Ulsrud, 2011: When not every
response to climate change is a good one: identifying principles for sustainable
adaptation. Climate and Development, 3(1), 7-20.
EU, 2008: Climate Change and International Security. Paper from the High
Representative and the European Commission to the European Council,
European Commission, Brussels, Belgium, 11 pp.
Evans, L. and J. Oehler-Stinnett, 2006: Children and natural disasters: a primer for
school psychologists. School Psychology International, 27(1), 33-55.
Ewing, R. and F. Rong, 2008: The impact of urban form on US residential energy use.
Housing Policy Debate, 19(1), 1-30.
Fankhauser, S., J.B. Smith, and R.S.J. Tol, 1999: Weathering climate change:
some simple rules to guide adaptation decisions. Ecological Economics, 30(1),
67-78.
Farber, D., 2007: Basic compensation for victims of climate change. University of
Pennsylvania Law Review, 155(6), 1605-1656.
Fargione, J.E., T.R. Cooper, D.J. Flaspohler, J. Hill, C. Lehman, T. McCoy, S. Mcleod, E.J.
Nelson, K.S. Oberhauser, and D. Tilman, 2009: Bioenergy and wildlife: threats
and opportunities for grassland conservation. BioScience, 59(9), 767-777.
Fearnside, P.M., 2001: Status of South American natural ecosystems. In: Encyclopedia
of Biodiversity [Levin, S.A. (ed.)]. Academic Press, San Diego, CA, pp. 345-359.
FEMA, 2009: Dam Safety in the United States: A Progress Report on the National
Dam Safety Program Fiscal Years 2006 and 2007. Federal Emergency
Management Agency, Washington, DC, 60 pp.
Feng, S., A.B. Krueger, and M. Oppenheimer, 2010: Linkages among climate change,
crop yields and Mexico-US cross-border migration. Proceedings of the National
Academy of Sciences, 107(32), 14257-14262.
Fernandez-Gimenez, M.E., H.L. Ballard, and V.E. Sturtevant, 2008: Adaptive
management and social learning in collaborative and community-based
monitoring: a study of five community-based forestry organizations in the
western USA. Ecology and Society, 13(2), 4.
Ferris, E., 2011: Protecting Civilians in Disasters and Conflicts. Brookings Policy Brief
182, Brookings Institute, Washington, DC, 4 pp.
Fløysand, A. and S.-E. Jakobsen, 2010: The complexity of innovation: A relational
turn. Progress in Human Geography, 35(1), 328-344.
Foley, J.A., R. DeFries, G.P. Asner, C. Barford, G. Bonan, S.R. Carpenter, F.S. Chapin,
M.T. Coel, G.C. Daily, H.K. Gibbs, T. Helkowski, E.A. Howard, C.J. Kucharik, C.
Monfreda, J.A. Patz, I.C. Prentice, N. Ramankutty, and P.K. Snyder, 2005: Global
consequences of land use. Science, 309, 570-574.
Folke, C., 2006: Resilience: The emergence of a perspective for social-ecological
systems analyses. Global Environmental Change, 16(3), 253-267.
Folke, C., S. Carpenter, B. Walker, M. Scheffer, T. Elmqvist, L. Gunderson, and C.S. Holling,
2004: Regime shifts, resilience, and biodiversity in ecosystem management.
Annual Review of Ecology, Evolution and Systematics, 35, 557-581.
Folke, C., F.S. Chapin III, and P. Olsson, 2009: Transformations in ecosystem
stewardship. In: Principles of Ecosystem Stewardship: Resilience-Based Natural
Resource Management in a Changing World [Chapin III, F.S., G.P. Kofinas, and
C. Folke (eds.)]. Springer, New York, NY, pp. 103-125.
Folke, C., S.R. Carpenter, B.H. Walker, M. Scheffer, F.S. Chapin III, and J. Rockström, 2010:
Resilience thinking: integrating resilience, adaptability and transformability.
Ecology and Society, 15(4), 20.
Ford, J.D. and L.B. Ford (eds.), 2011: Climate Change Adaptation in Developed
Nations: From Theory to Practice. Springer, Dordrecht, The Netherlands, 295 pp.
Foster, V. and C. Briceño-Garmendia (eds.), 2011: Africa’s Infrastructure. A Time for
Transformation. World Bank and French Development Agency, Washington, DC,
44 pp.
Frittaion, C., P.N. Duinker, and J. Grant, 2010: Suspending disbelief: Influencing
engagement in scenarios of forest futures. Technological Forecasting and Social
Change, 78(3), 421-436.
Fuentes, C.J. and H.G. Brauch, 2009: The human security network: A global North-
South coalition. In: Facing Global Environmental Change: Environmental, Human,
Energy, Food, Health and Water Security Concepts [Brauch, H.G., Ú. Oswald
Spring, J. Grin, C. Mesjasz, P. Kameri-Mbote, N. Chadha Behera, B. Chourou, and
H. Krummenacher (eds.)]. Springer, Berlin, Germany, pp. 991-1001.
Gachathi, F. and S. Eriksen, 2011: Gums and resins: the potential for supporting
sustainable adaptation in Kenya’s drylands. Climate and Development, 3(1),
59-70.
Gaffin, S, X. Xing, and G. Yetman, 2004: Downscaling and geospatial gridding of
socio-economic projections from the IPCC special report on emissions scenarios
(SRES). Global Environmental Change, 14, 105-123.
Gaillard, J.C., E. Clavé, and I. Kelman, 2008: Wave of peace? Tsunami disaster
diplomacy in Aceh, Indonesia. Geoforum, 39(1), 511-526.
Galer, G., 2004: Preparing the ground? Scenarios and political change in South
Africa. Development, 47(4), 26-34.
Gao, Z. and Y. Hu, 2011: Coping with population growth, climate change, water
scarcity and growing food demand in China in the 21st century. In: Coping with
Global Environmental Change, Disaster and Security [Brauch, H.G., Ú, Oswald
Spring, C. Mesjasz, J. Grin, P. Kameri-Mbote, B. Chourou, P. Dunay, and J.
Birkmann (eds.)] Springer, Berlin, Germany, pp. 957-968.
Chapter 8 Toward a Sustainable and Resilient Future
476
Gardiner, S. 2010: Climate change: a global test for contemporary political
institutions and theories. In: Climate Change, Ethics and Human Security
[O’Brien, K., A. Lera St. Clair, and B. Kristoffersen (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 131-153.
Garg, A., R.C. Dhiman, S. Bhattacharya, and P.R. Shukla, 2009: Development, malaria
and adaptation to climate change: A case study from India. Environmental
Management Journal, 43(5), 779-789.
Gawith, M., R. Street, R. Westaway, and A. Steynor, 2009: Application of the UKCIP02
climate change scenarios: reflections and lessons learnt. Global Environmental
Change, 19(1), 113-121.
GECHS, 2000: Science Plan: Global Environmental Change and Human Security
Project. IHDP Report 11, International Human Dimensions Programme, Bonn,
Germany, 60 pp.
Geels, F.W., 2002: Technological transitions as evolutionary reconfiguration processes:
A multi-level perspective and a case-study. Research Policy, 31(8/9), 1257-1274.
Geels, F.W. and J. W. Schot, 2007: Typology of sociotechnical transition pathways.
Research Policy, 36(3), 399-417.
Gero, A., K. Méheux, and D. Dominey-Howes, 2011: Integrating community based
disaster risk reduction and climate change adaptation: examples from the
Pacific. Natural Hazards and Earth Systems Science, 11, 101-113.
Ghesquiere, F., L. Jamin, and O. Mahul, 2006: Earthquake Vulnerability Reduction
Program in Colombia: A Probabilistic Cost-Benefit Analysis. Policy Research
Working Paper 3939, World Bank, Washington, DC, 22 pp.
Gilbert, A. and P. Ward, 1984: Community participation in upgrading irregular
settlements: the community response. World Development, 12(9), 913-922.
Glaeser, E.L. and M.E. Kahn, 2010: The greenness of cities: Carbon dioxide emissions
and urban development. Journal of Urban Economics, 67(3), 404-418.
Goklany, I.M., 2007: Integrated strategies to reduce vulnerability and advance
adaptation, mitigation, and sustainable development. Mitigation and
Adaptation Strategies for Global Change, 12(5), 755-786.
Goldstein, B.E., 2009: Resilience to surprises through communicative planning.
Ecology and Society, 14(2), 33.
Goodess, C.M., J.W. Hall, M. Best, R. Betts, L. Cabantous, P.D. Jones, C.G. Kilsby, A.
Pearman, and C.J. Wallace, 2007: Climate scenarios and decision making under
uncertainty. Built Environment, 33(1), 10-30.
Gordon, P., H. Richardson, and B. Davis, 1998: Transport related impacts of the
Northridge earthquake. Journal of Transportation and Statistics, 1, 21-36.
Gostin, L.O., E.A. Friedman, G. Ooms, T. Gebauer, N. Gupta, D. Sridhar, W. Chenguang,
J-A. Røttingen, and D. Sanders, 2011: The joint action and learning initiative:
towards a global agreement on national and global responsibilities for health.
PLoS Medicine, 8(5)
, e1001031.
Grasso, M., 2009: An ethical approach to climate adaptation finance. Global
Environmental Change, 20(1), 74-81.
Grasso, M., 2010: Justice in Funding Adaptation Under the International Climate
Change Regime. Springer, Dordrecht, The Netherlands, 184 pp.
Grazi, F., J.C.J.M. van den Bergh, and J. N van Ommeren, 2008: An empirical analysis of
urban form, transport, and global warming. The Energy Journal, 29(4), 97-122.
Groves, D.G. and R.J. Lempert, 2007: A new analytic method for finding policy-
relevant scenarios. Global Environmental Change, 17(1), 73-85.
Groves, D.G., D. Knopman, R.J. Lempert, S.H. Berry, and L. Wainfan, 2008: Presenting
Uncertainty about Climate Change to Water-Resource Managers. A Summary of
Workshops with the Inland Empire Utilities Agency. Technical Reports 505, Rand
Corporation, Santa Monica, CA, 100 pp.
Grubb, M., C. Vrolijk, and D. Brack, 1999: The Kyoto Protocol: A Guide and
Assessment. Royal Institute of International Affairs and Earthscan, London, UK,
342 pp.
Grübler, A., B. O’Neill, K.Riahi, V. Chirkov, A. Goujon, P. Kolp, I. Prommer, S. Scherbov,
and E. Slentoe, 2007: Regional, national, and spatially explicit scenarios of
demographic and economic change based on SRES. Technological Forecasting
and Social Change, 74(7), 980-1029.
Gunderson, L., 2009: Comparing Ecological and Human Community Resilience.
CARRI Research Paper 5, Community and Regional Resilience Initiative,
National Security Directorate, Oak Ridge, TN, 35 pp.
Gunderson, L.H. and C.S. Holling (eds.), 2002: Panarchy: Understanding
Transformations in Human and Natural Systems. Island Press, Washington, DC,
450 pp.
Guojie, C., 2003: Ecological reconstruction of the upper reaches of the Yangtze River.
In: Natural Disasters and Development in a Globalizing World [Pelling, M. (ed.)].
Routledge, London, UK, pp. 214-230.
Gupta, J., C. Termeer, J. Klostermann, S. Meijerink, M. van den Brink, P. Jong, S.
Nooteboom, and E. Mbergsma, 2010: The adaptive capacity wheel: a method
to assess the inherent characteristics of institutions to enable the adaptive
capacity of society. Environmental Science and Policy, 13(6), 459-471.
Gusdorf, F., S. Hallegatte, and A. Lahellec, 2008: Time and space matter: how urban
transitions create inequality. Global Environment Change, 18(4), 708-719.
Ha Duong, M.,M.J. Grubb, and J.-C. Hourcade, 1997: Influence of socioeconomic
inertia and uncertainty on optimal CO
2
-emission abatement. Nature, 390,
270-274.
Haines, A., R.S. Kovats, D. Campbell-Lendrum, and C. Corvalan, 2006: Climate change
and human health: impacts, vulnerability and public health. Public Health, 120,
585-596.
Hall, J.W., 2007: Probabilistic climate scenarios may misrepresent uncertainty and
lead to bad adaptation decisions. Hydrological Processes, 21(8), 1127-1129.
Hallegatte, S., 2006: A Cost-Benefit Analysis of the New Orleans Flood Protection
System. Regulatory Analysis 2, American Enterprise Institute–Brookings Joint
Center, Washington, DC, 18 pp.
Hallegatte, S., 2008a: An adaptive regional input-output model and its application
to the assessment of the economic cost of Katrina. Risk Analysis, 28(3), 779-799.
Hallegatte, S., 2008b: A Note on Including Climate Change Adaptation in an
International Scheme. Idées pour le Débat 18, Institut du Développement
Durable et des Relations Internationals, Paris, France, 15 pp.
Hallegatte, S., 2009: Strategies to adapt to an uncertain climate change. Global
Environmental Change, 19(2), 240-247.
Hallegatte, S. and P. Dumas, 2008: Can natural disasters have positive consequences?
Investigating the role of embodied technical change. Ecological Economics,
68(3), 777-786.
Hallegatte, S. and V. Przyluski, 2010: The Economics of Natural Disasters: Concepts
and Methods. Policy Research Working Paper 5507, World Bank, Washington,
DC, 29 pp.
Hallegatte, S., J.-C. Hourcade, and P. Dumas, 2007: Why economic dynamics matter
in assessing climate change damages: illustration on extreme events. Ecological
Economics, 62(2), 330-340.
Hallegatte, S., P. Dumas, and J.-C. Hourcade, 2010: A Note on the Economic Cost of
Climate Change and the Rationale to Limit it Below 2°C. Background Paper to
the 2010 World Development Report, Policy Research Working Paper 5179,
World Bank, Washington, DC, 19 pp.
Hallegatte, S., F. Henriet, and J. Corfee-Morlot, 2011a: The economics of climate change
impacts and policy benefits at city scale: A conceptual framework. Climatic
Change, 104(1), 51-87.
Hallegatte, S., F. Lecocq, and C. De Perthuis, 2011b: Designing Climate Change
Adaptation Policies: An Economic Framework. Policy Research Working Paper
5568, World Bank, Washington, DC, 41 pp.
Hamin, E.M. and N. Gurran, 2009: Urban form and climate change: Balancing
adaptation and mitigation in the U.S. and Australia. Habitat International,
33(3), 238-245.
Handy, S., X. Cao, and P. Mokhtarian, 2005: Correlation or causality between the
built environment and travel behavior? Evidence from Northern California.
Transportation Research Part D: Transport and Environment, 10(6)
,
427-444.
Hanson, S., R. Nicholls, N. Ranger, S. Hallegatte, J. Corfee-Morlot, C. Herweijer, and J.
Chateau, 2011: A global ranking of port cities with high exposure to climate
extremes. Climatic Change, 104(1), 89-111.
Haque, C.E. and D. Blair, 1992: Vulnerability to tropical cyclones: evidence from the
April 1991 cyclone in coastal Bangladesh. Disasters, 16(3), 217-229.
Harberger, A.C., 1978: On the use of distributional weights in social cost benefit
analysis. Journal of Political Economy, 86(2), 87-120.
Chapter 8Toward a Sustainable and Resilient Future
477
Harberger, A.C., 1984: Basic needs versus distributional weights in social cost-
benefit analysis. Economic Development and Cultural Change, 32(3), 455-474.
Hardoy, J.E., D. Mitlin, and D. Satterthwaite, 2001: Environmental Problems in an
Urbanizing World. Earthscan, London, UK.
Harrington, L.M.B., 2005: Vulnerability and sustainability concerns for the US High
Plains. In: Rural Change and Sustainability: Agriculture, the Environment and
Communities [Essex, S.J., A.W. Gilg, R.B. Yarwood, J. Smithers, and R. Wilson
(eds.)]. CABI, Wallingford, UK, pp. 169-184.
Harrington, L.M.B., 2010: The U.S. Great Plains, change, and place development. In:
The Next Rural Economies: Constructing Rural Place in Global Economies
[Halseth, G., S. Markey, and D. Bruce (eds.)]. CABI, Wallingford, UK, pp. 32-44.
Harrington, L.M.B., M. Lu, and J.A. Harrington, Jr., 2009: Fossil water and agriculture
in southwestern Kansas. In: Sustainable Communities on a Sustainable Planet:
The Human-Environment Regional Observatory Project [Yarnal, B., C. Polsky,
and J. O’Brien (eds.)]. Cambridge University Press, Cambridge, UK, pp. 269-291.
Harvard Business Essentials, 2003: Managing Change and Transition. Harvard
Business School Press, Boston, MA.
Harvey, D., 2010: The Enigma of Capital: And the Crisis of Capitalism. Profile,
London, UK.
Hazell, P. and U. Hess, 2010: Drought insurance for agricultural development and
food security in dryland areas. Food Security, 2(4), 395-405.
Hedlund-de Witt, A., 2011: The rising culture and worldview of contemporary
spirituality: A sociological study of potentials and pitfalls for sustainable
development. Ecological Economics, 70, 1057-1065.
Hedrén, J. and B.O. Linnér, 2009: Utopian thought and the politics of sustainable
development. Futures, 41(4), 210-219.
Heifetz, R., 2010: Leadership. In: Political and Civic Leadership: A Reference
Handbook [Couto, R.A. (ed.)]. Sage Publications, London, UK, pp. 12-23.
Heifetz, R., A. Grashow, and M. Linsky, 2009: The Practice of Adaptive Leadership:
Tools and Tactics for Changing Your Organization and the World. Harvard
Business Press, Boston, MA.
Hewitt, K. (ed.), 1983: Interpretations of Calamity. Allen and Unwin, London, UK,
304 pp.
Hewitt, K., 1997: Regions of Risk: Geographical Introduction to Disasters. Longman,
London, UK, 389 pp.
Heyd, T. and N. Brooks, 2009: Exploring cultural dimensions of adaptation to climate
change. In: Adapting to Climate Change: Thresholds, Values, Governance
[Adger, W.N., I. Lorenzoni, and K. O’Brien (eds.)]. Cambridge University Press,
Cambridge, UK, pp. 269-282.
Hine, D. and J.W. Hall, 2010: Information gap analysis of flood model uncertainties
and regional frequency analysis. Water Resources Research, 46, W01514.
Hochrainer, S., 2009: Assessing Macroeconomic Impacts of Natural Disasters: Are There
Any? Policy Research Working Paper 4968, World Bank, Washington, DC, 43 pp.
Hogarth, R. and H. Kunreuther, 1995: Decision making under ignorance: arguing
with yourself.
Journal of Risk and Uncertainty, 10(1), 15-36.
Holling, C.S., 1973: Resilience and stability of ecological systems. Annual Review of
Ecology and Systematics, 4, 1-23.
Hoogma, R., R. Kemp, J. Schot, and B. Truffer, 2002: Experimenting for Sustainable
Transport: The Approach of Strategic Niche Management. Routledge, London,
New York.
Hopwood, B., M. Mellor, and G. O’Brien, 2005: Sustainable development: mapping
different approaches. Sustainable Development, 13(1), 38-52.
Hughes, D., 2001: Cadastral politics: The making of community-based resource
management in Zimbabwe and Mozambique. Development and Change,
32(4), 741-768.
Hughes, T., A.H. Baird, D.R. Bellwood, M. Card, S.R. Connolly, C. Folke, R. Grosberg,
O. Hoegh-Guldberg, J.B.C. Jackson, J. Kleypas, J.M. Lough, P. Marshall, M. Nystro,
S.R. Palumbi, J.M. Pandolfi, B. Rosen, and J. Roughgarden, 2003: Climate change,
human impacts, and the resilience of coral reefs. Science, 301, 929-933.
Huitema, D., C. Cornelisse, and B. Ottow. 2010. Is the jury still out? Toward greater
insight in policy learning in participatory decision processes – the case of Dutch
citizens’ juries on water management in the Rhine Basin. Ecology and Society,
15(1), 16.
Hunt, A. and T. Taylor, 2009: Values and cost-benefit analysis: economic efficiency
criteria in adaptation. In: Adapting to Climate Change: Thresholds, Values,
Governance [Adger, W.N., I. Lorenzoni, and K.L. O’Brien (eds.)]. Cambridge
University Press, Cambridge, UK, pp. 197-211.
Hunter, L.M., 2005: Migration and environmental hazards. Population and
Environment, 26(4), 273-302.
ICOMOS, 1998: Report on Economics of Conservation: An Appraisal of Theories,
Principles and Methods. International Economics Committee, International
Council on Monuments and Sites, Paris, France, 123 pp.
ICSU, 2002: Science and Technology for Sustainable Development. Consensus Report
and Background Document, Mexico City Synthesis Conference, May 2002. ICSU
Series on Science for Sustainable Development 9, International Council for
Science, Paris, France, 37 pp.
IFRC, 2002: World Disasters Report: Focus on Reducing Risk. International Federation
of Red Cross and Red Crescent Societies, Eurospan, London, UK, 239 pp.
IFRC, 2005: World Disasters Report: Focus on Information in Disasters. International
Federation of Red Cross and Red Crescent Societies, Eurospan, London, UK, 246 pp.
IFRC, 2010: World Disasters Report 2010: Focus on Urban Risk. International Federation
of Red Cross and Red Crescent Societies, Geneva, Switzerland, 220 pp.
Inayatullah, S., 2006: Anticipatory action learning: Theory and practice. Futures,
38(6), 656-666.
IOM, 2007: Discussion Note: Migration and the Environment. MC/ING/288, 94th
Session, International Organization for Migration, Geneva, Switzerland, 9 pp.
IOM, 2009a: Migration, Climate Change and the Environment. IOM Policy Brief,
International Organization for Migration, Geneva, Switzerland, 9 pp.
IOM, 2009b: Compendium of IOM’s Activities on Migration, Climate Change and the
Envi ronment. International Organization for Migration, Geneva, Switzerland,
311 pp.
IRGC, 2009: Risk Governance Deficits: An Analysis and an Illustration of the Most
Common Deficits in Risk Governance. International Risk Governance Council,
Geneva, Switzerland, 92 pp.
ISET, 2010: The Shared Learning Dialogue: Building Stakeholder Capacity and
Engagement for Resilience Action. Climate Resilience in Concept and Practice,
Working Paper 1, Institute for Social and Environmental Transition, Boulder, CO,
29 pp.
Jabry, A. (ed.), 2003: Children in Disasters: After the Cameras Have Gone. Plan UK,
London, UK, 68 pp.
Jackson, T., 2009: Prosperity without Growth: Economics for a Finite Planet.
Earthscan, London, UK, 278 pp.
Jaramillo, C.R.H., 2009: Do Natural Disasters have Long-Term Effects on Growth?
Manuscript, Universidad de los Andes, Bogota, Colombia, 44 pp.
Jones, C. and P. Klenow, 2010: Beyond GDP? Welfare across Countries and Time.
NBER Working Paper No. 16352, National Bureau of Economic Research,
Cambridge, MA, 52 pp.
Jonkman, S.N., A. Lentz, and J.K. Vrijling, 2010: A general approach for the estimation
of loss of life due to natural and technological disasters. Reliability Engineering
and System Safety, 95(11), 1123-1133.
Kahane, A., 2004: Colombia: speaking up. Development, 47(4), 95-98.
Kahl, C.H., 2003: The political ecology of violence: lessons for the Mediterranean. In:
Security and Environment in the Mediterranean. Conceptualising Security and
Environmental Conflicts [Brauch, H.G., P.H. Liotta, A. Marquina, P.F. Rogers, and
M. El-Sayed Selim (eds.)]. Springer, Berlin, Germany, pp. 465-476.
Kahl, C.H., 2006: States, Scarcity and Civil Strife in the Developing World. Princeton
University Press, Princeton, NJ, 352 pp.
Kameri-Mbote, P. and K. Kindiki, 2009: Water and food security in the Nile River
Basin: perspectives of governments and NGOs of upstream countries. In: Facing
Global Environmental Change: Environmental, Human, Energy, Food, Health
and Water Security Concepts [Brauch, H.G., Ú. Oswald Spring, J. Grin, C. Mesjasz,
P. Kameri-Mbote, N. Chadha Behera, B. Chourou, and H. Krummenacher (eds.)].
Springer, Berlin, Germany, pp. 655-664.
Kapucu, N., 2009: Interorganizational coordination in complex environments of
disasters: the evolution of intergovernmental disaster response systems.
Journal of Homeland Security and Emergency Management, 6(1), 47.
Chapter 8 Toward a Sustainable and Resilient Future
478
Karim, M.F. and N. Mimura, 2008: Impacts of climate change and sea level rise on
cyclonic storm surge floods in Bangladesh. Global Environmental Change,
18(3), 490-500.
Keen, M., V.A. Brown, and R. Dyball, 2005: Social Learning in Environmental
Management: Towards a Sustainable Future. Earthscan, London, UK, 270 pp.
Keenan, H.T., S.W. Marshall, M.A. Nocera, and D.K. Runyan, 2004: Increased
incidence of inflicted traumatic brain injury in children after a natural disaster.
American Journal of Preventive Medicine, 26(3), 189-193.
Kegan, R. and L.L. Lahey, 2009: Immunity to Change: How to Overcome It and Unlock
the Potential in Yourself and Your Organization. Harvard Business Press, Boston,
341 pp.
Kelleher, A., 2005: A personal philosophy of anticipatory action-learning. Journal of
Future Studies, 10(1), 85-90.
Kelman, I., 2003: Beyond disaster, beyond diplomacy. In: Natural Disasters and
Development in a Globalizing World [Pelling, M. (ed.)]. Routledge, London, UK,
pp. 110-123.
Kelman, I. and T. Koukis, 2000: Disaster diplomacy. Cambridge Review of
International Affairs, 14(1), 214-294.
Kemp, R., 1994: Technology and the transition to environmental sustainability. The
problems of technological regime shifts. Futures, 26(10), 1023-1046.
Kemp, R., J. Schot, and R. Hoogma, 1998: Regime shifts to sustainability through
processes of niche formation: The approach of strategic niche management.
Technology Analysis & Strategic Management, 10(2), 175-195.
Kemp, R., D. Loorbach, and R. Rotmans, 2007: Transition management as a model to
manage processes of co-evolution towards sustainable development.
International Journal of Sustainable Development and World Ecology, 14(1),
78-91.
Kent, G., 2001: The human right to disaster mitigation and relief. Environmental
Hazards, 3(3-4), 137-138.
Kesby, M., 2005: Retheorizing empowerment-through-participation as a performance
in space: beyond tyranny to transformation. Signs: Journal of Women in Culture
and Society, 30(4), 2037-2065.
Kiehl, J.T., 2006: Geoengineering climate change: Treating the symptom over the
cause? Climatic Change, 77(3-4), 227-228.
Kingdon, J.W., 1995: Agendas, Alternatives and Public Policies. 2nd ed.
HarperCollins, New York, NY, 273 pp.
Kinzig, A.P., P. Ryan, M. Etienne, H. Allison, T. Elmqvist, and B.H. Walker, 2006:
Resilience and regime shifts: assessing cascading effects. Ecology and Society,
11(1), 20.
Klein, N., 2007: The Shock Doctrine: The Rise of Disaster Capitalism
. Penguin,
London, UK, 701 pp.
Klein, R.J.T., E.L.F. Schipper, and S. Dessai, 2005: Integrating mitigation and adaptation
into climate and development policy: three research questions. Environmental
Science & Policy, 8(6), 579-588.
Klein, R.J.T., S.E.H. Eriksen, L.O. Næss, A. Hammill, T.M. Tanner, C. Robledo, and K.L.
O’Brien, 2007: Portfolio screening to support the mainstreaming of adaptation
to climate change into development assistance. Climatic Change, 84(1), 23-44.
Knutson, T.R., J.L. McBride, J. Chan, K. Emanuel, G. Holland, C. Landsea, I. Held, J.P.
Kossin, A.K. Srivastava, and M. Sugi, 2010: Tropical cyclones and climate
change. Nature Geoscience, 3, 157-163.
Kolb, D.A., 1984: Experiential Learning. Englewood Cliffs, Prentice Hall, NJ, 256 pp.
Kolb, D.A. and R. Fry, 1975: Toward an applied theory of experiential learning. In:
Theories of Group Process [Cooper, C. (ed.)]. John Wiley, London, UK, pp. 33-57.
Kolmannskog, V., 2008: Future Floods of Refugees: A Comment on Climate Change,
Conflict and Forced Migration. Report by the Norwegian Refugee Council, Oslo,
Norway, 44 pp.
Kossin, J.P., K.R. Knapp, D.J. Vimont, R.J. Murnane, and B.A. Harper, 2007: A globally
consistent reanalysis of hurricane variability and trends. Geophysical Research
Letters, 34, L04815.
Kotter, J.P., 1995: Leading change: Why transformation efforts fail. Harvard Business
Review OnPoint, March-April, 1-10.
Kotter, J.P., 1996: Leading Change. Harvard Business School Press, Boston, MA,
200 pp.
Kovats, S., D. Campbell-Lendrum, and F. Matthies, 2005: Climate change and human
health: estimating avoidable deaths and disease. Risk Analysis, 25(6), 1409-1418.
Kristjanson, P., R.S. Reid, N. Dickson, W.C. Clark, D. Romney, R. Puskur, S. MacMillan,
and D. Grace, 2009: International agricultural research knowledge with action
for sustainable development. Proceedings of the National Academy of Sciences,
106(13), 5047-5052.
Kunreuther, H., R. Ginsberg, L. Miller, P. Sagi, P. Slovic, B. Borkan, and N. Katz, 1978:
Disaster Insurance Protection: Public Policy Lessons. John Wiley and Sons, New
York, NY.
Kwa, E., 2009: Climate change and indigenous peoples in the South Pacific: the need
for regional and local strategies. In: Climate Law and Developing Countries:
Legal and Policy Challenges for the World Economy [Richardson, B.J., Y. le
Bouthillier, H. McLeod-Kilmurray, and S. Wood (eds.)]. Edward Elgar,
Cheltenham, UK, pp. 102-124.
La Trobe, S. and I. Davis, 2005: Mainstreaming Disaster Risk Reduction: A Tool for
Development Organizations. Tearfund, London, 20 pp.
Lacambra, C. and K. Zahedi, 2011: Climate change, natural hazards and coastal
ecosystems in Latin-America: A framework for analysis. In: Coping with Global
Environmental Change, Disaster and Security [Brauch, H.G., Ú. Oswald Spring,
C. Mesjasz, J. Grin, P. Kameri-Mbote, B. Chourou, P. Dunay, and J. Birkmann
(eds.)]. Springer, Berlin, Germany, pp. 585-602.
Larsen, P., S. Goldsmith, O. Smith, M. Wilson, K. Strzepek, P. Chinowsky, and B. Saylor,
2007: Estimating Future Costs for Alaska Public Infrastructure at Risk from
Climate Change
. Institute of Social and Economic Research, University of
Alaska, Anchorage, AK, 108 pp.
Last, M., 1994: Putting children first. Disasters, 18(3), 192-202.
Lauten, A.W. and K. Lietz, 2008: A look at the standards gap: comparing child
protection responses in the aftermath of hurricane Katrina and the Indian
Ocean tsunami. Children, Youth and Environments, 18(1), 158-201.
Lavell, A., 1999: Desastres en América Latina: Avences Teóricos y Prácticos: 1990-1999.
Anuario Social y Politico de América Latina y el Caribe, FLACSO-Nueva
Sociedad, Caracas, Venezuela, 34 pp.
Lavell, A., 2009: Unpacking Climate Change Adaptation and Disaster Management:
Searching for the Links and Differences: A Conceptual and Epistemological
Critique and Proposal. FLACSO, Caracas, Venezuela.
Lawn, P., 2003: A theoretical foundation to support the Index of Sustainable
Economic Welfare (ISEW), Genuine Progress Indicator (GPI), and other related
indexes. Ecological Economics, 44, 105-118.
Le Billon, P. and A. Waizenegger, 2007: Peace in the wake of disaster? Secessionist
conflicts and the 2004 Indian Ocean tsunami. Transactions of the Institute of
British Geographers, 32(3), 411-427.
Le Roy Ladurie, E., 1971: Times of Feast, Times of Famine: A History ofClimateSince
the Year 1000. Doubleday, New York, NY.
Leary, N., J. Adequwon, V. Barros, I. Burton, J. Kukarni, and R. Lasco (eds.), 2008:
Climate Change and Adaptation. Earthscan, London, UK.
Lebel, L., J.M. Anderies, B. Campbell, C. Folke, S. Hatfield-Dodds, T.P. Hughes, and J.
Wilson, 2006: Governance and the capacity to manage resilience in regional
social-ecological systems. Ecology and Society, 11(1), 19.
Lebel, L., B.T. Sinh, P. Garden, S. Seng, L.A. Tuan, and D. Van Truc, 2009: The promise of
flood protection: dykes and dams, drains and diversions. In: Contested Waterscapes
in the Mekong Region: Hydropower, Livelihoods and Governance [Molle, F., T.
Foran, and M. Kakonen (eds.)]. Earthscan Publications, London, UK, pp. 283-306.
Lee, E.K.O., C. Shen, and T.V. Tran, 2009: Coping with Hurricane Katrina: psychological
distress and resilience among African Americans evacuees. Journal of Black
Psychology, 35(1), 5-23.
Lehner, B., P. Döll, J. Alcamo, T. Henrichs, and F. Kaspar, 2006: Estimating the impact
of global change on flood and drought risks in Europe: A continental, integrated
analysis. Climatic Change, 75(3), 273-299.
Leichenko, R.M. and K.L. O’Brien, 2008: Environmental Change and Globalization:
Double Exposures. Oxford University Press, New York, NY, 192 pp.
Leichenko, R. and W. Solecki, 2008: Consumption, inequity, and environmental
justice: the making of new metropolitan landscapes in developing countries.
Society and Natural Resources, 21(7), 611-624.
Chapter 8Toward a Sustainable and Resilient Future
479
Leiserowitz, A., 2006: Climate change risk perception and policy preferences: the
role of affect, imagery, and values. Climatic Change, 77(1-2), 45-72.
Leiserowitz, A.A., R.W. Kates, and T.M. Parris. 2005: Do global attitudes and behaviors
support sustainable development? Environment, 47(9), 22-38.
Leiserowitz, A.A., R. Kates, and T. M. Parris. 2006: Sustainability values, attitudes,
and behaviors: A review of multinational and global trends. Annual Review of
Environment and Resources, 31, 413-444.
Lemos, M.C., E. Boyd, E.L. Tompkins, H. Osbahr, and D. Liverman, 2007: Developing
adaptation and adapting development. Ecology and Society, 12(2), 26.
Lempert, R., 2007: Can scenarios help policymakers be both bold and careful? In:
Blindside: How to Anticipate Forcing Events and Wild Cards in Global Politics
[Fukuyama, F. (ed.)]. Brookings Institution Press, Washington, DC, pp. 109-179.
Lempert, R.J. and M.T. Collins, 2007: Managing the risk of uncertain thresholds
responses: comparison of robust, optimum, and precautionary approaches. Risk
Analysis, 27(4), 1009-1026.
Lenton, T.M., H. Held, E. Kriegler, J.W. Hall, W. Lucht, S. Rahmstorf, and H.J.
Schellnhuber, 2008: Tipping elements in the Earth’s climate system. Proceedings
of the National Academy of Sciences, 105(6), 1786-1793.
Levin, S.A., 1998: Ecosystems and the biosphere as complex adaptive systems.
Ecosystems, 1(5), 431-436.
Lin, J., 2009: Supporting adaptation in developing countries at the national and
global levels. In: Climate Law and Developing Countries: Legal and Policy
Challenges for the World Economy [Richardson, B.J., Y. le Bouthillier, H. McLeod-
Kilmurray, and S. Wood (eds.)]. Edward Elgar, Cheltenham, UK, pp. 127-150.
Lindley, S., D. McEvoy, and J. Handley, 2006: Adaptation and mitigation in urban
areas: synergies and conflicts. Municipal Engineer, 159(4), 185-191.
Linnenluecke, M. and A. Griffiths, 2010: Beyond adaptation: resilience for business in
light of climate change and weather extremes. Business & Society, 49(3), 477-511.
Linnerooth-Bayer, J. and R. Mechler, 2007: Disaster safety nets for developing
countries: extending public-private partnerships. Environmental Hazards, 7(1),
54-61.
Linnerooth-Bayer, J., R. Mechler, and G. Pflug, 2005: Refocusing Disaster Aid.
Science, 309(5737), 1044-1046.
List, D., 2006: Action Research cycles for multiple futures perspectives. Futures,
38(6), 673-684.
Loayza, N., E. Olaberria, J.Rigolini, and L. Christiansen, 2009:
Natural Disasters and
Growth – Going Beyond the Averages. World Bank Policy Research Working
Paper 4980, World Bank, Washington, DC, 42 pp.
Lobell, D.B., M.B. Burke, C. Tebaldi, M.D. Mastrandrea, W.P. Falcon, and R.L. Naylor,
2008: Prioritizing climate change adaptation needs for food security in 2030.
Science, 319(5863), 607-610.
Locatelli, B., V. Evans, A. Wardell, A. Andrade, and R. Vignola, 2010: Forests and
Climate Change in Latin America Linking Adaptation and Mitigation in Projects
and Policies. CIFOR Infobrief No. 31, Center for International Forestry Research,
Bogor, Indonesia, 8 pp.
Loorbach, D., 2010: Transition management for sustainable development: a
prescriptive, complexity-based governance framework. Governance, 23(1),
161-183.
Loorbach, D., R. van der Brugge, and M. Taanman, 2008: Governance in the energy
transition: practice of transition management in the Netherlands. International
Journal of Environmental Technology and Management, 9(2/3), 294-315.
López, R., 2009: Natural Disasters and the Dynamics of Intangible Assets. Policy
Research Working Paper 4874, Background paper for the UN–World Bank
Assessment on the Economics of Disaster Risk Reduction, World Bank,
Washington, DC, 79 pp.
Lovins, A. and H. Lovins, 1982: Brittle Power: Energy Strategy for National Security.
Brick House, Andover, MA, 486 pp.
Lyons, M., T. Schilderman, and C. Boano (eds.), 2010: Building Back Better: Delivering
People-Centred Reconstruction at Scale. Practical Action Publishing, Rugby, UK,
380 pp.
Magat, W., K.W. Viscusi, and J. Huber, 1987: Risk-dollar tradeoffs, risk perceptions,
and consumer behaviour. In: Learning About Risk [Viscusi, W. and W. Magat
(eds.)]. Harvard University Press, Cambridge, MA, pp. 83-97.
Mahul, O. and C.J. Stutley, 2010: Government Support to Agricultural Insurance:
Challenges and Options for Developing Countries. World Bank, Washington, DC,
19 pp.
Markard, J, and B. Truffer, 2008: Technological innovation systems and the multi-level
perspective: towards an integrated framework. Research Policy, 37(4), 596-615.
Marulanda, M.C., O.D. Cardona, and A.H. Barbat, 2010: Revealing the socioeconomic
impact of small disasters in Colombia using the DesInventar database.
Disasters, 34(2), 552-570.
Maskrey, A., 1989: Disaster Mitigation: A Community Based Approach. Oxfam,
Oxford, UK, 114 pp.
Maskrey, A., 1994: Comunidad y desastres en América Latina: Estrategias de
Intervención. In: Viviendo en Riesgo: Comunidades vulnerables y prevención de
desastres en América Latina [Lavell, A. (ed.)]. Red de Estudios Sociales en
Prevención de Desastres en América Latina, Bogota, Colombia, pp. 14-38.
Maskrey, A. (ed.), 1996: Terremotos en el tropico humedo:La gestion de los
desastres del Alto Mayo, Peru (1990-1991), Limon, Costa Rica (1991) y Atrato
Medio, Colombia (1992). LA RED, Tercer Mundo, Bogota, Colombia, 256 pp.
Maskrey, A., 2011: Revisiting community-based disaster risk management.
Environmental Hazards, 10(1), 42-52.
McCool, D., 2005: The river commons: a new era in U.S. water policy. Texas Law
Review, 83(7), 1903-1927.
McEvoy, D., S. Lindley, and J. Handley, 2006: Adaptation and mitigation in urban
areas: synergies and conflict. Municipal Engineer, 159(4), 185-191.
McGray, H., A. Hammill, R. Bradley, E.L. Schipper, and J.-E. Parry, 2007: Weathering
the Storm: Options for Framing Adaptation and Development. World Resources
Institute, Washington, DC, 57 pp.
MEA, 2005: Ecosystems and Human Well-being A Framework for Assessment.
Millennium Ecosystem Assessment Series, Island Press, Wash ington, DC, 266 pp.
Meadowcroft, J., 1997: Planning for sustainable development: insights from the
literatures of political science. European Journal of Political Research, 31(4),
427-454.
Medema, W., B.S. McIntosh, and P.J. Jeffrey, 2008: From premise to practice: a critical
assessment of integrated water resources management and adaptive
management approaches in the water sector. Ecology and Society, 13(2), 29.
Mercer, J., I. Kelman, K. Lloyd, and S. Suchet-Pearson, 2008: Reflections on use of
participatory research for disaster risk reduction. Area, 40(2), 172-183.
Metcalf & Eddy, 2005: Water Reuse: Issues, Technologies and Applications.
Mcgraw-Hill, New York, NY, 1570 pp.
Mezirow, J., 1995: Transformation theory in adult learning. In: In Defense of the Life
World [Welton, M.R. (ed.)]. State University of New York Press, Albany, NY, pp.
39-70.
Mgbako, C. and L.A. Smith, 2010: Sex work and human rights in Africa. Fordham
International Law Journal, 33(4), 1178-1220.
Michel-Kerjan, E., 2008: Disasters and public policy: Can market lessons help
address government failures? In: Proceedings of the 99th National Tax
Association Conference, Boston, MA, 16-18 Nov. 2006, 15 pp.
Midgley, J. J. and W. J. Bond, 2001: A synthesis of the demography of African
acacias. Journal of Tropical Ecology, 17(6), 871-886.
Miller, R., 2007: Futures literacy: A hybrid strategic scenario method. Futures, 39(4),
341-362.
Milligan, J., T. O’Riordan, S.A. Nicholson-Cole, and A.R. Watkinson, 2009: Nature
conservation for future sustainable shorelines: lessons from seeking to involve
the public. Land Use Policy, 26(2), 203-213.
Misselhorn, A.A., 2005: What drives food insecurity in southern Africa? A meta-
analysis of household economy studies. Global Environmental Change A, 15(1),
33-43.
Mitchell, J.K., 1999: Crucibles of Hazard: Mega-Cities and Disasters in Transition.
UNU Press, New York, NY, 535 pp.
Mitchell, J.K., 2001: Policy forum: human rights to disaster assistance and mitigation.
Environmental Hazards, 3(3-4), 123-124.
Mitchell, T., K. Haynes, N. Hall, W. Choong, and K. Oven, 2008: The role of children
and youth in communicating disaster risk. Children, Youth and Environments,
18(1), 254-279.
Chapter 8 Toward a Sustainable and Resilient Future
480
Moench, M. and A. Dixit, 2007: Working with the Winds of Change: Toward
Strategies for Responding to the Risks Associated with Climate Change and
other Hazards. Institute for Social and Environmental Transition, Kathmandu,
Nepal, 296 pp.
Moench, M., A. Dixit, S. Janakarajan, M.S. Rathore, and S. Mudrakartha, 2003:
The Fluid Mosaic: Water Governance in the Context of Variability, Uncertainty
and Change. Institute for Social and Environmental Transition, Boulder, CO,
71 pp.
Mol, A.P.J. and D.A. Sonnenfeld (eds.), 2000: Ecological Modernization around the
World.Frank Cass Publishers, Portland, OR, 312 pp.
Montgomery, S., M. Lucotte, and I. Rheault, 2000: Temporal and spatial influences
of flooding on dissolved mercury in boreal reservoirs. The Science of the Total
Environment, 260(1-3), 147-157.
Moreno, A. and O.D. Cardona, 2011: Efectos de los desastres naturales sobre el
crecimiento, el desempleo, la inflación y la distribución del ingreso: Una
evaluación de los casos de Colombia y México. Background Paper prepared
for the 2011 Global Assessment Report on Disaster Risk Reduction, United
Nations International Strategy for Disaster Reduction, Geneva, Switzerland,
30 pp.
Mortreux, C. and J. Barnett, 2009: Climate change, migration and adaptation in
Funafuti, Tuvalu. Global Environmental Change, 19(1), 105-112.
Moser, S., 2009: Now more than ever: The need for more socially relevant research
on vulnerability and adaptation to climate change. Applied Geography, 30(4),
464-474.
Moser, S. and J.A. Ekstrom, 2010: A framework to diagnose barriers to climate
change adaptation. Proceedings of the National Academy of Sciences, 107(51),
22026-22031.
Moss, R.H., J.A. Edmonds, K.A. Hibbard, M.R. Manning, S.K. Rose, D.P. van Vuuren, T.R.
Carter, S. Emori, M. Kainuma, T. Kram, G.A. Meehl, J.F.B. Mitchell, N. Nakicenovic,
K. Riahi, S.J. Smith, R.J. Stouffer, A.M. Thomson, J.P. Weyant and T.J. Wilbanks,
2010: A new paradigm for the next generation of climate change scenarios.
Nature, 463, 747-756.
Murray, C. and A. Lopez (eds.), 1996: The Global Burden of Disease. Harvard
University Press, Cambridge, MA, 1022 pp.
Mustafa, D., 2005: The production of an urban hazardscape in Pakistan: modernity,
vulnerability and the range of choice. The Annals of the Association of American
Geographers, 95(3), 566-586.
Nakicenovic, N., J. Alcamo, G. Davis, B. de Vries, J. Fenhann, S. Gaffin, K. Gregory, A.
Grübler, T.Y. Jung, T. Kram, E.L. La Rovere, L. Michaelis, S. Mori, T. Morita, W.
Pepper, H. Pitcher, L. Price, K. Riahi, A. Roehrl, H-H. Rogner, A. Sankovski, M.
Schlesinger, P. Shukla, S. Smith, R. Swart, S. van Rooijen, N. Victor, and Z. Dadi,
2000: Special Report on Emissions Scenarios: A Special Report of Working
Group III of the Intergovernmental Panel on Climate Change. Cambridge
University Press, Cambridge, UK, 599 pp.
Nelson, D.R., 2009: Conclusions: transforming the world. In: Adapting to Climate
Change: Thresholds, Values, Governance [Adger, W.N., I. Lorenzoni, and K.L.
O’Brien (eds.)]. Cambridge University Press, Cambridge, UK, pp. 491-500.
Nelson, D.R., W.N. Adger, and K. Brown, 2007: Adaptation to environmental change:
contributions of a resilience framework. Annual Review of Environment and
Resources, 32, 395-419.
Nelson, R. and S. Winter, 1982: An Evolutionary Theory of Economic Change. Harvard
University Press, Cambridge, MA, 454 pp.
Nemarundwe, N., 2003: Negotiating Resource Access; Institutional Arrangement
for Woodlands and Water Use in Southern Zimbabwe. Doctoral Thesis,
University of Uppsala, Sweden, 279 pp.
Nepstad, D.C., I.M. Tohver, D. Ray, P. Moutinho, and G. Cardinot, 2007: Mortality of
large trees and lianas following experimental drought in an Amazon forest.
Ecology, 88(9), 2259-2269.
New, M., D. Liverman, H. Schroeder, and K. Anderson, 2011: Four degrees and
beyond: the potential for a global temperature increase of four degrees and its
implications. Philosophical Transactions of the Royal Society A, 369, 6-19.
Newman, P. and J.R. Kenworthy, 1989: Cities and Automobile Dependence: A
Sourcebook. Gower Publishing Company, Farnham, UK, 388 pp.
Newman, P.W.G., B. Birrell, D. Holmes, C. Mathers, P. Newton, G. Oakley, A. O’Connor,
B. Walker, A. Spessa, and D. Tait, 1996: Human settlements. In: Australian State
of the Environment Report [Beeton, R.J.S., K.I> Buckley, G.J. Jones, D. Morgan,
R.E. Reichelt, and D. Trewin (eds.)]. Department of Environment, Sport and
Territories, Canberra, pp. 7-18.
Nicholls, R., A. Watkinson, M. Mokrech, S. Hanson, J. Richards, J. Wright, S. Jude, S.
Nicholson-Cole, M. Walkden, J. Hall, R. Dawson, P. Stansby, G.K. Jacoub, M.
Rounsvell, C. Fontaine, L. Acosta, J. Lowe, J. Wolf, J. Leake, and M. Dickson, 2007:
Integrated coastal simulation to support shoreline management planning. In:
Proceedings of the 42nd Flood and Coastal Management Conference, York, UK,
3-5 July 2007, 11 pp.
Nicholls, R.J., S. Hanson, C. Herweijer, N. Patmore, S. Hallegatte, J. Chateau, and R.
Muir-Wood, 2008: Ranking Port Cities with High Exposure and Vulnerability to
Climate Extremes: Exposure Estimates. Environment Working Papers 1,
Organisation for Economic Co-operation and Development, Paris, France, 61 pp.
Njuki, J.M., M.T. Mapila, S. Zingore, and R. Delve, 2008: The dynamics of social capital
in influencing use of soil management options in the Chinyanja Triangle of
southern Africa. Ecology and Society, 13(2), 9.
Nobre, C.A., P. Sellers, and J. Shukla, 1991: Amazonian deforestation model and
regional climate change. Journal of Climate, 4, 957-988.
Nobre, C.A., E.D. Assad, and M.D. Oyama, 2005: Mudança ambiental no Brasil: o
impacto do aquecimento global nos ecossistemas da Amazônia e na agricultura.
Scientific American Brasil, Special Issue: A Terra na Estufa, 70-75.
Norris, F.H., 2005: Range, Magnitude, and Duration of the Effects of Disasters on
Mental Health: Review Update 2005. Dartmouth Medical School and National
Center for PTSD, Hanover, NH, 23 pp.
Norris, F., 2010: Behavioral Science Perspectives on Resilience. CARRI Research
Paper 10, Community and Regional Resilience Institute, Oak Ridge, TN, 50 pp.
Norris, F.H., M. J. Friedman, P.J. Watson, C.M. Byrne, E. Diaz, and K. Kaniasty, 2002:
60,000 disaster victims speak: Part I. An empirical review of the empirical
literature, 1981-2001. Psychiatry, 65(3), 207-239.
Noy, I., 2009: The macroeconomic consequences of disasters. Journal of Development
Economics, 88(2), 221-231.
Noy, I. and A. Nualsri, 2007: What do Exogenous Shocks Tell Us about Growth
Theories? Working Paper 07-28, University of Hawaii at Manoa, Department of
Economics, Manoa, HI, 31 pp.
Noy, I. and T.B. Vu, 2010: The economics of natural disasters in a developing country:
The case of Vietnam. Journal of Asian Economics, 21(4), 345-354.
NRC, 1999: Our Common Journey: A Transition Towards Sustainability. National
Research Council, National Academies Press, Washington, DC, 384 pp.
NRC, 2005: Thinking Strategically: The Appropriate Use of Metrics for the Climate
Change Science Program. National Research Council, National Academies
Press, Washington, DC, 162 pp.
NRC, 2008: Disaster Risk Management in an Age of Climate Change. A Summary of
the April 3, 2008 Workshop of the Disasters Roundtable [Anderson, W.A. (ed.)],
National Research Council, National Academies Press, Washington, DC, 18 pp.
NRC, 2010: Adapting to the Impacts of Climate Change. Panel on Adapting to
Impacts of Climate Change, National Research Council, National Academies
Press, Washington, DC, 292 pp.
O’Brien, G., P. O’Keefe, J. Rose, and B. Wisner, 2006: Climate change and disaster
management. Disasters, 30(1), 64-80.
O’Brien, K., 2009: Do values subjectively define the limits to climate change
adaptation? In: Adapting to Climate Change: Thresholds, Values, Governance
[Adger, W.N., I. Lorenzoni, and K. O’Brien (eds.)]. Cambridge University Press,
Cambridge, UK, pp. 164-180.
O’Brien, K. and R. Leichenko, 2003: Winners and losers in the context of global
change. Annals of the Association of American Geographers, 93(1), 89-103.
O’Brien, K. and J. Wolf. 2010: A values-based approach to vulnerability and adaptation
to climate change. Wiley Interdisciplinary Reviews: Climate Change, 1(2), 232-242.
O’Brien, K.L., R. Leichenko, U. Kelkar, H. Venema, G. Aandahl, H. Tompkins, A. Javed,
S. Bhadwal, S. Barg, L. Nygaard, and J. West, 2004: Mapping vulnerability to
multiple stressors: climate change and globalization in India. Global
Environmental Change, 14(4), 303-313.
Chapter 8Toward a Sustainable and Resilient Future
481
O’Brien, K. L., S. Eriksen, L. Sygna, and L.O. Næss, 2006: Questioning complacency:
climate change impacts, vulnerability, and adaptation in Norway. Ambio,
35(2), 50-56.
O’Brien, K., L. Sygna, R. Leichenko, W.N. Adger, J. Barnett, T. Mitchell, L. Schipper, T.
Tanner, C. Vogel, and C. Mortreux, 2008: Disaster Risk Reduction, Climate
Change Adaptation and Human Security. A Commissioned Report for the
Norwegian Ministry of Foreign Affairs, GECHS Report 2008:3, University of
Oslo, Oslo, Norway, 76 pp.
O’Brien, K., B. Hayward, and F. Berkes, 2009: Rethinking social contracts: building
resilience in a changing climate. Ecology and Society, 14(2), 12.
O’Brien, K., A. St Clair, and B. Kristoffersen (eds.), 2010a: Climate Change, Ethics and
Human Security. Cambridge University Press, Cambridge, UK, 246 pp.
O’Brien, K., A. St Clair, and B. Kristoffersen, 2010b: The framing of climate change:
Why it matters. In: Climate Change, Ethics and Human Security [O’Brien, K., A.
St Clair, and B. Kristoffersen (eds.)]. Cambridge University Press, Cambridge, UK,
pp. 3-22.
Obrist, B., C. Pfeiffer, and R. Henley, 2010: Multi-layered social resilience: a new
approach in mitigation research. Progress in Development Studies, 10(4),
283-293.
OECD, 2006: Declaration On Integrating Climate Change Adaptation into
Development Cooperation. Organisation for Economic Co-operation and
Development, Paris, France, 7 pp.
OECD, 2007: Climate Change in the European Alps: Adapting Winter Tourism and
Natural Hazards Management [Agrawala, S. (ed.)]. Organisation for
Econonomic Co-operation and Development, Paris, France.
OECD, 2009a: Measuring and Fostering the Progress of Societies. Document prepared
for OECD Meeting of the Council at Ministerial Level, 27-28 May 2010.
C/MIN(2010)13, 12 May 2010, Organisation for Econonomic Co-operation and
Development, Paris, France, 17 pp.
OECD, 2009b: Innovation in Country Risk Management. Organisation for Economic
Co-operation and Development, Paris, France, 47 pp.
Okuyama, Y., 2004: Modeling spatial economic impacts of an earthquake: Input-
output approaches. Disaster Prevention and Management, 13(4), 297-306.
Oliver-Smith, A., 1994: Peru’s five hundred year earthquake: vulnerability in historical
context. In: Disasters, Development and Environment [Varley, A. (ed.)]. John
Wiley and Sons, London, UK, pp. 31-48.
Oliver-Smith, A., 2007: Communities after catastrophes: reconstructing the material,
reconstituting the social. In: Community Building in the Twenty First Century
[Hyland, S.E. (ed.)]. School of American Research Press, Santa Fe, NM, pp.
49-75.
Olson, R.S., 2000: Toward a politics of disaster losses, values, agendas, and blame.
International Journal of Mass Emergencies and Disasters, 18(2), 265-287.
Olson, R.S. and V. Gawronski, 2003: Disasters as ‘critical junctures’ Managua,
Nicaragua 1972 and Mexico City 1985. International Journal of Mass
Emergencies and Disasters
, 21(1), 5-35.
Olsson, P., C. Folke, and T. Hahn, 2004: Social–ecological transformation for ecosystem
management: the development of adaptive co-management of a wetland
landscape in southern Sweden. Ecology and Society, 9(4), 2.
ONERC, 2009: Climate Change: Costs of Impacts and Lines of Adaptation. Report to
the Prime Minister and Parliament, Observatoire National sur les Effets du
Réchauffement Climatique, Paris, France, 136 pp.
Orlove, B., 2009: Glacier retreat: reviewing the limits of human adaptation to climate
change. Environment: Science and Policy for Sustainable Development, 51(3),
22-34.
Oswald Spring, Ú., 1991: Estrategias de Supervivencia en la Ciudad de México.
Centro Regional de Investigaciones Multidisciplinarias, Universidad Nacional
Autónoma de México, Cuernavaca, México, 219 pp.
Oswald Spring, Ú., 2009a: A HUGE gender security approach: towards human,
gender and environmental security. In: Facing Global Environmental Change:
Environmental, Human, Energy, Food, Health and Water Security Concepts
[Brauch, H.G., Ú. Oswald Spring, J. Grin, C. Mesjasz, P. Kameri-Mbote, N. Chadha
Behera, B. Chourou, and H. Krummenacher (eds.)]. Springer, Berlin, Germany,
pp.1165-1190.
Oswald Spring, Ú., 2009b: Food as a new human and livelihood security challenge.
In: Facing Global Environmental Change: Environmental, Human, Energy, Food,
Health and Water Security Concepts [Brauch, H.G., Ú. Oswald Spring, J. Grin, C.
Mesjasz, P. Kameri-Mbote, N. Chadha Behera, B. Chourou, and H.
Krummenacher (eds.)]. Springer, Berlin, Germany, pp. 471-500.
Oswald Spring, Ú., 2011a: Environmentally-forced migration in rural areas: security
risks and threats in Mexico. In: Climate Change, Human Security and Violent
Conflict: Challenges for Societal Stability [Scheffran, J., M. Brzoska, H.G. Brauch,
P.M. Link, and J. Schilling (eds.)]. Hexagon Series on Human and Environmental
Security and Peace 8, Springer, Heidelberg, Germany (in press).
Oswald Spring, Ú. (ed.), 2011b: Retos de la Investigación del Agua en México.
Centro Regional de Investigaciones Multidisciplinarias, Universidad Nacional
Autónoma de México / Scientific Network on Water – National Council on
Science and Technology, Cuernavaca, México.
Oswald Spring, Ú. and H.G. Brauch, 2011: Coping with global environmental
change – sustainability revolution and sustainable peace. In: Coping with
Global Environmental Change, Disaster and Security [Brauch, H.G., Ú. Oswald
Spring, C. Mesjasz, J. Grin, P. Kameri-Mbote, B. Chourou, P. Dunay, and J.
Birkmann (eds.)]. Springer, Berlin, Germany, pp. 1487-1504.
Owuor, B., W. Mauta, and S. Eriksen, 2011: Strengthening sustainable adaptation:
examining interactions between pastoral and agropastoral groups in dryland
Kenya. Climate and Development, 3(1), 42-58.
Oxfam, 2010: Gender, Disaster Risk Reduction and Climate Change Adaptation, A
Learning Companion. Oxfam GB, Oxford, UK, 15 pp.
Paavola, J., 2005: Seeking justice: international environmental governance and
climate change. Globalizations, 2(3), 309-322.
Paavola, J. and W.N. Adger, 2006: Fair adaptation to climate change. Ecological
Economics, 56(1), 594-609.
Paavola, J., W.N. Adger, and S. Huq, 2006: Multifaceted justice in adaptation to
climate change. In: Fairness in Adaptation to Climate Change [Adger, W.N., J.
Paavola, S. Huq, and M.J. Mace (eds.)]. MIT Press, Cambridge, MA, 263-277.
Pahl-Wostl, C., 2009: A conceptual framework for analysing adaptive capacity and
multi-level learning processes in resource governance regimes. Global
Environmental Change,
19(3), 354-365.
Pahl-Wostl, C., M. Craps, A. Dewulf, E. Mostert, D. Tabara, and T.Taillieu, 2007: Social
learning and water resources management. Ecology and Society, 12(2), 5.
Parry, M.L., O.F. Canziani, J.P. Palutikof and Co-authors, 2007: Technical summary. In:
Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of
Working Group II to the Fourth Assessment Report of the Intergovernmental Panel
on Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. van der Linden, and
C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK, pp. 23-78.
Patel, P. and K. Pavit, 1995: Patterns of technological activity: their measurement and
interpretation. In: Handbook of Economics of Innovation and Technological
Change [Stoneman, P. (ed.)]. Wiley-Blackwell, Oxford, UK, pp. 14-51.
Pearce, L., 2003: Disaster management and community planning, and public
participation: how to achieve sustainable hazard mitigation. Natural Hazards,
28(2-3), 211-228.
Peek, L., 2008: Children and disasters: understanding vulnerability, developing
capacities and promoting resilience – an introduction. Children, Youth and
Environments, 18(1), 1-29.
Peeples, M.A., C.M. Barton, and S. Schmich, 2006: Resilience lost: intersecting land
use and landscape dynamics in the prehistoric southwestern United States.
Ecology and Society, 11(2), 22.
Pelling, M., 1999: The political ecology of flood hazard in urban Guyana. Geoforum,
30, 249-261.
Pelling, M., 2003: Vulnerability of Cities, Natural Disasters and Social Resilience.
Earthscan, London, UK, 212 pp.
Pelling, M., 2007: Learning from others: scope and challenges for participatory
disaster risk assessment. Disasters, 31(4), 373-385.
Pelling, M., 2010a: Adaptation to Climate Change: From Resilience to
Transformation. Taylor and Francis, London, UK, 224 pp.
Pelling, M., 2010b: Systematisation Review of Urban Disaster Risk Reduction
Projects in the Caribbean Region. Oxfam, Oxford, UK, 90 pp.
Chapter 8 Toward a Sustainable and Resilient Future
482
Pelling, M. and K. Dill, 2009: Disaster politics: tipping points for change in the
adaptation of socio-political regimes. Progress in Human Geography, 34(1),
21-37.
Pelling, M. and A. Holloway, 2006: Legislation for Mainstreaming Disaster Risk
Reduction. Tearfund, London, UK, 36 pp.
Pelling, M. and D. Manuel-Navarrete, 2011: From resilience to transformation: the
adaptive cycle in two Mexican urban centers. Ecology and Society, 16(2), 11.
Pelling, M. and J.I. Uitto, 2001: Small island developing states: natural disaster
vulnerability and global change. Environmental Hazards: Global Environmental
Change B, 3(2), 49-62.
Pelling, M. and B. Wisner (eds.), 2009: Disaster Risk Reduction: Cases from Urban
Africa. Earthscan, London, UK.
Pelling, M., A. Özerdem, and S. Barakat, 2002: The macro-economic impact of
disasters. Progress in Development Studies, 2(4), 283-305.
Pelling, M., C. High, J. Dearing, and D. Smith, 2007: Shadow spaces for social
learning: a relational understanding of adaptive capacity to climate change
within organizations. Environment and Planning A, 40(4), 867-884.
Penrose, A. and M. Takaki, 2006: Children’s rights in emergencies and disasters. The
Lancet, 367, 698-699.
Perez, C., 2002: Technological Revolutions and Financial Capital. The Dynamics of
Bubbles and Golden Ages. Edward Elgar, Cheltenham, UK, 198 pp.
Peterson, G., 2009: Ecological limits of adaptation to climate change. In: Adapting
to Climate Change: Thresholds, Values, Governance [Adger, W.N., I. Lorenzoni,
and K. O’Brien (eds.)]. Cambridge University Press, Cambridge, UK, pp. 25-41.
Peterson, G., G.A. De Leo, J.J. Hellmann, M.A. Janssen, A. Kinzig, J.R. Malcolm, K.L.
O’Brien, S.E. Pope, D.S. Rothman, E. Shevliakova, and R.R.T Tinch, 1997:
Uncertainty, climate change, and adaptive management. Conservation Ecology,
1(2), 4.
Pettengell, C., 2010: Climate Change Adaptation: Enabling People Living in Poverty
to Adapt. Oxfam International Research Report, Oxfam, Oxford, UK, 48 pp.
Pielke, Jr., R., J. Rubiera, C. Landsea, M. Fernández, and R. Klein, 2003: Hurricane
vulnerability in Latin America and the Caribbean: Normalized damage and loss
potentials. Natural Hazards Review, 4(3), 101-114.
Pielke, Jr., R., G. Prins, S. Rayner, and D. Sarewitz, 2007: Climate change 2007: lifting
the taboo on adaptation. Nature, 445, 597-598.
Pielke, Jr., R.A., J. Gratz, C.W. Landsea, D. Collins, M. Saunders, and R. Musulin,
2008: Normalized hurricane damages in the United States: 1900-2005. Natural
Hazards Review, 9(1), 29-42.
Pigou, A.C., 1920: The Economics of Welfare. Macmillan, London, UK, 976 pp.
Piguet, E., 2008: Climate Change and Forced Migration. Research Paper No 153,
Evaluation and Policy Analysis Unit, UN High Commissioner for Refugees,
Geneva, Switzerland, 15 pp.
Platt, R.H., 1999: Disasters and Democracy: the Politics of Extreme Natural Events.
Island Press, Washington, DC, 320 pp.
Plummer, R. and D.R. Armitage, 2007: Charting the new territory of adaptive co-
management: a Delphi study. Ecology and Society, 12(2), 10.
Quarantelli, E.L., 1984: Perceptions and reactions to emergency warnings of sudden
hazards. Ekistics, 309,511-515.
Quarantelli, E.L., 1995: What is a Disaster? Routledge, London, UK, 312 pp.
Raddatz, C., 2007: Are external shocks responsible for the instability of output in
low-income countries? Journal of Development Economics, 84(1), 155-187.
Raleigh, C., L. Jordan, and I. Salehyan, 2008: Assessing the Impact of Climate
Change on Migration and Conflict. World Bank, Washington DC, 57 pp.
Ramos, J.M., 2006: Dimensions in the confluences of futures studies and action
research. Futures, 38(6), 642-655.
Raven, R.P.J.M., S. van den Bosch, and R. Weterings, 2010: Transitions and strategic
niche management: towards a competence kit for practitioners. International
Journal of Technology Management, 51(1), 57-74.
Rawls, J., 1971: A Theory of Justice. Harvard University Press, Cambridge, MA,
607 pp.
Reaser, J.K., R. Pomerance, and P.O. Thomas, 2000: Coral bleaching and global climate
change: scientific findings and policy recommendations. Conservation Biology,
14(5), 1500-1511.
Redclift, M., 1987: Sustainable Development: Exploring the Contradictions.
Methuen, London, UK, 221 pp.
Redclift, M., 1992: The meaning of sustainable development. Geoforum, 25(3),
395-403.
Reeder, T., J. Wicks, L. Lovell, and O. Tarrant, 2009: Protecting London from tidal
flooding: limits to engineering adaptation. In: Adapting to Climate Change:
Thresholds, Values, Governance [Adger, W.N., I. Lorenzoni, and K. O’Brien (eds.)].
Cambridge University Press, Cambridge, UK, pp. 54-78.
Reid, H., 2006: Climatic change and biodiversity in Europe. Conservation and Society,
4(1), 84-101.
Repetto
, R., 2008: The Climate Crisis and the Adaptation Myth. Working Paper 13,
Yale School of Forestry and Environmental Studies, New Haven, CT, 24 pp.
Reuveny, R., 2007: Climate change induced migration and violent conflict. Political
Geography, 26(6), 656-673.
Reyos, J., 2010: Community-driven Disaster Intervention: Experiences of the
Homeless People’s Federation in the Philippines. Homeless Peoples Federation
Philippines, Philippines Action for Community-led Shelter Initiatives, and
International Institute for Environment and Development, Manila, Philippines
and London, UK, 70 pp.
Richardson, B.J., Y. Le Bouthillier, H. McLeod-Kilmurray, and S. Wood (eds.), 2009:
Climate Law and Developing Countries: Legal and Policy Challenges for the
World Economy. Edward Elgar, Cheltenham, UK, 448 pp.
Rip, A. and R. Kemp, 1998: Technological change. In: Human Choice and Climate Change
[Rayner, S. and L. Malone (eds.)]. Batelle Press, Washington, DC, pp. 327-399.
Rist, S., 2000: Linking ethics and market – Campesino economic strategies in the
Bolivian Andes. Mountain Research and Development, 20(4), 310-315.
Roberts, D., 2008: Thinking globally, acting locally – institutionalizing climate
change at the local government level in Durban, South Africa. Environment and
Urbanization, 20(2), 521-537.
Roberts, D., 2010: Prioritizing climate change adaptation and local level resilience
in Durban, South Africa. Environment and Urbanization, 22(2), 397-413.
Robinson, J., 2003: Future subjunctive: backcasting as social learning. Futures,
35(8), 839-856.
Rodriguez-Oreggia, E., A. de la Fuente, R. de la Torre, H. Moreno, and C. Rodriguez,
2009: The Impact of Natural Disasters on Human Development and Poverty at
the Municipal Level in Mexico. CID Working Paper 43, Center for International
Development at Harvard University, Cambridge, MA, 35 pp.
Rohan, M.J., 2000: A rose by any name? The values construct. Personal and Social
Psychology Review, 4(3), 255-277.
Rohracher, H., 2008: Energy systems in transition: contributions from social sciences.
International Journal of Environmental Technology and Management, 9(2-3),
144-161.
Rokeach, M., (ed.), 1979: Understanding Human Values: Individual and Societal. The
Free Press, New York, NY, 322 pp.
Rose, A. and S.-Y. Liao, 2005: Modeling regional economic resilience to disasters: A
computable general equilibrium analysis of water service disruptions. Journal of
Regional Science, 45(1), 75-112.
Rose, A., J. Benavides, S.E. Chang, P. Szczesniak, and D. Lim, 1997: The regional
economic impact of an earthquake: Direct and indirect effects of electricity
lifeline disruptions. Journal of Regional Science, 37(3), 437-458.
Rosegrant, M.W., T. Zhu, S. Msangi, and T. Sulser, 2011: Global scenarios for biofuels:
impacts and implications. Applied Economic Perspectives and Policy,
30(3),
595-505.
Rosenberg, S.W. (ed.), 2007: Deliberation, Participation and Democracy: Can the
People Govern? Palgrave Macmillan, New York, NY, 322 pp.
Rosenzweig, C., K.M. Strzepek, D.C. Major, A. Iglesias, D.N. Yates, A. McCluskey,
and D. Hillel, 2004: Water resources for agriculture in a changing climate:
international case studies. Global Environmental Change, 14(4), 345-360.
Roshetko, J.M., R.D. Lasco, and M.S.D. Angeles, 2007: Smallholder agroforestry
systems for carbon storage. Mitigation and Adaptation Strategies for Global
Change, 12, 219-242.
Royal Society, 2009: Geoengineering The Climate: Science, Governance and
Uncertainty. The Royal Society, London, UK, 98 pp.
Chapter 8Toward a Sustainable and Resilient Future
483
Sacoby, M., R.R. Wilson, J. Lesley, and E. Williams, 2010: Climate change,
environmental justice, and vulnerability: an exploratory spatial analysis.
Environmental Justice, 3(1), 13-19.
Salagnac, J.-L., 2007: Lessons from the 2003 heat wave: a French perspective.
Building Research and Information, 35(4), 450-457.
Santos, I., 2007: Disentangling the Effects of Natural Disasters on Children: 2001
Earthquakes in El Salvador. Doctoral Dissertation, Kennedy School of
Government, Harvard University, Cambridge, MA.
Satterthwaite, D., 1997: Sustainable cities or cities that contribute to sustainable
development? Urban Studies, 34(10), 1167-1691.
Satterthwaite, D., 1998: Meeting the challenge of urban disasters. In: World
Disasters Report 1998. International Federation of Red Cross and Red Crescent
Societies, Oxford University Press, Oxford, UK, pp. 9-19.
Satterthwaite, D. (ed.), 2007: The Transition to a Predominantly Urban World and
its Underpinnings. International Institute for Environment and Development,
London, UK, 91 pp.
Satterthwaite, D., 2010: Slum upgrading. Economic and Political Weekly, 45(10),
12-16.
Schär, C., P.L. Vidale, D. Lüthi, C. Frei, C.Häberli, M.A. Liniger, and C. Appenzeller,
2004: The role of increasing temperature variability in European summer
heatwaves. Nature, 427, 332-336.
Scharmer, C.O., 2009: Theory U: Leading from the Future as it Emerges. The Society
for Organizational Learning, Cambridge, MA, 533 pp.
Scheffer, M., 2009. Critical Transitions in Nature and Society. Princeton University
Press, Princeton, NJ, 400 pp.
Scheffer, M., S.R. Carpenter, J.A. Foley, C. Folke, and B. Walker, 2001: Catastrophic
shifts in ecosystems. Nature, 413, 591-596.
Scheffer, M., J. Bascompte, W.A. Brock, V. Brovkin, S.R. Carpenter, V. Dakos, H. Held,
E.H. van Nes, M. Rietkerk, and G. Sugihara, 2009: Early-warning signals for
critical transitions. Nature, 461, 53-59.
Scheffran, J., 2011: Security risks of climate change: vulnerabilities, threats, conflicts
and strategies. In: Coping with Global Environmental Change, Disaster and
Security [Brauch, H.G., Ú. Oswald Spring, C. Mesjasz, J. Grin, P. Kameri-Mbote, B.
Chourou, P. Dunay, and J. Birkmann (eds.)]. Springer, Berlin, Germany, pp. 735-756.
Schipper, E.L.F., 2007: Climate Change Adaptation and Development: Exploring the
Linkages. Working Paper 107, Tyndall Centre for Climate Change Research,
Norwich, UK, 20 pp.
Schipper, L., 2009: Meeting at the crossroads? Exploring the linkages between climate
change adaptation and disaster risk reduction. Climate and Development, 1(1),
16-30.
Schipper, L. and M. Pelling, 2006: Disaster risk, climate change and international
development: scope and challenges for integration. Disasters, 30(1), 19-38.
Schlenker, W. and M.J. Roberts, 2006: Nonlinear effects of weather on corn yields.
Review of Agricultural Economics, 28(3), 391-398.
Scholze, M., W. Knorr, N.W. Arnell, and I.C. Prentice, 2005: A climate change risk
analysis for world ecosystems. Proceedings of the National Academy of
Sciences, 103, 13116-13120.
Schumacher, I. and E. Strobl, 2008: Economic Development and Losses due to
Natural Disasters: The Role of Risk. Cahiers 2008-32, Département d’Economie,
Ecole Polytechnique, Palaiseau, France, 39 pp.
SCBD, 2010: Global Biodiversity Outlook 3. Secretariat of the Convention on
Biological Diversity, Montréal, Canada.
Scott, J.C., 2003: The Moral Economy of the Peasant: Rebellion and Subsistence in
Southeast Asia. Yale University Press, New Haven, CT, 246 pp.
Sen, A, 2000: Un Nouveau Modèle Économique. Développement, Justice, Liberté.
Editions Odile Jacob, Paris, France, 356 pp.
Sen, A, 2003: Human security now. Soka Gakkai International Quarterly, 33, 1-7.
Senge, P.M., 1990: The Fifth Discipline: The Art and Practice of the Learning
Organization. Currency/Doubleday, New York, NY, p. 445.
Silvestri, S. and F. Kershaw (eds.), 2010: Framing the flow: Innovative Approaches to
Understand, Protect and Value Ecosystem Services across Linked Habitats.
United Nations Environment Programme World Conservation Monitoring
Centre, Cambridge, UK, 66 pp.
Simmie, J. and R. Martin. 2010: The economic resilience of regions: towards an
evolutionary approach. Economy and Society, 3(1), 27-43.
Skidmore, M. and H. Toya, 2002: Do natural disasters promote long-run growth?
Economic Inquiry, 40(4), 664-687.
Slovic, P., 1987: Perception of risk. Science, 236, 280-285.
Smith, A. and A. Stirling, 2010: The politics of social–ecological resilience and
sustainable sociotechnical transitions. Ecology and Society, 15(1), 11.
Smith, A., A. Stirling, and F. Berkhout, 2005: The governance of sustainable socio-
technical transitions: a quasi-evolutionary model. Research Policy, 34(10),
1491-1510.
Smith, D.M. and S. Barchiesi, 2009: Environment as Infrastructure: Resilience to
Climate Change impacts on Water through Investments in Nature. International
Union for Conservation of Nature, Gland, Switzerland, 13 pp.
Smith
, S.K. and C. McCarty, 2006: Florida’s 2004 Hurricane Season: Demographic
Response and Recovery. Bureau of Economic and Business Research, University
of Florida, Gainesville, FL, 40 pp.
Smithers, J. and B. Smit, 1997: Agricultural system response to environmental stress.
In: Agricultural Restructuring and Sustainability: A Geographical Perspective
[Ilbery, B., Q. Chiotti, and T. Rickard (eds.)]. CABI, Wallingford, UK, pp. 167-183.
Solecki, W. and R. Leichenko, 2006: Urbanization and the metropolitan environment:
lessons from New York and Shanghai. Environment, 48(4), 8-23.
Sperling, F. and F. Szekely, 2005: Disaster Risk Management in Changing Climate.
Discussion Paper prepared for the World Conference on Disaster Risk Reduction
on behalf of the Vulnerability and Adaptation Resource Group (VARG). Reprint
with Addendum on Conference Outcomes, VARG, Washington, DC, 45 pp.
Sperling, F., C. Valdivia, R. Quiroz, R. Valdivia, L. Angulo, A. Seimon, and I. Noble,
2008: Transitioning to Climate Resilient Development: Perspectives from
Communities in Peru. World Bank Environment Department Papers 115, World
Bank, Washington, DC, 103 pp.
Sphere, 2004: Humanitarian Charter and Minimum Standards in Humanitarian
Response. The Sphere Project, Belmont Press Ltd., Northampton, UK, 402 pp.
St. Clair, A., 2010: Global poverty: towards the responsibility to protect. In: Climate
Change, Ethics and Human Security [O’Brien, K., A. St. Clair, and B. Kristoffersen
(eds.)]. Cambridge University Press, Cambridge, UK, pp. 180-198.
Stafford Smith, M., L. Horrocks, A. Harvey, and C. Hamilton, 2011: Rethinking
adaptation for a 4°C world. Philosophical Transactions of the Royal Society A,
369, 196-216.
Stehr, N. and H. Von Storch, 2005: Introduction to papers on mitigation and
adaptation strategies for climate change: protecting nature from society or
protecting society from nature? Environmental Science Policy, 8(6), 537-540.
Stern, P. and W. Easterling (eds.), 1999: Making Use of Seasonal Climate Forecasts.
National Academies Press, Washington, DC, 175 pp.
Stevenson, T., 2002: Anticipatory action learning: conversations about the future.
Futures, 34(5), 417- 425.
Stevenson, T., 2008: Enacting the vision for sustainable development. Futures,
41(4), 246-252.
Strobl, E., 2011: The economic growth impact of hurricanes: evidence from US
coastal counties. The Review of Economics and Statistics, 93(2), 575-589.
Swart, R., J. Robinson, and S. Cohen, 2003: Climate change and sustainable
development: expanding the options. Climate Policy, 391, 810-849.
Swift, J., 1989: Why are rural people vulnerable to famine? IDS Bulletin, 20(2), 8-15.
Tanner, T.M., 2010: Shifting the narrative: child-led responses to climate change
and disasters in El Salvador and the Philippines. Children and Society, 24(4),
339-351.
Tanner, T.M., M. Garcia, J. Lazcano, F. Molina, G. Molina, G. Rodríguez, B. Tribunalo,
and F. Seballos, 2009: Children’s participation in community-based disaster risk
reduction and adaptation to climate change. Participatory Learning and Action,
60, 54-64.
Telford, J., J. Cosgrave, and R. Houghton, 2006: Joint Evaluation of the International
Response to the Indian Ocean Tsunami: Synthesis Report. Tsunami Evaluation
Coalition, London, UK, 178 pp.
Thaler, R., 1999: Mental accounting matters. Journal of Behavioral Decision Making,
12, 183-206.
Chapter 8 Toward a Sustainable and Resilient Future
484
Theobald, D., 2005: Landscape patterns of exurban growth in the USA from 1980 to
2020. Ecology and Society, 10(1), 32.
Thomas, D.S.G. and C. Twyman, 2005: Equity and justice in climate change adaptation
amongst natural-resource dependent societies. Global Environmental Change
Part A, 15(2), 115-124.
Thomas-Slayter, B., R. Polestico, A.L. Esser, O. Taylor, and E. Mutua, 1995: A Manual
for Socio-Economic and Gender Analysis: Responding to the Development
Challenge. Clark ECOGEN Publication, Worcester, MA, 278 pp.
Thompson, I., B. Mackey, S. McNulty, and A. Mosseler, 2009: Forest Resilience,
Biodiversity, and Climate Change. A Synthesis of the Biodiversity/Resilience/
Stability Relationship in Forest Ecosystems. Technical Series no. 43, Secretariat
of the Convention on Biological Diversity, Montreal, Canada, 67 pp.
Tierney, K. and M. Bruneau, 2007: Conceptualizing and measuring resilience: A key
to disaster loss reduction. TR News 250, May-June, 14-17.
Tol, R. and G. Yohe, 2007: The weakest link hypothesis for adaptive capacity: An
empirical test. Global Environmental Change, 17(2), 218-227.
Tol, R.S.J., T.E. Downing, O.J. Kuik, and J.B. Smith: 2004. Distributional aspects of
climate change impacts. Global Environmental Change, 14(3), 259-272.
Tompkins, E.L., M.C. Lemos, and E. Boyd, 2008: A less disastrous disaster: Managing
response to climate-driven hazards in the Cayman Islands and NE Brazil. Global
Environmental Change, 18(4), 736-745.
Törnquist, O., S.A. Prasetyo, and T. Birks (eds.), 2010: Aceh: The Role of Democracy
for Peace and Reconstruction. PCD Press Indonesia, Jogjakarta, Indonesia,
431 pp.
Tran, P., R. Shaw, G. Chantry, and J. Norton, 2009: GIS and local knowledge in disaster
management: a case study of flood risk mapping in Viet Nam. Disasters, 33(1),
152-169.
Trawick, P., 2001: The moral economy of water: equity and antiquity in the Andean
commons. American Anthropologist, 103(2), 361-379.
Tribe, M. and A. Lafon, 2009: After 2015: Promoting Pro-Poor Policy after the MDGS:
The Plenary Presentations and Discussion. Report on the EADI – DSA – IDS–
ActionAid – DIFID High Level Policy Forum, June 2009, 17 pp.
Trope, Y. and N. Liberman, 2003: Temporal construal. Psychological Review, 110(3),
403-421.
Tschakert, P. and K.A. Dietrich, 2010: Anticipatory learning for climate change
adaptation and resilience. Ecology and Society, 15(2), 11.
Tschakert
, P., R. Sagoe, G. Darko, and S.N. Codjoe, 2010: Floods in the Sahel: an
analysis of anomalies, memory, and anticipatory learning. Climatic Change,
103(3-4), 471-502.
Tsuchiya, S., H. Tatano, and N. Okada, 2007: Economic loss assessment due to railroad
and highway disruptions. Economic Systems Research, 19(2), 147–162.
Turner, N.J. and H. Clifton, 2009: It’s so different today: climate change and
indigenous lifeways in British Columbia. Global Environmental Change, 19(2),
180-190.
Tversky, A. and E. Shafir, 1992: Choice under conflict: the dynamics of deferred
decision. Psychological Science, 3(6), 358-361.
Twigg, J., 2001: Corporate Social Responsibility and Disaster Reduction: A Global
Overview. Benfield Greig Hazard Research Centre, University College London,
UK, 84 pp.
Tyler, S., S.O. Reed, K. Macclune, and S. Chopde, 2010: Planning for Urban Climate
Resilience: Framework and Examples from the Asian Cities Climate Change
Resilience Network (ACCCRN). Climate Resilience in Concept and Practice
Working Paper Series 3, Institute for Social and Environmental Transition,
Boulder, CO, 168 pp.
UK Environment Agency, 2009: The Thames estuary is changing. In: Managing
Flood Risk through London and the Thames Estuary. TE2100 Plan Consultation
Document, Environment Agency, London, UK, pp. 20-23.
Ulsrud, K., L. Sygna, and K.L. O’Brien, 2008: More than Rain, Identifying Sustainable
Pathways for Climate Adaptation and Poverty Reduction. Report prepared for
the Development Fund, Oslo, Norway, 68 pp.
UN, 2002: World Population Ageing 1950 – 2050. Executive Summary. United
Nations, Department of Economic and Social Affairs, Population Division, New
York, NY.
UN-HABITAT, 2009: Planning Sustainable Cities: Global Report on Human
Settlements 2009. Earthscan Publications Ltd., London, UK, 306 pp.
UNDP, 1994: Human Development Report 1994: New Dimensions of Human
Security. United Nations Development Programme, Oxford University Press,
New York, NY.
UNDP, 2004: Reducing Disaster Risk: A Chal lenge for Development. United Nations
Development Programme, Oxford University Press, New York, NY, 161 pp.
UNDP, 2007a: Thailand Human Development Report 2007: Sufficiency Economy and
Human Development. United Nations Development Programme, Bangkok,
Thailand, 152 pp.
UNDP, 2007b: Human Development Report 2007/2008: Fighting Climate Change:
Human Solidarity in a Divided World. United Nations Development Programme,
Palgrave Macmillan, New York, NY, 399 pp.
UNDP, 2009: Human Development Report 2009: Overcoming Barriers: Human
Mobility and Development. United Nations Development Programme, Oxford
University Press, Oxford, UK.
UNFCCC, 2006:
Technologies for Adaptation to Climate Change. Adaptation,
Technology and Science Programme of the United Nations Framework
Convention on Climate Change Secretariat, Bonn, Germany, 40 pp.
UNGA, 2009: General Assembly, Expressing Deep Concern, Invites Major United
Nations Organs: to Intensify Efforts in Addressing Security Implications of
Climate Change. GA/10830, 3 June 2009, United Nations General Assembly,
New York, NY.
UNISDR, 2004: Living with Risk. A Global Review of Disaster Reduction Initiatives.
United Nations International Strategy for Disaster Reduction Secretariat,
Geneva, Switzerland, 429 pp.
UNISDR, 2008: Gender Perspectives: Integrating Disaster Risk Reduction into
Climate Change Adaptation. Good practices and Lessons Learned. United
Nations International Strategy for Disaster Reduction, Geneva, Switzerland,
87 pp.
UNISDR, 2009: Global Assessment Report on Disaster Risk Reduction. United
Nations International Strategy for Disaster Reduction Secretariat, Geneva,
Switzerland, 207 pp.
UNISDR, 2011: Global Assessment Report on Disaster Risk Reduction. United
Nations International Strategy for Disaster Reduction, Geneva, Switzerland.
UNSC, 2007: Security Council Holds First-Ever Debate on Impact of Climate Change
on Peace, Security, Hearing over 50 Speakers. UN Security Council, 5663rd
Meeting, 17 April 2007. UN Security Council, New York, NY.
UNSG, 2009: Climate Change and its Possible Security Implications. Report of the
Secretary-General. United Nations, New York, NY.
USGBC, 2010: USGBC Celebrates Five Years of Green Building, Economic &
Educational Progress in New Orleans. Press Release, US Green Building Council,
Washington, DC.
van den Berg, M., 2010: Household income strategies and natural disasters:
dynamic livelihoods in rural Nicaragua. Ecological Economics, 69(3), 592-602.
van Vuuren, D.P., P.L. Lucas, and H. Hilderink, 2006: Downscaling drivers of global
environmental change scenarios: enabling use of the IPCC-SRES scenarios at
the national and grid level. Global Environmental Change, 17(1), 114-130.
Velásquez, L.S., 1998: Agenda 21: A form of joint environmental management in
Manizales, Colombia. Environment and Urbanization, 10(2), 9-36.
Velásquez, L.S., 2005: The bioplan: decreasing poverty in Manizales, Colombia,
through shared environmental management. In: Reducing Poverty and
Sustaining the Environment [Bass, S., H. Reid, D. Satterthwaite, and P. Steele
(eds.)]. Earthscan Publications, London, UK, pp. 44-72.
Verchot, L.V., M. van Noordijk, S. Kandji, T. Tomich, C. Ong, A. Albrecht, J. Mackensen,
C. Bantilan, K.V. Anupama, and C. Palm, 2007: Climate change: linking adaptation
and mitigation through agroforestry. Mitigation and Adaptation Strategies for
Global Change, 12(5), 901-918.
Vergara, W. and S.M. Scholz (eds.), 2011: Assessment of the Risk of Amazon
Dieback. World Bank, Washington, DC, 99 pp.
Verweij, M., M. Douglas, R. Ellis, C. Engel, F. Hendriks, S. Lohmann, S. Ney, S. Rayner,
and M. Thompson, 2006: Clumsy solutions for a complex world: the case of
climate change. Public Administration, 84(4), 817-843.
Chapter 8Toward a Sustainable and Resilient Future
485
Villagrán de León, J.C., 2011: Risks in Central America: bringing them under
control. In: Coping with Global Environmental Change, Disaster and Security
[Brauch, H.G., Ú. Oswald Spring, C. Mesjasz, J. Grin, P. Kameri-Mbote, B.
Chourou, P. Dunay, and J. Birkmann (eds.)]. Springer, Berlin, Germany, pp.
1147-1158.
Viratkapan, V. and R. Perera, 2006: Slum relocation projects in Bangkok: what has
contributed to their success or failure? Habitat International, 30(1),
157-174.
Virgoe, J., 2002: International governance of a possible geoengineering intervention
to combat climate change. Climatic Change, 95(1-2), 103-119.
Viscusi, W.K. and J. Aldy, 2003: The value of a statistical life: a critical review of
market estimates throughout the world. Journal of Risk and Uncertainty, 27(1),
5-76.
Vogel, C., S.C. Moser, R.E. Kasperson, and G. Dabelko. 2007: Linking vulnerability,
adaptation, and resilience science to practice: pathways, players, and partnerships.
Global Environmental Change, 17(3-4), 349-364.
Volkery, A. and T. Ribeiro, 2009: Scenario planning in public policy: understanding
use, impacts and the role of institutional context factors. Technological
Forecasting and Social Change, 76(9), 1198-1207.
von Hesse, M., J. Kamiche, and C. de la Torre, 2008: Contribución Temática de
America Latina al Informe Bienal de Evaluación Mundial Sobre la Reducción de
Riesgo 2009. Contribution to the GTC-UNDP Background Paper prepared for
the 2009 Global Assessment Report on Disaster Risk Reduction, United Nations
International Strategy for Disaster Reduction, Geneva, Switzerland, 131 pp.
Voss, M. and K. Wagner, 2010: Learning from (small) disasters. Natural Hazards, 55,
657-669.
Wack, P., 1985: Scenarios: shooting the rapids. Harvard Business Review, 63(6),
139-150.
Walker, B. and D. Salt, 2006: Resilience Thinking: Sustaining Ecosystems and People
in a Changing World. Island Press, Washington, DC, 176 pp.
Walker, B., S. Carpenter, J. Anderies, N. Abel, G.S. Cumming, M. Janssen, L. Lebel, J.
Norberg, G.D. Peterson, and R. Pritchard, 2002: Resilience management in
social-ecological systems: a working hypothesis for a participatory approach.
Conservation Ecology, 6(1), 14.
Walker, B., C.S. Holling, S.R. Carpenter, and A. Kinzig, 2004: Resilience, adaptability
and transformability in social–ecological systems. Ecology and Society, 9(2), 5.
Wandel, J. and B. Smit, 2000: Agricultural risk management in light of climate
variability and change. In: Agricultural and Environmental Sustainability in the
New Countryside [Millward, H., K. Beesley, B. Ilbery, and L. Harrington (eds.)].
Hignell Printing Limited, Winnipeg, Canada, pp. 30-39.
Warner, K., M. Hamza, A. Oliver-Smith, F. Renaud, and A. Julca, 2010: Climate
change, environmental degradation and migration. Natural Hazards, 55(3),
689-715.
Waterson, T., 2006: Climate change – the greatest crisis for children?
Journal of
Tropical Pediatrics, 52(6), 383-385.
WCC-3, 2009: Global Framework for Climate Services. Background paper prepared by
WMO secretariat, World Climate Conference 3, World Meteorological
Organization, Geneva, Switzerland, 6 pp.
WCED, 1987: Our Common Future. World Commission on Environment and
Development, Oxford University Press, Oxford, UK.
Weber, E.U., 2010: What shapes perceptions of climate change? Wiley
Interdisciplinary Reviews: Climate Change, 1(3), 332-342.
Weissbecker, I., S.E. Sephton, M.B. Martin, and D.M. Simpson, 2008: Psychological
and physiological correlates of stress in children exposed to disaster: review of
current research and recommendations for intervention. Children, Youth and
Environments, 18(1), 30-70.
Weitzman, M.L., 2009: On modeling and interpreting the economics of catastrophic
climate change. Review of Economics and Statistics, 91(1), 1-19.
West, C.T. and D.G. Lenze, 1994: Modeling the regional impact of natural disasters
and recovery: a general framework and an application to hurricane Andrew.
International Regional Science Review, 17(2), 121-150.
WGBU, 2008: Climate Change as a Security Risk. German Advisory Council on
Climate Change (WGBU), Earthscan, London, UK, 271 pp.
Whitehead, J., R. Bacon, G.Carbone, K. Dow, J. Thigpen, and D. Tufford, 2010: Using
climate extension to assist coastal decision-makers with climate adaptation. In:
22nd International Conference of the Coastal Society, Shifting Shorelines:
Adapting to the Future, Wilmington, NC, 13-16 June 2010.
Wilbanks, T.J., 1994: “Sustainable development” in geographic context. Annals of
the Association of American Geographers, 84(4), 541-557.
Wilbanks, T., 2003: Integrating climate change and sustainable development in a
place-based context. Climate Policy, 3(1), 147-154.
Wilbanks, T., 2007: Scale and sustainability. Climate Policy, 7(4), 278-287.
Wilbanks, T., 2009: How Geographic Scale Matters in Seeking Community
Resilience. CARRI Research Report 7, Community and Resilience Research
Institute, Oak Ridge, TN, 28 pp.
Wilbanks, T., 2010: Research and development priorities for climate change mitigation
and adaptation. In: Dealing with Climate Change: Setting a Global Agenda for
Mitigation and Adaptation [Pachauri, R. (ed.)]. The Energy and Resources
Institute, New Delhi, India, pp. 77-99.
Wilbanks, T. and R. Kates, 2010: Beyond adapting to climate change: embedding
adaptation in responses to multiple threats and stresses. Annals of the
Association of American Geographers, 100(4), 719-728.
Wilbanks, T. and J. Sathaye, 2007: Integrating mitigation and adaptation as responses
to climate change: a synthesis. Mitigation and Adaptation Strategies for Global
Change, 12(5), 957-962.
Wilbanks, J. and T. Wilbanks, 2010: Science, open communication, and sustainable
development. Sustainability, 2(4), 993-1015.
Wilbanks, T., P. Leiby, R. Perlack, J.T. Ensminger, and S.B. Wright, 2007: Toward an
integrated analysis of mitigation and adaptation: Some preliminary findings.
Mitigation and Adaptation Strategies for Global Change, 12(5), 713-725.
Wisner, B., 2001: Disasters: what the United Nations and its world can do.
Environmental Hazards, 3(3-4), 125-127.
Wisner, B., 2003: Changes in capitalism and global shifts in the distribution of hazard
and vulnerability. In: Natural Disasters and Development in a Globalizing World
[Pelling, M. (ed.)]. Routledge, London, UK, pp. 43-56.
Wisner, B., 2006: Let Our Children Teach Us! A Review of the Role of Education and
Knowledge in Disaster Risk Reduction. Books for Change, Bangalore, India, 148 pp.
Wisner, B., 2011: Are we there yet? Reflections on integrated disaster risk management
after ten years. Journal of Integrated Disaster Risk Management, 1(1),
doi:10.5595/idrim.2011.0015.
Wisner, B., P. Blaikie, T. Cannon, and I. Davis, 2004: At Risk: Natural Hazards, People’s
Vulnerability and Disaster. 2nd ed. Routledge, London, UK, 472 pp.
Wolf, J., 2011: Climate change adaptation as a social process. In: Climate Change
Adaptation in Developed Nations: From Theory to Practice [Ford, J. and L. Ford
Berrang (eds.)]. Springer, Dordrecht, The Netherlands, pp. 21-32.
Wolsink, M., 2006: River basin approach and integrated water management:
Governance pitfalls for the Dutch Space-Water-Adjustment Management
Principle. Geoforum, 37(4), 473-487.
Woodward, G. (ed.), 2010: Advances in Ecological Research 42: Ecological
Networks. Academic Press, London, UK, 454 pp.
World Bank, 2005: Managing Agricultural Production Risk. World Bank,
Washington, DC, 129 pp.
World Bank, 2008: World Development Report 2009: Reshaping Economic
Geography. World Bank, Washington, DC.
World Bank, 2010a: Natural Hazards, Unnatural Disasters, The Economics of
Effective Prevention. World Bank and United Nations, Washington, DC, 254 pp.
World Bank, 2010b: Vulnerability of Bangladesh to Cyclones in a Changing Climate.
Potential Damages and Adaptation Cost. Policy Research Working Paper 5280.
World Bank, Washington, DC, 54 pp.
WWF, 2010: Living Planet Report. World Wide Fund for Nature, Gland, Switzerland,
119 pp.
Yohe, G. and R. Leichenko, 2010: Adopting a risk-based approach. Annals of the New
York Academy of Sciences, 1196(1), 29-40.
Yohe, G. and R.S.J. Tol, 2002: Indicators for social and economic coping capacity –
moving toward a working definition of adaptive capacity. Global Environmental
Change, 12, 25-40.
Chapter 8 Toward a Sustainable and Resilient Future
486
Yohe, G.W., R.D. Lasco, Q.K. Ahmad, N.W. Arnell, S.J. Cohen, C. Hope, A.C. Janetos, and
R.T. Perez, 2007: Perspectives on climate change and sustainability. In: Climate
Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working
Group II to the Fourth Assessment Report of the Intergovernmental Panel on
Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der Linde, and
C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK, pp. 811-841.
Young, H. and S. Jaspars, 1995: Nutrition Matters. People Food and Famine.
Intermediate Technology Publications, London, UK, 151 pp.
Zahran, S., L. Peek, and S.D. Brody, 2008: Youth mortality by forces of nature.
Children, Youth and Environments, 18(1), 371-388.
Zimmerman, F.J. and M.R. Carter, 2003: Asset smoothing, consumption smoothing
and the reproduction of inequality under risk and subsistence constraints.
Journal of Development Economics, 71(2), 233-260.
Chapter 8Toward a Sustainable and Resilient Future
487
Coordinating Lead Authors:
Virginia Murray (UK), Gordon McBean (Canada), Mihir Bhatt (India)
Lead Authors:
Sergey Borsch (Russian Federation), Tae Sung Cheong (Republic of Korea), Wadid Fawzy Erian (Egypt),
Silvia Llosa (Peru), Farrokh Nadim (Norway), Mario Nunez (Argentina), Ravsal Oyun (Mongolia),
Avelino G. Suarez (Cuba)
Review Editors:
John Hay (New Zealand), Mai Trong Nhuan (Vietnam), Jose Moreno (Spain)
Contributing Authors:
Peter Berry (Canada), Harriet Caldin (UK), Diarmid Campbell-Lendrum (UK / WHO), Catriona Carmichael
(UK), Anita Cooper (UK), Cherif Diop (Senegal), Justin Ginnetti (USA), Delphine Grynzspan (France),
Clare Heaviside (UK), Jeremy Hess (USA), James Kossin (USA), Paul Kovacs (Canada), Sari Kovats (UK),
Irene Kreis (Netherlands), Reza Lahidji (France), Joanne Linnerooth-Bayer (USA), Felipe Lucio
(Mozambique), Simon Mason (USA), Sabrina McCormick (USA), Reinhard Mechler (Germany),
Bettina Menne (Germany / WHO), Soojeong Myeong (Republic of Korea), Arona Ngari (Cook Islands),
Neville Nicholls (Australia), Ursula Oswald Spring (Mexico), Pascal Peduzzi (Switzerland), Rosa Perez
(Philippines), Caroline Rodgers (Canada), Hannah Rowlatt (UK), Sohel Saikat (UK), Sonia Seneviratne
(Switzerland), Addis Taye (UK), Richard Thornton (Australia), Sotiris Vardoulakis (UK), Koko Warner
(Germany), Irina Zodrow (Switzerland / UNISDR)
This chapter should be cited as:
Murray, V., G. McBean, M. Bhatt, S. Borsch, T.S. Cheong, W.F. Erian, S. Llosa, F. Nadim, M. Nunez, R. Oyun, and A.G. Suarez,
2012: Case studies. In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation
[Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen,
M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate
Change (IPCC). Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 487-542.
9
Case Studies
Case Studies
488
Executive Summary .................................................................................................................................489
9.1. Introduction..............................................................................................................................490
9.2. Case Studies .............................................................................................................................492
9.2.1. European Heat Waves of 2003 and 2006.........................................................................................................................492
9.2.2. Response to Disaster Induced by Hot Weather and Wildfires .........................................................................................496
9.2.3. Managing the Adverse Consequences of Drought ..........................................................................................................498
9.2.4. Recent Dzud Disasters in Mongolia .................................................................................................................................500
9.2.5. Cyclones: Enabling Policies and Responsive Institutions for Community Action ............................................................502
9.2.6. Managing the Adverse Consequences of Floods .............................................................................................................505
9.2.7. Disastrous Epidemic Disease: The Case of Cholera .........................................................................................................507
9.2.8. Coastal Megacities: The Case of Mumbai ........................................................................................................................510
9.2.9. Small Island Developing States: The Challenge of Adaptation .......................................................................................512
9.2.10. Changing Cold Climate Vulnerabilities: Northern Canada...............................................................................................514
9.2.11. Early Warning Systems: Adapting to Reduce Impacts .....................................................................................................517
9.2.12. Effective Legislation for Multilevel Governance of Disaster Risk Reduction and Adaptation .......................................519
9.2.13. Risk Transfer: The Role of Insurance and Other Instruments
in Disaster Risk Management and Climate Change Adaptation in Developing Countries..............................................522
9.2.14. Education, Training, and Public Awareness Initiatives for Disaster Risk Reduction and Adaptation..............................526
9.3. Synthesis of Lessons Identified from Case Studies .................................................................529
References ...............................................................................................................................................530
Chapter 9
Table of Contents
489
Case studies contribute more focused analyses which, in the context of human loss and damage, demonstrate the
effectiveness of response strategies and prevention measures and identify lessons about success in disaster risk reduction
and climate change adaptation. The case studies were chosen to complement and be consistent with the information
in the preceding chapters, and to demonstrate aspects of the key messages in the Summary for Policymakers and the
Hyogo Framework for Action Priorities.
The case studies were grouped to examine types of extreme events, vulnerable regions, and methodological approaches.
For the extreme event examples, the first two case studies pertain to events of extreme temperature with moisture
deficiencies in Europe and Australia and their impacts including on health. These are followed by case studies on drought
in Syria and dzud, cold-dry conditions in Mongolia. Tropical cyclones in Bangladesh, Myanmar, and Mesoamerica, and
then floods in Mozambique are discussed in the context of community actions. The last of the extreme events case
studies is about disastrous epidemic disease, using the case of cholera in Zimbabwe, as the example.
The case studies chosen to reflect vulnerable regions demonstrate how a changing climate provides significant concerns
for people, societies, and their infrastructure. These are: Mumbai as an example of a coastal megacity; the Republic of
the Marshall Islands, as an example of small island developing states with special challenges for adaptation; and
Canada’s northern regions as an example of cold climate vulnerabilities focusing on infrastructures.
Four types of methodologies or approaches to disaster risk reduction (DRR) and climate change adaptation (CCA) are
presented. Early warning systems; effective legislation; risk transfer in developing countries; and education, training,
and public awareness initiatives are the approaches demonstrated. The case studies demonstrate that current disaster
risk management (DRM) and CCA policies and measures have not been sufficient to avoid, fully prepare for, and
respond to extreme weather and climate events, but these examples demonstrate progress.
A common factor was the need for greater information on risks before the events occur, that is, early warnings. The
implementation of early warning systems does reduce loss of lives and, to a lesser extent, damage to property and
was identified by all the extreme event case studies (heat waves, wildfires, drought, dzud, cyclones, floods, and
epidemic disease) as key to reducing impacts from extreme events. A need for improving international cooperation
and investments in forecasting was recognized in some of the case studies but equally the need for regional and local
early warning systems was heavily emphasized, particularly in developing countries.
A further common factor identified overall was that it is better to invest in preventative-based DRR plans, strategies,
and tools for adaptation than in response to extreme events. Greater investments in proactive hazard and vulnerability
reduction measures, as well as development of capacities to respond and recover from the events were demonstrated to
have benefits. Specific examples for planning for extreme events included increased emphasis on drought preparedness;
planning for urban heat waves; and tropical cyclone DRM strategies and plans in coastal regions that anticipate these
events. However, as illustrated by the small island developing states case study, it was also identified that DRR planning
approaches continue to receive less emphasis than disaster relief and recovery.
One recurring theme is the value of investments in knowledge and information, including observational and monitoring
systems, for cyclones, floods, droughts, heat waves, and other events from early warnings to clearer understanding of
health and livelihood impacts. In all cases, the point is made that with greater information available it would be
possible to know the risks better and ensure that response strategies were adequate to face the coming threat.
Research improves our knowledge, especially when it integrates the natural, social, health, and engineering sciences
and their applications. The case studies have reviewed past events and identified lessons which could be considered
for the future. Preparedness through DDR and DRM can help to adapt to climate change and these case studies offer
examples of measures that could be taken to reduce the damage that is inflicted as a result of extreme events.
Investment in increasing knowledge and warning systems, adaptation techniques, and tools and preventive measures
will cost money now but will save money and lives in the future.
Chapter 9 Case Studies
Executive Summary
490
9.1. Introduction
In this chapter, case studies are used as examples of how to gain a better
understanding of the risks posed by extreme weather and climate-related
events while identifying lessons and best practices from past responses
to such occurrences. Using the information in Chapters 1 to 8, it was
possible to focus on particular examples to reflect the needs of the
whole Special Report. The chosen case studies are illustrative of an
important range of disaster risk reduction, disaster risk management,
and climate change adaptation issues. They are grouped to examine
representative types of extreme events, vulnerable regions, and
methodological approaches.
For the extreme event examples, the first two case studies pertain to
extreme temperature with moisture deficiencies: the European heat waves
of 2003 and 2006 and response to disaster induced by hot weather and
wildfires in Australia. Managing the adverse consequences of drought is
the third case study, with the focus on Syrian droughts. The combination
of drought and cold is examined through two recent dzud disasters in
Mongolia, 1999-2002 and 2009-2010. Tropical cyclones in Bangladesh,
Myanmar, and Mesoamerica are used as examples of how a difference
can be made via enabling policies and responsive institutions for
community action. The next case study shifts the geographical focus to
floods in Mozambique in 2000 and 2007. The last of the extreme events
case studies is about disastrous epidemic disease, using the case of cholera
in Zimbabwe, as the example.
The case studies chosen to reflect a few vulnerable regions all demonstrate
how a changing climate provides significant concerns for people,
societies, and their infrastructure. The case of Mumbai is used as an
example of a coastal megacity and its risks. Small island developing
states have special challenges for adaptation, with the Republic of the
Marshall Islands being the case study focus. Cold climate vulnerabilities,
particularly the infrastructure in Canada’s northern regions, provide the
final vulnerable region case study.
Following examples of extreme events and vulnerable regions, this
chapter presents case study examination of four types of methodologies
or approaches to DRR and CCA. Early warning systems provide the
opportunity for adaptive responses to reduce impacts. Effective legislation
to provide multi-level governance is another way of reducing impacts.
The case study on risk transfer examines the role of insurance and other
instruments in developing countries. The final case study is on education,
training, and public awareness initiatives. This selection provides a good
basis of information and serves as an indicator of the resources needed
for future DRR and CCA. Additionally, it allows good practices to be
identified and lessons to be extracted.
The case studies provide the opportunity for connecting with common
elements across the other chapters. Each case study is presented in a
consistent way to enable better comparison of approaches. After an
introduction, authors provide background to the event, vulnerable region,
or methodology. Then the description of the events, vulnerability, or
strategy is given as appropriate. Next is the discussion of interventions,
followed by the outcomes and/or consequences. Each case study
concludes with a discussion of lessons identified. These case studies
relate to the key messages of the Summary for Policymakers and also to
the Hyogo Framework for Action Priorities (see Table 9-1).
Case studies are widely used in many disciplines including health care
(Keen and Packwood, 1995; McWhinney, 2001), social science (Flyvbjerg,
2004), engineering, and education (Verschuren, 2003). In addition, case
studies have been found to be useful in previous IPCC Assessment
Reports, including the 2007 Working Group II report (Parry et al., 2007).
Case studies offer records of innovative or good practices. Specific
problems or issues experienced can be documented as well as the actions
taken to overcome these. Case studies can validate our understanding
and encourage re-evaluation and learning. It is apparent that (i) case
studies capture the complexity of disaster risk and disaster situations;
(ii) case studies appeal to a broad audience; and (iii) case studies should
be fully utilized to provide lessons identified for DRR and DRM for
adaptation to climate change (Grynszpan et al., 2011). Several projects
have identified lessons from case studies (Kulling et al., 2010). The
Forensic Investigation of Disaster (Burton, 2011; FORIN, 2011) Project of
the Integrated Research on Disaster Risk (ICSU, 2008) program has
developed a methodology and template for future case study investigations
to provide a basis for future policy analysis and literature for assessments.
The FORIN template lays out the elements: (a) critical cause analysis; (b)
meta-analysis; (c) longitudinal analysis; and (d) scenarios of disasters.
Case studies included in this chapter have been extracted from a variety
of literature sources from many disciplines. As a result, an integrated
approach examining scientific, social, health, and economic aspects of
disasters was used where appropriate and included different spatial and
temporal scales, as needed. The specialized insights they provide can be
useful in evaluating some current disaster response practices.
This chapter addresses events whose impacts were felt in many
dimensions. A single event can produce effects that are felt on local,
regional, national, and international levels. These effects could have
been the direct result of the event itself, from the response to the event,
or through indirect impacts such as a reduction in food production or a
decrease in available resources. In addition to the spatial scales, this
chapter also addresses temporal scales, which vary widely in both
event-related impacts and responses. However, the way effects are felt
is additionally influenced by social, health, and economic factors. The
resilience of a society and its economic capacity to allay the impact of
a disaster and cope with the after effects has significant ramifications
for the community concerned (UNISDR, 2008a). Developing countries with
fewer resources, experts, equipment, and infrastructure have been shown
to be particularly at risk (see Chapter 5). Developed nations are usually
better equipped with technical, financial, and institutional support to enable
better adaptive planning including preventive measures and/or quick and
effective responses (Gagnon-Lebrun and Agrawala, 2006). However, they
still remain at risk of high-impact events, as exemplified by the European
heat wave of 2003 and by Hurricane Katrina (Parry et al., 2007).
Chapter 9Case Studies
491
Chapter 9 Case Studies
••
••
`
Exposure and vulnerability are key determinants of disaster risk.
A changing climate leads to changes in the frequency, intensity,
spatial extent, duration, and timing of extreme weather and
climate events, and can result in unprecedented extreme weather
and climate events.
Exposure and vulnerability are dynamic, varying across temporal
and spatial scales, and depend on economic, social, geographic,
demographic, cultural, institutional, governance, and
environmental factors.
Settlement patterns, urbanization, and changes in socioeconomic
status have all influenced observed trends in exposure and
vulnerability to climate extremes.
Trends in exposure and vulnerability are major drivers of changes
in disaster risk.
Inequalities influence local coping and adaptive capacity, and
pose disaster risk management and adaptation challenges from
the local to national levels.
Humanitarian relief is often required when disaster risk reduction
measures are absent or inadequate.
Post-disaster recovery and reconstruction provide an opportunity
for reducing weather and climate-related disaster risk and for
improving adaptive capacity.
Risk sharing and transfer mechanisms at local, national, regional,
and global scales can increase resilience to climate extremes.
Attention to the temporal and spatial dynamics of exposure and
vulnerability is particularly important given that the design and
implementation of adaptation and disaster risk management
strategies and policies can reduce risk in the short term, but may
increase exposure and vulnerability over the longer term.
Closer integration of disaster risk management and climate
change adaptation, along with the incorporation of both into
local, subnational, national, and international development
policies and practices, could provide benefits at all scales.
Models project substantial warming in temperature extremes by
the end of the 21st century.
It is likely that the frequency of heavy precipitation or the
proportion of total rainfall from heavy falls will increase in the
21st century over many areas of the globe.
There is medium confidence that droughts will intensify in the
21st century in some seasons and areas, due to reduced
precipitation and/or increased evapotranspiration.
A. Context
B. Observations
of Exposure,
Vulnerability,
Climate
Extremes,
Impacts, and
Disaster Losses
C. Disaster Risk
Management and
Adaptation to
Climate Change:
Past Experience
with Climate
Extremes
D. Future Climate
Extremes,
Impacts, and
Disaster Losses
Continued next page
Heat-
waves
Hot
weather
and
wildfires
Drought Dzud Cyclones Floods Epidemic
Disease
Mega-
cities
SIDS Cold
Climate
EWS Legislation Risk
Transfer
Education
Key Message
9.2.1 9.2.2 9.2.3 9.2.4 9.2.5 9.2.6 9.2.7 9.2.8 9.2.9 9.2.10 9.2.11 9.2.12 9.2.13 9.2.14
Table 9-1 | Matrix demonstrating the connectivity between the case studies (9.2.1-9.2.14) and the Summary for Policymakers messages. Those with the strongest relationship are shown. Connectivity between the case studies and the
Hyogo Framework for Action Priority Areas (UNISDR, 2005b) are also shown.
492
Most importantly, this chapter highlights
the complexities of disasters in order to
encourage effective solutions that address
these complexities rather than just one issue
or another. The lessons of this chapter provide
examples of experience that can help develop
strategies to adapt to climate change.
9.2. Case Studies
9.2.1. European Heat Waves
of 2003 and 2006
9.2.1.1. Introduction
Extreme heat is a prevalent public health
concern throughout the temperate regions of
the world and extreme heat events have been
encountered recently in North America, Asia,
Africa, Australia, and Europe. It is very likely
that the length, frequency, and/or intensity
of warm spells, including heat waves, will
continue to increase over most land areas
(Section 3.3.1). As with other types of
hazards, extreme heat can have disastrous
consequences, particularly for the most
vulnerable populations. Risk from extreme heat
is a function of hazard severity and population
exposure and vulnerability. Extreme heat
events do not necessarily translate into
extreme impacts if vulnerability is low. It is
important, therefore, to consider factors that
contribute to hazard exposure and population
vulnerability. Recent literature has identified
a host of factors that can amplify or dampen
hazard exposure. Experience with past heat
waves and public health interventions suggest
that it is possible to manipulate many of
these variables to reduce both exposure and
vulnerability and thereby limit the impacts of
extreme heat events. This case study, which
compares the European heat wave of 2003
with 2006, demonstrates developments in
disaster risk management and adaptation to
climate change.
9.2.1.2. Background/Context
Extreme heat is a prevalent public health
concern throughout the temperate regions of
the world (Kovats and Hajat, 2008), in part
Chapter 9Case Studies
`
Measures that provide benefits under current climate and a range
of future climate change scenarios, called low-regrets measures,
are available starting points for addressing projected trends in
exposure, vulnerability, and climate extremes. They have the
potential to offer benefits now and lay the foundation for
addressing projected changes.
Effective risk management generally involves a portfolio of
actions to reduce and transfer risk and to respond to events and
disasters, as opposed to a singular focus on any one action or
type of action.
Multi-hazard risk management approaches provide opportunities
to reduce complex and compound hazards.
Integration of local knowledge with additional scientific and
technical knowledge can improve disaster risk reduction and
climate change adaptation.
Appropriate and timely risk communication is critical for effective
adaptation and disaster risk management.
1: Ensure that disaster risk reduction is a national and a local
priority with a strong institutional basis for implementation.
2: Identify, assess and monitor disaster risks and enhance early
warning.
3: Use knowledge, innovation and education to build a culture of
safety and resilience at all levels.
4: Reduce the underlying risk factors.
5: Strengthen disaster preparedness for effective response at all
levels.
E. Managing
Changing Risk
of Climate
Extremes and
Disasters
Hyogo
Framework for
Action –
Priorities for
Action
Heat-
waves
Hot
weather
and
wildfires
Drought Dzud Cyclones Floods Epidemic
Disease
Mega-
cities
SIDS Cold
Climate
EWS Legislation Risk
Transfer
Education
Key Message
9.2.1 9.2.2 9.2.3 9.2.4 9.2.5 9.2.6 9.2.7 9.2.8 9.2.9 9.2.10 9.2.11 9.2.12 9.2.13 9.2.14
Table 9-1 (continued)
493
because heat-related extreme events are projected to result in increased
mortality (Peng et al., 2010). Extreme heat events have been encountered
recently in North America (Hawkins-Bell and Rankin, 1994; Klinenberg,
2002), Asia (Kumar, 1998; Kalsi and Pareek, 2001; Srivastava et al., 2007),
Africa (NASA, 2008), Australia (DSE, 2008b), and Europe (Robine et al.,
2008; Founda and Giannakopoulos, 2009). This concern may also be
present in non-temperate regions, but there is little research on this.
As with other types of hazards, extreme heat events can have disastrous
consequences, partly due to increases in exposure and particular types
of vulnerabilities. However, it is important to note that reducing the
impacts of extreme heat events linked to climate change will necessitate
further actions, some of which may be resource intensive and further
exacerbate climate change.
9.2.1.2.1. Vulnerabilities to heat waves
Physiological: Several factors influence vulnerability to heat-related
illness and death. Most of the research related to such vulnerability is
derived from experiences in industrialized nations. Several physiological
factors, such as age, gender, body mass index, and preexisting health
conditions, play a role in the body’s ability to respond to heat stress. Older
persons, babies, and young children have a number of physiological and
social risk factors that place them at elevated risk, such as decreased
ability to thermoregulate (the ability to maintain temperature within the
narrow optimal physiologic range; Havenith, 2001). Preexisting chronic
disease – more common in the elderly – also impairs compensatory
responses to sustained high temperatures (Havenith, 2001; Shimoda,
2003). Older adults tend to have suppressed thirst impulse resulting in
dehydration and increased risk of heat-related illness. In addition, multiple
diseases and/or drug treatments increase the risk of dehydration
(Hodgkinson et al., 2003; Ebi and Meehl, 2007).
Social: A wide range of socioeconomic factors are associated with
increased vulnerability (see Sections 2.3 and 2.5). Areas with high crime
rates, low social capital, and socially isolated individuals had increased
vulnerability during the Chicago heat wave in 1995 (Klinenberg, 2002).
People in areas of low socioeconomic status are generally at higher risk
of heat-related morbidity and mortality due to higher prevalence of
chronic diseases – from cardiovascular diseases such as hypertension to
pulmonary disease, such as chronic obstructive pulmonary disease and
asthma (Smoyer et al., 2000; Sheridan, 2003). Minorities and communities
of low socioeconomic status are also frequently situated in higher heat
stress neighborhoods (Harlan et al., 2006). Protective measures are
often less available for those of lower socioeconomic status, and even if
air conditioning, for example, is available, some of the most vulnerable
populations will choose not to use it out of concern over the cost (O’Neill
et al., 2009). Other groups, like the homeless and outdoor workers, are
particularly vulnerable because of their living situation and being more
acutely exposed to heat hazards (Yip et al., 2008). Older persons may also
often be isolated and living alone, and this may increase vulnerability
(Naughton et al., 2002; Semenza, 2005).
9.2.1.2.2. Impact of urban infrastructure
Addressing vulnerabilities in urban areas will benefit those at risk.
Around half of the world’s population live in urban areas at present, and
by 2050, this figure is expected to rise to about 70% (UN, 2008). Cities
across the world are expected to absorb most of the population growth
over the next four decades, as well as continuing to attract migrants from
rural areas (UN, 2008). In the context of a heat-related extreme event,
certain infrastructural factors can either amplify or reduce vulnerability
of exposed populations. The built environment is important since local heat
production affects the urban thermal budget (from internal combustion
engines, air conditioners, and other activities). Other factors also play a
role in determining local temperatures, including surface reflectivity or
albedo, the percent of vegetative cover, and thermal conductivity of
building materials. The urban heat island effect, caused by increased
absorption of infrared radiation by buildings and pavement, lack of
shading, evapotranspiration by vegetation, and increased local heat
production, can significantly increase temperatures in the urban core by
several degrees Celsius, raising the likelihood of hazardous heat exposure
for urban residents (Clarke, 1972; Shimoda, 2003). Street canyons where
building surfaces absorb heat and affect air flow are also areas where
heat hazards may be more severe (Santamouris et al., 1999; Louka et
al., 2002). The restricted air flow within street canyons may also cause
accumulation of traffic-related air pollutants (Vardoulakis et al., 2003).
Research has also identified that, at least in the North American and
European cities where the phenomenon has been studied, these factors
can have a significant impact on the magnitude of heat hazards on a
neighborhood level (Harlan et al., 2006). One study in France has shown
that higher mortality rates occurred in neighborhoods in Paris that were
characterized by higher outdoor temperatures (Cadot et al., 2007). High
temperatures can also affect transport networks when heat damages
roads and rail tracks. Within cities, outdoor temperatures can vary
significantly (Akbari and Konopacki, 2004), resulting in the need to
focus preventive strategies on localized characteristics.
Systems of power generation and transmission partly explain vulnerability
since electricity supply underpins air conditioning and refrigeration – a
significant adaptation strategy particularly in developed countries, but
one that is also at increased risk of failure during a heat wave (Sailor
and Pavlova, 2003). It is expected that demand for electricity to power
air conditioning and refrigeration units will increase with rising ambient
temperatures. Areas with lower power capacities face increased risk of
disruptions to generating resources and transmission under excessive
heat events.
In addition to increased demand, there can be a risk of reduced output
from power generating plants (UNEP, 2004). The ability of inland thermal
power plants, both conventional and nuclear, to cool their generators is
restricted by rising river temperatures. Additionally, fluctuating levels of
water availability will affect energy outputs of hydropower complexes.
During the summer of 2003 in France, six power plants were shut down
and others had to control their output (Parry et al., 2007).
Chapter 9 Case Studies
494
9.2.1.2.3. Heat waves and air pollution
Concentrations of air pollutants such as particulate matter and ozone are
often elevated during heat waves due to anticyclonic weather conditions,
increased temperatures, and light winds. Photochemical production of
ozone and emissions of biogenic ozone precursors increase during hot,
sunny weather, and light winds do little to disperse the build-up of air
pollution. Air pollution has well-established acute effects on health,
particularly associated with respiratory and cardiovascular illness, and can
result in increased mortality and morbidity (WHO, 2006a). Background
ozone levels in the Northern Hemisphere have doubled since preindustrial
times (Volz and Kley, 1988) and increased in many urban areas over the
last few decades (Vingarzan, 2004). Air quality standards and regulations
are helping to improve air quality, although particles and ozone are still
present in many areas at levels that may cause harm to human health,
particularly during heat waves (Royal Society, 2008; EEA, 2011). The effects
of climate change (particularly temperature increases) together with a
steady increase in background hemispheric ozone levels is reducing the
efficacy of measures to control ozone precursor emissions in the future
(Derwent et al., 2006). The increased frequency of heat waves in the
future will probably lead to more frequent air pollution episodes (Stott
et al., 2004; Jones et al., 2008).
9.2.1.3. Description of Events
9.2.1.3.1. European heat wave of 2003
During the first two weeks of August 2003, temperatures in Europe
soared far above historical norms. The heat wave stretched across much
of Western Europe, but France was particularly affected (InVS, 2003).
Maximum temperatures recorded in Paris remained mostly in the range
of 35 to 40°C between 4 and 12 August, while minimum temperatures
recorded by the same weather station remained almost continuously
above 23°C between 7 and 14 August (Météo France, 2003). The European
heat wave had significant health impacts (Lagadec, 2004). Initial estimates
were of costs exceeding €13 billion, with a death toll of over 30,000
across Europe (UNEP, 2004). It has been estimated that mortality over
the entire summer could have reached about 70,000 (Robine et al.,
2008) with approximately 14,800 excess deaths in France alone (Pirard
et al., 2005). The severity, duration, geographic scope, and impact of the
event were unprecedented in recorded European history (Grynszpan,
2003; Kosatsky, 2005; Fouillet et al., 2006) and put the event in the
exceptional company of the deadly Beijing heat wave of 1743, which killed
at least 11,000, and possibly many more (Levick, 1859; Bouchama,
2004; Lagadec, 2004; Pirard et al., 2005; Robine et al., 2008).
During the heat wave of August 2003, air pollution levels were high
across much of Europe, especially surface ozone (EEA, 2003). A rapid
assessment was performed for the United Kingdom after the heat wave,
using published exposure-response coefficients for ozone and PM
10
(particulate matter with an aerodynamic diameter of up to 10 μm). The
assessment associated 21 to 38% of the total 2,045 excess deaths in
the United Kingdom in August 2003 to elevated ambient ozone and
PM
10
concentrations (Stedman, 2004). The task of separating health
effects of heat and air pollution is complex; however, statistical and
epidemiological studies in France also concluded that air pollution was
a factor associated with detrimental health effects during August 2003
(Dear et al., 2005; Filleul et al., 2006).
9.2.1.3.2. European heat wave of 2006
Three years later, between 10 and 28 July 2006, Europe experienced
another major heat wave. In France, it ranked second only to the one in
2003 as the most severe heat wave since 1950 (Météo France, 2006;
Fouillet et al., 2008). The 2006 heat wave was longer in duration than
that of 2003, but was less intense and covered less geographical area
(Météo France, 2006). Ozone levels were high across much of southern and
northwestern Europe in July 2006, with concentrations reaching levels
only exceeded in 2003 to date (EEA, 2007). Across France, recorded
maximum temperatures soared to 39 to 40°C, while minimum recorded
temperatures reached 19 to 23°C (compared with 23 to 25°C in 2003)
(Météo France, 2006). Based on a historical model, the temperatures
were expected to cause around 6,452 excess deaths in France alone, yet
only around 2,065 excess deaths were recorded (Fouillet et al., 2008).
9.2.1.4. Interventions
Efforts to minimize the public health impact for the heat wave in 2003
were hampered by denial of the event’s seriousness and the inability of
many institutions to instigate emergency-level responses (Lagadec,
2004). Afterwards several European countries quickly initiated plans to
prepare for future events (WHO, 2006b). France, the country hit hardest,
developed a national heat wave plan, surveillance activities, clinical
treatment guidelines for heat-related illness, identification of vulnerable
populations, infrastructure improvements, and home visiting plans for
future heat waves (Laaidi et al., 2004).
9.2.1.5. Outcomes/Consequences
The difference in impact between the heat waves in 2003 and 2006
may be at least partly attributed to the difference in the intensity and
geographic scope of the hazard. It has been considered that in France
at least, some decrease in 2006 mortality may also be attributed to
increased awareness of the ill effects of a heat wave, the preventive
measures instituted after the 2003 heat wave, and the heat health
watch system set up in 2004 (Fouillet et al., 2008). While the mortality
reduction may demonstrate the efficacy of public health measures, the
persistent excess mortality highlights the need for optimizing existing
public health measures such as warning and watch systems (Hajat et al.,
2010), health communication with vulnerable populations (McCormick,
2010a), vulnerability mapping (Reid et al., 2009), and heat wave
response plans (Bernard and McGeehin, 2004). It also highlights the
Chapter 9Case Studies
495
need for other, novel measures such as modification of the urban form
to reduce exposure (Bernard and McGeehin, 2004; O’Neill et al., 2009;
Reid et al., 2009; Hajat et al., 2010; Silva et al., 2010). Thus the outcomes
from the two European heat waves of 2003 and 2006 are extensive and
are considered below. They include public health approaches to reducing
exposure, assessing heat mortality, communication and education, and
adapting the urban infrastructure.
9.2.1.5.1. Public health approaches to reducing exposure
A common public health approach to reducing exposure is the Heat
Warning System (HWS) or Heat Action Response System. The four
components of the latter include an alert protocol, community response
plan, communication plan, and evaluation plan (Health Canada, 2010).
The HWS is represented by the multiple dimensions of the EuroHeat
plan, such as a lead agency to coordinate the alert, an alert system, an
information outreach plan, long-term infrastructural planning, and
preparedness actions for the health care system (WHO, 2007). The
European Network of Meteorological Services has created Meteoalarm
as a way to coordinate warnings and to differentiate them across
regions (Bartzokas et al., 2010). There are a range of approaches used
to trigger alerts and a range of response measures implemented once
an alert has been triggered. In some cases, departments of emergency
management lead the endeavor, while in others public health-related
agencies are most responsible (McCormick, 2010b).
As yet, there is not much evidence on the efficacy of heat warning
systems. A few studies have identified an effect of heat preparedness
programming. For example, the use of emergency medical services during
heat wave events dropped by 49% in Milwaukee, Wisconsin, between
1995 and 1999; an outcome that may be partly due to heat preparedness
programming or to differences between the two heat waves (Weisskopf
et al., 2002). Evidence has also indicated that interventions in Philadelphia,
Pennsylvania, are likely to have reduced mortality rates by 2.6 lives per
day during heat events (Ebi et al., 2004). An Italian intervention program
found that caretaking in the home resulted in decreased hospitalizations
due to heat (Marinacci et al., 2009). However, for all these studies, it is
not clear whether the observed reductions were due to the interventions.
Questions remain about the levels of effectiveness in many circumstances
(Cadot et al., 2007).
Heat preparedness plans vary around the world. Philadelphia,
Pennsylvania – one of the first US cities to begin a heat preparedness
plan – has a ten-part program that integrates a ‘block captain’ system
where local leaders are asked to notify community members of dangerous
heat (Sheridan, 2006; McCormick, 2010b). Programs like the Philadelphia
program that utilize social networks have the capacity to shape behavior
since networks can facilitate the sharing of expertise and resources
across stakeholders; however, in some cases the influence of social
networks contributes to vulnerability (Crabbé and Robin, 2006). Other
heat warning systems, such as that in Melbourne, Australia, are based
solely on alerting the public to weather conditions that threaten older
populations (Nicholls et al., 2008). Addressing social factors in
preparedness promises to be critical for the protection of vulnerable
populations. This includes incorporating communities themselves into
understanding and responding to extreme events. It is important that top-
down measures imposed by health practitioners account for community-
level needs and experiences in order to be more successful. Greater
attention to and support of community-based measures in preventing heat
mortality can be more specific to local context, such that participation
is broader (Semenza et al., 2007). Such programs can best address the
social determinants of health outcomes.
9.2.1.5.2. Assessing heat mortality
Assessing excess mortality is the most widely used means of assessing
the health impact of heat-related extreme events. Mortality represents
only the ‘tip of the iceberg’ of heat-related health effects; however, it is
more widely and accurately reported than morbidity, which explains its
appeal as a data source. Nonetheless, assessing heat mortality presents
particular challenges. Accurately assessing heat-related mortality faces
challenges of differences in contextual variations (Hémon and Jougla, 2004;
Poumadere et al., 2005), and coroner’s categorization of deaths (Nixdorf-
Miller et al., 2006). For example, there are a number of estimates of
mortality for the European heat wave that vary depending on geographic
and temporal ranges, methodological approaches, and risks considered
(Assemblée Nationale, 2004). The different types of analyses used to
assess heat mortality, such as certified heat deaths and heat-related
mortality measured as an excess of total mortality over a given time
period, are important distinctions in assessing who is affected by the
heat (Kovats and Hajat, 2008). Learning from past and other countries’
experience, a common understanding of definitions of heat waves and
excess mortality, and the ability to streamline death certification in the
context of an extreme event could improve the ease and quality of
mortality reporting.
9.2.1.5.3. Communication and education
One particularly difficult aspect of heat preparedness is communicating
risk. In many locations populations are unaware of their risk and heat
wave warning systems go largely unheeded (Luber and McGeehin, 2008).
Some evidence has even shown that top-down educational messages
do not result in appropriate resultant actions (Semenza et al., 2008). The
receipt of information is not sufficient to generate new behaviors or the
development of new social norms. Even when information is distributed
through pamphlets and media outlets, behavior of at-risk populations
often does not change and those targeted by such interventions have
suggested that community-based organizations be involved in order to
build on existing capacity and provide assistance (Abrahamson et al.,
2008). Older people, in particular, engage better with prevention
campaigns that allow them to maintain independence and do not focus
on their age, as many heat warning programs do (Hughes et al., 2008).
More generally, research shows that communication about heat
Chapter 9 Case Studies
496
preparedness centered on engaging with communities results in
increased awareness compared with top-down messages (Smoyer-
Tomic and Rainham, 2001).
9.2.1.5.4. Adapting the urban infrastructure
Several types of infrastructural measures can be taken to prevent
negative outcomes of heat-related extreme events. Models suggest that
significant reductions in heat-related illness would result from land use
modifications that increase albedo, proportion of vegetative cover,
thermal conductivity, and emissivity in urban areas (Yip et al., 2008;
Silva et al., 2010). Reducing energy consumption in buildings can improve
resilience, since localized systems are less dependent on vulnerable
energy infrastructure. In addition, by better insulating residential
dwellings, people would suffer less effect from heat hazards. Financial
incentives have been tested in some countries as a means to increase
energy efficiency by supporting those who are insulating their homes.
Urban greening can also reduce temperatures, protecting local populations
and reducing energy demands (Akbari et al., 2001).
9.2.1.6. Lessons Identified
With climate change, heat waves are very likely to increase in frequency
and severity in many parts of the world (Section 3.3.1). Smarter urban
planning, improvements in existing housing stock and critical infrastructure,
along with effective public health measures will assist in facilitating climate
change adaptation.
Through understanding local conditions and experiences and current and
projected risks, it will be possible to develop strategies for improving
heat preparedness in the context of climate change. The specificity of
heat risks to particular sub-populations can facilitate appropriate
interventions and preparedness.
Communication and education strategies are most effective when they
are community-based, offer the opportunity for changing social norms,
and facilitate the building of community capacity.
Infrastructural considerations are critical to reducing urban vulnerability
to extreme heat events. Effective preparedness includes building techniques
that reduce energy consumption and the expansion of green space.
Heat wave preparedness programs may be able to prevent heat mortality;
however, testing and development is required to assess the most effective
approaches.
Further research is needed on the efficacy of existing plans, how to
improve preparedness that specifically focuses on vulnerable groups,
and how to best communicate heat risks across diverse groups. There
are also methodological difficulties in describing individual vulnerability
that need further exploration.
9.2.2. Response to Disaster Induced by
Hot Weather and Wildfires
9.2.2.1. Introduction
Climate change is expected to increase global temperatures and change
rainfall patterns (Christensen et al., 2007). These climatic changes will
increase the risk of temperature- and precipitation-related extreme
weather and climate events. The relative effects will vary by regions and
localities (Sections 3.3.1, 3.3.2, and 3.5.1). In general, an increase in
mean temperature, and a decrease in mean precipitation can contribute
to increased fire risk (Flannigan et al., 2009). When in combination with
severe droughts and heat waves, which are also expected to increase
in many fire regions (Sections 3.3.1 and 3.5.1), fires can become
catastrophic (Bradstock et al., 2009). Wildfires occur in many regions of
the world, and due to their extreme nature, authorities and the public in
general are acquainted with such extreme situations, and plans have
been enacted to mitigate them. However, at times, the nature of fire
challenges these plans and disasters emerge. This case study uses the
example from Victoria, Australia, in 2009. The goal is to present hot
weather and wildland fire hazards and their effects and potential
impacts, and to provide an overview of experience to learn to manage
these extreme risks, as well as key lessons for the future.
9.2.2.2. Background
Wildfire risk occurs in many regions of the globe; however embodying this
risk in a single and practical universal index is difficult. The relationships
between weather and wildfires have been studied for many areas of the
world; in some, weather is the dominant factor in ignitions, while in
others, human activities are the major cause of ignition, but weather and
environmental factors mainly determine the area burned (Bradstock et al.,
2009). Wildfire behavior is also modified by forest and land management
and fire suppression (Allen et al., 2002; Noss et al., 2006). Wildfires do
not burn at random in the landscape (Nunes et al., 2005), and occur at
particular topographic locations or distances from towns or roads
(Mouillot et al., 2003; Badia-Perpinyà and Pallares-Barbera, 2006;
Syphard et al., 2009). The intensity and rate of spread of a wildfire is
dependent on the amount, moisture content, and arrangement of fine
dead fuel, the wind speed near the burning zone, and the terrain and
slope where it is burning. Wildfire risk is a combination of all factors that
affect the inception, spread, and difficulty of fire control and damage
potential (Tolhurst, 2010).
9.2.2.3. Description of Events
An episode of extreme heat waves began in South Australia on 25
January 2009. Two days later they had become more widespread over
southeast Australia. The exceptional heat wave was caused by a slow-
moving high-pressure system that settled over the Tasman Sea, in
combination with an intense tropical low located off the northwest
Chapter 9Case Studies
497
Australian coast and a monsoon trough over northern Australia.
This produced ideal conditions for hot tropical air to be directed over
southeastern Australia (National Climate Centre, 2009).
In Melbourne the temperature was above 43°C for three consecutive
days (28 to 30 January 2009), reaching a peak of 45.1°C on 30 January
2009. This was the second-highest temperature on record. The extremely
high day and night temperatures combined to produce a record high
daily mean temperature of 35.4°C on 30 January (Victorian Government,
2009). The 2008 winter season was characterized by below-average
precipitation across much of Victoria. While November and December
2008 experienced average and above average rainfall, respectively, in
January and February the rainfall was substantially below average
(Parliament of Victoria, 2010a). During the 12 years between 1998 and
2007, Victoria experienced warmer than average temperatures and a
14% decline in average rainfall (DSE, 2008a). In central Victoria the 12-
year rainfall totals were approximately 10 to 20% below the 1961 to
1990 average (Australian Government, 2009).
This heat wave had a substantial impact on the health of Victorians,
particularly the elderly (National Climate Centre, 2009; Parliament of
Victoria, 2009). A 25% increase in total emergency cases and a 46%
increase over the three hottest days were reported for the week of the
heat wave. Emergency departments reported a 12% overall increase in
presentations, with a greater proportion of acutely ill patients and a
37% increase in patients 75 years or older (Victorian Government,
2009). Attribution of mortality to a heat wave can be difficult, as deaths
tend to occur from exacerbations of chronic medical conditions as well
as direct heat-related illness; this is particularly so for the frail and
elderly (Kovats and Hajat, 2008). However, excess mortality can provide
a measure of the impact of a heat wave. With respect to total all-cause
mortality, there were 374 excess deaths with a 62% increase in total
all-cause mortality. The total number of deaths during the four days of
the heat wave was 980, compared to a mean of 606 for the previous
five years. Reported deaths in people 65 years and older more than
doubled compared to the same period in 2008 (Victorian Government,
2009).
On 7 February 2009, the temperatures spiked again. The Forest Fire
Danger Index, which is calculated using variables such as temperature,
precipitation, wind speed, and relative humidity (Hennessy et al., 2005),
this time reached unprecedented levels, higher than the fire weather
conditions experienced on Black Friday in 1939 and Ash Wednesday in
1983 (National Climate Centre, 2009) – the two previous worse fire
disasters in Victoria.
By the early afternoon of 7 February, wind speeds were reaching their
peak, resulting in a power line breaking just outside the town of Kilmore,
sparking a wildfire that would later generate extensive pyrocumulus
cloud and become one of the largest, deadliest, and most intense
firestorms ever experienced in Australia’s history (Parliament of Victoria,
2010a). The majority of fire activity occurred between midday and
midnight on 7 February, when wind speeds and temperature were at
their highest and humidity at its lowest. A major wind change occurred
late afternoon across the fire ground, turning the northeastern flank
into a new wide fire front, catching many people by surprise. This was
one of several hundred fires that started on this day, most of which
were quickly controlled; however, a number went on to become major
fires resulting in much loss of life. The worst 15 of these were examined
in detail by the Victorian Bushfires Royal Commission (Parliament of
Victoria, 2010a). A total of 173 people died as a result of the Black
Saturday bushfires (Parliament of Victoria, 2010a). They also destroyed
almost 430,000 ha of forests, crops, and pasture, and 61 businesses
(Parliament of Victoria, 2009). The Victorian Bushfires Royal Commission
conservatively values the cost of the 2009 fire at AUS$ 4.4 billion
(Parliament of Victoria, 2010a).
9.2.2.4. Interventions
The Victorian Government had identified the requirement to respond to
predicted heat events in the Sustainability Action Statement and Action
Plan (released in 2006 and revised in January 2009), which committed to
a Victorian Heat Wave Plan involving communities and local governments.
As a part of this strategy, the Victorian Government has established the
heat wave early warning system for metropolitan Melbourne and is
undertaking similar work for regional Victoria. The government is also
developing a toolkit to assist local councils in preparing for a heat wave
response that could be integrated with existing local government public
health and/or emergency management plans (Victorian Government,
2009).
The ‘Prepare, Stay and Defend, or Leave Early’ (SDLE) approach instructs
that residents decide well before a fire whether they will choose to
leave when a fire threatens but is not yet in the area, or whether they
will stay and actively defend their property during the fire.SDLE also
requires residents to make appropriate preparations in advance for
either staying or leaving. Prior to 7 February 2009, the Victorian State
Government devoted unprecedented efforts and resources to informing
the community regarding fire risks. The campaign clearly had benefits,
but there were a number of weaknesses and failures with Victoria’s
information and warning systems (Bushfire CRC, 2009; Parliament of
Victoria, 2010b).
Another key focus during the wildfire season is protecting the reservoirs,
especially the Upper Yarra and Thomson catchments that provide the
majority of Melbourne’s water supply (Melbourne Water, 2009a). During
the February 2009 fires, billions of liters of water were moved from
affected reservoirs to other safe reservoirs to protect Melbourne’s drinking
water from contamination with ash and debris (Melbourne Water, 2009b).
The Victorian Bushfires Royal Commission made wide-ranging
recommendations about the way fire is managed in Victoria. These have
included proposals to replace all single-wire power lines in Victoria, and
new building regulations for bushfire-prone areas (Parliament of Victoria,
2010c).
Chapter 9 Case Studies
498
9.2.2.5. Outcomes/Consequences
Following the findings from the various inquiries into the 2009 Victorian
Bushfires, which found failings in assumptions, policies, and
implementation, a number of far-reaching recommendations were
developed (Parliament of Victoria, 2010c). National responses have been
adopted through the National Emergency Management Committee,
including: (i) revised bushfire safety policies to enhance the roles of
warning and personal responsibility; (ii) increased fuel reduction burning
on public lands; (iii) community refuges established in high-risk areas; (iv)
improved coordination and communication between fire organizations;
(v) modifying the ‘Prepare, stay and defend, or leave early’ approach (now
‘Prepare, act, survive’) to recognize the need for voluntary evacuations
on extreme fire days; and (vi) further ongoing investment in bushfire
research, including a national research center.
9.2.2.6 Lessons Identified
Australia has recognized the need for strengthening risk management
capacities through measures including: (i) prior public campaigns for
risk awareness; (ii) enhanced information and warning systems; (iii)
translation of messages of awareness and preparedness into universal
action; (iv) sharing responsibility between government and the people;
(v) development of integrated plans; and (vi) greater investment in risk
mitigation and adaptation actions.
Predicted changes in future climate will only exacerbate the impact of
other factors through increased likelihood of extreme fire danger days
(Hennessy et al., 2005). We are already seeing the impact of many factors
on wildfires and heat waves, for example, demographic and land use
changes. In the future, a better understanding of the interplay of all the
causal factors is required. The Victorian Bushfires Royal Commission
stated “It would be a mistake to treat Black Saturday as a ‘one off’
event. With populations at the rural-urban interface growing and the
impact of climate change, the risks associated with bushfire are likely to
increase” (Parliament of Victoria, 2010c).
9.2.3. Managing the Adverse Consequences of Drought
9.2.3.1. Introduction
Water is a critical resource throughout the world (Kundzewicz et al.,
2007). Drought can increase competition for scarce resources, cause
population displacements and migrations, and exacerbate ethnic
tensions and the likelihood of conflicts (Barnett and Adger, 2007;
Reuveny, 2007; UNISDR, 2011a). Mediterranean countries are prone to
droughts that can heavily impact agricultural production, cause economic
losses, affect rural livelihoods, and may lead to urban migration
(UNISDR, 2011b). This case study focuses on Syria as one of the
countries that has been affected by drought in recent years (2007-2010)
(Erian et al., 2011).
9.2.3.2. Background
The Eastern Mediterranean region is subject to frequent soil moisture
droughts, and in areas where annual rainfall ranges between 120-150
and 400 mm, rain-fed crops are strongly affected (Erian et al., 2011).
During the period from 1960 to 2006, a severe decrease in annual rainfall
has been documented in some major cities in Syria. These reductions were
related to decreases in spring and winter rainfall (Skaff and Masbate,
2010). The negative trend in precipitation in Syria during the past
century and the beginning of the 21st century is of a similar magnitude
to that predicted by most general circulation models for the Mediterranean
region in the coming decades (Giannakopoulos et al., 2009).
9.2.3.3. Description of Events
Syria is considered to be a dry and semi-arid country (FAO and NAPC,
2010). Three-quarters of the cultivated land depends on rainfall and the
annual rate is less than 350 mm in more than 90% of the overall area
(FAO, 2009; FAO and NAPC, 2010). Syria has a total population of 22
million people, 47% of whom live in rural areas (UN, 2011). The
National Programme for Food Security in the Syrian Arab Republic
reported that in the national economy of Syria, the agricultural and rural
sector is vital, but with the occurrence of frequent droughts, this sector is
less certain of maintaining its contribution of about 20 to 25% of GDP
and employment of about 38 to 47% of the work force (UNRCS and
SARPCMSPC, 2005; FAO and NAPC, 2010).
The prolonged drought, that in 2011 was in its fourth consecutive year,
has affected 1.3 million people, and the loss of the 2008 harvest has
accelerated migration to urban areas and increased levels of extreme
poverty (UN,
2009, 2011; Sowers et al., 2011). During the 2008-2009
winter grain-growing season there were significant losses of both rain-fed
and irrigated winter grain crops (USDA, 2008a). This was exacerbated by
abnormally hot spring temperatures (USDA, 2008a). Wheat production
decreased from 4,041 kt in 2007 to 2,139 kt in 2008, an almost 50%
reduction
(SARPMETT, 2010). Of the farmers who depended on rain-fed
production, most suffered complete or near-total loss of crops (FAO,
2009). Approximately 70% of the 200,000 affected farmers in the rain-
fed areas have produced minimal to no yields because seeds were not
planted due to poor soil moisture conditions or failed germination
(USDA, 2008b; FAO, 2009).
Herders in the region were reported to have lost around 80% of their
livestock due to barren grasslands, and animal feed costs rose by 75%,
forcing sales at 60 to 70% below cost (FAO, 2008). Many farmers and
herders sold off productive assets, eroding their source of livelihoods,
with only few small-scale herders retaining a few animals, possibly as
few as 3 to 10% (FAO, 2009).
Drought has affected the livelihoods of small-scale farmers and herders,
threatening food security and having negative consequences for entire
families living in affected areas (FAO, 2009; UN, 2009). It is estimated
Chapter 9Case Studies
499
that 1.3 million people have been affected by drought with up to
800,000 (75,641 households) being severely affected (FAO, 2009; UN,
2009). Of those severely affected, around 20% (160,000 people) are
considered to be highly vulnerable, including female-headed households,
pregnant women, children 14 and under, those with illness, the elderly,
and the disabled (UN, 2009).
The United Nations (UN) estimates that a large number of the severely
affected population are living below the poverty line (US$ 1/person/day)
(UN, 2009). When combined with an increase in the price of food and basic
resources, this reduced income has resulted in negative consequences
for whole households (FAO, 2009). Many cannot afford basic supplies or
food, which has led to a reduction in their food intake, the selling of assets,
a rise in the rate of borrowing money, the degradation of land, urban
migration, and children leaving school (FAO, 2009; UN, 2009; Solh, 2010).
The UN assessment mission stated that the reasons for removing children
from school included financial hardship, increased costs of transport,
migration to cities, and the requirement for children to work to earn
extra income for families (UN, 2009). Consequently, due to poor food
consumption, the rates of malnutrition have risen between 2007 and 2008,
with the Food and Agriculture Organization (FAO) estimating a doubling
of malnutrition cases among pregnant women and children under five
(FAO, 2008). Due to inadequate consumption of micro- and macronutrients
in the most-affected households, it has been estimated that the average
diet contains less than 15% of the recommended daily fat intake and
50% of the advised energy and protein requirements (UN, 2009).
One of the most visible effects of the drought was the large migration of
between 40,000 and 60,000 families from the affected areas (UN, 2009;
Solh, 2010; Sowers and Weinthal, 2010). In June 2009, it was estimated
that 36,000 households had migrated from the Al-Hasakah Governorate
(200,000-300,000 persons) to the urban centers of Damascus, Dara’a,
Hama, and Aleppo (UN, 2009; Solh, 2010). Temporary settlements and
camps were required, creating further strains on resources and public
services, that had already been attempting to support approximately
1 million Iraqi refugees (UN, 2009; Solh, 2010). In addition, migration
leads to worse health, educational, and social indicators among the
migrant population (IOM, 2008; Solh, 2010).
Deficits in water resources exceeding 3.5 billion cubic meters have arisen
in recent years due to growing water demands and drought
(FAO and
NAPC, 2010;
SARPMETT, 2010). Interventions by a project further
upstream to control the flow of the Euphrates and Tigris rivers have been
initiated and these have had a significant impact on water variability
downstream in Iraq and Syria, which, added to the severe drought, have
caused these rivers to flow at well-below normal levels (USDA, 2008a;
Daoudy, 2009; Sowers et al., 2011).
9.2.3.4. Interventions
The UN Syria Drought Response Plan was published in 2009. It was
designed to address the emergency needs of, and to prevent further impact
on, the 300,000 people most affected by protracted drought (FAO,
2009). The Response Plan identified as its strategic priorities the rapid
provision of humanitarian assistance, the strengthening of resilience to
future drought and climate change, and assisting in the return process
and ensuring socioeconomic stability among the worst-affected groups
(UN, 2009). Syria also welcomed international assistance provided to
the drought-affected population through multilateral channels (Solh,
2010). Various loans to those affected, including farmers and women
entrepreneurs, were provided (UN, 2009).
9.2.3.5. Outcomes/Consequences
A combination of actions including food and agriculture assistance,
supplemented by water and health interventions, and measures aimed
at increasing drought resilience, were identified as required to allow
affected populations to remain in their villages and restart agricultural
production (UN, 2009). Ongoing interventions with the aim of reducing
vulnerability and increasing resilience to drought were summarized by
the UN Syrian Drought Response Plan (UN, 2009) and FAO (FAO, 2009).
These interventions were aimed at providing support through four main
approaches: (1) the rapid distribution of wheat, barley, and legume seeds
to 18,000 households in the affected areas, potentially assisting 144,000
people; (2) sustaining the remaining asset base of the approximately
20,000 herders by providing animal feed and limited sheep restocking
to approximately 1,000 herders; (3) the development of a drought early
warning system to facilitate the government taking early actions before
serious and significant losses occur and to develop this to ensure
sustainability; and (4) building capability to implement the national
drought strategy by developing and addressing all stages of the disaster
management cycle (FAO, 2009). Conservation agriculture (which has been
defined as no-tillage, direct drilling/seeding, and drilling/seeding through
a vegetative cover) is considered to be a way forward for sustainable
land use (Stewart et al., 2008; Lalani 2011). However, how to take this
forward has caused considerable debate (Stewart et al., 2008)
.
9.2.3.6. Lessons Identified
The need for the UN Syrian Drought Response Plan was identified and
has facilitated the understanding of the work programs and links to the
interventions listed in Case Study 9.2.3.4 (UN, 2009). Other response
strategies that have been considered include:
Development of capacities to identify, assess, and monitor drought
risks through national and local multi-hazard risk assessment;
building systems to monitor, archive, and disseminate data (Lalani,
2011), taking into account decentralization of resources, community
participation, and regional early warning systems and networks
(UNISDR, 2011a).
Integrating activities in the national strategy for CCA and DRR,
including drought risk loss insurance; improved water use efficiency;
adopting and adapting existing water harvesting techniques;
integrating use of surface and groundwater; upgrading irrigation
Chapter 9 Case Studies
500
practices at both the farm level and on the delivery side; developing
crops tolerant to salinity and heat stress; changing cropping patterns;
altering the timing or location of cropping activities; diversifying
production systems into higher value and more efficient water use
options; and capacity building of relevant stakeholders in vulnerable
national and local areas (Abou Hadid, 2009; El-Quosy, 2009).
Building resilience through knowledge, advocacy, research, and
training by making information on drought risk accessible (UNISDR,
2007a), and having any adaptation measures be developed as
part of, and closely integrated into, overall and country-specific
development programs and strategies that should be understood
as a ‘shared responsibility’ (Easterling et al., 2007). This could be
achieved through educational material and training to enhance
public awareness (UN, 2009).
9.2.4. Recent Dzud Disasters in Mongolia
9.2.4.1. Introduction
This case study introduces dzud disaster: the impacts, intervention
measures, and efforts toward efficient response using the example of
two events that occurred in Mongolia in 1999-2002 and 2009-2010.
Mongolia is a country of greatly variable, highly arid and semi-arid climate,
with an extensive livestock sector dependent upon access to grasslands
(Batima and Dagvadorj, 2000; Dagvadorj et al., 2010; Marin, 2010). The
Mongolian term dzud denotes unusually extreme weather conditions
that result in the death of a significant number of livestock over large
areas of the country (Morinaga et al., 2003; Oyun, 2004). Thus, the term
implies both exposure to such combinations of extreme weather
conditions but also the impacts thereof (Marin, 2010).
9.2.4.2. Background
The climate of Mongolia is continental with sharply defined seasons,
high annual and diurnal temperature fluctuations, and low rainfall
(Batima and Dagvadorj, 2000). Summer rainfall seldom exceeds 380 mm
in the mountains and is less than 50 mm in the desert areas (Dagvadorj
et al., 2010). Dzud is a compound hazard (see Section 3.1.3 for discussion
of compound events) occurring in this cold dry climate, and encompasses
drought, heavy snowfall, extreme cold, and windstorms. It can last all year
round and can cause mass livestock mortality and dramatic socioeconomic
impacts – including unemployment, poverty, and mass migration from
rural to urban areas, giving rise to heavy pressure on infrastructure and
social and ecosystem services (Batima and Dagvadorj, 2000; Batjargal
et al., 2001; Oyun, 2004; AIACC, 2006; Dagvadorj et.al., 2010).
There are several types of dzud. If there is heavy snowfall, the event is
known as a white dzud, conversely if no snow falls, a black dzud occurs,
which results in a lack of drinking water for herds (Morinaga et al., 2003;
Dagvadorj et al., 2010). The trampling of plants by passing livestock
migrating to better pasture or too high a grazing pressure leads to a
hoof dzud, and a warm spell after heavy snowfall resulting in an icy
crust cover on short grass blocking livestock grazing causes an iron dzud
(Batjargal et al., 2001; Marin, 2008). Livestock have been the mainstay
of Mongolian agriculture and the basis of its economy and culture for
millennia (Mearns, 2004; Goodland et al., 2009). This sector is likely to
continue to be the single most important sector of the economy in terms
of employment (Mearns, 2004; Goodland et al., 2009).
In the last decades, dzuds occurred in 1944-1945, 1954-1955, 1956-1957,
1967-1968, 1976-1977, 1986-1987, 1993-1994, and 1996-1997, with
further dzuds discussed below (Morinaga et al., 2003; Sternberg, 2010).
The dzud of 1944-1945 was a record for the 20th century with mortality
of one-third of Mongolia’s total livestock (Batjargal et al., 2001). The
2009-2010 dzud caused similarly high animal mortality (NSO, 2011). The
large losses of animals in dzud events demonstrates that Mongolia as a
whole has low capacity to combat natural disaster (Batjargal et al.,
2001). These potential losses are considered to be beyond the financial
capacity of the government and the domestic insurance market
(Goodland et al., 2009).
9.2.4.3. Description of Events –
Dzud of 1999-2002 and of 2009-2010
Dzud disasters occurred in 1999-2002 and 2009-2010, causing social and
economic impacts. These disasters occurred as a result of environmental
and human-induced factors. The environmental factors included drought
resulting in very limited pasture grass and hay with additional damage
to pasture by rodents and insects (Batjargal et al., 2001; Begzsuren et al.,
2004; Saizen et al., 2010). Human factors included budgetary issues for
preparedness in both government and households, inadequate pasture
management and coordination, and lack of experience of new and/or
young herders (Batjargal et al., 2001). Climatic factors contributing to
both dzuds were summer drought followed by extreme cold and snowfall
in winter. However the autumn of 1999-2000 brought heavy snowfall
and unusual warmth with ice cover, while the winter and spring of
2009-2010 also brought windstorms. Summer drought was a more
significant contributor to the 1999-2000 dzud (Batjargal et al., 2001),
while winter cold was more extreme in the 2009-2010 dzud.
9.2.4.3.1. Dzud of 1999-2002
The dzud began with summer drought followed by heavy snowfall and
unusual warmth with ice cover in the autumn and extreme cold and
snowfall in the winter. The sequence of events was as follows (Batjargal
et al., 2001):
Drought: In the summer of 1999, 70% of the country suffered
drought. Air temperature reached 41 to 43°C, exceeding its highest
value recorded at meteorological stations since the 1960s. The
condition persisted for a month and grasslands dried up. As a
result, animals were unfit for the winter, with insufficient haymaking
for winter preparedness.
Chapter 9Case Studies
501
Iron dzud: Autumn brought early snowfall and snow depth
reached 30 to 40 cm, even 70 to 80 cm in some places. Heavy
snowfall exceeding climatic means was recorded in October.
Moreover, a warming in November and December by 1.7 to 3.5°C
above the climatic mean resulted in snow cover compaction and
high density, reaching 0.37 g cm
-3
, and ice cover formation, both
of which blocked livestock pasturing.
White dzud: In January, air temperature dropped down to -40 to
-50°C over the western and northern regions of the country. The
monthly average air temperature was lower than climatic means
by 2 to 6°C. The cold condition persisted for two months. Abundant
snowfall resulted with 80% of the country’s territory being covered
in snow of 24 to 46 cm depth.
Black dzud: Lack of snowfall in the Gobi region and Great Lake
depression caused water shortages for animals.
Hoof dzud: The improper pasture management led to unplanned
concentration of a great number of livestock in a few counties in
the middle and south Gobi provinces that were not affected by
drought and snowfall.
Animals were weakened as a result of long-lasting climatic hardship
and forage shortage of this dzud (Batjargal et al., 2001). After three
years of dzuds that occurred in sequence, the country had lost nearly
one-third (approximately 12 million) of its livestock and Mongolia’s
national gross agricultural output in 2003 decreased by 40% compared
to that in 1999 (Mearns, 2004; Oyun, 2004; AIACC, 2006; Lise et al.,
2006; Saizen et al., 2010). It was reported that in 1998 there were an
estimated 190,000 herding households but as a result of the dzud,
11,000 families lost all their livestock (Lise et al., 2006). Thus the dzud
had severe impacts on the population and their livelihoods, including
unemployment, poverty, and negative health impacts (Batjargal et al.,
2001; Oyun, 2004; AIACC, 2006; Morris, 2011).
9.2.4.3.2. Dzud of 2009-2010
In the summer of 2009, Mongolia suffered drought conditions, restricting
haymaking and foraging (UNDP Mongolia, 2010; Morris, 2011). Rainfall
at the end of November became a sheet of ice, and, in late December,
19 of 21 provinces recorded temperatures below -40°C; this was followed
by heavy and continuous snowfall in January and February 2010
(Sternberg, 2010; UNDP Mongolia, 2010). Over 50% of all the country
herders’ households and their livestock were affected by the dzud
(Sternberg, 2010). By April, 75,000 herder families had lost all or more
than half their livestock (Sternberg, 2010).
9.2.4.4. Interventions
9.2.4.4.1. Dzud of 1999-2002
The government of Mongolia issued the order for intensification of winter
preparedness in August 1999, but allocated funding for its implementation
in January 2000, by which time significant animal mortality had already
occurred (Batjargal et al., 2001). The government then appealed to its
citizens and international organizations for assistance and relief, including
distribution of money, fodder, medicine, clothes, flour, rice, high energy
and high protein biscuits for children, health and veterinary services,
medical equipment, and vegetable seeds (Batjargal et al., 2001).
Capacity-building activities through mass media campaigns were also
carried out, focused on providing advice on methods of care and feeding
for weak animals (Batjargal et al., 2001).
Herders rely upon traditional informal coping mechanisms and ad hoc
support from government and international agencies (Mahul and Skees,
2005). For affected areas, after immediate relief, the main longer term
support has conventionally been through restocking programs (Mahul
and Skees, 2005). Evaluation has shown that these can be expensive,
relatively inefficient, and fail to provide the right incentives for herders
(Mahul and Skees, 2005). Restocking in areas with drought, poor pasture
condition, and unfit animals can actually increase livestock vulnerability
in the following year (Mahul and Skees, 2005) as a result of greater
competition for scarce resources.
The government has prioritized the livestock sector with parliament-
approved state policy (MGH, 2003) and, with support from donors,
responded to dzud disasters with reforms that include greater flexibility
in pasture land tenure, coupled with increased investment in rural
infrastructure and services (Mahul and Skees, 2005). However, livestock
sector reforms and approaches have not yet proved sufficient to cope
with catastrophic weather events (Mahul and Skees, 2005). Although
the State Reserves Agency is working to reduce the effects of dzud,
catastrophic livestock mortality persists (Mahul and Skees, 2005).
9.2.4.4.2. Dzud of 2009-2010
At a local level: The National Climate Risk Management Strategy and
Action Plan (MMS, 2009) sets a goal of building climate resilience at the
community level through reducing risk and facilitating adaptation by: (i)
improving access to water through region-specific activities such as
rainwater harvesting and creation of water pools from precipitation and
flood waters, for use for animals, pastureland, and crop irrigation purposes;
(ii) improving the quality of livestock by introducing local selective
breeds with higher productivity and more resilient to climate impacts;
(iii) strengthened veterinarian services to reduce animal diseases and
parasites and cross-border epidemic infections; and (iv) using traditional
herding knowledge and techniques for adjusting animal types and herd
structure to make appropriate for the carrying capacity of the pastureland
and pastoral migration patterns. The formation of herders’ community
groups and establishment of pasture co-management teams (Ykhanbai et
al., 2004), along with better community-based disaster risk management,
could also facilitate effective DRR and CCA (Baigalmaa, 2010).
At a national level: Mongolia’s millennium development priorities
clearly state an aim to adapt to climate change and desertification and
Chapter 9 Case Studies
502
to implement strategies to minimize negative impacts (Mijiddorj,
2008; UNDP Mongolia, 2009a). The recent national CCA report outlines
government strategy priorities as: (i) education and awareness campaigns
among the decisionmakers, rural community, herders, and the general
public; (ii) technology and information transfer to farmers and herdsmen;
(iii) research and technology to ensure the development of agriculture
that could successfully deal with various environmental problems; and
(iv) improved coordination of stakeholders’ activities based on research,
inventory, and monitoring findings (Dagvadorj et al., 2010). The
management of risk in the livestock sector requires a combination of
approaches. Traditional herding and pastoral risk reduction practices
can better prepare herders for moderate weather events. For country-
wide dzud events, however, high levels of livestock mortality are often
unavoidable, even for the most experienced herders, and pasture
resource and herd management must be complemented by risk-financing
mechanisms that provide herders with instant liquidity in the aftermath
of a disaster (Goodland et al., 2009).
At an international level: As Mongolia is a country extremely prone
to natural disasters, addressing climate change risks is a priority in
Mongolia. In 2009, the Mongolian Government undertook the project
‘Strengthening the Disaster Mitigation and Management Systems in
Mongolia’ under the National Emergency Management Agency (UNDP
Mongolia, 2009b; Sternberg, 2010).
9.2.4.5. Consequences
The most critical consequences of dzud are increased poverty and mass
migration from rural to urban and from remote to central regions (Oyun,
2004; Dagvadorj et al., 2010). According to national statistics there has
been a continuous increase in poverty over the last decade (NSO, 2011).
In response to the climatic hardship, a growing proportion of the rural
population has migrated to urban areas and the central region (Dagvadorj
et al., 2010; UNDP Mongolia, 2010). Livestock-herding families are
forced to migrate because of poverty caused by loss of livestock from
catastrophic weather events (Sternberg, 2010). Besides poverty, there
are reasons why members of herding families may wish to leave the
livestock sector, including obtaining a better education for their children
and access to health care (Mahul and Skees, 2005). Many migrants
travel from Western Mongolia to the capital city Ulaanbaatar (Saizen et
al., 2010; Sternberg, 2010).
9.2.4.6. Lessons Identified
Current policies and measures are mainly limited to post-disaster
government relief and restocking activities with donors’ funding and
individual herder’s traditional knowledge and practices (Batjargal et al.,
2001; AIACC, 2006). These can be insufficient to avoid, prepare for, and
respond to a dzud (Goodland et al., 2009). Various practices have been
identified as effective for DRR, and could further contribute to promote
CCA. These include localized seasonal climate prediction and improvement
of early warning (Morinaga et al., 2003; MMS, 2009), risk-insuring systems
(Skees and Enkh-Amgalan, 2002; Mahul and Skees, 2005), and policy
(Batjargal et al., 2001; AIACC, 2006; Goodland et al., 2009).
At present, adaptation occurs through increased mobility of herders in
search of better pasture for their animals in dzud disasters (Batjargal et
al., 2001), and as a response to changed rain patterns occurring over
small areas, which the herders call ‘silk embroidery rain’ (Marin, 2010).
Livelihood diversification to create resilient livelihoods for herders has
also been seen as being effective for building climate resilience (Mahul
and Skees, 2005; Borgford-Parnell, 2009; MMS, 2009, Dagvadorj et al.,
2010).
9.2.5. Cyclones: Enabling Policies and
Responsive Institutions for Community Action
9.2.5.1. Introduction
Tropical cyclones, also called typhoons and hurricanes, are powerful
storms generated over tropical and subtropical waters. Their extremely
strong winds damage buildings, infrastructure and other assets; the
torrential rains often cause floods and landslides; and high waves and
storm surge often lead to extensive coastal flooding and erosion – all of
which have major impacts on people. Tropical cyclones are typically
classified in terms of their intensity, based on measurements or estimates
of near-surface wind speed (sometimes categorized on a scale of 1 to 5
according to the Saffir-Simpson scale). The strongest storms (Categories
3, 4, and 5) are comparatively rare but are generally responsible for the
majority of damage (Section 3.4.4).
The focus of this case study is the comparison between the response to
Indian Ocean cyclones in Bangladesh (Sidr in 2007) and in Myanmar
(Nargis in 2008) in the context of the developments in preparedness
and response in Bangladesh resulting from experiences with Cyclone
Bhola in 1970, Gorky in 1991, and other events. To provide a more
global context, the impacts and responses to 2005 Hurricanes Stan and
Wilma in Central America and Mexico are also discussed. These clearly
demonstrate that climate change adaptation efforts can be effective in
limiting the impacts from extreme tropical cyclone events by use of
disaster risk reduction methods.
Changes in tropical cyclone activity due to anthropogenic influences are
discussed in Section 3.4.4. There is low confidence that any observed long-
term increases in tropical cyclone activity are robust, after accounting for
past changes in observing capabilities. The uncertainties in the historical
tropical cyclone records, the incomplete understanding of the physical
mechanisms linking tropical cyclone metrics to climate change, and the
degree of tropical cyclone variability provide only low confidence for the
attribution of any detectable changes in tropical cyclone activity to
anthropogenic influences. There is low confidence in projections of
changes in tropical cyclone genesis, location, tracks, duration, or areas
Chapter 9Case Studies
503
of impact. Based on the level of consistency among models, and physical
reasoning, it is likely that tropical cyclone-related rainfall rates will
increase with greenhouse warming. It is likely that the global frequency
of tropical cyclones will either decrease or remain essentially
unchanged. An increase in mean tropical cyclone maximum wind speed
is likely, although increases may not occur in all tropical regions. While
it is likely that overall global frequency will either decrease or remain
essentially unchanged, it is more likely than not that the frequency of
the most intense storms will increase substantially in some ocean
basins. Although there is evidence that surface sea temperature (SST) in
the tropics has increased due to increasing greenhouse gases, the
increasing SST does not yet have a fully understood physical link to
increasingly strong tropical cyclones (Section 3.4.4).
9.2.5.2. Indian Ocean Cyclones
Although only 15% of world tropical cyclones occur in the North Indian
Ocean (Reale et al., 2009), Bangladesh and India account for 86% of
mortality from tropical cyclones (UNISDR, 2009c). The 2011 Global
Assessment Report (UNISDR, 2011b) provides strong evidence that
weather-related mortality risk is highly concentrated in countries with
low GDP and weak governance. Many of the countries exposed to
tropical cyclones in the North Indian Ocean are characterized by high
population density and vulnerability and low GDP.
9.2.5.2.1. Description of events – Indian Ocean cyclones
In 2007, Cyclone Sidr made landfall in Bangladesh on 15 November and
caused over 3,400 fatalities (Paul, 2009). Cyclone Nargis hit Myanmar
on 2 May 2008 and caused over 138,000 fatalities (CRED, 2009; Yokoi
and Takayabu, 2010), making it the eighth deadliest cyclone ever recorded
(Fritz et al., 2009). Sidr and Nargis were both Category 4 cyclones of
similar severity; affecting coastal areas with a comparable number of
people exposed (see Table 9-2). Although Bangladesh and Myanmar
both are considered least developed countries (Giuliani and Peduzzi,
2011), these two comparable events had vastly different impacts. The
reasons for the differences follow.
9.2.5.2.2. Interventions – Indian Ocean cyclones
Bangladesh has a significant history of large-scale disasters (e.g., Cyclones
Bhola in 1970 and Gorky in 1991; see Table 9-2). The Government of
Bangladesh has made serious efforts aimed at DRR from tropical cyclones.
It has worked in partnership with donors, nongovernmental organizations
(NGOs), humanitarian organizations and, most importantly, with
coastal communities themselves (Paul, 2009).
First, they constructed multi-storied cyclone shelters with capacity for
500 to 2,500 people (Paul and Rahman, 2006) that were built in coastal
regions, providing safe refuge from storm surges for coastal populations.
Also, killas (raised earthen platforms), which accommodate 300 to 400
livestock, have been constructed in cyclone-prone areas to safeguard
livestock from storm surges (Paul, 2009).
Second, there has been a continued effort to improve forecasting and
warning capacity in Bangladesh. A Storm Warning Center has been
established in the Meteorological Department. System capacity has been
enhanced to alert a wide range of user agencies with early warnings
and special bulletins, soon after the formation of tropical depressions in
the Bay of Bengal. Periodic training and drilling practices are conducted
at the local level for cyclone preparedness program (CPP) volunteers for
effective dissemination of cyclone warnings and for raising awareness
among the population in vulnerable communities.
Third, the coastal volunteer network (established under the CPP) has
proved to be effective in disseminating cyclone warnings among the
coastal communities. These enable time-critical actions on the ground,
including safe evacuation of vulnerable populations to cyclone shelters
(Paul, 2009). With more than sevenfold increase in cyclone shelters and
twofold increase in volunteers from 1991 to 2007, 3 million people were
safely evacuated prior to landfall of Sidr in 2007 (Government of
Bangladesh, 2008).
In addition, a coastal reforestation program, including planting in the
Sundarbans, was initiated in Bangladesh in the late 1960s, covering
riverine coastal belt and abandoned embankments (Saenger and Siddiqi,
1993). Sidr made landfall on the western coast of Bangladesh, which is
Chapter 9 Case Studies
Bhola 1970 223 3 1 Unknown
Gorky 1991 260 4 15.4 138,000 1.8
Sidr 2007 245 4 4,200 2.3
Nargis 2008 ~ 4 235 4 138,000 4.0
Stan
a
2005 Negligible 130 1 1,726 3.9
Wilma 2005 12.8 295 5 10
b
62 (8 in Mexico) 29 (7.5 in Mexico)
6 - 9
6 - 7.5
5 - 6
8 - 10
2 - 8
3 - 8
300,000 - 500,000
Notes:
a
Most of damage and mortality caused by landslides and river flooding.
b
Affecting Jamaica, Bahamas, Haiti, Cayman Islands, Belize, Honduras, El Salvador, Nicaragua, Honduras, Yucatán Peninsula (Mexico), and Florida (USA).
Cyclone
Event
Year Storm Surge (m)
Maximum Wind
Speed (km h
-1
)
Category
(Saffir-
Simpson)
Number of Affected
People
(approximate in millions)
Mortality
(approximate)
Damages
(US$ billion)
Table 9-2 | Key data for extreme cyclones in Bangladesh, Myanmar, and Mexico. Sources: García et al., 2006; National Hurricane Center, 2006; Government of Bangladesh, 2008;
Karim and Mimura, 2008; Webster, 2008; CRED, 2009; Paul, 2009; Giuliani and Peduzzi, 2011.
504
lined by the world’s largest mangrove forest, the Sundarbans. This
region is the least-populated coastal area in the country and has been
part of a major reforestation effort in recent years (Hossain et al., 2008).
The Sundarbans provided an effective attenuation buffer during Sidr,
greatly reducing the impact of the storm surge (Government of
Bangladesh, 2008).
In contrast to Bangladesh, Myanmar has very little experience with
previous powerful tropical cyclones. The landfall of Nargis was the first
time in recorded history that Myanmar experienced a cyclone of such a
magnitude and severity (Lateef, 2009) and little warning was provided.
9.2.5.2.3. Outcomes – Indian Ocean cyclones
Despite Nargis being both slightly less powerful and affecting fewer
people than Sidr, it resulted in human losses that were much higher.
Bangladesh and Myanmar are both very poor countries with low levels
of HDI (World Bank, 2011a). The relatively small differences in poverty
and development cannot explain the discrepancy in the impacts of Sidr
and Nargis. However, the governance indicators developed by the World
Bank (Kaufmann et al., 2010) suggest significant differences between
Bangladesh and Myanmar in the quality of governance, notably in voice
and accountability, rule of law, regulatory quality, and government
effectiveness. Low quality of governance, and especially voice and
accountability, has been highlighted as a major vulnerability component
for human mortality due to tropical cyclones (Peduzzi et al., 2009).
9.2.5.3. Mesoamerican Hurricanes
9.2.5.3.1. Description of events – Mesoamerican hurricanes
Central America and Mexico (Mesoamerica) are heavily affected by
strong tropical storms. From 1-13 October 2005, Hurricane Stan affected
the Atlantic coast of Central America and the Yucatan Peninsula in Mexico.
Stan was a relatively weak storm that only briefly reached hurricane
status. It was associated with a larger non-tropical storm system that
resulted in torrential rains and caused debris flows, rockslides, and
widespread flooding. Guatemala reported more than 1,500 fatalities and
thousands of missing people. El Salvador reported 69 fatalities while
Mexico reported 36 (CRED, 2009). Wilma hit one week later (19-24
October). It was an intense cyclone in the Atlantic (National Hurricane
Center, 2006; Table 9-2), with winds reaching a speed of 297 km h
-1
.
Wilma caused 12 fatalities in Haiti, 8 in Mexico, and 4 in the United States.
Most residents in western Cuba, and tourists and local inhabitants in
the Yucatan Peninsula in Mexico, were evacuated (CRED, 2009).
9.2.5.3.2. Interventions – Mesoamerican hurricanes
While Stan mainly affected the poor indigenous regions of Guatemala,
El Salvador, and Chiapas, Wilma affected the international beach resort
of Cancun. A joint study of Mexico’s response to the hurricanes funded
by the World Bank and conducted through the Economic Commission
for Latin America and the Caribbean (ECLAC, 2006) and its Commission
for Latin American and the Caribbean and the Mexican National Center
for Prevention of Disasters (García et al., 2006) showed that Stan
caused about US$ 2.2 billion damage in that country, 65% of which
were direct losses and 35% due to future impacts on agricultural
production. About 70% of these damages were reported in the state of
Chiapas (Oswald Spring, 2011), representing 5% of the GDP of the state.
Comparing the management of the two hurricanes by the Mexican
authorities, in the same month and year, highlights important issues in
DRM. Evacuation of areas in Mexico affected by Stan only started during
the emergency phase, when floods in 98 rivers had already affected
800 communities. 100,000 people fled from the mountain regions to
improvised shelters – mostly schools – and ‘guest families’ (Oswald
Spring, 2010). In comparison, following the early alert for Wilma, people
were evacuated from dangerous places, most tourists were moved to
safe areas, and local inhabitants and remaining tourists were taken to
shelters (García et al., 2006). Before the hurricane hit the coast, heavy
machines and emergency groups were mobilized in the region, to
reestablish water, electricity, communications, and health services
immediately after the event. After the disaster, all ministries were
involved in reopening the airport and tourist facilities as quickly as
possible. By December, most hotels were operating, and the sand lost
from the beaches had been reestablished (Oswald Spring, 2011).
9.2.5.3.3. Outcomes – Mesoamerican hurricanes
Comparing government responses to these two hurricanes in the same
month, it is possible to note vastly different official actions in terms of
early warning, evacuation, and reconstruction (Oswald Spring, 2011).
The federal institutions in charge of DRM functioned well during
Hurricane Wilma. A massive recovery support strategy restored almost
all services and hotels in Cancun within two months, with a significant
portion of costs being covered by insurance companies (García et al.,
2006). The government response to Stan left the poor indigenous
population with limited advice, insufficient disaster relief, and scant
reconstruction support, especially among the most marginal groups
(Oswald Spring, 2011).
9.2.5.4. Lessons Identified
Comparative studies of disaster risk management practices for tropical
cyclones demonstrate that choices and outcomes for response to climatic
extreme events are triggered by multiple interacting processes, and
competing priorities. Indigenous, poor, and illiterate people have low
resilience, limited resources, and are highly exposed without early
warnings and DRM. Government response to similar extreme events
may be quite different in neighboring countries, or even within the same
country.
Chapter 9Case Studies
505
Tropical cyclone DRM strategies in coastal regions that create protective
measures, and anticipate and plan for extreme events, along with
continuing changes in vulnerability and in causal processes, increase the
resilience of potentially exposed communities. International cooperation
and investment in the following measures are essential in improving the
capacity of developing nations in coping with extreme tropical cyclone
events:
Improvement of forecasting capacity and implementation of
improved early warning systems (including evacuation plans and
infrastructure)
Protection of healthy ecosystems
Post-disaster support service to dispersed communities
Transparent management of recovery funds directly with the victims.
Awareness, early warnings and evacuation, hurricane experience, disaster
funds, and specialized bodies reduce the impact of tropical cyclones on
socially vulnerable people. Good governance and participation of
people at risk in the decisionmaking process may overcome conflicting
governmental priorities. Disaster risk management is most effectively
pursued by understanding the diverse ways in which social processes
contribute to the creation, management, and reduction of disaster risk
with the involvement of people at risk. A development planning
perspective that includes disaster risk management as an integral part
of the development framework is the key to a coherent strategy for the
reduction of risk associated with extreme weather events.
9.2.6. Managing the Adverse Consequences of Floods
9.2.6.1. Introduction
Floods are a major natural hazard in many regions of the world (Ahern
et al., 2005). Averaged over 2001 to 2010, floods and other hydrological
events accounted for over 50% of the disasters (Guha-Sapir et al.,
2011); for example, it was reported that, in 2007, flooding worldwide
accounted for four of the top five deadliest natural disasters (Subbarao
et al., 2008). Currently about 800 million people live in flood-prone areas
and about 10% are annually exposed to floods (Chapter 4; Peduzzi et
al., 2011; UNISDR, 2011b). Causes of floods are varied, but may occur
as a result of heavy, persistent, and sustained rainfall or as a result of
coastal flooding (Ahern et al., 2005; see also Section 3.5.2). Flooding
impacts are wide ranging, potentially interrupting food and water
supplies, affecting economic development, and causing acute as well as
subsequent long-term health impacts (Ahern et al., 2005; Subbarao et
al., 2008). It is important to study flooding events to develop or enhance
reliable approaches to risk reduction as well as systems for forecasting and
informing the population, in order to help minimize negative consequences
(ICSU, 2008). This case study examines the impacts on the population
and economy of Mozambique from the 2000 and 2007 flooding events.
Effective functioning of DRR and DRM programs at all levels can help to
reduce the risks from extreme events including floods (UNISDR, 2005a).
These programs operate best with a combination of local, national, and
international strategies (Hellmuth et al., 2007; UNISDR, 2011a). A
variety of strategies have been used to reduce the impact of floods. For
example, dams and sea walls prevent flooding of coastal areas but are
expensive and difficult to maintain and these facilities can be breached
(ProAct Network, 2008). Furthermore, urban drainage systems are
recognized as an important tool to reduce urban flood risk, but less
than half (46%) of low-income countries have invested in drainage
infrastructure in flood-prone areas (UNISDR, 2011b). Timely flood
warnings in many countries have been developed as part of DRR and
DRM programs (Case Study 9.2.11).
The Global Assessment Report (UNISDR, 2011b) reported that the 2000
floods in Mozambique are one of the four examples of large disasters
that have highlighted DRM capacity gaps that have led to institutional
and legislative changes.
9.2.6.2. Background
Mozambique has high socioeconomic vulnerability with approximately
50% of its population of 21 million living below the poverty line (see
Sections 2.3 and 2.5; WMO, 2011a; World Bank, 2011b). Its development
has been restricted by previous civil war and conflict with neighboring
South Africa. Further examples of its vulnerability include rising
HIV/AIDS rates, an almost 70% female illiteracy rate, and most of the
population depending on subsistence farming (Hellmuth et al., 2007;
World Bank, 2011b).
Geographic position and climatic factors contribute to Mozambique’s
high physical vulnerability. Mozambique has a 2,700-km coastline and
the whole country and neighboring countries are subjected to cyclones
and resultant flooding (Hellmuth et al., 2007; WMO, 2011a; World Bank,
2011b). Nine of the 11 rivers in Mozambique are transboundary (Hellmuth
et al., 2007) making its location downstream more susceptible to rainfall
events across a large region such that increases in river levels and flows
in neighboring countries can result in or exacerbate floods. Therefore
the development and operating of early warning and flood control systems
in Mozambique depend on a close collaboration with other countries of
the Southern Africa Development Community and its protocol on shared
watercourse systems (SADC, 2000).
The World Bank (2005a) reported that Mozambique experienced 12 major
floods, 9 major droughts, and 4 major cyclone events between 1965 and
1998. In 1999, a national government policy on disaster management
was articulated and a National Institute for Disaster Management
(NIDM), with an emphasis on coordination rather than delivery, created
(World Bank, 2005a).
9.2.6.3. Description of Events – 2000 Floods in Mozambique
In February 2000, catastrophic floods caused the loss of more than 700
lives with over half a million people losing their homes, and more than
Chapter 9 Case Studies
506
4.5 million affected (Mirza, 2003; Hellmuth et al., 2007; WMO, 2011a;
World Bank, 2011b).
The flooding was the result of a cascade of events. It started with
above-
average rainfall in southern Mozambique and adjacent countries
from October to December 1999 (Hellmuth et al., 2007). Exacerbating
the situation was the series of cyclones Astride, Connie, Eline, and Gloria
with the main impact coming from cyclone Eline (UNESC, 2000; Asante
et al., 2007; Hellmuth et al., 2007). Cyclone Eline, after tracking over
7,000 km west across the tropical south Indian Ocean (Reason and Keibel,
2004), made landfall on 22 February 2000, crossing the Mozambique
coastline and moving over the headwater basins of the Limpopo River,
making a critical situation worse.
The rainfall that occurred over Mozambique and the northeastern parts
of South Africa and Zimbabwe was exceptional; record flooding ensued
downstream on the Limpopo and Zambezi rivers (Carmo Vaz, 2000;
Kadomura, 2005), and in parts of the Sabie catchment the return period
was in excess of 200 years (Smithers et al., 2001).
As a result of the floods it was reported that many small towns and
villages remained under water for approximately two months (Hellmuth
et al., 2007). Access roads were rendered impassable with railways, bridges,
water management systems (including water intake and treatment
plants), and more than 600 primary schools damaged or destroyed
(UNESC, 2000; Dyson and van Heerden, 2001; Reason and Keibel, 2004).
The UN World Food Programme reported that Mozambique lost 167,000
ha of agricultural land (FAO and WFP, 2000). Dams were overwhelmed;
for example, the total inflow to Massingir reservoir between January
and March was approximately eight times the storage capacity of the
reservoir at that time (Carmo Vaz, 2000).
Although floodwaters can wash away breeding sites and, hence, reduce
mosquito-borne disease transmission (Sidley, 2000), the collection of
emergency clinic data and interviews of 62 families found that the
incidence of malaria was reported as increasing by a factor of 1.5 to 2.0.
Diarrhea also increased by a factor of 2 to 4 (Kondo et al., 2002).
The government declared an emergency, mobilized its disaster response
mechanisms, and made appeals for assistance from other countries
(Hellmuth et al., 2007).The enormous material damage and human
losses during the floods in Mozambique in 2000 were associated with
the following problems:
Institutional problems: It was only in 1999 that the National
Policy on Disaster Management in Mozambique began to shift
from a reactive to a proactive approach, with an aim to develop a
culture of prevention (Asante et al., 2005; Hellmuth et al., 2007).
Technological problems: In 2000, in Mozambique, there were
problems with the installation and maintenance of in situ gauging
equipment due to financial constraints. In addition, the hydrological
and precipitation gauges were washed away and many key stations
were destroyed, leaving Mozambican water authorities with no
source of information on the actual magnitude of floodwaters
(Dyson and van Heerden, 2001; Smithers et al., 2001; Asante et
al., 2005).
Financial problems:The UN Economic and Social Council (UNESC,
2000) reported that the Government of Mozambique responded to
the emergency despite limited means, but due to the extensive
international financial support requested help in its coordination
from the UN. The World Bank estimates that the direct losses as a
result of the 2000 floods amounted to US$ 273 million (UNESC,
2000).
9.2.6.4. Interventions
After the catastrophic floods in 2000, the Government of Mozambique
took a range of measures to improve the effectiveness of disaster risk
management. In 2001, an Action Plan for the Reduction of Absolute
Poverty (PARPA I) was adopted (Republic of Mozambique, 2001); and
this was revised for the period 2006 to 2009 (PARPA II) (Republic of
Mozambique, 2006a,b; Foley, 2007). In 2006, the government also
adopted a Master Plan, which provides a comprehensive strategy for
dealing with Mozambique’s vulnerability to natural disasters (Republic
of Mozambique, 2006a).
After the 2000 floods, Mozambique implemented intensive programs to
move people to safe areas (World Bank, 2005a). Since the 2000 floods,
a large resettlement program for communities affected by the floods
and tropical cyclones was initiated, with about 59,000 families resettled
although a lack of funds for improved livelihoods has reduced the
success of this program (WMO, 2011a).
Success and effectiveness of warnings depend not only on the accuracy
of the forecast, but also their delivery in adequate time before the
disaster to put in place prevention strategies. From November 2006 to
November 2007 the Severe Weather Forecasting Demonstration Project,
conducted by the World Meteorological Organization in southeastern
Africa, tested a new concept for capacity building and this service
contributed to the forecasting and warnings about Cyclone Favio in
February 2007 (Poolman et al., 2008). The demonstration phase was found
to be valuable, and the implementation phase – with training, supported
with efficient and effective forecasting and warning of tropical cyclones
in developing countries – continues (WMO, 2011b).
Besides high-level alerting it is important that a warning is received
by each person in the disaster zone, in an easily understandable way
(UNISDR, 2010). In 2005 and 2006 the German Agency for Technical
Cooperation developed a simple but effective early warning system along
the River Búzi (
Bollin et al., 2005; Loster and Wolf, 2007). This warning
system was adapted to the specific needs and skills of the people. The
village officials receive daily precipitation and water levels at strategic
points along the Búzi River basin. If precipitation is particularly heavy or
the river reaches critical levels, this information is passed on by radio
and blue, yellow, or red flags are raised depending on the flood alert
level
(Bollin et al., 2005; UNISDR, 2010).
Chapter 9Case Studies
507
9.2.6.5. Outcomes/Consequences – 2007 Floods in Mozambique
Seven years after the catastrophic floods of 2000, similar flooding
occurred in Mozambique, but the country was prepared to a greater
extent than before. Between December 2006 and February 2007, heavy
rains across northern and central Mozambique together with a severe
downpour in neighboring countries led to flooding in the Zambezi River
basin (IFRC, 2007). Additional flooding was caused by the approach of
tropical cyclone Favio, which struck the Búzi area at the end of February
2007 (Poolman et al., 2008). During the flood period on the southern
coast of Mozambique, 29 people were killed, 285,000 people affected,
and approximately 140,000 displaced (Kienberger, 2007; World Bank,
2011b). The heavy rains and floods damaged health centers, public
buildings, drug stocks, and medical equipment and affected safe water
and sanitation facilities (UNOCHA, 2007). In total, the floods and cyclone
caused approximately US$ 71 million of damage to local infrastructure
and destroyed 277,000 ha of crops (USAID, 2007).
During the course of January 2007, it became clear that there was an
imminent threat of severe flooding in the Zambezi River basin valley
(Foley, 2007). A multinational flood warning covering Zambia, Malawi,
and Mozambique was issued on 26 January 2007. With forecasts and
warnings increasing over the next week, NIDM increased the flood
warning until a ‘Red Alert’ was issued (UNISDR, 2010). This was a test
of the earlier work undertaken by Bollin et al. (2005); when the rivers
rose rapidly, it was reported that approximately 12,800 people who
were at risk had been well prepared by prior training (Loster and Wolf,
2007). The district’s disaster mitigation committee had alerted threatened
villages two days previously (blue-flag alert) and now with a red-flag
alert announced evacuations, which were completed in less than two
days, with approximately 2,300 going to accommodation centers (Loster
and Wolf , 2007).
In the emergency period, NIDM, with local and international partner
organizations, established networks with local centers to coordinate the
emergency operations. The International Federation of Red Cross and
Red Crescent Societies and its local partners, the US Agency for
International Development, and other organizations worked to distribute
basic goods, food, and medical assistance during the emergency period
(IFRC, 2007; USAID, 2007).
A resettlement program, although a policy of last resort, to move
inhabitants from flood-prone areas to safer areas was initiated (Stal,
2011; WMO, 2011a). Resettlement is not an easy option. Although brick-
built housing was provided in flood-safe areas with new (or nearby)
schools and health facilities, these have not been as well received as
intended as these flood-safe resettlements suffer from water scarcity
and drought, and growing crops is therefore difficult (Stal, 2011).
The floods of 2000 and 2007 along with other natural hazards are
considered to have undone years of development efforts (Sietz et al., 2008)
and to have undermined national efforts in realizing Mozambique’s
poverty reduction strategy (IMF, 2011).
9.2.6.6. Lessons Identified
This comparison of the two floods events that occurred in Mozambique
in 2000 and 2007 shows:
Floods, as one of the most dangerous natural phenomena, are a
real threat to the sustainable development of nations (Ahern et al.,
2005; Guha-Sapir et al., 2011)
The consequences of floods depend on the long-term adaptation
to extremes of climate, and associated hydrologic extremes require
further understanding. After the 2000 floods in Mozambique, national
and international organizations updated their strategies to include
disaster preparedness, risk management, and contingency and
response capacities according to the lessons of catastrophic floods.
The Government of Mozambique introduced new DRM structures
between 2000 and 2007, illustrating the flexibility needed to
accommodate the scientific and communication systems that need
to be in place to adapt to a climate change-driven disaster and that
this can be done in liaison with and with guidance from external
agencies. Realization of the new program of DRM led to a reduction
in consequences from the floods in 2007 (Republic of Mozambique,
2006a).
Experience in Mozambique shows that creation and development
of effective and steadily functioning systems of hydrological
monitoring and early warning systems at a local, regional, and
national level as key components of DRM allowing more realistic
warnings of flooding threats (WMO, 2011a).
The implementation of resettlement programs in periodically flooded
areas in 2007 has reduced flood damage, but these measures are
not easy to implement (WMO, 2011a).
Limited available resources are one of the most important problems
for both disaster preparedness and disaster response. The extreme
poverty of the people makes them highly vulnerable to floods and
other natural disasters, despite the best efforts of the government
to protect them (World Bank, 2011b).
The example of Mozambique shows that climate change adaptation
needs to be achieved through the understanding of vulnerability in
all sectors (social, infrastructure, production, and environmental)
and this knowledge needs to be used for the formulation of
preparedness and response mechanisms (Sietz et al., 2008).
9.2.7. Disastrous Epidemic Disease: The Case of Cholera
9.2.7.1. Introduction
Weather and climate have a wide range of health impacts and play a
role in the ecology of many infectious diseases (Patz et al., 2000). The
relationships between health and weather, climate variability, and long-
term climate change are complex and often indirect (McMichael et al.,
2006). As with other impacts explored in this report, not all extreme
health impacts associated with weather and climate result from
extreme events; some result instead from less dramatic events unfolding
in the context of high population vulnerability. In such cases, impacts
Chapter 9 Case Studies
508
are typically indirect and are mediated by a constellation of factors, as
opposed to the direct health impacts of severe weather, for example,
traumatic injuries resulting directly from exposure to kinetic energy
associated with storms (Noji, 2000).
Commonly, underdeveloped health and other infrastructure, poverty,
political instability, and ecosystem disruptions interact with weather to
impact health adversely, sometimes to a disastrous degree (Myers
and Patz, 2009). For example, cholera is an infectious disease that is
perpetuated by poverty and associated factors, though outbreaks are
commonly associated with rainy season onset. Research in the last
decade has demonstrated that cholera is also sensitive to climate
variability (Rodó et al., 2002; Koelle et al., 2005a; Constantin de Magny
et al., 2007). Assuming persistence of these vulnerability factors, cholera
outbreaks may become more widespread as the climate continues to
change (Lipp et al., 2002) due to the projected likely increase in frequency
of heavy precipitation over many areas of the globe, and tropical
regions in particular (see Table 3-1). Insights into the disease’s ecology,
however, including its climate sensitivity, may one day inform early warning
systems and other interventions that could blunt its disastrous impact.
Equally, if not more important, poverty reduction and improvements in
engineering, critical infrastructure, and political stability and transparency
can reduce vulnerability among exposed populations to the degree that
cholera could be contained.
9.2.7.2. Background
Cholera has a long history as a human scourge. The world is in the midst
of the seventh global pandemic, which began in Indonesia in 1961 and
is distinguished by continued prevalence of the El Tor strain of the Vibrio
cholerae bacterium (Zuckerman et al., 2007; WHO, 2010). Primarily driven
by poor sanitation, cholera cases are concentrated in areas burdened by
poverty, inadequate sanitation, and poor governance. Between 1995 and
2005, the heaviest burden was in Africa, where poverty, water source
contamination, heavy rainfall and floods, and population displacement
were the primary risk factors (Griffith et al., 2006).
V. cholerae is flexible and ecologically opportunistic, enabling it to cause
epidemic disease in a wide range of settings and in response to climate
forcings (Koelle et al., 2005b). Weather, particularly seasonal rains, has
long been recognized as a risk factor for cholera epidemics.
Cholera is one of a handful of diseases whose incidence has been directly
associated with climate variability and long-term climate change (Rodó
et al., 2002). One driver of cholera’s presence and pathogenicity is the
El Niño-Southern Oscillation (ENSO), which brings higher temperatures,
more intense precipitation, and enhanced cholera transmission. ENSO
has been associated with cholera outbreaks in coastal and inland
regions of Africa (Constantin de Magny et al., 2007), South Asia
(Constantin de Magny et al., 2007), and South America (Gil et al., 2004).
There is concern that climate change will work synergistically with
poverty and poor sanitation to increase cholera risk.
As with other disasters, the risk of disastrous cholera epidemics can be
deconstructed into hazard probability, exposure probability, and population
vulnerability, which can be further broken down into population
susceptibility and adaptive capacity. As noted in the introduction, some
disastrous cholera epidemics are not associated with discrete extreme
weather events, but extreme impacts are triggered instead by exposure
to a less dramatic weather event in the context of high population
vulnerability. We focus on factors affecting exposure and vulnerability in
general, then apply this discussion to the Zimbabwe cholera epidemic
that began in 2008.
9.2.7.2.1. Exposure
Cholera epidemics occur when susceptible human hosts are brought
into contact with toxigenic strains of V. cholerae serogroup O1 or
serogroup O139. A host of ecological factors affect V. cholerae’s
environmental prevalence and pathogenicity (Colwell, 2002) and the
likelihood of human exposure (Koelle, 2009). In coastal regions, there is
a commensal relationship between V. cholerae, plankton, and algae
(Colwell, 1996). Cholera bacteria are attracted to the chitin of zooplankton
exoskeletons, which provide them with stability and protect them from
predators. The zooplankton feed on algae, which bloom in response to
increasing sunlight and warmer temperatures. When there are algal
blooms in the Bay of Bengal, the zooplankton prosper and cholera
populations grow, increasing the likelihood of human exposure.
Precipitation levels, sea surface temperature, salinity, and factors affecting
members of the marine and estuarine ecosystem, such as algae and
copepods, affect exposure probability (Huq et al., 2005). Many of these
factors appear to be similar across regions, although their relative
importance varies, such as the association of V. cholerae with chitin
(Pruzzo et al., 2008) and the importance of precipitation and sea level
(Emch et al., 2008). For example, marine and estuarine sources were the
source of pathogenic V. cholerae strains responsible for cholera epidemics
in Mexico in recent El Niño years (Lizarraga-Partida et al., 2009).
Other variables are associated with increased likelihood of exposure,
including conflict (Bompangue et al., 2009), population displacement,
crowding (Shultz et al., 2009), and political instability (Shikanga et al.,
2009). Many of these factors are actually mediated by the more
conventional cholera risk factors of poor sanitation and lack of access
to improved water sources and sewage treatment.
9.2.7.2.2. Population susceptibility
Population susceptibility includes both physiological factors that increase
the likelihood of infection after cholera exposure, as well as social and
structural factors that drive the likelihood of a severe, persistent epidemic
once exposure has occurred. Physiologic factors that affect cholera risk
or severity include malnutrition and co-infection with intestinal parasites
(Harris et al., 2009) or the bacterium Helicobacter pylori. Infections are
more severe for people with blood group O, for children, and for those
Chapter 9Case Studies
509
with increased health-related vulnerability. Waxing and waning immunity
as a result of prior exposure has a significant impact on population
vulnerability to cholera over long periods (Koelle et al., 2005b).
While physiologic susceptibility is important, social and economic drivers
of population susceptibility persistently seem to drive epidemic risk.
Poverty is a strong predictor of risk on a population basis (Ackers et al.,
1998; Talavera and Perez, 2009), and political factors, as illustrated by
the Zimbabwe epidemic, are often important drivers of epidemic severity
and persistence once exposure occurs. Many recent severe epidemics
exhibit population susceptibility dynamics similar to Zimbabwe, including
in other poor communities (Hashizume et al., 2008), in the aftermath
of political unrest (Shikanga et al., 2009), and following population
displacement (Bompangue et al., 2009).
9.2.7.2.3. Adaptive capacity
Cholera outbreaks are familiar sequelae of complex emergencies. The
DRM community has much experience with prevention efforts to reduce
the likelihood of cholera epidemics, containing them once they occur,
and reducing the associated morbidity and mortality among the infected.
Best practices include guidelines for water treatment and sanitation and
for population-based surveillance (Sphere Project, 2004).
9.2.7.3. Description of Event
Zimbabwe has had cholera outbreaks every year since 1998, with the
2008 epidemic the worst the world had seen in two decades, affecting
approximately 92,000 people and killing over 4,000 (Mason, 2009). The
outbreak began on 20 August 2008, slightly lagging the onset of
seasonal rains, in Chitungwiza city, just south of the capital Harare
(WHO, 2008a). In the initial stages, several districts were affected. In
October, the epidemic exploded in Harare’s Budiriro suburb and soon
spread to include much of the country, persisting well into June 2009
and ultimately seeding outbreaks in several other countries. Weather
appears to have been crucial in the outbreak, as recurrent point-source
contamination of drinking water sources (WHO, 2008a) was almost
certainly amplified by the onset of the rainy season (Luque Fernandez et
al., 2009). In addition to its size, this epidemic was distinguished by its
urban focus and relatively high case fatality rate (CFR; the proportion of
infected people who die) ranging from 4 to 5% (Mason, 2009). Most
outbreaks have CFRs below 1% (Alajo et al., 2006). Underlying structural
vulnerability with shortages of medicines, equipment, and staff at
health facilities throughout the country compounded the effects of the
cholera epidemic (WHO, 2008b).
9.2.7.4. Intervention
There are several risk management considerations for preventing
cholera outbreaks and minimizing the likelihood that an outbreak
becomes a disastrous epidemic (Sack et al., 2006). Public health has a
wide range of interventions for preventing and containing outbreaks,
and several other potentially effective interventions are in development
(Bhattacharya et al., 2009). As is the case in managing all climate-
sensitive risks, the role of institutional learning is becoming ever more
important in reducing the risk of cholera and other epidemic disease as
the climate shifts.
9.2.7.4.1. Conventional public health strategies
The conventional public health strategies for reducing cholera risk
include a range of primary, secondary, and tertiary prevention strategies
(Holmgren, 1981).
Primary prevention, or prevention of contact between a hazardous
exposure and susceptible host, includes promoting access to clean water
and reducing the likelihood of population displacement; secondary
prevention, or prevention of symptom development in an exposed host,
includes vaccination; and tertiary prevention, or containment of symptoms
and prevention of complications once disease is manifest, includes
dehydration treatment with oral rehydration therapy.
9.2.7.4.2. Newer developments
Enhanced understanding of cholera ecology has enabled development
of predictive models that perform relatively well (Matsuda et al., 2008)
and fostered hope that early warning systems based on remotely sensed
trends in sea surface temperature, algal growth, and other ecological
drivers of cholera risk can help reduce risks of epidemic disease,
particularly in coastal regions (Mendelsohn and Dawson, 2008).
Strategies to reduce physiologic susceptibility through vaccination have
shown promise (Calain et al., 2004; Chaignat et al., 2008; Lopez et al.,
2008; Sur et al., 2009) and mass vaccination campaigns have potential
to interrupt epidemics (WHO, 2006c), and may be cost effective in
resource-poor regions or for displaced populations where provision of
sanitation and other services has proven difficult (Jeuland and
Whittington, 2009). Current World Health Organization policy on cholera
vaccination holds that vaccination should be used in conjunction with
other control strategies in endemic areas and be considered for
populations at risk for epidemic disease, and that cholera immunization
is a temporizing measure while more permanent sanitation improvements
can be pursued (WHO, 2010). Ultimately, given the strong association
with poverty, continued focus on development may ultimately have the
largest impact on reducing cholera risk.
9.2.7.5. Outcomes
Managing the risk of climate-sensitive disease, like risk management of
other climate-sensitive outcomes, will necessarily become more iterative
and adaptive as climate change shifts the hazard landscape and
Chapter 9 Case Studies
510
heightens vulnerability in certain populations. Learning is an important
component of this iterative process (see Sections 1.4 and 8.6.3.2).
There are multiple opportunities for learning to enhance risk management
related to epidemic disease. First, while reactive containment processes
can be essential for identifying and containing outbreaks, this approach
often glosses over root causes in an effort to return to the status quo.
As the World Health Organization states, “Current responses to cholera
outbreaks are reactive, taking the form of a more or less well-organized
emergency response, and prevention is lacking (WHO, 2006c). Without
losing the focus on containment, institutional learning could incorporate
strategies to address root causes, reducing the likelihood of future
outbreaks. This includes continued efforts to better understand cholera’s
human ecology to explore deeper assumptions, structures, and policy
decisions that shape how risks are constructed. In the case of cholera,
such exploration has opened the possibility of devising warning systems
and other novel risk management strategies. Another equally important
conclusion – one that experts on climate’s role in driving cholera risk
have emphasized (Pascual et al., 2002) – is that poverty and political
instability are the fundamental drivers of cholera risk, and that emphasis
on development and justice are risk management interventions as well.
9.2.7.6. Lessons Identified
The 2008 cholera epidemic epitomized the complex interactions
between weather events and population vulnerability that can interact
to produce disastrous epidemic disease. Recent studies of cholera,
including its basic and human ecology, demonstrate the potential for
early warning and potential points of leverage that may be useful for
interventions to contain future epidemics. The key messages from this
work include:
Variability in precipitation and temperature canaffect important
epidemic diseases such as choleraboth through direct effects on
thetransmission cycle, but alsopotentially through indirect effects,
for example through problems arising from inadequate basic water
and sanitation services.
If other determinants remained constant, climate change would be
expected to increase risk by increasing exposure likelihood –
through increased variability in precipitation and gradually rising
temperatures and by increasing population vulnerability.
The health impacts of cholera epidemics are strongly mediated
through individual characteristics such as age and immunity, and
population-level social determinants, such as poverty, governance,
and infrastructure.
Experience from multiple cholera epidemics demonstrates that
non-climatic factors can either exacerbate or override the effects of
weather or other infection hazards.
The processes of DRM and preventive public health are closely
linked, and largely synonymous. Strengthening and integrating
these measures, alongside economic development, should increase
resilience against the health effects of extreme weather and gradual
climate change.
9.2.8. Coastal Megacities: The Case of Mumbai
9.2.8.1. Introduction
In July 2005, Mumbai, India, was struck by an exceptional storm (Revi,
2005). In one 24-hour span alone, the city received 94 cm of rain, and
the storm left more than 1,000 dead, mostly in slum settlements (De
Sherbinin et al., 2007; Sharma and Tomar, 2010). A week of heavy rain
disrupted water, sewer, drainage, road, rail, air transport, power, and
telecommunications systems (Revi, 2005). As a result of this ‘synchronous
failure,’ Mumbai-based automated teller machine banking systems ceased
working across much of the country, and the Bombay and National
Stock Exchanges were temporarily forced to close (Revi, 2005; UNISDR,
2011b). This demonstrates that within megacities, risk and loss are both
concentrated and also spread through networks of critical infrastructure
as well as connected economic and other systems.
9.2.8.2. Background
At present, Mumbai is the city with the largest population exposed to
coastal flooding – estimated at 2,787,000 currently, and projected to
increase to more than 11 million people exposed by 2070 (Hanson et
al., 2011). During that same period, exposed assets are expected to
increase from US$ 46.2 billion to nearly US$ 1.6 trillion (Hanson et al.,
2011).
Mumbai’s significant, and increasing, exposure of people and assets –
both within the urban fabric but also outside, connected to the city’s
functions through networks of critical infrastructure, financial, and
resource flows – will be affected by changes in climate means and
climate extremes (Nicholls et al., 2007; Revi, 2008; Fuchs et al., 2011;
Ranger et al., 2011). It is difficult to associate a single extreme event
with climate change, but it may be possible to discuss the changed
probability of an event’s occurrence in relation to a particular cause,
such as global warming (see FAQ 3.2). For the Indian monsoon, for
example, extreme rain events have an increasing trend between 1901
and 2005, with the trend being stronger since 1950 (see Section 3.4.1).
9.2.8.3. Description of Vulnerability
Attributing causes of changes in monsoons is difficult due to substantial
differences between models, and the observed maximum rainfall on
India’s west coast, where Mumbai is located, is poorly simulated by
many models (see Section 3.4.1). That being said, increases in precipitation
are projected for the Asian monsoon, along with increased interannual
seasonally averaged precipitation variability (see Section 3.4.1).
Furthermore, extreme sea levels can be expected to change in the future
as a result of mean sea level rise and changes in atmospheric storminess,
and it is very likely that sea level rise will contribute to increases in
extreme sea levels in the future (see Section 3.5.3).
Chapter 9Case Studies
511
The development failures that have led to an accumulation of disaster
risk in Mumbai and allowed its transmission beyond the urban core are
common to many other large urban centers. The IPCC Fourth
Assessment Report (AR4) stated with very high confidence that the
impact of climate change on coasts is exacerbated by increasing
human-induced pressures, with subsequent studies being consistent
with this assessment (see Section 3.5.5).
The AR4 also reported with very high confidence that coasts will be
exposed to increasing risks, including coastal erosion, over coming
decades due to climate change and sea level rise, both of which will be
exacerbated by increasing human-induced pressures (see Section 3.5.5).
The July 2005 flooding in Mumbai underscores the fact that coastal
megacities are already at risk due to climate-related hazards (De
Sherbinin et al., 2007; McGranahan et al., 2007). Refuse and debris
commonly clog storm drains, causing flooding even on the higher
ground in Mumbai’s slums, and landslides are another threat to squatter
communities that are near or on the few hillsides in the city (De Sherbinin
et al., 2007). Urban poor populations often experience increased rates
of infectious disease after flood events, and after the July 2005 floods
the prevalence of leptospirosis rose eightfold in Mumbai (Maskey et al.,
2006; Kovats and Akhtar, 2008).
To the present, drivers of flood risk have been largely driven by
socioeconomic processes and factors, such as poverty, ecosystem
degradation, and poorly governed rapid urbanization (Revi, 2005, 2008;
De Sherbinin et al., 2007; Huq et al., 2007; UNISDR, 2009c, 2011b;
Hanson et al., 2011; Ranger et al., 2011). These processes are interrelated,
and within these cities, vulnerability is concentrated in the poorest
neighborhoods, which often lack access to sanitation, health care, and
transportation infrastructure, and whose homes and possessions are
unprotected by insurance (Revi, 2005; De Sherbinin et al., 2007; UNISDR,
2009c; Ranger et al., 2011).
Slum settlements are often located in sites with high levels of risk due
to environmental and social factors. For example, they are often located
in floodplains or on steep slopes, which means their residents suffer from
a considerable degree of physical exposure and social vulnerability to
losses from flood events (Huq et al., 2007; McGranahan et al., 2007;
Chatterjee, 2010).
Mumbai is one of many coastal megacities that have been built in part
on reclaimed land, a process that increases flood risks to low-lying
areas where slums are frequently located (Chatterjee, 2010). Its slums
do not benefit from structural flood protection measures and are located
in low-lying areas close to marshes and other marginal places and
are frequently flooded during monsoon season, especially when heavy
rainfall occurs during high tides (McGranahan et al., 2007; Chatterjee,
2010). A rise in sea level of 50 cm, together with storm surges, would
render uninhabitable the coastal and low-lying areas (De Sherbinin et
al., 2007) where many of Mumbai’s informal settlements are currently
located.
9.2.8.4. Outcomes/Consequences
India’s 2001 census indicated that in Mumbai 5,823,510 people (48.9%
of the population) lived in slums (Government of India, 2001). In 2005,
the global slum population was nearly 1 billion, and it is projected
to reach 1.3 to 1.4 billion by 2020, mostly concentrated in cities in
developing countries (UN-HABITAT, 2006). In addition to Mumbai,
Hanson et al. (2011) found that the following cities will have the greatest
population exposure to coastal flooding in 2070: Kolkata, Dhaka,
Guangzhou, Ho Chi Minh City, Shanghai, Bangkok, Rangoon, Miami, and
Hai Phòng. Many of these cities are already characterized by significant
population and asset exposure to coastal flooding, and all but Miami
are located in developing countries in Asia.
Africa does not have a large share of the world’s biggest coastal cities
but most of its largest cities are on the coast and large sections of their
population are at risk from flooding (Awuor et al., 2008; Adelekan, 2010).
Compared to Asia, Europe, and the Americas, a greater percentage of
Africa’s population lives in coastal cities of 100,000 to 5 million people,
which is noteworthy because Africa’s medium-to-large cities tend to be
poor and many are growing at much higher rates than cities on the
other continents (McGranahan et al., 2007).
The amount of vulnerability concentrated within these cities will define
their risks, and in the absence of adaptation there is high confidence
that locations currently experiencing adverse effects, such as coastal
erosion and inundation, will continue to do so in the future (see Section
3.5.5).
However, there is a certain limit to adaptation given that these cities are
fixed in place and some degree of exposure to hazards is ‘locked in’ due
to the unlikelihood of relocation (Hanson et al., 2011). For example, India’s
large infrastructure investments, which have facilitated Mumbai’s rapid
growth, were built to last 50 to 150 years (Revi, 2008). This forecloses
some adaptation and DRR strategies, such as risk avoidance.
Furthermore, all large coastal cities are centers of high population
density, infrastructure, investments, networking, and information
(McGranahan et al., 2007; Chatterjee, 2010). This concentration and
connectivity make them important sources of innovation and economic
growth, especially in developing countries where these ingredients may
be absent elsewhere. This underscores the importance of governance
and economic relations, including insurance and more general basic
needs of health and education, in allowing urban systems and those at
risk to build resilience if they cannot avoid hazard.
9.2.8.5. Lessons Identified
Measures to reduce exposure to existing weather-related hazards can
also serve as means of adapting to climate change (McGranahan et al.,
2007; UNISDR, 2009c, 2011b; Chapters 1 and 2). At the time of the
2005 flood, Mumbai lacked the capacity to address a complex portfolio
of (interrelated) risks (De Sherbinin et al., 2007; Revi, 2008), and its
Chapter 9 Case Studies
512
multi-hazard risk plan from 2000 was not well implemented (Revi,
2005). Risk protection in most other megacities in developing countries
was also found to be more informal than robust (Hanson et al., 2011).
Multi-hazard risk models, based on probabilistic analysis, can help
governments better reduce risks and facilitate better management of
and preparedness for risks that cannot be reduced cost effectively
(Revi, 2008; Ranger et al., 2011; UNISDR, 2011b).
Given that up to US$ 35 trillion (approximately 9% of projected global
GDP) may be exposed to climate-related hazards in port cities by 2070
(in purchasing power parity, 2001 US$) (Hanson et al., 2011), managing
– and reducing – these risks represents a high-leverage policy area for
adaptation. The scale of economic assets at risk is impressive and to
this must be added the livelihoods and health of the poor that may be
disproportionately impacted by disaster events but have partial visibility
in macroeconomic assessments.
The need to adapt is especially acute in developing countries in Asia
given that 14 of the top 20 urban agglomerations projected to have the
greatest exposure of assets in 2070 are in developing countries in this
region (Hanson et al., 2011). This suggests that scaled-up financing for
adaptation may be needed to safeguard the residents and economic
activity in these cities to a level comparable to that of other coastal
megacities that face similar population and asset exposure, such as
New York or Tokyo. Two critical distinctions are the degree of poverty
and the less complete reach of local government in those cities most at
risk.
Despite efforts to assess the impacts of climate change at the city scale,
analysis of the economic impacts of climate change at this scale has
received relatively little attention to date (Hallegatte and Corfee-
Morlot, 2011). In developing countries, the sometimes incomplete
understanding of climate risks and the limited institutional capacity
have meant that analysis of climate change impacts at the city scale has
generally considered only flood risks and not yet assessed additional
potential impacts (Hunt and Watkiss, 2011). A standardized, multi-
hazard impact analysis at the city scale would be useful and facilitate
comparisons between cities (Hunt and Watkiss, 2011).
9.2.9. Small Island Developing States:
The Challenge of Adaptation
9.2.9.1. Introduction
Small Island Developing States (SIDS) are defined as those that are small
island nations, have low-lying coastal zones, and share development
challenges (UNCTAD, 2004). Strengthening of SIDS technical capacities
to enable resilience building has been recommended (UNESC, 2011).
This case study explores the critical vulnerabilities of the Republic of the
Marshall Islands (RMI). Additional data from the Maldives, also highly
vulnerable to sea level rise and extreme weather events and where the
tsunami caused significant damage, and Grenada, which is a country
with a small open economy vulnerable to external shocks and natural
disasters, are used to develop the full context of the limits of adaptation.
Specifically, the RMI highlights the availability of fresh water as a major
concern. “There is strong evidence that under most climate change
scenarios, water resources in small islands are likely to be seriously
compromised (very high confidence)” (Mimura et al., 2007).
9.2.9.2. Background
SIDS can be particularly vulnerable to hazards and face difficulties when
responding to disasters (TDB, 2007). SIDS share similar development
challenges including small but growing populations, economic
dependence on international funders, and lack of resources (e.g.,
freshwater, land) (World Bank, 2005b; UNFCCC, 2007a). The IPCC
(Mimura et al., 2007) concluded that “small islands, whether located in
the tropics or higher latitudes, have characteristics which make them
especially vulnerable to the effects of climate change, sea level rise, and
extreme events (very high confidence). Many SIDS share vulnerabilities
with high levels of poverty and are reported to suffer serious
environmental degradation and to have weak human and institutional
capacities for land management that is integrated and sustainable (GEF,
2006). The range of physical resources available to states influences their
options to cope with disasters and the relatively restricted economic
diversity intrinsic to SIDS minimizes their capacity to respond in
emergencies with measures such as shelter or evacuation (Boruff and
Cutter, 2007). Hence SIDS are among the most vulnerable states to the
impacts of climate change (UNFCCC, 2007b). As of 2010, 38 UN member
nations and 14 non-UN Members/Associate Members of the Regional
Commissions were classified as SIDS (UN-OHRLLS, 2011).
The RMI provides an example of critical vulnerabilities. It is made up of
five islands and 29 atolls that are spread across more than 1.9 million
square kilometers of Pacific Ocean (World Bank, 2006a). The country
has a population of 64,522, approximately two-thirds of which is
concentrated in urban areas on just two atolls (UNDESA, 2010; World
Bank, 2011b). The other one-third live on the even more remote outer
islands and atolls (World Bank, 2006a). Even the main inhabited islands
remain extremely isolated; the nearest major port is over 4,500 km from
Majuro, the capital atoll (World Bank, 2005b).
The Maldives and Grenada both provide other examples of vulnerability
to extreme events and disasters and climate change adaptation needs:
The Maldives consist of 1,192 small islands. 80% are 1 m or less
above sea level (Quarless, 2007), of which only three islands have a
surface area of more than 500 ha (De Comarmond and Payet, 2010).
These characteristics make them highly vulnerable to damage from
sea level rise and extreme weather events. The economic and
survival challenges of the people of the Maldives were evident
after the 2004 tsunami caused damage equivalent to 62% of
national GDP (World Bank, 2005c). As of 2009, the country still
faced a deficit of more than US$ 150 million for reconstruction.
Chapter 9Case Studies
513
Such devastation in a SIDS might be countered with further disaster
preparation and efforts to maintain emergency funds to rebuild
their economies (De Comarmond and Payet, 2010).
Grenada is a small tri-island state in the Eastern Caribbean with a
population of 102,000 and a per capita GDP of US$ 4,601 in 2004
(IMF, 2011). It is a small open economy, vulnerable to external
shocks and natural disasters as seen by the effects of Hurricane
Ivan, which created large fiscal and balance of payments financing
needs in 2004, and Hurricane Emily, which struck in 2005 (IMF, 2004,
2006). Hurricane Ivan brought major disruption to an economic
recovery process, eventually costing the island an estimated US$ 3
billion (Boruff and Cutter, 2007). It is projected that Ivan reduced
the country’s forecasted growth rate from 5.7% to -1.4% (Quarless,
2007). Hurricane Emily followed 10 months later, virtually completing
the trail of destruction started by Ivan. The impact was seen in
every sector of the economy. Capital stock was severely damaged
and employment was significantly affected (UNDP, 2006).
9.2.9.3. Description of Vulnerability
Many SIDS face specific disadvantages associated with their small size,
insularity, remoteness, and susceptibility to natural hazards. SIDS are
particularly vulnerable to climate change because their key economic
sectors such as agriculture, fisheries, and tourism are all susceptible
(Barnett and Adger, 2003; Read, 2010) (a more extensive discussion is
provided in Chapter 4, especially Sections 4.2.1, 4.4.4, 4.5.2, and 4.5.3).
The hazards of extreme weather events are coupled with other long-
term climate change impacts, especially sea level rise (see Box 3-4).
Low-lying atoll communities, such as the Maldives and Cook Islands, are
especially vulnerable (Ebi et al., 2006; Woodroffe, 2008; Kelman and
West, 2009) and are expected to lose significant portions of land
(Mimura et al., 2007). Small island states and particularly atoll countries
may experience erosion, inundation, and saline intrusion resulting in
ecosystem disruption, decreased agricultural productivity, changes in
disease patterns, economic losses, and population displacement – all of
which reinforce vulnerability to extreme weather events (Pernetta,
1992; Nurse and Sem, 2001; Mimura et al., 2007).
SIDS suffer higher relative economic losses from natural hazards and
are less resilient to those losses so that one extreme event may have the
effect of countering years of development gains (UN, 2005; Kelman, 2010).
The distances between many SIDS and economic centers make their
populations among the most isolated in the world (World Bank, 2005b).
Underdevelopment and susceptibility to disasters are mutually reinforcing:
disasters not only cause heavy losses to capital assets, but also disrupt
production and the flow of goods and services in the affected economy,
resulting in a loss of earnings (Pelling et al., 2002). In both the short and
the long term, those impacts can have sharp repercussions on the
economic development of a country, affecting GDP, public finances, and
foreign trade, thus increasing levels of poverty and public debt
(Mirza, 2003; Ahrens and Rudolph, 2006). Climate change threatens to
exacerbate existing vulnerabilities and hinder socioeconomic development
(UNFCCC, 2007b).
The RMI faces major climate-related natural hazards including sea level
rise, tropical storms or typhoons with associated storm surges, and
drought. These hazards should be considered within the context of
additional hazards and challenges such as ecosystem degradation,
pollution of the marine environment, and coastal erosion as well as
food security. The RMI faces physical and economic challenges that
amplify the population’s vulnerability to climate hazards, including high
population density, high levels of poverty, low elevation, and fragile
freshwater resources (World Bank, 2011b). The Global Facility for Disaster
Reduction and Recovery report concludes that the hazard that poses the
most threat is sea level rise, as the highest point on RMI is only 10 m
above sea level (World Bank, 2011b). Consequently, multilateral donors
considered the RMI ‘high risk’ and the Global Facility for Disaster
Reduction and Recovery has identified it as a priority country for
assistance (World Bank, 2011b).
9.2.9.4. Outcomes/Consequences
A range of both local and donor-supported actions have endeavored to
build resilience among SIDS. The example of the RMI shows the benefits
that risk reduction and climate change adaptation efforts may offer
other island states.
Freshwater availability is a major concern for many SIDS (Quarless,
2007), including the RMI. Since SIDS are especially vulnerable to
extreme weather events, their water supplies face rapid salinization due
to seawater intrusion and contamination (PSIDS, 2009). According to
one study, countries such as the RMI lack the financial and technical
resources to implement seawater desalination for their populations
(UNDESA, 2011). Some disaster and climate risk management gains
may come from simple technology (UNDESA, 2010). New scavenger
technology for wells has been introduced (UNESC, 2004) as one of the
ways forward. Simple abstraction of freshwater from thin groundwater
lenses (a typical practice in oceanic atolls) often results in upward coning
of saltwater, which, in turn, causes contamination of the water supplies.
The RMI has benefited from its use of new, pioneering technology to
limit the effects of extreme weather events on its water supply (UNDESA,
2011). The improvement of climate sensitivity knowledge, particularly in
the context of risk management, is key to adaptation to climate change.
Climate and disaster risk are closely entwined and, for example,
resilience to drought and resilience to climate change both stand to be
enhanced through a single targeted program.
In addition to project-oriented development assistance, the RMI
receives substantial financial assistance from the United States through
a Compact of Free Association (Nuclear Claims Tribunal, Republic of the
Marshall Islands, 2007). Grants and budget support provided under
Compact I over the period of 1987 to 2001 totaled an average of over
30% of GDP, not including any other form of bilateral assistance (World
Chapter 9 Case Studies
514
Bank, 2005b). The RMI stands out among other lower middle income
countries, receiving average aid per capita of US$ 1,183, compared with
the average of US$ 8 for other lower middle income countries (World
Bank, 2005b). This assistance, buttressed by national disaster management
policies dating back to the RMI’s independence in 1986 and including
the Global Facility for Disaster Reduction and Recovery’s role in assessing
the RMI’s systems and noting existing gaps for future development
partner projects, has resulted in a range of national and regional disaster
and climate risk management initiatives (World Bank, 2009, 2011b).
9.2.9.5. Lessons Identified
The physical, social, and economic characteristics by which SIDS and
developing countries are defined (education, income, and health, for
example) increase their vulnerability to extreme climate events.
Experiences from the Marshall Islands, the Maldives, and Grenada
indicate that limited freshwater supplies and inadequate drainage
infrastructure are key vulnerability factors. These examples also indicate
an important difference between risk of frequent smaller hazards and
catastrophic risk of infrequent but extreme events.
The cases of Grenada and the Maldives demonstrate the high relative
financial impact that a hazard can have on a small island state. For
the RMI, financial support from donors has enabled a range of risk
management programs. Although the importance of disaster risk
reduction strategies is apparent, preventive approaches continue to
receive less emphasis than disaster relief and recovery (Davies et al., 2008).
Considering the range of challenges facing policymakers in some SIDS,
preventive climate adaptation policies can seem marginal compared
with pressing issues of poverty, affordable energy, affordable food,
transportation, health care, and economic development.
National policymaking in this context remains a major challenge and
availability of funding for preventive action – such as disaster and climate
risk management – may continue to be limited for many countries
(Ahmad and Ahmed, 2002; Jegillos, 2003; Huq et al., 2006; Yohe et al.,
2007). Although most developing countries participate in various
international protocols and conventions relating to climate change and
sustainable development and most have adopted national environmental
conservation and natural disaster management policies (Yohe et al.,
2007), the policy agendas of many developing countries do not yet fully
address all aspects of climate change (Beg et al., 2002).
9.2.10. Changing Cold Climate Vulnerabilities:
Northern Canada
9.2.10.1. Introduction
In cold climate regions all over the world, climate change is occurring
more rapidly than over most of the globe (Anisimov et al., 2007). These
changes have implications for the built environment. The vulnerability of
residents of the Canadian North is complex and dynamic. In addition to
the increasing risks from extreme weather events, there are climate
impacts upon travel, food security, and infrastructural integrity, which in
turn affect many other aspects of everyday life (Pearce et al., 2009; Ford
et al., 2010). Additionally, the relative isolation of these northern
communities makes exposure to climate-related risk more difficult to
adapt to, thus increasing their level of vulnerability (Ford and Pearce,
2010). This case study will examine the increased vulnerabilities in
regions of the Canadian North due to climate change’s effect on
infrastructure through changes in permafrost thaw and snow loading.
The study illustrates existing and projected risks and governmental
responses to them at the municipal, provincial/territorial, and national
levels. Canada has three territories: the Yukon (YT); the Northwest
Territories (NWT); and Nunavut (NU); this study deals with all three and,
to a much lesser extent, the northern regions of the provinces, such as
Nunavik in northern Quebec. Though both permafrost thaw and
changing snow loads are slowly progressing events, as opposed to one-
time extreme events, their impacts can result in disasters. Future
protection relies upon risk reduction and adaptation. Sections 3.3.1 and
3.5.7 discuss changes in cold extremes and other climate variables at
high latitudes.
9.2.10.2. Background
Over the past few decades, the northern regions of Canada have
experienced a rate of warming about twice that of the rest of the world
(McBean et al., 2005; Field et al., 2007; Furgal and Prowse, 2008). In
northern Canada, winter temperatures are expected to rise by between
3.5 and 12.5°C by 2080, with smaller changes projected for spring and
summer; in more southerly regions of northern Canada, temperatures
could warm to be above freezing for much longer periods (Furgal and
Prowse, 2008). For example, whereas it was estimated that the
Northwest Passage would be navigable for ice-strengthened cargo
ships in 2050 (Instanes et al., 2005), it has already been navigable in
2007 (Barber et al., 2008). Recent studies have suggested that some
communities in northern Canada will be vulnerable to the accelerated
rate of climate change (Ford and Smit, 2004; Ford and Furgal, 2009).
Higher temperatures have several implications for infrastructure that
plays an important role in maintaining the social and economic functions
of a community (CSA, 2010). Permafrost thaw and changing snow
loads have the potential to affect the structural stability of essential
infrastructure (Nelson et al., 2002; Couture and Pollard, 2007). Design
standards in northern Canada were based on permafrost and snow load
levels of a previous climate regime (CSA, 2010). Adaptation is essential
to avoid higher operational and maintenance costs for structures and to
ensure that the designed long lifespan of each structure remains viable
(Allard et al., 2002). Addressing these impacts of climate change is a
complex task. Naturally each structure will be differently affected and
the resulting damage can exacerbate existing weaknesses and create
new vulnerabilities. For example, although increasing snow loads
alone can have negative impacts on infrastructure, the fact that many
Chapter 9Case Studies
515
buildings have been structurally weakened by permafrost thaw adds to
the damage potential during any snow event (CSA, 2010).
9.2.10.3. Description of Vulnerability
9.2.10.3.1. Permafrost thaw
Permafrost thaw is one of the leading factors increasing climate-related
vulnerability. Permafrost is by definition dependent on a sub-zero
temperature to maintain its state (CSA, 2010; NRCAN, 2011a). With a
changing climate, it is difficult to predict where permafrost is most likely
to thaw, but about half of Canada’s permafrost zones are sensitive to
small, short-term increases in temperature, compromising the ability of
the ground to support infrastructure (Nielson, 2007; NRTEE, 2009; CSA,
2010). The rate of thaw (and hence implications for infrastructure stability)
is also dependent on soil type within the permafrost zone (Nielson,
2007). Areas that have ice-rich soil are much more likely to be affected
than those with a lower ice-soil ratio or those that are underlain by
bedrock (Nelson et al., 2002). Municipalities in discontinuous or sporadic
permafrost zones may feel the impacts of a warming climate more
intensely since the permafrost is thinner than it would be in continuous
zones where ice has built up over time (Nelson et al., 2002).
Though some infrastructure maintenance will always be required, climate-
related permafrost thaw will increase the needs for infrastructure
maintenance and the rate of damage that is inflicted (Allard et al., 2002).
Permafrost thaw affects different types of infrastructure in radically
different ways. In northern Canada, municipalities have experienced
many different climate-related impacts on physical infrastructure
including the following (Infrastructure Canada, 2006; Nielson, 2007;
NRTEE, 2009):
Nunavik, in northern Quebec, reported that local roads and airport
runways have suffered from severe erosion, heaving, buckling, and
splitting (Nielson, 2007; Fortier et al., 2011).
In Iqaluit, in Nunavut, 59 houses have required foundation repair
and/or restoration and buildings with shallow foundation systems
have been identified as needing attention in the near future
(Nielson, 2007). In Inuvik, in the Northwest Territories, a recent
study estimated that 75% of the buildings in the municipality
would experience structural damage (Bastedo, 2007) depending
on the rate of permafrost thaw.
The Tibbitt to Contwoyto winter road (Northwest Territories)
experienced climate-related closures in 2006, remaining open for
only 42 days compared to 76 in 2005 (Bastedo, 2007). This resulted
in residents and businesses having to airlift materials to their
communities instead. In particular, the Diavik Diamond Mine was
forced to spend millions of dollars flying in materials (Bastedo,
2007; Governments of Northwest Territories, Nunavut, and Yukon,
2010).
The Northwest Territories reported that the airport runway in
Yellowknife required extensive retrofitting when the permafrost
below it began to thaw (Infrastructure Canada, 2006).
The impacts of permafrost thaw on infrastructure have implications for
the health, economic livelihood, and safety of northern Canadian
communities. The costs of repairing and installing technologies to adapt
to climate change in existing infrastructure can range from several
million to many billions of dollars, depending on the extent of the
damage and the type of infrastructure that is at risk (Infrastructure
Canada, 2006). Lessons from municipalities in the United States have
proven that these costs can be large. For instance, while the Yukon
had financial difficulties with CDN$ 4,000 km
-1
yr
-1
costs related to
permafrost damage to highways, Alaska is experiencing costs of up to
CDN$ 30,000 km
-1
yr
-1
for an annual cost of over CDN$ 6 million over
a 200-km stretch (Governments of Northwest Territories, Nunavut, and
Yukon, 2010). In the future, as infrastructure needs to be replaced, costs
will multiply rapidly (Larsen et al., 2008).
9.2.10.3.2. Snow loading
In most northern Canadian communities, buildings and roadways are
built using historical snow load standards (Nielson, 2007; Auld, 2008).
This makes them particularly vulnerable to climate change since snow
loads are expected to increase with higher levels of winter precipitation
(Christensen et al., 2007; NRTEE, 2009). Already in the Northwest
Territories, 10% of public access buildings have been retrofitted since
2004 to address critical structural malfunctions. An additional 12% of
buildings are on high alert for snow load-related roof collapse (Auld et
al., 2010). In Inuvik, NWT, a local school suffered a complete roof collapse
under a particularly heavy snowfall (Bastedo, 2007). As permafrost
continues to thaw, resulting in a loss of overall structural integrity, greater
impacts will be linked to the increase in snow loads as previously
weakened or infirm structures topple under larger or heavier snowfalls.
9.2.10.4. Outcomes
In response to these vulnerabilities, government and community leaders
have put emphasis on action and preparedness (Government of Northwest
Territories, 2008; Governments of Northwest Territories, Nunavut, and
Yukon, 2010). The social impacts of relocating communities or complete
restoration after a major disaster, as well as the financial costs, provide
a strong deterrent to complacency and relocation will be utilized where
necessary as a last resort (USARC, 2003). Though each government tier,
from federal to municipal levels, tackles the issue from a different angle,
their approaches are proving complementary as is demonstrated below.
This section explores adaptation efforts from each level of government
and the contribution they make to adaptive capacity in northern
Canadian communities.
9.2.10.4.1. Federal level
The Canadian government contributes to numerous adaptation efforts
at different levels and through various programs (Lemmen et al., 2008).
Chapter 9 Case Studies
516
Some federal-level climate change adaptation programs are reactive; for
example, at the most basic level, the federal government is responsible
for the provision of assistance after a disaster or in order to relocate
structures and communities (Henstra and McBean, 2005). Other programs
are more proactive, designed to prevent disasters from occurring; for
example, climate change is currently being incorporated into the 2015
version of the National Building Code (Environment Canada, 2010),
which would help ensure that future infrastructure is built to a more
appropriate standard and that adaptive measures are incorporated into
the design and building of any new infrastructure. This could also help
ensure that adaptation measures are implemented in a uniform way
across the country.
In addition, several federal-level departments have programs specially
designed to prevent damage from climate-related impacts. As part of
the Climate Change Adaptation Program offered by the Aboriginal Affairs
and Northern Development Canada, the Assisting Northerners in Assessing
Key Vulnerabilities and Opportunities helps to support aboriginal and
northern communities, organizations and territories in addressing the
urgent climate-related risks (INAC, 2010). For example, the program
offers risk assessments for existing infrastructure, water quality, and
management programs and helps to identify new infrastructure designs
to reduce risk from climate change (INAC, 2010).
Similarly, the Regional Adaptation Collaborative (RAC) funding provided
by Natural Resources Canada was designed to assist communities that
are adapting to climate change (NRCAN, 2011b). The Northern RAC
initiatives are focused on identifying vulnerabilities in the mining sector.
Permafrost thaw and snow loading are examples of factors that the
program will examine (NRCAN, 2011b).
Another adaptation initiative that has come from the federal level is the
site selection guidelines developed by the Canadian Standards
Association (CSA, 2010). Though voluntary, this set of guidelines
encourages engineers, land use planners, and developers to consider
environmental factors including the rate of permafrost thaw and type of
soil when building (CSA, 2010). Additionally, it strongly encourages the
use of projections and models in the site selection process, instead of
relying on extrapolated weather trends (CSA, 2010).
Similarly, federal-level design requirements such as the Canadian
environmental assessment process are required to account for climate
change in the design phase of significant new projects such as tailings
containment, water retention, pipelines, or roads (Furgal and Prowse,
2008). Facilitating use of the guidelines and environmental assessment
requirements are proactive responses that aim to prevent future
permafrost-related damage to infrastructure.
9.2.10.4.2. Provincial/territorial level
The territorial governments are contributing to the protection of
infrastructure in several ways, including conducting and funding
research to identify vulnerable areas and populations (INAC, 2010). The
Yukon transportation department has undertaken several adaptation
initiatives including the design and implementation of road embankments
to minimize melting; construction of granular blankets on ice-rich slopes
to provide for stability and to prevent major slope failure; and the
installation of culverts in thawed streambeds (Government of Yukon,
2010). Ground-penetrating radar and resistivity to assess permafrost
conditions underground are being used in Nunavik, Quebec (Fortier et al.,
2011). To protect existing permafrost, light-colored pavement on roadways
is being used to reflect greater amounts of sunlight and prevent heat
absorption (Walsh et al., 2009). Collaborations with federal-level
departments to address community infrastructure resilience are being
conducted with, for example, the Nunavut Climate Change Partnership,
which involves the Government of Nunavut, Natural Resources Canada,
Aboriginal Affairs and Northern Development Canada, and the Canadian
Institute of Planners (NRCAN, 2011c). These programs help communities
to develop action plans that detail suitable options for addressing issues
related to climate change. The Yukon government is providing funding
for municipalities to develop their own climate change adaptation plans
through the Northern Strategy Trust Fund to the Northern Climate
ExChange (Government of Yukon, 2009).
About 85 flat-loop thermosyphons, a sort of ground-source heat pump,
which extract heat from the ground (through convection) during the
winter and reduce thawing, have been constructed in territorial-owned
buildings including schools and hospitals, prisons, and visitor centers in
Nunavut, the Northwest Territories, and the Yukon (Holubec, 2008; CSA,
2010). The installation of thermosyphon technology is not, in itself, a long-
term strategy but merely prolongs the lifetime of most infrastructures
(CSA, 2010). Finally, screw jack foundations, a technology that helps to
stabilize vulnerable foundations and has been used to prevent damage
due to permafrost thaw and related shifting of house foundations, have
been implemented in new buildings built by the Northwest Territories
Housing Corporation (Government of Northwest Territories, 2008).
9.2.10.4.3. Municipal level
The municipal level is often most involved in building adaptive capacity
and implementing adaptation strategies (Black et al., 2010) because
municipal governments feel the effects of damaged infrastructure more
keenly than higher levels of government (Richardson, 2010).
Municipalities, community groups, and businesses all over the three
territories have contributed in many ways. Some examples include:
Urban planning and design are being used to reduce exposure to
wind and snowdrifts as well as minimize heat loss from buildings
in Iqaluit, NU (NRCAN, 2010).
Insulated lining was placed underneath a 100-m section of runway
to prevent damage from permafrost thaw in Yellowknife, NWT
(Infrastructure Canada, 2006).
Ice-rich soil under important infrastructure has been replaced with
gravel and heat-absorbing pavement in Yellowknife, NWT
(Bastedo, 2007).
Chapter 9Case Studies
517
Wind deflection fins are being used to prevent snow loading on
roofs and obstructions around exits in NWT (Waechter, 2005).
In Tuktoyaktuk, NWT, important buildings, including the police
station and a school at risk of severe damage or loss have been
moved inland (Governments of Northwest Territories, Nunavut, and
Yukon, 2010) and concrete mats bound together with chains are
being used to limit erosion (Johnson et al., 2003).
Shims or pillars to elevate buildings are being used to make them
less vulnerable to permafrost thaw (USARC, 2003).
Construction of new bridges and all-weather roads to replace ice
roads that are no longer stable is underway (Infrastructure Canada,
2006).
9.2.10.5. Lessons Identified
Northern Canada can be considered a vulnerable region given the
expected climate-related risks. As the climate continues to warm in the
North, infrastructure in many remote communities will become more
vulnerable as well.
More research, especially into vulnerabilities in northern regions of the
globe, and the identification of adaptation options for established
communities would be of benefit for adaptation. Additionally, while
governmental programs and support are available, a significant portion
of it has been devoted to adaptation planning and strategizing. An
important issue is the funding needed to help northern Canadian
communities implement adaptation actions.
Finally, codes and standards are an integral part of addressing climate
impacts on infrastructure. Given the importance of this task, building
codes in vulnerable regions need more review and attention to protect
communities. An evaluation and monitoring program that focuses on
codes and structures as well as adaptation options is noticeably lacking.
Despite the complexity of these risks, however, a concerted effort from
three tiers of government and community can work to reduce the
vulnerability of infrastructure and northern communities.
9.2.11. Early Warning Systems: Adapting to Reduce Impacts
9.2.11.1. Introduction
It is recognized that vulnerability and exposure can never be reduced to
zero but risk can be reduced by effective systems for early warning of
extreme events that may occur in the near- through to longer-term
future (Broad and Agrawala, 2000; Da Silva et al., 2004; Haile, 2005; Patt
et al., 2005; Hansen et al., 2011). This sense of ‘seeing the future’ by
understanding current and projected risks is essential to effectively
prepare for, respond to, and recover from extreme events and disasters.
It is important to recognize that a changing climate poses additional
uncertainty and therefore early warning systems can contribute to climate-
smart disaster risk management. Effective disaster risk management in
a changing climate is facilitated by strong coordination within and
between sectors to realize adaptation potentials through assessing
vulnerabilities and taking anticipatory actions (Choularton, 2007;
Braman et al., 2010).
9.2.11.2. Background
The Hyogo Framework for Action (UNISDR, 2010; Chapter 7) stresses that
early warning systems should be “people centered” and that warnings
need to be “timely and understandable to those at risk” and need to
“take into account the demographic, gender, cultural, and livelihood
characteristics of the target audiences. “Guidance on how to act
upon warnings” should be included. An early warning system is thus
considerably more than just a forecast of an impending hazard.
In 2006, the United Nations International Strategy for Disaster
Reduction completed a global survey of early warning systems. The
executive summary opened with the statement that “If an effective
tsunami early warning system had been in place in the Indian Ocean
region on 26 December 2004, thousands of lives would have been
saved. ... Effective early warning systems not only save lives but also
help protect livelihoods and national development gains” (Basher,
2006; UN, 2006). Improved early warning systems have contributed to
reductions in deaths, injuries, and livelihood losses over the last 30 years
(IFRC, 2009). Early warning systems are important at local (Chapter 5),
national (Chapter 6), and international scales (Chapter 7). Toward the
achievement of sustainable development, early warning systems provide
important information for decisionmaking and in avoiding tipping
points (Chapter 8).
9.2.11.3. Description of Strategy of Early Warning Systems
Early warning systems are to alert and inform citizens and governments
of changes on time scales of minutes to hours for immediate threats
requiring urgent evasive action; weeks for more advanced preparedness;
and seasons and decades for climate variations and changes (Brunet et
al., 2010). To date most early warning systems have been based on
weather predictions, which provide short-term warnings often with
sufficient lead time and accuracy to take evasive action. However, the
range of actions that can be taken is limited. Weather predictions often
provide less than 24 hours notice of an impending extreme weather
event and options in resource-poor areas may not extend beyond the
emergency evacuations of people (Chapter 5). Thus although lives may
be saved, livelihoods can be destroyed, especially those of the poorest
communities.
While most of the successfully implemented early warning systems to
date have focused on shorter time scales, for example, for tornadoes
(Doswell et al., 1993), benefits of improved predictions on sub-seasonal
to seasonal scales are being addressed (Nicholls, 2001; Brunet et al.,
2010; Webster et al., 2010). While hazardous atmospheric events can
Chapter 9 Case Studies
518
develop in a matter of minutes (in the case of tornadoes), it can be
across seasons and decades that occurrence of extremes can change
climatically (McBean, 2000). Since planning for hazardous events
involves decisions across a full range of time scales, An Earth-system
Prediction Initiative for the 21st Century’ covering all scales has been
proposed (Shapiro et al., 2007, 2010).
With the rapid growth in the number of humanitarian disasters, the
disaster risk management community has become attentive to changes
in extreme events possibly attributed to climate change including
floods, droughts, heat waves, and storms that cause the most frequent
and economically damaging disasters (Gall et al., 2009; Munich Re,
2010; Vos et al., 2010). Early warning systems provide an adaptation
option to minimize damaging impacts resulting from projected severe
events. Such systems also provide a mechanism to increase public
knowledge and awareness of natural risks and may foster improved
policy- and decisionmaking at various levels.
Important developments in recent years in the area of sub-seasonal and
seasonal-to-interannual prediction have led to significant improvements
in predictions of weather and climate extremes (Nicholls, 2001; Simmons
and Hollingsworth, 2002; Kharin and Zwiers, 2003; Medina-Cetina and
Nadim, 2008). Some of these improvements, such as the use of soil
moisture initialization for weather and (sub-) seasonal prediction (Koster
et al., 2010), have potential for applications in transitional zones between
wet and dry climates, and in particular in mid-latitudes (Koster et al.,
2004). Such applications may potentially be relevant for projections of
temperature extremes and droughts (Lawrimore et al., 2007; Schubert
et al., 2008; Koster et al., 2010). Decadal and longer time scale predictions
are improving and could form the basis for early warning systems in the
future (Meehl et al., 2007, 2009; Palmer et al., 2008; Shukla et al., 2009,
2010).
Developing resiliency to weather and climate involves developing
resiliency to its variability on a continuum of time scales, and in an ideal
world, early warnings would be available across this continuum (Chapters
1 and 2; McBean, 2000; Hellmuth et al., 2011). However, investments in
developing such resiliency are usually primarily informed by information
only over the expected lifetime of the investment, especially among
poorer communities. For the decision of which crops to grow next season,
some consideration may be given to longer-term strategies but the
more pressing concern is likely to be the expected climate over the next
season. Indeed, there is little point in preparing to survive the impacts of
possible disasters a century in the future if one is not equipped to survive
more immediate threats. Thus, within the disaster risk management
community, preparedness for climate change must involve preparedness
for climate variability (Chapters 3 and 4).
Improving prediction methods remains an active area of research and it
is hoped significant further progress will be reached in coming years
(Brunet et al., 2010; Shapiro et al., 2010). However for such predictions
to be of use to end users, improved communication will be required to
develop indices appropriate for specific regional impacts. A better
awareness of such issues in the climate modeling community through
greater feedback from the disaster risk management community (and
other user communities) may lead to the development of additional
applications for weather and climate hazard predictions. Prediction
systems, if carefully targeted and sufficiently accurate, can be useful
tools for reducing the risks related to climate and weather extremes
(Patt et al., 2005; Goddard et al., 2010).
Despite an inevitable focus on shorter-term survival and hence interest
in shorter-term hazard warnings, the longer time scales cannot be
ignored if reliable predictions are to be made. Changing greenhouse gas
concentrations are important even for seasonal forecasting, because
including realistic greenhouse gas concentrations can significantly
improve forecast skill (Doblas-Reyes et al., 2006; Liniger et al., 2007).
Similarly, adaptation tools traditionally based on long-term records
(e.g., stream flow measurements over 50 to 100 years) coupled with the
assumption that the climate is not changing may lead to incorrect
conclusions about the best adaptation strategy to follow (Milly et al.,
2008). Thus reliable prediction and successful adaptation both need a
perspective that includes consideration of short to long time scales
(days to decades).
While there are potential benefits of early warning systems (NRC, 2003;
Shapiro et al., 2007) that span a continuum of time scales, for much of
the disaster risk management community the idea of preparedness
based on predictions is a new concept. Most communities have largely
operated in a reactive mode, either to disasters that have already occurred
or in emergency preparedness for an imminent disaster predicted with
high confidence (Chapter 5). The possibility of using weather and climate
predictions longer than a few days to provide advanced warning of
extreme conditions has only been a recent development (Brunet et al.,
2010; Shapiro et al., 2010). Despite over a decade of operational seasonal
predictions in many parts of the globe, examples of the use of such
information by the disaster risk management community are scarce, due
to the uncertainty of predictions and comprehension of their implications
(Patt et al., 2005; Meinke et al., 2006; Hansen et al., 2011). Most
seasonal rainfall predictions, for example, are presented as probabilities
that total rainfall over the coming few (typically three) months will be
amongst the highest or lowest third of rainfall totals as measured over
a historical period and these are averaged over large areas (typically
tens of thousands of square kilometers). Not only are the probabilities
lacking in precision but the target variable – seasonal rainfall total –
does not necessarily map well onto flood occurrence. Although higher-
than-normal seasonal rainfall will often be associated with a higher risk
of floods, it is possible for the seasonal rainfall total to be unusually
high yet no flooding occurs. Alternatively, the total may be unusually
low, yet flooding might occur because of the occurrence of an isolated
heavy rainfall event (Chapter 3). Thus even when seasonal predictions
are understood properly, it may not be obvious how to utilize them.
These problems emphasize the need for the development of tools to
translate such information into quantities directly relevant to end users.
Better communication between modeling centers and end users is
needed (Chapters 5 and 6). Where targeted applications have been
Chapter 9Case Studies
519
developed, some success has been reported (e.g., for malaria prediction)
(Thomson et al., 2006; Jones et al., 2007). Nonetheless there may be
additional obstacles such as policy constraints that restrict the range of
possible actions.
9.2.11.4. Interventions
There are many examples of interventions of early warning systems
outlined in the other case studies of this chapter and also in Chapters 5, 6,
and 7. As a part of their strategy of reducing risk, the Victorian Government
in Australia has established the heat wave early warning system for
metropolitan Melbourne and is undertaking similar work for regional
Victoria (see Case Study 9.2.2). A Storm Warning Center and associated
coastal volunteer network has been established in Bangladesh and has
been proven effective (Case Study 9.2.5). The absence of a storm warning
system in Myanmar contributed to the tragedy of that event (Case Study
9.2.5). The benefits of early warning systems are also discussed with
respect to floods (Case Study 9.2.6), heat waves (Case Study 9.2.1),
epidemic disease (Case Study 9.2.7), and drought (Case Study 9.2.3).
9.2.11.5. Outcomes
There have been examples of major benefits of early warning systems
(Einstein and Sousa, 2007). Assessments of community capacity to
respond to cyclone warnings have been performed for India (Sharma et
al., 2009), Florida (Smith and McCarty, 2009), New Orleans (Burnside et
al., 2007), New South Wales, Australia (Cretikos et al., 2008), and China
(Wang et al., 2008). Predictions of landfall for tropical cyclones are
important (Davis et al., 2008). In Bangladesh (Case Study 9.2.5; Paul,
2009), the implementation of an early warning system enabled people
to evacuate a hazardous area promptly (Paul and Dutt, 2010; Stein et
al., 2010). If forecasts are frequently incorrect, the response of people is
affected (Chapter 5; Dow and Cutter, 1998). Public health impacts of
hazards also depend on the preparedness of the local community (Vogt
and Sapir, 2009) and this can be improved by early warnings. However,
accurate predictions alone are insufficient for a successful early warning
system, as is demonstrated by the case in the United Kingdom – a country
that regularly experiences flooding (Parker et al., 2009). Severe damage
and health problems followed flooding in 2007 due to insufficiently clear
warning communication, issued too late and inadequately coordinated,
so that people, local government, and support services were unprepared
(UNISDR, 2009c). Heat-health warnings (Case Study 9.2.1) have proved
more effective (Fouillet et al., 2008; Hajat et al., 2010; Michelozzi et al.,
2010; Rubio et al., 2010) although improvements are still needed
(Kalkstein and Sheridan, 2007).
Notwithstanding the difficulties outlined for use of seasonal predictions in
disaster risk management, the successful use of such predictions has been
possible (IRI, 2011). Since all preparative actions have some direct cost
and it is impractical to be always prepared for all eventualities, seasonal
predictions can help to choose priorities from a list of actions.
9.2.11.6. Lessons Identified
Early warning systems for extreme weather- or climate-related events,
such as heat waves, floods, and storms have been implemented to
provide warnings on time scales of hours to days. The skill of warnings
beyond a few days ahead is improving as seasonal predictions are now
demonstrating benefits for drought, floods, and other phenomena, and
decadal forecasts of increased numbers of intense precipitation events
and heat waves are now being factored into planning decisions (NRC,
2003; Lazo et al., 2009; Goddard et al., 2010). It is expected that early
warning systems will enable the implementation of DRR and CCA. Early
warning systems rely on the ability of people to factor information on the
future into plans and strategies and need to be coupled with education
programs, legislative initiatives, and scientific demonstrations of the
skill and cost value benefits of these systems.
9.2.12. Effective Legislation for Multilevel Governance
of Disaster Risk Reduction and Adaptation
9.2.12.1. Introduction
This case study, through focus on South Africa’s disaster risk management
law and comparable legal arrangements in other states, such as the
Philippines and Colombia, explores critical provisions for effective
legislation. South Africa’s legislation has served as a model for others
(Pelling and Holloway, 2006; Van Niekerk, 2011) because it focuses on
prevention, decentralizes DRR governance, mandates the integration of
DRR into development planning, and requires stakeholder inclusiveness.
Implementation has proven challenging, however, particularly at the
local level (NDMC, 2007, 2010; Visser and Van Niekerk, 2009; Botha et
al., 2011; Van Niekerk, 2011) as is the case for most states (GNDR, 2009;
UNISDR, 2011b). Through analysis of South Africa’s legislation and the
difficulties that it faces in implementation, this study provides relevant
information to other governments as they assess whether their own
national legislation to reduce and manage disaster risk is adequate for
adapting to climate change.
9.2.12.2. Background
A legal framework establishes legal authority for programs and
organizations that relate to hazards, risk, and risk management. These
laws may dictate – or encourage – policies, practices, processes, the
assignment of authorities and responsibilities to individuals and/or
institutions, and the creation of institutions or mechanisms for
coordination or collaborative action among institutions (Mattingly, 2002).
Law can be used to provide penalties and incentives by enforcing
standards, to empower existing agencies or establish new bodies with
new responsibilities, and to assign budget lines (Pelling and Holloway,
2006). In short, legislation enables and promotes sustainable engagement,
helps to avoid disjointed action at various levels, and provides recourse
for society when things go wrong.
Chapter 9 Case Studies
520
Most states have some form of disaster risk management legislation or
are in the process of enacting it (UNDP 2005; UNISDR 2005b). In 2011,
48 countries reported substantial achievements in developing national
policy and legislation; importantly, almost half are low or lower-middle
income countries (UNISDR, 2011b). An increasing number of countries
have been adopting or updating existing legislation modeled on Hyogo
Framework for Action principles. Countries with new or updated laws
include India and Sri Lanka in 2005; El Salvador, Saint Lucia, Saint Vincent,
and the Grenadines in 2006; Anguilla (United Kingdom) and Gambia in
2007; Indonesia in 2008; Egypt and the Philippines in 2009; and Zambia
and Papua New Guinea in 2010 (UNISDR, 2011b). As yet some of the new
laws addressing disaster risk have not been harmonized with preexisting
legislative frameworks in relevant sectors, such as water, agriculture,
and energy (UNISDR, 2011b). Although these national legislations for
disaster management do not necessarily include a disaster risk reduction
orientation (Pelling and Holloway 2006), the evidence suggests a global
paradigm shift from the former responsive approach to disaster
management toward more long-term, sustainable preventive action
(Britton, 2006; Benson, 2009). India, Pakistan, Indonesia, South Africa,
and several Central American states have enacted such paradigm-
changing amendments to disaster management legislation and, in
Ecuador, the “notion of risk-focused disaster management was rooted
directly into its new constitution adopted in 2008” (IFRC, 2011).
In the case of South Africa, the country was impacted by floods and
droughts, and there was a high motivation for change in the post-
Apartheid era (Pelling and Holloway, 2006; NDMC, 2007), which starting
in 1994 led to legislative reform for disaster risk reduction. A “Green
Paper” first solicited public input and debate, and a second “White
Paper” translated responses into policy options for further technical and
administrative deliberations. These documents are noteworthy for their
consultative approach and their emphasis on disaster risk reduction
rather than traditional response (Pelling and Holloway, 2006; NDMC,
2007). Thereafter the Government passed three disaster management
bills that culminated in the promulgation of the Disaster Management
Act No. 57 of 2002 and of the National Disaster Management Framework
in 2005 (Pelling and Holloway, 2006; NDMC, 2007).
9.2.12.3. Description of Strategy
South Africa’s 2002 Disaster Management Act and its National Disaster
Management Policy Framework of 2005 (Republic of South Africa, 2002,
2005) are noteworthy because they were among the first to focus on
prevention, decentralize DRR governance, mandate the integration of
DRR into development planning, and require stakeholder inclusiveness.
The Act and Framework define the hierarchical institutional structure
that governs disaster risk reduction at national, provincial, and municipal
levels. They effectively decentralize DRR by mandating each level of
government to create:
A disaster risk management framework – a policy focused on the
prevention and mitigation of risk
A disaster risk management center – inter alia, to promote an
integrated and coordinated system of management, to integrate
DRR into development plans, to maintain disaster risk management
information, to monitor implementation, and to build capacity
A disaster risk management advisory forum – for government and
civil society stakeholders in DRR to coordinate their actions
An interdepartmental disaster risk management committee – for
government departments to coordinate or integrate activities for
DRR, to compile disaster risk management plans, and to provide
interdepartmental accountability (Republic of South Africa, 2002,
2005; Van Niekerk, 2006).
The Act further details each entity’s responsibilities. South Africa’s
legislation makes a legal connection between disaster risk reduction
and development planning. Some other countries adopting this
approach include Comoros, Djibouti, Ethiopia, Hungary, Ivory Coast,
Mauritius, Romania, and Uganda (Pelling and Holloway, 2006).
The Act requires that municipalities include risk management plans in
their integrated development plans (Republic of South Africa, 2002; Van
Niekerk, 2006). Municipal-level requirements are supported with a
mandate for provincial governments to ensure that their disaster risk
management plans “form an integral part of development planning” and
for the National Centre of Disaster Management to develop guidelines
for the integration of plans and strategies into development plans
(Republic of South Africa, 2002).
Closely related to the ability to influence development planning is the
authority to lead coordinated government action for DRR across
government agencies. The interdepartmental committees mandated by
the Act for each level provide the opportunity to communicate plans
and develop strategies across ministries and departments, avoiding
unilateral action that may increase risk. The forums established by the
Act similarly give voice to additional stakeholders to participate in DRR
decisionmaking. South Africa, Colombia, and the Philippines’ DRR laws,
for example, include provisions for the involvement of NGOs, traditional
leaders, volunteers, community members, and the private sector in
disaster risk reduction.
9.2.12.4. Outcomes
Implementation of South Africa’s benchmark legislative provisions has
proven challenging. Many district municipalities have not yet established
the disaster management centers required by the Act or these are not
yet functioning adequately (Van Riet and Diedericks, 2010; Botha et al.,
2011). The majority of local municipalities (which are subdivisions of
district municipalities) have not yet established advisory forums
although it should be noted that the Act does not require their creation
at this level (Botha et al., 2011). A greater percentage of metropolitan
districts have established advisory forums. Similarly, interdepartmental
committees, which facilitate cross-sectoral governmental collaboration
and the integration of DRR into development planning, have also not yet
Chapter 9Case Studies
521
been established in a majority of municipalities. Although municipalities
reported good progress in integrating their disaster management plans
into integrated development plans (Botha et al., 2011), such plans as
yet contain little evidence of integration (Van Niekerk, 2011).
Provincial and municipal levels attribute the lack of progress in
implementing the Act and Framework to inadequate resources for
start-up costs for municipalities as well as for the continuous operations
of disaster risk reduction projects (Visser and Van Niekerk, 2009; NDMC,
2010). There is also the continuing need for resources for response
recovery and rehabilitation activities. Reasons for the lack of funding
include a lack of clarity of the Act on funding sources and confusion
regarding the processes to access various sources of funding (Visser and
Van Niekerk, 2009). Although the mechanisms to obtain funding exist,
not all municipalities and provinces are using them, hence the perception
of inadequate funding persists (Van Niekerk, 2011). In some cases there
appears to be an absolute lack of funds, such as in district municipalities,
which are more rural and less densely populated, and have a narrower
tax base to fund DRR (Van Riet and Diedericks, 2010). This situation is
similar to that in other countries, such as Colombia, where more than
80% of municipalities are able to assign only 20% of their own non-
earmarked resources to risk reduction and disaster response. Because
the law does not stipulate percentages and amounts, municipalities
allocate minimal sums for disaster risk reduction (MIJ, 2009) given
competing infrastructure and social spending needs (Cardona and
Yamín, 2007).
South Africa and Colombia’s experiences are replayed around the world.
Governments informed in their 2011 Hyogo Framework progress reports
that the lack of efficient and appropriate budget allocations remains one
of the major challenges for effective disaster risk reduction legislation
(UNISDR, 2011a). Even in countries in which funding for disaster risk
management is mandated by law, actual resource allocation for disaster
risk reduction remains low and is concentrated in preparedness and
response (UNDP, 2007). The Philippines has new legislation that
attempts to address these issues. The Philippines new Disaster Risk
Reduction and Management Act 10121 renames the Local Calamity
Fund as the Local Disaster Risk Reduction and Management Fund and
stipulates that no less than 5% shall be set aside for risk management
and preparedness (Republic of the Philippines, 2010). Further, to carry
out the provisions of the Act, the Commission allocated one billion
pesos or US$ 21.5 million (Republic of the Philippines, 2010). Unspent
money will remain in the fund to promote risk reduction and disaster
preparedness. The adequacy of this provision has yet to be tested as the
Act is recent and implementation has yet to begin.
In South Africa, all relevant national departments have yet to undertake
required DRR activities or identified sectoral focal points; consequently,
the advisory committee at the national level is not yet functioning
optimally (Van Niekerk, 2011). Similarly, at provincial and municipal levels,
departmental representatives are absent or too junior to make decisions
at meetings, reflecting lack of understanding about their department’s
role in DRR and about DRR generally (Van Riet and Diedericks, 2010).
Moreover, as mentioned above, between 55 and 73% of municipalities
have not established a committee, which Botha et al. (2011) point out
hampers local government’s ability to implement integrated multi-
sectoral DRM.
The Philippines’ Climate Change Act, enacted in 2009, addresses the
challenge of inter-sectoral government collaboration by creating a
commission to be chaired by the president and attached to that office, thus
ensuring highest-level political support for collaborative implementation
of the law (Republic of the Philippines, 2009). The commission is composed
of the secretaries of all relevant departments as well as the Secretary of
the Department of National Defense, as Chair of the National Disaster
Coordinating Council, and representatives from the disaster risk reduction
community. The main functions of the Commission are to “ensure the
mainstreaming of climate change, in synergy with disaster risk reduction,
into the national, sectoral, and local development plans and programs”
and to create a panel of technical experts, “consisting of practitioners
in disciplines that are related to climate change, including disaster risk
reduction” (Republic of the Philippines, 2009).
Implementing the multi-sectoral DRM envisioned by South Africa’s
legislation may be hindered by the placement of the National Disaster
Management Centre within a line ministry (the Department of
Cooperative Governance) (Van Niekerk, 2011). Sub-national levels have
likewise placed their centers within sectors with insufficient political
authority; consequently, local municipal and district levels rate current
interdepartmental collaboration as low (Botha et al., 2011). This placement
allows other departments to disregard DRM, as the National Disaster
Management Centre cannot enforce punitive measures (Van Niekerk,
2011).
Similar to South Africa’s arrangements, the Philippines’ highest
policymaking and coordinating body for disaster risk management, the
National Disaster Risk Reduction and Management Council (formerly
called the National Disaster Coordinating Council), sits within the
Department of National Defense. As such, it is focused on disaster
preparedness and response and does not have sustainable development
and poverty reduction responsibilities. The Philippines’ new Disaster
Risk Reduction and Management Act of 2010 attempts to redress this
issue by including experts from all relevant fields as members of the
Council and expressly defining its mandate on mainstreaming disaster
risk reduction into sustainable development and poverty reduction
strategies, policies, plans, and budgets at all levels (Republic of the
Philippines, 2010).
Positioning DRR institutions within the highest levels of government has
proven effective because this position often determines the amount of
political authority of the national disaster risk management body
(UNDP, 2007; UNISDR, 2009a). National disaster risk management
offices attached to prime ministers’ offices usually can take initiatives
affecting line ministries, while their colleagues operating at the sub-
ministerial level often face administrative bottlenecks (UNDP, 2007).
High-level support is particularly important to enable disaster risk
Chapter 9 Case Studies
522
reduction legislation to provide a framework for strategies to build risk
reduction into development and reconstruction (Pelling and Holloway,
2006).
9.2.12.5. Lessons Identified
The main lesson that emerges from this case study is that carefully crafted
legislation buttresses DRR activities, thus avoiding a gap between the
law’s vision and its implementation. The experiences of South Africa and
the Philippines in implementing their DRR legislation (as described by
Visser and Van Niekerk, 2009; NDMC, 2010; Van Riet and Diedericks,
2010; Botha et al., 2011; Van Niekerk, 2011) and the literature on DRR
legislation (Mattingly, 2002; Britton, 2006; Pelling and Holloway, 2006;
UNDP, 2007; Benson, 2009; UNISDR, 2009c) point to the following
elements of effective legislation and implementation:
The law allocates adequate funding for implementation at all levels
with clarity about the generation of funds and procedures for
accessing resources at every administrative level.
The institutional arrangements provide both access to power for
facilitating implementation and opportunities to ‘mainstream’
disaster risk reduction and adaptation into development plans.
The law includes provisions that increase accountability and enable
coordination and implementation, that is, the clear identification of
roles and responsibilities and access to participate in decisionmaking.
An additional element is the need for periodic assessment and revision
to ensure that legislation for disaster risk reduction and adaptation is
dynamic and relevant (Llosa and Zodrow, 2011). For instance, the
Philippines’ Disaster Risk Reduction Management (DRRM) Act calls for
the development of a framework to guide disaster risk reduction and
management efforts to be reviewed “on a five-year interval, or as may
be deemed necessary, in order to ensure its relevance to the times”
(Republic of the Philippines, 2010). The DRRM Act also calls for the
development of assessments on hazards and risks brought about by
climate change (Republic of the Philippines, 2010). Likewise, the
Philippines Climate Change Act calls for the framework strategy that will
guide climate change planning, research and development, extension,
and monitoring of activities to be reviewed every three years or as
necessary (Republic of the Philippines, 2009). Similarly, the United
Kingdom’s Climate Change Act establishes the preparation of a report
informing parliament on risks of current and predicted impact of climate
change no later than five years after the previous report (United
Kingdom, 2008). Thus an additional element for effective DRR-adaptation
legislation may be that the law be based on up-to-date risk assessment
and mandates periodic reassessment as risks evolve and knowledge of
climate change impacts improves.
Developing and enacting legislation takes considerable time and political
capital. It took South Africa and the Philippines about a decade to enact
comprehensive disaster risk reduction frameworks. Linking the
development of disaster risk reduction legislation to the politically
prominent climate change discussion could substantially increase the
sense of urgency and thus speed of parliamentary processes (Llosa and
Zodrow, 2011).
Another method for hastening the legislative process would be to first
assess the adequacy of existing disaster risk reduction legislation and
strengthen these laws rather than starting a wholly new drafting and
negotiations process for adaptation that may create a parallel legal and
operational system (Llosa and Zodrow, 2011). As frequently reported
(e.g., UNDP, 2007; UNISDR, 2009c), an overload of laws and regulations
without a coherent and comprehensive framework, clear competencies,
and budget allocations hinders the effective implementation of disaster
risk reduction legislation.
9.2.13. Risk Transfer: The Role of Insurance and Other
Instruments in Disaster Risk Management and
Climate Change Adaptation in Developing Countries
9.2.13.1. Introduction
The human and economic toll from disasters can be greatly amplified by
the long-term loss in incomes, health, education, and other forms of
capital resulting from the inability of communities to restore infrastructure,
housing, sanitary conditions, and livelihoods in a timely way (Mechler,
2004; Mills, 2005). By providing timely financial assistance following
extreme event shocks, insurance and other risk-transfer instruments
contribute to DRR by reducing the medium- and long-term consequences
of disasters. These instruments are widespread in developed countries,
and are gradually becoming part of disaster management in developing
countries, where novel micro-insurance programs are helping to put
cash into the hands of affected poor households so they can begin
rebuilding livelihoods (Bhatt et al., 2010). These mechanisms can also
contribute to reducing vulnerability and advancing development even
before disasters strike by providing the requisite security for farmers and
firms to undertake higher-return, yet more risky investments in the face
of pervasive risk. Governments also engage in risk transfer. Investors
can be encouraged to invest in a country if there is evidence that the
government has reduced its risks (Gurenko, 2004).
9.2.13.2. Background
This case study focuses on instruments for risk transfer in order to manage
catastrophe risk in developing countries (see also Sections 5.5.2, 6.5.3,
and 7.4.4). Table 9-3 provides an overview of financial instruments and
arrangements, including risk transfer, as they are employed by households,
farmers, small- and medium-sized enterprises (SMEs), and governments,
as well as international organizations and donors. Typically, losses are
reimbursed on an ad hoc basis after disasters strike through appeals to
solidarity, for example, from neighbors, governments, and international
donors. Households and other agents also rely on savings and credit,
and many governments set aside national or sub-national level reserve
funds. Alternatively, agents can engage in risk transfer (the shaded cells
Chapter 9Case Studies
523
in Table 9-3), which is defined by UNISDR as “the process of formally or
informally shifting the financial consequences of particular risks from
one party to another whereby a household, community, enterprise or
state authority will obtain resources from the other party after a disaster
occurs, in exchange for ongoing or compensatory social or financial
benefits provided to that other party” (UNISDR, 2009b). Risk sharing
can be considered synonymous with risk transfer, although the latter is
often used to connote more informal forms of shifting risk without
explicit compensation or payment, for example, mutual non-market
arrangements among family or community. Insurance is the best known
form of market risk transfer; yet, risks can be transferred with many
formal and informal instruments as described in the following section.
Traditional channels for financing disaster relief and recovery, although
in many cases less costly than risk transfer, have in the past proved to
be inadequate for managing large-scale weather-related events in highly
vulnerable countries (Cohen and Sebstad, 2003; Cardenas et al., 2007;
Barnett et al., 2008). In poor countries, households and businesses usually
do not have the resources to purchase commercial insurance to cover
their risks with the additional difficulty that in many developing countries
the commercial insurance providers do not exist. If there is no support
from family or government, disasters can lead to a worsening of poverty
in the absence of insurance. The victims then must either obtain high-
interest loans (or default on existing loans), sell their important and valued
assets and livestock, or engage in low-risk, low-yield farming to lessen
their exposure to extreme events (Varangis et al., 2002). In recognition of
these issues and to reduce the overall costs of disasters, investments in
disaster risk reduction and proactive risk transfer are strongly encouraged
by governments, the insurance sector, and the donor community (Kreimer
and Arnold, 2000; Gurenko, 2004; Linnerooth-Bayer et al., 2005).
9.2.13.3. Description of Strategy – Catastrophe Risk Transfer
Mechanisms and Instruments
As shown by the shaded cells in Table 9-3, risk transfer includes a range
of pre-disaster mechanisms and instruments (Cummins and Mahul,
2009), the most important of which are briefly described below:
Informal mutual arrangements involve pre-agreed non-market
exchanges of post-disaster support (informal risk sharing).
Insurance is a “well-known form of risk transfer, where coverage of
a risk is obtained from an insurer in exchange for ongoing premiums
paid to the insurer” (UNISDR, 2009b). A contractual transaction
based on a premium is used to guarantee financial protection
against potentially large loss; contracts typically cover losses to
property, productive assets, commercial facilities, crops and
livestock, public infrastructure (sovereign insurance), and business
interruption.
Micro-insurance, based on the same principles as insurance, is
aimed, most often, at lower-income individuals who cannot afford
traditional insurance and hence the premiums are lower but also
the coverage may be restricted. In some cases, the individuals are
unable to access more traditional insurance (Mechler et al., 2006).
Often it is provided in innovative partnerships involving
communities, NGOs, self-help groups, rural development banks,
insurers, government authorities, and donors.
Alternative risk transfer denotes a range of arrangements that
hedge risk (Mechler et al., 2006). These include catastrophe bonds,
which are instruments where the investor receives an above-
market return when a pre-specified catastrophe does not occur
within a specified time interval. However, the investor sacrifices
interest or part of the principal following the event.
Weather derivatives typically take the form of a parametric
(indexed-based) transaction, where payment is made if a chosen
weather index, such as 5-day rainfall amounts, exceeds some
predetermined threshold.
Contingent credit (also called deferred drawdown option) is a
prearranged loan contingent on a specified event; it can be provided
by the insurance industry to other insurers, or by international
financial institutions to governments.
Risk pools aggregate risks regionally (or nationally) allowing
individual risk holders to spread their risk geographically.
9.2.13.4. Interventions – Examples of Local, National, and
International Risk Transfer for Developing Countries
Development organizations working together with communities,
governments, insurers, and NGOs have initiated or supported many
recent pilot programs offering risk transfer solutions in developing
Chapter 9 Case Studies
Weather derivatives
Property insurance; crop and livestock insurance;
micro-insurance
Bilateral and multilateral assistance, regional
solidarity funds
Contingent credit; emergency liquidity funds
Re-insurance; regional catastrophe insurance
pools
Catastrophe bonds; risk swaps, options, and loss
warranties
Catastrophe bonds
National insurance programs; sovereign risk
transfer
Reserve funds; domestic bonds
Government post-disaster assistance;
government guarantees/bailouts
Solidarity
Informal risk transfer (sharing)
Savings, credit, and storage
(inter-temporal risk spreading)
Insurance instruments
Alternative risk transfer
Savings; micro-savings; fungible assets; food
storage; money lenders; micro-credit
Kinship and other reciprocity obligations,
semi-formal micro-finance, rotating savings and
credit arrangements, remittances
Help from neighbors and local organizations
Local
Households, Farmers, SMEs
National
Governments
International
Development organizations, donors, NGOs
Table 9-3 | Examples of risk financing mechanisms (shaded cells) at different scales. Source: adapted from Linnerooth-Bayer and Mechler, 2009.
524
countries. Three examples at the local, national, and international scales
are briefly discussed below.
9.2.13.4.1. Covering local risks: index-based micro-insurance
for crop risks in India
Micro-insurance to cover, for example, life and health is widespread in
developing countries, but applications for catastrophic risks to crops
and property are in the beginning phases (see Morelli et al., 2010 for a
review, and Loster and Reinhard, 2010 for a focus on micro-insurance
and climate change). Typically a micro-insurance company, often
operating on a not-for-profit basis, evolves from an organization that
has developed insurance products for a community. Most are based on
the expectation that the pool of participants will provide payments that
cover the costs incurred, including expected damage claims (which are
generally low because of infrequent and small claims), administrative
costs (which are reduced through group contracts or linking contracts
to loans), taxes, and regulatory fees. Many depend on the support of
government subsidies and international development organizations and
participation of NGOs (Mechler et al., 2006).
An innovative insurance program set up in India in 2003 covers non-
irrigated crops in the state of Andhra Pradesh against the risk of
insufficient rainfall during key times during the cropping season. The
index-based policies are offered by a commercial insurer and marketed
to growers through microfinance banks. In contrast to conventional
insurance, which is written against actual losses, this index-based
(parametric) insurance is written against a physical or economic trigger,
in this case rainfall measured by a local rain gauge. The scheme owes its
existence to technical assistance provided by the World Bank (Hess and
Syroka, 2005). Schemes replicating this approach are currently targeting
700,000 exposed farmers in India (Cummins and Mahul, 2009).
One advantage of index-based insurance is the substantial decrease in
transaction costs due to eliminating the need for expensive post-event
claims handling, which has impeded the development of insurance
mechanisms in developing countries (Varangis et al., 2002). A disadvantage
is basis risk, which is the lack of correlation of the trigger with the loss
incurred. If the rainfall measured at the weather station is sufficient, but
for isolated farmers insufficient, they will not receive compensation for
crop losses. Similar schemes are implemented or underway, for instance,
in Malawi, Ukraine, Peru, Thailand, and Ethiopia (Hellmuth et al., 2009). A
blueprint for insuring farmers in developing countries who face threats to
their livelihoods from adverse weather has been developed (World Bank,
2005d). Overcoming major institutional and other barriers must be done
in order for these programs to achieve this target (Hellmuth et al., 2009).
Weather insurance and especially index-based contracts contribute, in at
least two ways, to climate change adaptation and disaster risk reduction.
Since farmers will receive payment based on rainfall and thus have an
incentive to plant weather-resistant crops, indexed contracts eliminate
moral hazard, which is defined as the disincentive for risk prevention
provided by the false perception of security when purchasing insurance
coverage. Second, an insurance contract renders high-risk farmers more
creditworthy, which enables them to access loans for agricultural inputs.
This was illustrated in the pilot program in Malawi, where farmers
purchased index-based drought insurance linked to loans to cover costs
of hybrid seed, with the result that their productivity was doubled
(Linnerooth-Bayer et al., 2009). Increased productivity decreases
vulnerability to weather extremes, thus contributing to climate change
adaptation (to the extent that risks of weather extremes are increased
by climate change). In another innovative micro-insurance project in
Ethiopia, farmers can pay their premiums by providing labor on risk-
reducing projects (Suarez and Linnerooth-Bayer, 2010).
9.2.13.4.2. Covering national risks:
the Ethiopian weather derivative
The World Food Programme (WFP), to supplement and partly replace its
traditional food-aid approach to famine, has recently supported the
Ethiopian government-sponsored Productive Safety Net Programme
(PSNP). The WFP is now insuring it against extreme drought (World Bank,
2006b). When there is a food emergency, the PSNP is able to provide
immediate cash payments that may be sufficient to save lives even in
the case of very severe droughts (Hess et al., 2006). However, these
payments may not be sufficient to restore livelihoods (World Bank, 2006b).
To provide extra capital in the case of extreme drought, an index-based
contract, sometimes referred to as a weather derivative, was designed
by the WFP. The amount of capital is based on contractually specified
catastrophic shortfalls in precipitation based on the Ethiopia Drought
Index (EDI). The EDI depends on rainfall amounts that were measured at
26 weather stations that represent the various agricultural areas of
Ethiopia. In 2006, the WFP successfully obtained an insurance contract
based on the EDI through an international reinsurer (Hess et al., 2006).
A drawback of this arrangement, in contrast to the micro-insurance
programs in India and Malawi, is that it perpetuates dependence on
post-drought government assistance with accompanying moral hazard.
9.2.13.4.3. Intergovernmental risk sharing: the Caribbean
Catastrophe Risk Insurance Facility (CCRIF)
The world’s first regional catastrophe insurance pool was launched in
2007 in the Caribbean region; this is the Caribbean Catastrophe Risk
Insurance Facility (discussed in Section 7.4.4). Sixteen participating
governments secured insurance protection against costs associated
with catastrophes such as hurricanes and earthquakes (Ghesquiere et
al., 2006; World Bank, 2007). Several of the participating countries
represent the countries experiencing the greatest economic losses from
disasters in the last few decades, when measured as a share of GDP
(CCRIF, 2010).
The aim of the Caribbean facility is to provide immediate liquidity to
cover part of the costs that participating governments expect to incur
Chapter 9Case Studies
525
while they provide relief and assistance for recovery and rehabilitation.
Because it does not cover all costs, CCRIF provides an incentive for
governments to invest in risk reduction and other risk transfer tools. The
cost of participation is based on estimates of the respective countries’ risk
(measured as probability and cost). The advantage of pooling is that due
to diversifying risk it greatly reduces the costs of reinsurance compared
to the price each government would have paid individually. Funding
for the program, although mainly the responsibility of participating
countries, has been supported by a donor conference hosted by the
World Bank.
Insofar as weather extremes are increased by climate change, the CCRIF
contributes directly to disaster risk reduction and climate change
adaptation. By providing post-event capital it enables governments to
restore critical infrastructure so important for reducing the long-term
human and economic impacts from hurricanes. Experience with CCRIF
also shows the importance of designing programs that reflect the needs
of the participating countries. Finally, it demonstrates how international
assistance can support disaster management in tandem with national
responsibility.
9.2.13.4.4. Outcomes – the role of risk transfer for advancing
disaster risk reduction and climate change adaptation
As these examples illustrate, risk-transfer instruments and especially
insurance can promote disaster risk reduction and climate change
adaptation by enabling recovery and productive activities. By providing
means to finance relief, recovery of livelihoods, and reconstruction,
insurance reduces long-term indirect losses – even human losses – that
do not show up in the disaster statistics. Risk transfer arrangements
thus directly lead to the reduction of post-event losses from extreme
weather events, what is commonly viewed as adaptation. Moreover,
insured households and businesses can plan with more certainty, and
because of the safety net provided by insurance, they can take on cost-
effective, yet risky, investments. This ultimately reduces vulnerability
to weather extremes and by so doing contributes to climate change
adaptation.
Experience in developed countries has demonstrated additional ways in
which insurance and other risk-transfer instruments have promoted
DRR and CCA as listed below:
Because risk-transfer instruments require detailed analysis of risk,
they can both raise awareness and provide valuable information
for its response and reduction; for example, in some developed
countries insurers with other partners have made flood and other
hazard maps publicly available (Botzen et al., 2009; Warner et al.,
2009). Potential challenges include the technical difficulties related
to risk assessment, dissemination of appropriate information, and
overcoming education and language barriers in some areas.
By pricing risk, insurance can provide incentives for investments
and behavior that reduce vulnerability and exposure, especially if
premium discounts are awarded. Differential premium pricing has
been effective in discouraging construction in high-risk areas; for
example, UK insurers price flood policies according to risk zones,
but insurers are reluctant to award premium discounts for other
types of mitigation measures, such as reinforcing windows and
doors to protect against hurricanes (Kunreuther and Roth, 1998;
Kunreuther and Michel-Kerjan, 2009). The incentive effect of
actuarial risk pricing should be weighed against the benefits of
increasing insurance penetration to those unable to afford risk-
based premiums. The positive incentives provided by insurance
should also not overshadow the potential for negative incentives
or moral hazard.
Insurers and other providers can make risk reduction a contractual
stipulation, for example, by requiring fire safety measures as a
condition for insuring a home or business (Surminski, 2010). The US
National Flood Insurance Program requires communities to reduce
risks as a condition for offering subsidized policies to their residents
(Kunreuther and Roth, 1998; Linnerooth-Bayer et al., 2007). It was
noted above that the WFP might require risk-reducing activities as
a condition for its support for weather derivatives.
Providers can partner with government and communities to
establish appropriate regulatory frameworks and promote, for
instance, land use planning, building codes, emergency response,
and other types of risk-reducing policies. Ungern-Sternberg (2003)
has shown that Swiss cantons having public monopolies that provide
disaster insurance outperform cantons with private systems in
reducing risks and premiums, mainly because the public monopolies
have better access to land use planning institutions, fire departments,
and other public authorities engaged in risk reduction. In many
countries, insurers have co-financed research institutes and disaster
management centers, and in other cases, have partnered with
government to achieve changes in the planning system and
investment in public protection measures (Surminski, 2010).
9.2.13.5. Lessons Identified
Governments, households, and businesses can experience liquidity gaps
limiting their ability to recover from disasters (high confidence). There is
robust evidence to suggest that risk-transfer instruments can help
reduce this gap, thus enabling recovery.
There are a range of risk-transfer instruments, where insurance is the
most common. With support from the international community, risk
transfer is becoming a reality in developing countries at the local,
national, and international scales, but the future is still uncertain. Index-
based contracts greatly reduce transaction costs and moral hazard
(medium confidence); while more costly than many traditional financing
measures, insurance has benefits both before disasters (by enabling
productive investment) and after disasters (by enabling reconstruction
and recovery) (medium confidence). Insurance and other forms of risk
transfer can be linked to disaster risk reduction and climate change
adaptation by enabling recovery, reducing vulnerability, and providing
knowledge and incentives for reducing risk (medium confidence).
Chapter 9 Case Studies
526
9.2.14. Education, Training, and Public Awareness Initiatives
for Disaster Risk Reduction and Adaptation
9.2.14.1. Introduction
Disasters can be substantially reduced if people are well informed and
motivated to prevent risk and to build their own resilience (UNISDR,
2005b). Disaster risk reduction education is broad in scope: it encompasses
primary and secondary schooling, training courses, academic programs,
and professional trades and skills training (UNISDR, 2004), community-
based assessment, public discourse involving the media, awareness
campaigns, exhibits, memorials, and special events (Wisner, 2006). Given
the breadth of the topic, this case study illustrates just a few practices
in primary school education, training programs, and awareness-raising
campaigns in various countries.
9.2.14.2. Background
The Hyogo Framework calls on states to “use knowledge, innovation,
and education to build a culture of safety and resilience at all levels”
(UNISDR, 2005b). States report minor progress in implementation,
however (UNISDR, 2009c). Challenges noted include the lack of capacity
among educators and trainers, difficulties in addressing needs in poor
urban and rural areas, the lack of validation of methodologies and tools,
and little exchange of experiences. On the positive side, the 2006-2007
international campaign “Disaster Risk Reduction Begins at School
(UNISDR, 2006) raised awareness of the importance of education with
55 governments undertaking awareness-raising activities and 22
governments reporting success in making schools safer (e.g., 175 schools
developed disaster plans in Gujarat, India) by developing educational
and training materials, introducing school drills, and implementing DRR
teacher trainings (UNISDR, 2008b). Furthermore, the implementation
scheme of the United Nations Decade of Education for Sustainable
Development 2005-2014 seeks to improve the knowledge base on
disaster reduction as one of the keys to sustainable development.
A related emerging trend is to engage children in disaster risk reduction
and adaptation, as children are increasingly understood as effective agents
of change (Mitchell et al., 2009). Children’s inclusion also increases the
likelihood that they will maintain their own DRR and adaptation learning
(Back et al., 2009). A report from five NGOs (Twigg and Bottomley, 2011)
states that their DRR work with children and young people involves risk
identification and action planning for preparedness; training of school
teachers and students; DRR curriculum development; youth-led prevention
and risk reduction actions, such as mangrove and tree conservation;
awareness raising (e.g., through peer-to-peer community exchanges
and children’s theater); and “lobbying and networking in promoting
and supporting children’s voice and action.
Effective DRR education initiatives seek to elicit behavioral change not
only by imparting knowledge of natural hazards but also by engaging
people in identifying and reducing risk in their surroundings. In formal
education, disaster risk education should not be confined within the
school but promoted to family and community (Shaw et al., 2004).
Lectures can create knowledge, particularly if presented with visual aids
and followed up with conversation with other students. Yet it is family,
community, and self learning, coupled with school education, that
transform knowledge into behavioral change (Shaw et al., 2004).
9.2.14.3. Description of Strategies
9.2.14.3.1. School curriculum
States are increasingly incorporating DRR in the curriculum (UNISDR,
2009c) and have set targets for so doing in all school curricula by 2015
(UNISDR, 2009c). Initiatives to integrate the teaching of climate change
and DRR are also emerging, such as the described Philippines program.
Importantly, the new Philippines disaster risk reduction and climate
change laws mandate the inclusion of DRR and climate change,
respectively, in school curricula; the following example predates these
laws, however.
The Asian Disaster Preparedness Centre and UN Development
Programme, with the National Disaster Coordinating Council and support
from the European Commission Humanitarian Aid and Civil Protection,
assisted the Ministry of Education in the Philippines, Cambodia, and Lao
People’s Democratic Republic to integrate disaster risk reduction into
the secondary school curriculum. Each country team developed its own
draft module, adapting it to local needs.
The Philippines added climate change and volcanic hazards into its
disaster risk reduction curriculum. The relevant lessons addressed ‘what
is climate change,’ they then asked ‘what is its impact,’ and finally ‘how
can you reduce climate change impact?’ Other lessons focused on the
climate system, typhoons, heat waves, and landslides, among other
related topics (Luna et al., 2008). The Philippines’ final disaster risk
reduction module was integrated into 12 lessons in science and 16
lessons in social studies for the first year of secondary school (Grade 7)
(Luna et al., 2008). Each lesson includes group activities, questions to be
asked of the students, the topics that the teacher should cover in the
lecture, and a learning activity in which students apply knowledge
gained and methodology for evaluation of learning by the students
(Luna et al., 2008). The project reports that it reached 1,020 students,
including 548 girls, who learned about disaster risk reduction and climate
change. Twenty-three teachers participated in the four-day orientation
session. An additional 75 teachers and personnel were trained to train
others and replicate the experience across the country (Luna et al., 2008).
9.14.2.3.2. Training for disaster risk reduction and adaptation
In order to effectively include disaster risk reduction and adaptation in
the curriculum, teachers require (initial and in-service) training on
the substantive matter as well as the pedagogical tools (hands-on,
Chapter 9Case Studies
527
experiential learning) to elicit change (Shiwaku et al., 2006; Wisner,
2006). Education program proponents might have to overcome teachers’
resistance to incorporate yet another topic into overburdened curricula.
To enlist teachers’ cooperation, developing a partnership with the
ministry of education and school principals can be helpful (UNISDR,
2007b; World Bank et al., 2009). The following program in Indonesia
and the evaluation results from Nepal demonstrate the importance of
engaging teachers for effective education. The subsequent example from
Nepal, Pakistan, and India focuses on training builders through extensive
hands-on components in which new techniques are demonstrated and
participants practice these techniques under expert guidance (World
Bank et al., 2009).
The Disaster Awareness in Primary Schools project, which provides
teacher training, was launched in Indonesia in 2005 with German support
and is ongoing. By 2007, through this project, 2,200 school teachers had
received DRR training. Project implementers found that existing teaching
methods were not conducive to active learning. Students listened to
teacher presentations, recited facts committed to memory, and were not
encouraged to understand concepts and processes. The training took
teachers’ capabilities into account by emphasizing the importance of
clarity and perseverance in delivering lessons so as to avoid passing on
faulty life-threatening information (e.g., regarding evacuation routes).
Scientific language was avoided and visual aids and activities encouraged.
Teachers were asked to take careful notes and to participate in practical
activities such as first-aid courses, thus modeling proactive learning.
Continuity with the teachers’ traditional teaching methods was
maintained by writing training modules in narrative form and following
the established lesson plan model. Moreover, to avoid further burdening
teachers’ heavy lesson requirements and schedules, the modules were
designed to be integrated into many subjects, such as language and
physical education, and to require minimum preparation (UNISDR, 2007b).
In Nepal, Kyoto University researchers evaluated the knowledge and
perceptions of 130 teachers in 40 schools, most of whom were imparting
disaster education (Shiwaku et al., 2006). Through responses to a survey,
the researchers found that the content of the disaster risk education
being imparted depended on the awareness of individual teachers.
Teachers focused lessons on the effects of disasters with which they
could relate from personal experience. The researchers concluded that
teacher training is the most important step to improve disaster risk
education in Nepal. Most social studies teachers reported a need for
teacher training but the survey analysis recommended that training
programs be designed to integrate DRR into any subject rather than
taught in special classes (Shiwaku et al., 2006).
The National Society for Earthquake Technology in Nepal conducted
large-scale training for masons, carpenters, bar benders, and construction
supervisors in 2007 over a five-month period to impart risk-resilient
construction practices and materials. Participants from Kathmandu
and five other municipalities formed working groups to train other
professionals. As the project was successful, a mason-exchange program
was designed with the Indian NGO Seeds. Nepali masons were sent to
Gujarat, India, to mentor local masons in the theory and practice of safer
construction. Also in India, the government of Uttar Pradesh trained two
junior engineers in the rural engineering service in each district to carry
out supervisory inspection functions and delegated the construction
management to school principals and village education committees.
Similarly, the Department of Education of the Philippines mandated
principals to take charge of the management of the repair and/or
construction of typhoon-resistant classrooms after the 2006 typhoons.
Assessment, design, and inspection functions were provided by the
Department’s engineers, who also assist with auditing procurement
(World Bank et al., 2009).
9.2.14.3.3. Raising public awareness
In addition to the insights on the psychological and sociological aspects
of risk perception, risk reduction education has benefited from lessons in
social marketing. These include involving the community and customizing
for audiences using cultural indicators to create ownership; incorporating
local community perspectives and aggressively involving community
leaders; enabling two-way communications and speaking with one
voice on messages (particularly if partners are involved); and evaluating
and measuring performance (Frew, 2002).
According to the UNISDR Hyogo Framework Mid-Term Review (UNISDR,
2011a), few DRR campaigns have translated into public action and
greater accountability. However, successful examples include Central
America and the Caribbean, where the media played an important role,
including through radio soap operas. The UNISDR review also found a
high level of risk acceptance, even among communities demonstrating
heightened risk awareness. In some cultures, the spreading of alarming
or negative news – such as information on disaster risks – is frowned
upon (UNISDR, 2011a). The following examples from Brazil, Japan, and
the Kashmir region illustrate good practice in raising awareness for risk
reduction.
Between 2007 and 2009, the Brazilian Santa Catarina State Civil
Defence Department, with the support of the Executive Secretariat and
the state university, undertook a public awareness initiative to reduce
social vulnerability to disasters induced by natural phenomena and
human action (SCSCDD, 2008a,b). During the two-year initiative, 2,000
educational kits were distributed free of charge to 1,324 primary
schools. Students also participated in a competition of drawings and
slogans that were made into a 2010 calendar. As the project’s goal was
public awareness of risk, the project jointly launched a communications
network in partnership with media and social networks to promote better
dissemination of risk and disasters (SCSCDD, 2008a,b). The initiative
also focused on the most vulnerable populations. A pilot project for 16
communities precariously perched on a hill prone to landslides featured
a 44-hour course on risk reduction. Community participants elaborated
risk maps and reduction strategies, which they had to put to use
immediately. Shortly into the course, heavy rains battered the state,
triggering a state of emergency; 10 houses in the pilot project area had
Chapter 9 Case Studies
528
to be removed and over 50 remained at risk. The participants’ risk
reduction plans highlighted the removal of garbage and large rocks as
well as the building of barriers. The plans also identified public entities
for partnership and the costs for services required. The training closed
with a workshop on climate change and with the community leaders’
presentation of the major risk reduction lessons learned (SCSCDD,
2008c). On international disaster risk reduction day, representatives of
the community, Civil Defence, and other public entities visited the most
at-risk areas of the hill community, planted trees, installed signs pointing
out risky areas and practices, distributed educational pamphlets, and
discussed risk. One of the topics of discussion was improper refuse
disposal and the consequent blocking of drains, causing flooding
(SCSCDD, 2008d).
In 2004, typhoons resulted in flooding in urban areas of Saijo City
(Ehime Prefecture of Shikoku Island, Japan). There were also landslides
in the mountains. As a result, a public awareness campaign was
implemented. Saijo City, a small city with semi-rural mountainous areas,
faces challenges in disaster risk reduction that are relatively unique. In
Japan, young people have a tendency to leave smaller communities and
move to larger cities. The result is that Japanese smaller towns have
older than the national average populations. Since younger, able-
bodied people are important for community systems of mutual aid and
emergency preparedness, there is a special challenge. Saijo City has an
urban plain, semi-rural and isolated villages on hills and mountains, and
a coastal area and, hence, is spread over a mix of geographic terrains
(Yoshida et al., 2009; ICTILO et al., 2010); this brings another challenge.
In 2005, the Saijo City Government launched a risk awareness program
to meet both of these challenges through a program targeted at school
children. The project for 12-year olds has a ‘mountain-watching’ focus
for the mountainside and a ‘town-watching’ focus for the urban area
(ICTILO et al., 2010). The students are taken, accompanied by teachers,
forest workers, local residents, and municipal officials, on risk-education
field trips. In the mountains, the young urban dwellers meet with the
elderly and they learn together about the risks the city faces. Part of the
process is to remember the lessons learned from the 2004 typhoons.
Additionally, a ‘mountain and town watching’ handbook has been
developed, a teachers’ association for disaster education was formed, a
kids’ disaster prevention club started, and a disaster prevention forum
for children was set up (Yoshida et al., 2009; ICTILO et al., 2010). This is
an example of a local government both conceiving and implementing the
program. The city government led a multi-stakeholder and community-
based disaster risk awareness initiative that then became self-sustaining.
Professionals from disaster reduction and education departments were
provided through government support. The government also funds the
town and mountain watching and puts on an annual forum (ICTILO et
al., 2010).
The Centre for Environment Education (CEE) Himalaya is undertaking a
disaster risk reduction and climate change education campaign in 2,000
schools and 50 Kashmir villages in the Himalayas. In the schools, teachers
and students are involved in vulnerability and risk mapping through
rapid visual risk assessment and in preparing a disaster management
plan for their school. Disaster response teams formed in selected
schools have been trained in life-saving skills and safe evacuation (CEE
Himalaya, 2009).
CEE Himalaya celebrated International Mountain Day 2009 with educators
by conducting a week-long series of events on climate change adaptation
and disaster risk reduction. About 150 participants including teachers
and officials of the Department of Education, Ganderbal, participated in
these events (CEE Himalaya, 2009). Participants worked together to
identify climate change impacts in the local context, particularly in terms
of water availability, variation in microclimate, impact on agriculture/
horticulture and other livelihoods, and vulnerability to natural disasters.
The concept of School Disaster Management Plans (SDMPs) was
introduced. Participants actively prepared SDMPs for their schools
through group exercises, and discussed their opinions about village
contingency plans (CEE Himalaya, 2009). Some of the observations on
impacts of climate change in the area discussed by participants included
the melting, shrinking, and even disappearance of some glaciers and
the drying up of several wetlands and perennial springs. Heavy
deforestation, decline and extinction of wildlife, heavy soil erosion,
siltation of water bodies, fall in crop yields, and reduced availability of
fodder and other non- timber forest produce were some of the other
related issues discussed (CEE Himalaya, 2009). Participants watched
documentaries about climate change and played the Urdu version of
‘Riskland: Let’s Learn to Prevent Disasters.’ They received educational
kits on disaster risk reduction and on climate change, translated and
adapted for Kashmir (CEE Himalaya, 2009).
9.2.14.4. Lessons Identified
The main lesson that can be drawn from the various initiatives
described above is that effective DRR education does not occur in a silo.
As the examples from Japan, Brazil, and the Himalayas illustrate,
successful programs actively engage participants and their wider
communities to elicit risk-reducing behavioral change (Shaw et al., 2004;
Wisner, 2006; Bonifacio et al., 2010). Lessons on actively engaging
participants include:
Assessing community risk, discussing risk with others, and joining
a risk-reducing activity in school or community forums provide
opportunities for active learning. Engaging children and community
members in vulnerability and capacity assessments has been found
to be effective in disaster risk reduction and adaptation programs
(Twigg and Bottomley, 2011; see Himalaya example).
Interactive lectures with visual aids can be effective in building
knowledge (Shaw et al., 2004; see teacher training in Indonesia
example) and should be followed up with discussion with peers
and family – and action – beyond the classroom (Shaw et al., 2004;
Wisner, 2006).
Additional lessons of good practice illustrated above include:
Integrating climate change information into DRR education and
integrating both into various subject matters is simple and effective.
Chapter 9Case Studies
529
The Philippines example shows that such integration is underway,
and the teacher training in Indonesia example concludes that such
integration can be helpful in avoiding overburdening full curricula.
Training of teachers and professionals in all relevant sectors can have
a positive multiplier effect. As the Nepalese teachers’ evaluation
example shows, teacher training is critical to address risk self-
perception and ensure that teachers pass on appropriate DRR
knowledge. The training of builders example in Nepal, India, and
the Philippines illustrates the successful dissemination of DRR
methods and tools within a critical sector across borders.
As well as providing further examples of current adaptation and DRR
initiatives, a United Nations Framework Convention on Climate Change
synthesis report of initiatives undertaken by Nairobi Work Programme
partners concludes that the integration of activities relating to education,
training, and awareness-raising into relevant ongoing processes and
practices is key to the long-term success of such activities (UNFCCC, 2010).
9.3. Synthesis of Lessons
Identified from Case Studies
This chapter examined case studies of extreme climate events, vulnerable
regions, and methodological management approaches in order to glean
lessons and good practices. Case studies are provided to add context
and value to this report. They contribute to a focused analysis and convey,
in part, the reality of an event: the description of how certain extreme
events develop; the extent of human loss and financial damage; the
response strategies and interventions; the DRR, DRM, and CCA measures
and their effect on the overall outcomes; and cultural or region-specific
factors that may influence the outcome. Most importantly, case studies
provide a medium through which to learn practical lessons about
successes in DRR that are applicable for adaptation to climate change.
The lessons identified will prove useful at various levels from the
individual to national and international organizations as people try to
respond to extreme events and disasters and adapt to climate change.
The case studies highlight several recurring themes and lessons.
A common factor was the need for greater amounts of useful information
on risks before the events occur, including early warnings. The
implementation of early warning systems does reduce loss of lives and
to a lesser extent damage to property. Early warning was identified by
all the extreme event case studies – heat waves, wildfires, drought,
dzud, cyclones, floods, and epidemics – as key to reducing the impacts
from extreme events. A need for improving international cooperation and
investments in forecasting was recognized in some of the case studies,
but equally the need for regional and local early warning systems was
heavily emphasized, particularly in developing countries.
A further common factor identified overall was that it is better to invest
in preventive-based DRR plans, strategies, and tools for adaptation than
in response to extreme events. Greater investments in proactive hazard
and vulnerability reduction measures, as well as development of capacities
to respond and recover from the events were demonstrated to have
benefits. Specific examples for planning for extreme events included
increased emphasis on drought preparedness; planning for urban heat
waves; and tropical cyclone DRM strategies and plans in coastal regions
that anticipate these events. However, as illustrated by the SIDS case
study, it was also identified that DRR planning approaches continue to
receive less emphasis than disaster relief and recovery.
It was also identified that DRM and preventive public health are closely
linked and largely synonymous. Strengthening and integrating these
measures, along with economic development, should increase resilience
against the health effects of extreme weather and facilitate adaptation
to climate change. Extreme weather events and population vulnerability
can interact to produce disastrous epidemic disease through direct
effects on thetransmission cycle andalsopotentially through indirect
effects, such as population displacement.
Another lesson is that in order to implement a successful DRR or CCA
strategy, legal and regulatory frameworks are beneficial in ensuring
direction, coordination, and effective use of funds. The case studies are
helpful in this endeavor as effective and implemented legislation can
create a framework for governance of disaster risks. While this type of
approach is mainly for national governments and the ways in which
they devolve responsibilities to local administrations, there is an
important message for international governance and institutions as well.
Frameworks that facilitate cooperation with other countries to attain
better analysis of the risks will allow institutions to modify their focus
with changing risks and therefore maintain their effectiveness. This
cooperation could be at the local through national to international
levels. Here and in other ways, civil society has an important role.
Insurance and other forms of risk transfer can be linked to disaster risk
reduction and climate change adaptation by providing knowledge and
incentives for reducing risk, reducing vulnerability, and enabling recovery.
A lesson identified by many case studies was that effective DRR education
contributes to reduce risks and losses, and is most effective when it is
not done in isolation, but concurs with other policies. The integration of
activities relating to education, training, and awareness-raising into
relevant ongoing processes and practices is important for the long-term
success of DRR and DRM activities. Investing in knowledge at primary
to higher education levels produces significant DRR and DRM benefits.
Research improves our knowledge, especially when it includes integration
of natural, social, health, and engineering sciences and their applications.
In all cases, the point was made that with greater information available
it would be possible to better understand the risks and to ensure that
response strategies were adequate to face the risks. It further poses a
set of questions to guide the investigations.
The case studies have reviewed past events and identified lessons that
could be considered for the future. Preparedness through DDR and DRM
Chapter 9 Case Studies
530
can help to adapt for climate change and these case studies offer
examples of measures that could be taken to reduce the damage that is
inflicted as a result of extreme events. Investment in increasing
knowledge and warning systems, adaptation techniques and tools, and
preventive measures will cost money now but they will save money and
lives in the future.
Asante, K.O., J.S. Famiglietti, and J.P. Verdin, 2005: Current and future applications of
remote sensing for routine monitoring of surface water controllers. In:
Proceedings Pecora 16 Conf., Sioux Falls, SD, 23-27 October 2005, pp. 1-9.
Asante, K.O., R.D. Macuacua, G.A. Artan, R.W. Lietzow, and J.P. Verdin, 2007:
Developing a flood monitoring system from remotely sensed data for the
Limpopo Basin. IEEE Transactions on Geoscience and Remote Sensing, 45(6),
1709-1714.
Assemblée Nationale, 2004: Rapport de la Commission D’Enquête sur les
Conséquences Sanitaires et Sociales de la Canicule. Assemblée Nationale,
France, vols. 1 and 2.
Auld, H.E., 2008: Adaptation by design: the impact of changing climate on
infrastructure. Journal of Public Works and Infrastructure, 3, 276-288.
Auld, H., J. Waller, S. Eng, J. Klaassen, R. Morris, S. Fernandez, V. Cheng, and D.
MacIver, 2010: The Changing Climate and National Building Codes and
Standards. Environment Canada, Ottawa, Canada, ams.confex.com/ams/
pdfpapers/174517.pdf.
Australian Government, 2009: Metrological aspects of the 7 February 2009
Victorian fires, an overview. Bureau of Meteorology Report for the 2009
Victorian Bushfires Royal Commission. www.royalcommission.vic.gov.au/
getdoc/f1eaba2f-414f-4b24-bcda-d81ff680a061/WIT.013.001.0012.pdf.
Awuor, C.B., V.A. Orindi, and A. Adwerah, 2008: Climate change and coastal cities:
The case of Mombasa, Kenya. Environment and Urbanization, 20(1), 231-242.
Back, E., C. Cameron, and T. Tanner, 2009: Children and Disaster Risk Reduction:
Taking stock and moving forward. Children in a Changing Climate, Institute of
Development Studies and UNICEF, Brighton, UK, 42 pp.
Badia-Perpinyà, A. and M. Pallares-Barbera, 2006: Spatial distribution of ignitions
in Mediterranean periurban and rural areas: the case of Catalonia. International
Journal of Wildland Fire, 15, 187-196.
Baigalmaa, P., 2010: Community base disaster risk management, Self Sharpen – 2,
guideline. In: Strengthening disaster management system in Mongolia,
MON/08/305 project [Boldbaatar, S. (ed.)]. Phase III, United Nations Development
Programme, New York, NY, 137 pp., www.un-mongolia.mn/publication/
UNDP_gamshig_04-14.pdf.
Barber, D., J. Lukovich, J. Keogak, S. Baryluk, L. Fortier, and G. Henry, 2008: Changing
climate of the arctic. Arctic, 61(1), 7-26.
Barnett, B.J., C.B. Barrett, and J.R. Skees, 2008: Poverty traps and index-based risk
transfer products. World Development, 36(10), 1766-1785.
Barnett, J. and W.N. Adger, 2003: Climate dangers and atoll countries. Climatic
Change, 61, 321-337.
Barnett, J. and W.N. Adger, 2007: Climate change, human security and violent
conflict. Political Geography, 26(6), 639-655.
Bartzokas, A., J. Azzopardi, L. Bertotti, A. Buzzi, L. Cavaleri, D. Conte, S. Davolio, S.
Dietrich, A. Drago, O. Drofa, A. Gkikas, V. Kotroni, K. Lagouvardos, C.J. Lolis, S.
Michaelides, M. Miglietta, A. Mugnai, S. Music, K. Nikolaides, F. Porcù, K.
Savvidou, and M. I. Tsirogianni, 2010: The RISKMED project: philosophy,
methods and products. Natural Hazards and Earth System Sciences, 10,
1393-1401.
Basher, R., 2006: Global early warning systems for natural hazards: Systematic and
people-centered. Philosophical Transactions of the Royal Society A, 364(1845),
2167-2182.
Bastedo, J., 2007: On the Frontlines of Climate Change: What’s Really Happening in
the Northwest Territories. Government of the Northwest Territories, Yellowknife,
Canada, sen.parl.gc.ca/nsibbeston/Final%20for%20WEB%20or%20Email.pdf.
Batima, P. and D. Dagvadorj (eds.), 2000: Climate change and its Impacts in
Mongolia. Mongolian National Agency for Meteorology, Hydrology and
Environmental Monitoring (NAMHEM). JEMR, Ulaanbaatar, Mongolia, 227 pp.,
ISBN 99929-70-35-9.
Batjargal, Z., R. Oyun, S. Sangidansranjav, and N. Togtokh, 2001: Lessons Learned
from Dzud 1999-2000 [Batjargal, Z., S. Sangidansranjav, R. Oyun, and N.Togtokh
(eds.)]. Case study funded by UNDP, conducted by joint team of National
Agency of Meteorology, Hydrology and Environmental Monitoring, Civil
Defense Agency, Ministry of Agriculture and JEMR, Ulaanbaatar, Mongolia, 347
pp., ISBN 99929-70-54-7.
Chapter 9Case Studies
References
A digital library of non-journal-based literature cited in this chapter that
may not be readily available to the public has been compiled as part of
the IPCC review and drafting process, and can be accessed via either the
IPCC Secretariat or IPCC Working Group II web sites.
Abou Hadid, A.F., 2009: Food Production. In: Arab Environment: Climate Change.
Impact of Climate Change on Arab Countries [Tolba, M.K. and N.W. Saab (eds.)].
Arab Forum for Environment and Development, Beirut, Lebanon, pp. 63-74,
www.afedonline.org/afedreport09/Full%20English%20Report.pdf.
Abrahamson, V., J. Wolf, B. Fenn, S. Kovats, P. Wilkinson, W.N. Adger, and R. Raine,
2008: Perceptions of heatwave risks to health: interview-based study of older
people in London and Norwich, UK. Journal of Public Health, 31(1), 119-126.
Ackers, M.-L., R.E. Quick, C.J. Drasbek, L. Hutwagner, and R.V. Tauxe, 1998: Are
there national risk factors for epidemic cholera? The correlation between
socioeconomic and demographic indices and cholera incidence in Latin America.
International Journal of Epidemiology, 27(2), 330-334.
Adelekan, I.O., 2010: Vulnerability of poor urban coastal communities to flooding in
Lagos, Nigeria. Environment and Urbanization, 22(2), 433-450.
Ahern, M, R.S. Kovats, P. Wilkinson, R. Few, and F. Matthies, 2005: Global health
impacts of floods: Epidemiologic evidence. Epidemiologic Reviews, 27, 36-46.
Ahmad, Q.K. and A.U. Ahmed (eds.), 2002: Bangladesh: citizen’s perspective on
sustainable development. Bangladesh Unnayan Parishad, Dhaka, Bangladesh,
181 pp.
Ahrens, J., and P.M. Rudolph, 2006: The importance of governance in risk reduction
and disaster management. Journal of Contingencies and Crisis Management,
14(4), 207-220.
AIACC, 2006: Climate Change Vulnerability and Adaptation in the Livestock Sector of
Mongolia. A Final Report Submitted by P. Batima to Assessments of Impacts and
Adaptations to Climate Change (AIACC), Project No. AS 06, International START
Secretariat, Washington, DC, 84 pp.
Akbari, H. and S. Konopacki, 2004: Energy effects of heat-island reduction strategies
in Toronto, Canada. Energy, 29(2), 191-210.
Akbari, H., M. Pomerantza, and H. Tahaa, 2001: Cool surfaces and shade trees to reduce
energy use and improve air quality in urban areas. Solar Energy, 70(3), 295-310.
Alajo, S.O., J. Nakavuma, and J. Erume, 2006: Cholera in endemic districts in Uganda
during El Nino rains: 2002-2003. African Health Sciences, 6(2), 93-97.
Allard, M., R. Fortier, C. Duguay, and N. Barrette 2002: A trend of fast climate warming
in Northern Quebec since 1993: impacts on permafrost and man-made
infrastructures. Eos Transactions, American Geophysical Union, 83(47), F258.
Allen, C.D., M. Savage, D. Falk, K.F. Suckling, T.W. Swetnam, T. Schulke, P.B. Stacey, P.
Morgan, M. Hoffman, and J. Klingel, 2002: Ecological restoration of southwestern
ponderosa pine ecosystems: a broad perspective. Ecological Applications, 12,
1418-1433.
Anisimov, O.A., D.G. Vaughan, T.V. Callaghan, C. Furgal, H. Marchant, T.D. Prowse, H.
Vilhjálmsson, and J.E. Walsh. 2007: Polar regions (Arctic and Antarctic). In: Climate
Change 2007. Impacts, Adaptation and Vulnerability. Contribution of Working
Group II to the Fourth Assessment Report of the Intergovernmental Panel on
Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der Linde, and
C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK, pp. 653-685.
531
Beg, N., J.C. Morlot, O. Davidson, Y. Afrane-Okesse, L. Tyani, F. Denton, Y.Sokona,
J.P. Thomas, E.L. La Rovere, J.K. Parikh, K. Parikh, and A.A. Rahman, 2002:
Linkages between climate change and sustainable development. Climate
Policy, 2, 129-144.
Begzsuren, S., J.E. Ellis, D.S. Ojima, M.B. Coughenour, T. Chuluun, 2004: Livestock
responses to droughts and severe winter weather in the Gobi Three Beauty
National Park, Mongolia. Journal of Arid Environments, 59(4), 785-796.
Benson, C., 2009: Mainstreaming Disaster Risk Reduction into Development:
Challenges and Experience in the Philippines. International Federation of Red
Cross and Red Crescent Societies and Provention Consortium, Geneva,
Switzerland, 60 pp.
Bernard, S.M. and M.A. McGeehin, 2004: Municipal heat wave response plans.
American Journal of Public Health, 94(9), 1520-1522.
Bhatt, M., T. Reynolds, and M. Pandya, 2010: Disaster insurance for the poor: All India
Disaster Mitigation Institute and Afat Vimo. In: Microinsurance – An innovative
tool for risk and disaster management [Morelli, E., G.A. Onnis, W.J. Ammann,
and C. Sutter (eds.)]. Global Risk Forum, Davos, Switzerland, pp. 341-353.
Bhattacharya, S., R. Black, L. Bourgeois, J. Clemens, A. Cravioto, J.L. Deen, G.
Dougan, R. Glass, R.F. Grais, M. Greco, I. Gust, J. Holmgren, S. Kariuki, P.-H.
Lambert, M.A. Liu, I. Longini, G.B. Nair, R. Norrby, G.J.V. Nossal, P. Ogra, P.
Sansonetti, L. von Seidlein, F. Songane, A.-M. Svennerholm, D. Steele, and R.
Walker, 2009: The cholera crisis in Africa, Science, 324(5929), 885.
Black, R.A., J.P. Bruce, and I.D.M. Egener, 2010: Managing the Risks of Climate
Change: A Guide for Arctic and Northern Communities. Centre for Indigenous
Environmental Resources, Winnipeg, Canada.
Bollin, C., N. Lamade, and J. Ferguson, 2005: Disaster Risk Management along the
Rio Búzi. Case Study on the Background, Concept and Implementation of
Disaster Risk, Management in the Context of the GTZ-Programme for Rural
Development (PRODER), German Gesellschaft für Technische Zusammenarbeit,
Governance and Democracy Division, Eschborn, Germany, 26 pp.
Bompangue, D., P. Giraudoux, M. Piarroux, G. Mutombo, R. Shamavu, B. Sudre, A.
Mutombo, V. Mondonge, and R. Piarroux, 2009: Cholera epidemics, war and
disasters around Goma and Lake Kivu: an eight-year survey. PLoS Neglected
Tropical Diseases, 3(5), e436.
Bonifacio, A.C., Y. Takeuchi, and R. Shaw, 2010: Mainstreaming climate change
adaptation and disaster risk reduction through school education: perspectives
and challenges. In: Climate Change Adaptation and Disaster Risk Reduction:
Issues and Challenges [Shaw, R., J. Pulhin, and J. Pereira (eds.)]. Emerald Group
Publishing, Bingley, UK, pp.143-169.
Borgford-Parnell, N., 2009: Mongolia: a case for economic diversification in the
face of a changing climate. Sustainable Development Law & Policy, 9(2), 54-56.
Boruff, B. and S. Cutter, 2007: The environmental vulnerability of Caribbean island
nations. The Geographical Review, 97(1), 24-45.
Botha, D., D. Van Niekerk, G. Wentink, C. Coetzee, K. Forbes, Y. Maartens, E.
Annandale, T. Tshona, and E. Raju, 2011: Disaster Risk Management Status
Assessment at Municipalities in South Africa. South African Local Government
Association, Pretoria, South Africa, 108 pp.
Botzen, W.J.W., J.C.J.H. Aerts, and J.C.J.M. van den Bergh, 2009: Willingness of
homeowners to mitigate climate risk through insurance. Ecological Economics,
68(8-9), 2265-2277, doi:10.1016/j.ecolecon.2009.02.019.
Bouchama, A., 2004: The 2003 European heat wave. Intensive Care Medicine, 30(1),
1-3.
Bradstock, R.A., J.S. Cohn, A.M. Gill, M. Bedward, and C. Lucas, 2009: Prediction and
probability of large fires in the Sydney Region of South-eastern Australia using
fire weather. International Journal of Wildland Fire, 18, 932-943.
Braman,L., P. Suarez, and M.K. van Aalst, 2010: Climate change adaptation:
integrating climate science into humanitarian work. International Review of the
Red Cross, 92(879).
Britton, N.R., 2006: Getting the foundation right: in pursuit of effective disaster
legislation for the Philippines. In: Proceedings of the Second Asian Conference
on Earthquake Engineering, Manila, Philippines, March 2006, 15 pp.
Broad, K. and S. Agrawala, 2000: The Ethiopia food crisis: uses and limits of climate
forecasts. Science, 289(5485), 1693-1694.
Brunet, G., M. Shapiro, B. Hoskins, M. Moncrieff, R. Dole, G.N. Kiladis, B. Kirtman, A.
Lorenc, B. Mills, R. Morss, S. Polavarapu, D. Rogers, J. Schaake, and J. Shukla,
2010: Collaboration of the weather and climate communities to advance
subseasonal-to-seasonal prediction. Bulletin of the American Meteorological
Society, 91, 1397-1406.
Burnside, R., D.S. Miller, and J.D. Rivera, 2007: The impact of information and risk
perception on the hurricane evacuation decision-making of greater New
Orleans residents. Sociological Spectrum, 27(6), 727-740.
Burton, I, 2011: Forensic disaster investigations in depth: A new case study model.
Environment Magazine, 52(5), 36-41.
Bushfire CRC, 2009: Household Mail Survey. Bushfire Co-operative Research Centre
report, Bushfire CRC, East Melbourne, Australia, www.bushfirecrc.com/managed/
resource/mail-survey-report-10-1-10-rt-2.pdf.
Cadot, E., V.G. Rodwin, and A. Spiral, 2007: In the heat of the summer: lessons from
the heat waves in Paris. Journal of Urban Health, 84(4), 466-468.
Calain, P., J.P. Chaine, E. Johnson, M.L. Hawley, M.J. O’Leary, H. Oshitani, and C.L.
Chaignat, 2004: Can oral cholera vaccination play a role in controlling a cholera
outbreak? Vaccine, 22(19), 2444-2451.
Cardenas, V., S. Hochrainer, R. Mechler, G. Pflug, and J. Linnerooth-Bayer, 2007:
Sovereign Financial Disaster Risk Management: The Case of Mexico.
Environmental Hazards, 7, 40-53.
Cardona, O.D. and L.E. Yamín, 2007: Información para la gestión de riesgo de desastres.
Estudio de caso de cinco países: Colombia. United Nations, Inter-American
Development Bank, and Economic Commission for Latin America and the
Caribbean (ECLAC), 2007, Mexico City, Mexico, 25 pp.
Carmo Vaz, Á., 2000: Coping with floods – The experience of Mozambique. In:
1st WARFSA/WaterNet Symposium: Sustainable Use of Water Resources,
Maputo, Mozambique, 1-2 November 2000, www.bvsde.paho.org/bvsacd/cd46/
coping.pdf.
CCRIF, 2010: Annual Report 2009-2010. Caribbean Catastrophe Risk Insurance
Facility, Grand Cayman, Cayman Islands.
CEE Himalaya, 2009: Disaster risk reduction in the mountains. Ceenario
, 24,
December 16-31, Centre for Environment Education, Ahmedabad, India, 4 pp.
Chaignat, C.L., V. Monti, J. Soepardi, G. Petersen, E. Sorensen, J. Narain, and M.P.
Kieny, 2008: Cholera in disasters: do vaccines prompt new hopes? Expert
Review of Vaccines, 7(4), 431-435.
Chatterjee, M., 2010: Slum dwellers response to flooding events in the megacities
of India. Mitigation and Adaptation Strategies for Global Change, 15, 337-353.
Choularton, R., 2007: Contingency planning and humanitarian action: a review of
practice. Humanitarian Practice Network, 59.
Christensen, J.H., B. Hewitson, A. Busuioc, A. Chen, X. Gao, I. Held, R. Jones, R.K.
Kolli, W.-T. Kwon, R. Laprise, V. Magaña Rueda, L. Mearns, C.G. Menéndez, J.
Räisänen, A. Rinke, A. Sarr, and P. Whetton, 2007: Regional climate projections.
In: Climate Change 2007: The Physical Science Basis. Contribution of Working
Group I to the Fourth Assessment Report of the Intergovernmental Panel on
Climate Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B.
Averyt, M. Tignor and H.L. Miller (eds.)]. Cambridge University Press,
Cambridge, UK, and New York, NY, pp. 847-940.
Clarke, J., 1972: Some effects of the urban structure on heat mortality,
Environmental Research, 5, 93-104.
Cohen, M. and J. Sebstad, 2003: Reducing Vulnerability: the Demand for
Microinsurance. MicroSave-Africa, Nairobi, Kenya.
Colwell, R., 1996: Global climate and infectious disease: the cholera paradigm.
Science, 274, 2025-2031.
Colwell, R.R., 2002: A voyage of discovery: cholera, climate and complexity.
Environmental Microbiology, 4(2), 67-69.
Constantin de Magny, G., J.F. Guegan, M. Petit, and B. Cazelles, 2007: Regional-
scale climate-variability synchrony of cholera epidemics in West Africa. BMC
Infectious Diseases, 7, 20.
Couture, N.J. and W.H. Pollard, 2007: Modelling geomorphic response to climatic
change. Climatic Change, 85(3-4), 407-431.
Crabbé, P. and M. Robin, 2006: Institutional adaptation of water resource infrastructures
to climate change in Eastern Ontario. Climatic Change, 78, 103-133.
Chapter 9 Case Studies
532
CRED, 2009: EM-DAT, Emergency Events Database. Centre for Research on the
Epidemiology of Disasters (CRED), Université Catholique de Louvain, Brussels,
Belgium, www.emdat.be/.
Cretikos, M., K. Eastwood, C. Dalton, T. Merritt, F. Tuyl, L. Winn, and D. Durrheim,
2008: Household disaster preparedness and information sources: rapid
cluster survey after a storm in New South Wales, Australia. BMC Public Health,
8, 195.
CSA, 2010: Technical Guide – Infrastructure in Permafrost: A Guideline for Climate
Change Adaptation. Special Publication PLUS 4011-10, Canadian Standards
Association, Mississauga, Canada.
Cummins, D. and O. Mahul, 2009: Catastrophe Risk Financing in Developing
Countries: Principles for Public Intervention. World Bank, Washington, DC.
Da Silva, J., B. Garanganga, V. Teveredzi, S.M. Marx, S.J. Mason, and S.J. Connor, 2004:
Improving epidemic malaria planning, preparedness and response in Southern
Africa. Malaria Journal, 3, 37.
Dagvadorj, D., L., Natsagdorj, J. Dorjpurev, and B. Namkhainyam, 2010: Mongolia
Assessment Report on Climate Change 2009. Second publication with corrections,
Ministry of Nature, Environment and Tourism, Ulaanbaatar, Mongolia, 226 pp.
Daoudy, M., 2009: Asymmetric power : Negotiating water in the Euphrates and
Tigris. International Negotiation, 14, 361-391.
Davies, M., B. Guenther, J. Leavy, T. Mitchell, and T. Tanner, 2008:Adaptive social
protection’: synergies for poverty reduction. IDS Bulletin, 39(4), 105-112.
Davis, C., W. Wang, S.S. Chen, Y. Chen, K. Corbosiero, M. DeMaria, J. Dudhia, G.
Holland, J. Klemp, J. Michalakes, H. Reeves, R. Rotunno, C. Snyder, and Q. Xiao,
2008: Prediction of land falling hurricanes with the advanced hurricane WRF
model. Monthly Weather Review, 126, 1990-2005.
De Comarmond, A. and R. Payet, 2010: Small Island Developing States: incubators
of innovative adaptation and sustainable technologies? In: Coastal Zones and
Climate Change [Michel, D. and A. Pandya (eds.)]. The Henry L. Stimson Center,
Washington, DC, www.stimson.org/images/uploads/research-pdfs/Alain.pdf.
De Sherbinin, A., A. Schiller, and A. Pulsipher, 2007: The vulnerability of global cities
to climate hazards. Environment and Urbanization, 19(1), 39-64.
Dear, K., G. Ranmuthugala, T. Kjellström, C. Skinner, and I. Hanigan, 2005: Effects of
temperature and ozone on daily mortality during the August 2003 heat wave in
France. Archives of Environmental and Occupational Health, 60, 205-212.
Derwent, R.G., P.G. Simmonds, S. O’Doherty, D.S. Stevenson, W.J. Collins, M.G.
Sanderson, C.E. Johnson, F. Dentener, J. Cofala, R. Mechler and M. Amann, 2006:
External influences on Europe’s air quality: baseline methane, carbon monoxide
and ozone from 1990 to 2030 at Mace Head, Ireland. Atmospheric
Environment, 40, 844-855.
Doblas-Reyes, F.J., R. Hagedorn, T.N. Palmer, and J.J. Morcrette, 2006: Impact of
increasing greenhouse gas concentrations in seasonal ensemble forecasts.
Geophysical Research Letters, 33, L07708.
Doswell III, C.A., S.J. Weiss, and R.H. Johns, 1993: Tornado forecasting: a review. In:
The Tornado: Its Structure, Dynamics, Prediction, and Hazards [Church, C., D.
Burgess, C. Doswell, and R. Davies-Jones (eds.)]. Geophysical Monograph 79,
American Geophysical Union, Washington, DC, pp. 557-571.
Dow, K. and S.L. Cutter, 1998: Crying wolf: repeat responses to hurricane evacuation
orders. Coastal Management, 26(4), 237-252.
DSE, 2008a: Climate Change in Port Phillip and Westernport. Department of
Sustainability and Environment, East Melbourne, Victoria, Australia.
DSE, 2008b: Victoria’s State of the Forests Report 2008. Victorian Government,
Department of Sustainability and Environment, Melbourne, Australia.
Dyson, L.L. and J. van Heerden, 2001: The heavy rainfall and floods over the
northeastern Interior of South Africa during February 2000. South African
Journal of Science, 97, 80-86.
Easterling, W.E., P.K. Aggarwal, P. Batima, K.M. Brander, L. Erda, S.M. Howden, A.
Kirilenko, J. Morton, J.-F. Soussana, J. Schmidhuber, and F.N. Tubiello, 2007: Food,
fibre and forest products. In: Climate Change 2007. Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani,
J.P. Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 273-313.
Ebi, K.L. and G.A. Meehl, 2007: The Heat Is On: Climate Change And Heatwaves in
the Midwest. Pew Center on Global Climate Change, Arlington, VA, 14 pp.
Ebi, K.L., T.J. Teisberg, L.S. Kalkstein, L. Robinson, and R.F. Weiher, 2004: Heat watch/
warning systems save lives: estimated costs and benefits for Philadelphia 1995-
1998. Bulletin of the American Meteorological Society, 85, 1067-1073.
Ebi, K., N. Lewis, and C. Corvalan, 2006: Climate variability and change and their
potential health effects in small island states: Information for adaptation
planning in the health sector. Environmental Health Perspectives, 114(12),
1957-1963.
ECLAC, 2006: Características e impacto socioeconómico de los huracanes “Stan” y
“Wilma” en la República. Economic Commission for Latin America and the
Caribbean, Santiago, Chile, biblioteca.cepal.org/record=b1096781~S0.
EEA, 2003: Air Pollution by Ozone in Europe in Summer 2003. Topic report 3/2003,
European Environment Agency, Copenhagen, Denmark.
EEA, 2007: Air Pollution by Ozone in Europe in Summer 2006. Technical report
5/2007, European Environment Agency, Copenhagen, Denmark.
EEA, 2011: Air Pollution by Ozone across Europe during Summer 2010. Technical
report 6/2011, European Environment Agency, Copenhagen, Denmark.
Einstein, H.H. and R. Sousa, 2007: Warning systems for natural threats. Georisk:
Assessment and Management of Risk for Engineered Systems and Geohazards,
1(1), 3 -20.
El-Quosy, D., 2009: Fresh Water. In: Arab Environment: Climate Change. Impact of
Climate Change on Arab Countries [Tolba, M.K. and N.W. Saab (eds.)]. Arab
Forum for Environment and Development, Beirut, Lebanon, pp. 75-86,
www.afedonline.org/afedreport09/Full%20English%20Report.pdf.
Emch, M., C. Feldacker, M. Yunus, P.K. Streatfield, V. DinhThiem, D.G. Canh, and M. Ali,
2008: Local environmental predictors of cholera in Bangladesh and Vietnam.
American Journal of Tropical Medicine & Hygiene, 78(5), 823-832.
Environment Canada, 2010: Climate Information to Inform New Codes and
Standards. Environment Canada, Ottawa, Canada, www.ec.gc.ca/sc-cs/
default.asp?lang=En&n=20CD1ADB-1.
Erian
, W., B. Katlan, and O. Babah, 2011: Drought Vulnerability in the Arab Region,
Case Study – Drought in Syria. Arab Center for the Studies of Arid Zones and
Dry Lands and United Nations International Strategy for Disaster Reduction,
Geneva, Switzerland, www.preventionweb.net/english/hyogo/gar/2011/en/
bgdocs/Erian_Katlan_&_Babah_2010.pdf.
FAO, 2008: FAO’s role in the 2008 Syria drought appeal. Food and Agriculture
Organization, Rome, Italy, www.faoiraq.org/images/word/The%20FAO
%20proposed%20%20program%20under%20this%20Syria%20l.pdf.
FAO, 2009: FAO’s role in the Syria Drought Response Plan 2009. Food And Agriculture
Organization, Rome, Italy, www.fao.org/fileadmin/templates/tc/tce/pdf/
app_syriadrought2009.pdf.
FAO and NAPC, 2010: National Programme for Food Security in the Syrian Arab
Republic. Food and Agriculture Organization and National Agriculture Policy
Centre, Rome Italy and Damascus, Syria, www.napcsyr.org/events/ws/
f_s_program/NPFS_final_en.pdf.
FAO and WFP, 2000: FAO/WFP Crop and Food Supply Assessment Mission to
Mozambique. Food and Agriculture Organization and World Food Programme
Special Report, FAO, Rome, Italy.
Field, C.B., L.D. Mortsch, M. Brklacich, D.L. Forbes, P. Kovacs, J.A. Patz, S.W. Running,
and M.J. Scott, 2007: North America. In: Climate Change 2007. Impacts,
Adaptation and Vulnerability. Contribution of Working Group II to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change [Parry,
M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)].
Cambridge University Press, Cambridge, UK, pp. 617-652.
Filleul, L., S. Cassadou, S. Médina, P. Fabres, A. Lefranc, D. Eilstein, A. Le Tertre, L.
Pascal, B. Chardon, M. Blanchard, C. Declercq, J.F. Jusot, H. Prouvost, and M.
Ledrans, 2006: The relation between temperature, ozone, and mortality in nine
French cities during the heat wave of 2003. Environmental Health Perspectives,
114, 1344-1347.
Flannigan, M.D., M.A. Krawchuk, W.J. de Groot, B.M. Wotton, and L.M. Gowman,
2009: Implications of changing climate for global wildland fire. International
Journal of Wildland Fire, 18, 483-507.
Chapter 9Case Studies
533
Flyvbjerg, B., 2004: Five misunderstandings about case-study research. In:
Qualitative Research Practice [Seale, C., G. Gobo, J.F. Gubrium, D. Silverman
(eds.)]. Sage Publications, London, UK, pp. 420-434.
Foley, C., 2007: Mozambique: A Case Study In the Role of the Affected State in
Humanitarian Action. HPG Working Paper, Humanitarian Policy Group, Overseas
Development Institute, London, UK, www.odi.org.uk/resources/download/
2557.pdf.
Ford, J.D. and C. Furgal, 2009: Foreword to the special issue: climate change adaptation
and vulnerability in the arctic, Polar Research, 28, 1-9.
Ford, J.D. and T. Pearce, 2010: What we know, do not know, and need to know about
climate change vulnerability in the western Canadian Arctic: A systematic
literature review. Environmental Research Letters, 5, 014008.
Ford, J. and B. Smit, 2004: A framework for assessing the vulnerability of communities
in the Canadian Arctic to risks associated with climate change. Arctic, 57(4),
389-400.
Ford, J.D., T. Pearce, F. Duerden, C. Furgal, and B. Smit, 2010: Climate change policy
responses for Canada’s Inuit population: the importance of and opportunities
for adaptation. Global Environmental Change, 20,177-191.
FORIN, 2011: Forensic Disaster Investigations – The FORIN Project. Integrated
Research on Disaster Risk, Beijing, China, www.irdrinternational.org/about-
irdr/scientific-committee/working-group/forensic-investigations/.
Fortier, R., A-.M. LeBlanc, and W. Yu, 2011: Impacts of permafrost degradation on a
road embankment at Umiujaq in Nunavik (Quebec), Canada. Canadian
Geotechnical Journal, 48, 720-740, doi:10.1139/T10-101.
Fouillet, A., G. Rey, F. Laurent, G. Pavillon, S. Bellec, C. Guihenneuc-Jouyaux, J. Clavel,
E. Jougla, and D. Hemon, 2006: Excess mortality related to the August 2003
heat wave in France. International Archives of Occupational & Environmental
Health, 80(1), 16-24.
Fouillet, A.R., V. Wagner, K. Laaidi, P. Empereur-Bissonnet, A. Le Tertre, P. Frayssinet,
P. Bessemoulin, F. Laurent, P. De Crouy-Chanel, E. Jougla, and D. Hemon, 2008:
Has the impact of heat waves on mortality changed in France since the
European heat wave of summer 2003? A study of the 2006 heat wave.
International Journal of Epidemiology, 37(2), 309-317.
Founda, D. and C. Giannakopoulos, 2009: The exceptionally hot summer of 2007 in
Athens, Greece – A typical summer in the future climate? Global and Planetary
Change, 67(3-4), 227-236.
Frew, S.L., 2002: Public Awareness and Social Marketing. Proceedings of a workshop,
Asian Disaster Preparedness Center Regional Workshop on Best Practices in
Disaster Management, Bangkok, Thailand, pp. 381-393.
Fritz, H.M., C.D. Blount, S. Thwin, M.K. Thu, and N. Chan, 2009: Cyclone Nargis storm
surge in Myanmar. Nature Geoscience, 2, 448-449.
Fuchs, R., M. Conran, and E. Louis, 2011: Climate change and Asia’s coastal urban
cities. Environment and Urbanization Asia, 2(1), 13-28.
Furgal, C. and T. Prowse, 2008: Northern Canada. In: From Impacts to Adaptation:
Canada in a Changing Climate 2007 [Lemmen, D., F. Warren, E. Bush, and J.
Lacroix (eds.)]. Natural Resources Canada, Ottawa, Canada, pp. 57-118.
Gagnon-Lebrun, F. and S. Agrawala, 2006: Progress on Adaptation to Climate
Change in Developed Countries: An Analysis of Broad Trends. OECD, Paris,
France, 11 pp.
Gall, M., K.A. Borden, and S.L. Cutter, 2009: When do losses count? Six fallacies of
natural hazards loss data. Bulletin of the American Meteorological Society, 90,
799-809.
García, A., R.M.C. Norland, and K. Méndez Estrada, 2006: Características e impacto
socioeconómico de los huracanes “Stan” y “Wilma” en la República Mexicana
en 2005. Secretaria de Gobernación, Centro Nacional de Prevención de
Desastres, Mexico.
GEF, 2006: Sustainable Land Management in Least Developed Countries and Small
Island Developing States. Global Environment Facility, Washington, DC,
www.cndwebzine.hcp.ma/cnd_sii/IMG/pdf/SLM_Support_in_LDCs_SIDS.pdf.
Ghesquiere, F., O. Mahul, M. Forni, and R. Gartley, 2006: Caribbean Catastrophe Risk
Insurance Facility: A Solution to the Short-Term Liquidity Needs of Small Island
States in the Aftermath of Natural Disasters. IAT03-13/3, World Bank,
Washington, DC.
Giannakopoulos, C., P. Le Sager, M. Bindi, M. Moriondo, E. Kostopoulou, and C.M.
Goodess, 2009: Climatic changes and associated impacts in the Mediterranean
resulting from a 2°C global warming. Global and Planetary Change, 68(3),
209-224.
Gil, A.I., V.R. Louis, I.N. Rivera, E. Lipp, A. Huq, C.F. Lanata, D.N. Taylor, E. Russek-Cohen,
N. Choopun, R.B. Sack, and R.R. Colwell, 2004: Occurrence and distribution of
Vibrio cholerae in the coastal environment of Peru. Environmental
Microbiology, 6(7), 699-706.
Giuliani, G. and P. Peduzzi, 2011: The PREVIEW Global Risk Data Platform: a geoportal
to serve and share global data on risk to natural hazards. Natural Hazards and
Earth System Sciences, 11, 53-66, doi:10.5194/nhess-11-53-2011.
GNDR, 2009: Clouds but Little Rain - Views from the Frontline: A Local Perspective of
Progress towards Implementation of the Hyogo Framework for Action. Global
Network of Civil Society Organisations for Disaster Reduction, Teddington, UK,
59 pp.
Goddard, L., Y. Aichellouche, W. Baethgen, M. Dettinger, R. Graham, P. Hayman, M.
Kadi, R. Martinez, and H. Meinke, 2010: Providing seasonal-to-interannual
climate information for risk management and decision-making. Procedia
Environmental Sciences, 1, 81-101.
Goodland, A., D. Sheehy, and T. Shine, 2009: Mongolia Livestock Sector Study,
Volume I – Synthesis Report. Sustainable Development Department, East Asia
and Pacific Region, World Bank, Washington, DC.
Government of Bangladesh, 2008: Cyclone Sidr in Bangladesh: Damage, Loss
and Needs Assessment for Disaster Recovery and Reconstruction. Government
of Bangladesh, Dhaka, Bangladesh.
Government of India, 2001: Census of India: (Provisional) Slum Population in
Million Plus Cities (Municipal Corporations): Part A. Government of India,
Ministry of Home Affairs, Office of the Registrar General and Census
Commissioner, New Delhi, India, censusindia.gov.in/Tables_Published/
Admin_Units/Admin_links/slum1_m_plus.html.
Government of Northwest Territories, 2008: NWT Climate Change Impacts and
Adaptation Report. Government of Northwest Territories, Yellowknife, Canada,
www.enr.gov.nt.ca/_live/documents/content/NWT_Climate_Change_Impacts_
and_Adaptation_Report.pdf.
Government of Yukon, 2009: Yukon Government Climate Change Action Plan.
Government of Yukon, Whitehorse, Canada, www.env.gov.yk.ca/pdf/
YG_Climate_Change_Action_Plan.pdf.
Government of Yukon, 2010: Shawak and Permafrost. Government of Yukon,
Whitehorse, Canada, www.hpw.gov.yk.ca/trans/engineering/765.html.
Governments of Northwest Territories, Nunavut, and Yukon. 2010: A
Northern Vision: A Stronger North and a Better Canada Pan-Territorial
Adaptation Strategy. Governments of Northwest Territories, Nunavut, and
Yukon, Canada, www.anorthernvision.ca/strategy/.
Griffith, D.C., L.A. Kelly-Hope, and M.A. Miller, 2006: Review of reported cholera out-
breaks worldwide, 1995-2005. American Journal of Tropical Medicine &
Hygiene, 75(5), 973-977.
Grynszpan, D., 2003: Lessons from the French heatwave. Lancet, 362, 1169-1170.
Grynszpan, D., V. Murray, and S. Llossa, 2011: The value of case studies in disaster
assessment. Prehospital and Disaster Medicine, doi:10.1017/
S1049023X11006406.
Guha-Sapir, D., F. Vos, and R. Below, with S. Ponserre, 2011: Annual Disaster
Statistical Review 2010: The Numbers and Trends. Centre for Research on the
Epidemiology of Disasters, Université Catholique de Louvain, Brussels, Belgium,
42 pp. www.cred.be/sites/default/files/ADSR_2010.pdf.
Gurenko, E., 2004: Catastrophe Risk and Reinsurance: A Country Risk Management
Perspective. Risk Books, London, UK.
Haile, M., 2005: Weather patterns, food security and humanitarian response in sub-
Saharan Africa. Philosophical Transactions of the Royal Society B, 360(1463),
2169-2182.
Hajat, S., S.C. Sheridan, M. Allen, M. Pascal, K. Laaidi, A. Yagouti, U. Bickis, A.Tobias,
D. Bourque, B.G. Armstrong, and T. Kosatsky, 2010: Heat-health warning systems:
a comparison of the predictive capacity of different approaches to identifying
dangerously hot days. American Journal of Public Health, 100(6), 1137-1144.
Chapter 9 Case Studies
534
Hallegatte, S. and J. Corfee-Morlot, 2011: Understanding climate change impacts,
vulnerability and adaptation at city scale: an introduction. Climatic Change,
104, 1-12.
Hansen, J.W., S.J. Mason, L. Sun, and A. Tall, 2011: Review of seasonal climate
forecasting for agriculture in sub-Saharan Africa. Experimental Agriculture, 47,
205-240.
Hanson, S., R. Nicholls, N. Ranger, S. Hallegatte, J. Corfee-Morlot, and C. Herweijer,
2011: A global ranking of port cities with high exposure to climate extremes.
Climatic Change, 104, 89-111.
Harlan, S.L., A.J. Brazel, L. Prashad, W.L. Stefanov, and L. Larsen, 2006: Neighborhood
microclimates and vulnerability to heat stress. Social Science and Medicine, 63,
2847-2863.
Harris, J.B., M.J. Podolsky, T.R. Bhuiyan, F. Chowdhury, A.I. Khan, R.C. Larocque, T.
Logvinenko, J. Kendall, A. S. Faruque, C.R. Nagler, E.T. Ryan, F. Qadri, and S.B.
Calderwood, 2009: Immunologic responses to Vibrio cholerae in patients co-
infected with intestinal parasites in Bangladesh. PLoS Neglected Tropical
Diseases, 3(3), e403.
Hashizume, M., Y. Wagatsuma, A.S. Faruque, T. Hayashi, P.R. Hunter, B. Armstrong,
and D.A. Sack, 2008: Factors determining vulnerability to diarrhoea during and
after severe floods in Bangladesh. Journal of Water & Health, 6(3), 323-332.
Havenith, G., 2001: Individualized model of human thermoregulation for the
simulation of heat stress response. Journal of Applied Physiology, 90(5),
1943-1954.
Hawkins-Bell, L. and J.T. Rankin, 1994: Heat-related deaths – Philadelphia and United
States, 1993–1994. Morbidity and Mortality Weekly Report, 43(25), 453-455.
Health Canada, 2010: Communicating the Health Risks of Extreme Heat Events:
Toolkit for Public Health and Emergency Management Officials. Health Canada,
Ottawa, Canada.
Hellmuth, M.E., A. Moorhead, M.C. Thomson, and J. Williams (eds.), 2007: Climate
Risk Management in Africa: Learning from Practice. International Research
Institute for Climate and Society, Columbia University, New York, NY.
Hellmuth, M.E., D.E. Osgood, U. Hess, A. Moorhead, and H. Bhojwani (eds.), 2009:
Index Insurance and Climate Risk: Prospects for Development and Disaster
Management. International Research Institute for Climate and Society,
Columbia University, New York, NY.
Hellmuth, M., S.J. Mason, C. Vaughan, M.K. van Aalst, and R. Choularton (eds.),
2011: A Better Climate for Disaster Risk Management. International Research
Institute for Climate and Society, Columbia University, New York, NY, 118 pp.
Hémon, D. and E. Jougla, 2004: The heat wave in France in August 2003. Revue
d’Epidémiologie et de Sante Publique, 52(3-5).
Hennessy, K., C. Lucas, N. Nicholls, J. Bathols, R. Suppiah, and J. Ricketts, 2005:
Climate Change Impacts on Fire-Weather In South-East Australia. CSIRO Marine
and Atmospheric Research, Aspendale, Australia.
Henstra
, D. and G.A. McBean, 2005: Canadian Disaster Management Policy: Moving
Toward a Paradigm Shift? Canadian Public Policy, 31, 303-318.
Hess, U. and H. Syroka, 2005: Weather-based Insurance in Southern Africa. The Case
of Malawi. World Bank, Washington, DC.
Hess, U., W. Wiseman, and T. Robertson, 2006: Ethiopia: Integrated Risk Financing to
Protect Livelihoods and Foster Development. Discussion Paper, UN World Food
Programme, Rome, Italy.
Hodgkinson, R.D., J. Evans, and J. Wood, 2003: Maintaining oral hydration in older
adults: a systematic review. International Journal of Nursing Practice, 9, 19-28.
Holmgren, J., 1981: Actions of cholera toxin and the prevention and treatment of
cholera. Nature, 292(5822), 413-417.
Holubec, I., 2008: Flat Loop Thermosyphon Foundations in Warm Permafrost.
Government of Northwest Territories, Yellowknife, Canada.
Hossain, M.Z., M.T. Islam, T. Sakai, and M. Ishida, 2008: Impact of tropical cyclones
on rural infrastructures in Bangladesh. Agricultural Engineering International:
the CIGR Ejournal, X, Invited Overview No. 2, 13 pp.
Hughes, K., E. Van Beurden, E.G. Eakin, L.M. Barnett, E. Patterson, J. Backhouse, S.
Jones, D. Hauser, J.R. Beard, and B. Newman, 2008: Older persons’ perception of
risk of falling: implications for fall-prevention campaigns. American Journal of
Public Health, 98(2), 351-357.
Hunt, A., and P. Watkiss, 2011: Climate change impacts and adaptation in cities: a
review of the literature. Climatic Change, 104, 13-49.
Huq, A., R.B. Sack, A. Nizam, I.M. Longini, G.B. Nair, A. Ali, J.G. Morris, Jr., M.N. Khan,
A.K. Siddique, M. Yunus, M.J. Albert, D.A. Sack, and R.R. Colwell, 2005: Critical
factors influencing the occurrence of Vibrio cholerae in the environment of
Bangladesh. Applied & Environmental Microbiology, 71(8), 4645-4654.
Huq, S., H. Reid and L.A. Murray, 2006: Climate Change and Development Links.
Gatekeeper Series 123, IIED, London, UK, 24 pp.
Huq, S., S. Kovats, H. Reid, and D. Satterthwaite, 2007: Editorial: Reducing risks to cities
from disasters and climate change. Environment and Urbanization, 19(1), 3-15.
ICSU, 2008: A Science Plan for Integrated Research on Disaster Risk: Addressing the
Challenge Of Natural And Human-Induced Environmental Hazards. International
Council for Science, Paris, France, www.irdrinternational.org/reports.
ICTILO, UNDP, and UNISDR, 2010: Local Governments and Disaster Risk Reduction:
Good Practices and Lessons Learned. International Training Centre of the
International Labour Organization, UN Development Programme, and United
Nations International Strategy for Disaster Reduction, Geneva, Switzerland, 86 pp.
IFRC, 2007: Mozambique Floods and Cyclones. DREF Bulletin No. MDRMZ002,
Update no. 1, Glide no. FL-2006-000198-MOZ, International Federation of the
Red Cross and Red Crescent Societies, Geneva, Switzerland, 4 pp., www.ifrc.org/
docs/appeals/07/MDRMZ00201.pdf.
IFRC, 2009: World Disasters Report 2009 – Focus on Early Warning and Early Action.
International Federation of the Red Cross and Red Crescent Societies, Geneva,
Switzerland, 204 pp., www.ifrc.org/publicat/wdr2009/summaries.asp.
IFRC, 2011: Desk review on trends in the promotion of community-based disaster
risk reduction legislation. In: Global Assessment Report on Disaster Risk
Reduction: Revealing Risk, Redefining Development. United Nations
International Strategy for Disaster Reduction, Geneva, Switzerland.
IMF, 2004: Grenada: Use of Fund Resources-Request for Emergency Assistance-Staff
Report; Press Release on the Executive Board Discussion; and Statement by the
Executive Director for Grenada. International Monetary Fund, Washington, DC,
www.imf.org/external/pubs/ft/scr/2004/cr04405.pdf.
IMF, 2006: Grenada: Request for a Three-Year Arrangement under the Poverty
Reduction and Growth Facility – Staff Report; and Press Release on the
Executive Board Discussion. International Monetary Fund, Washington, DC,
www.imf.org/external/pubs/ft/scr/2006/cr06277.pdf.
IMF, 2011: World Economic Outlook Database. International Monetary Fund,
Washington, DC, www.imf.org/external/pubs/ft/weo/2011/01/weodata/index.aspx.
INAC, 2010: Report on Adaptation to Climate Change Activities in Arctic Canada.
Indian and Northern Affairs Canada, Gatineau, Canada.
Infrastructure Canada, 2006: Adapting Infrastructure to Climate Change in
Canada’s Cities and Communities: A Literature Review. Infrastructure Canada,
Ottawa, Canada, cbtadaptation.squarespace.com/storage/CdnInfrastructure
Adaptation-LiteratureReview.pdf.
Instanes, A., O.A. Anisimov, L. Brigham, D. Goering, L.N. Khrustalev, B. Ladanyi,
and J.O. Larsen, 2005: Infrastructure: buildings, support systems, and industrial
facilities. In: Arctic Climate Impact Assessment. Cambridge University Press,
Cambridge, UK, pp. 907-944.
InVS, 2003: Impact sanitaire de la vague de chaleur en France survenue en août
2003. Rapport d’étape (Health impact of the heatwave which took place in
France in August 2003. Progress Report). Institut de Veille Sanitaire (InVS),
Département Santé Environnement, Saint-Maurice, France.
IOM, 2008: Migration and Climate Change. IOM Research Series No. 31, International
Organization for Migration, Geneva, Switzerland, www.migrationdrc.org/
publications/resource_guides/Migration_and_Climate_Change/MRS-31.pdf.
IRI, 2011: Progress Report to National Oceanographic and Atmospheric
Administration. International Research Institute for Climate and Society,
Columbia University, New York, NY, portal.iri.columbia.edu/portal/server.pt/
gateway/PTARGS_0_2_7713_0_0_18/NOAAReport2011_FINAL_strip.pdf.
Jegillos, S.R., 2003: Methodology. In: Sustainability in Grass-Roots Initiatives: Focus
on Community Based Disaster Management [Shaw, R. and K. Okazaki (eds.)].
United Nations Centre for Regional Development, Disaster Management
Planning Hyogo Office, Hyogo, Japan, pp. 19-28.
Chapter 9Case Studies
535
Jeuland, M. and D. Whittington, 2009: Cost-benefit comparisons of investments in
improved water supply and cholera vaccination programs. Vaccine, 27(23),
3109-3120.
Johnson, K., S. Solomon, D. Berry, and P. Graham, 2003: Erosion Progression and
Adaptation Strategy in a Northern Coastal Community. Swets & Zeitlinger,
Lisse, The Netherlands.
Jones, A.E., U.U. Wort, A.P. Morse, I.M. Hastings, and A.S. Gagnon, 2007: Climate
prediction of El Niño malaria epidemics in north-west Tanzania. Malaria
Journal, 6, 162.
Jones, G.S., P.A. Stott, and N. Christidis, 2008: Human contribution to rapidly increasing
frequency of very warm Northern Hemisphere summers. Journal of Geophysical
Research, 113, D02109.
Kadomura, H., 2005: Climate anomalies and extreme events in Africa in 2003,
including heavy rains and floods that occurred during northern hemisphere
summer. African Study Monographs, S30, 165-181.
Kalkstein, A.J. and S.C. Sheridan, 2007: The social impacts of the heat-health
watch/warning system in Phoenix, Arizona: assessing the perceived risk and
response of the public. International Journal of Biometeorology, 52(1), 43-55.
Kalsi, S.R. and R.S. Pareek, 2001: Hottest April of the 20th century over north-west
and central India. Current Science, 80(7), 867-872.
Karim, M.F. and N. Mimura, 2008: Impacts of climate change and sea level rise on
cyclonic storm surge floods in Bangladesh. Global Environmental Change,
18(3), 490-500.
Kaufmann, D., A. Kraay, and M. Mastruzzi, 2010: The Worldwide Governance
Indicators: Methodology and Analytical Issues. Policy Research Working Paper
No. 5430, World Bank, papers.ssrn.com/sol3/ papers.cfm?abstract_id=1682130.
Kelman, I., 2010: Policy arena: introduction to climate, disasters and international
development. Journal of International Development, 22, 208-217.
Kelman, I. and J. West, 2009: Climate change and small island developing states: a
critical review. Ecological and Environmental Anthropology, 5(1), 1-16.
Keen, J. and T. Packwood, 1995. Qualitative research: Case study evaluation. BMJ,
311(1), 444-446.
Kharin, V.V. and F.W. Zwiers, 2003: Improved seasonal probability forecasts. Journal
of Climate, 16, 1684-1701.
Kienberger, S., 2007: Assessing the vulnerability to natural hazards on the provincial/
community level in Mozambique: the contribution of GIScience and remote
sensing. In: The 3rd International Symposium on Geo-information for Disaster
Management, Joint CIG/ISPRS Conference, Toronto, Canada, 23-25 May 2007.
Klinenberg, E., 2002: Heatwave: A Social Autopsy of Disaster in Chicago. University
of Chicago Press, Chicago, IL.
Koelle, K., 2009: The impact of climate on the disease dynamics of cholera. Clinical
Microbiology & Infection, 15(S1), 29-31.
Koelle, K., M. Pascual, and M. Yunus, 2005a: Pathogen adaptation to seasonal forcing
and climate change. Proc of the Royal Society of London B, 272(1566), 971-977.
Koelle, K., X. Rodo, M. Pascual, M. Yunus, and G. Mostafa, 2005b: Refractory periods
and climate forcing in cholera dynamics. Nature, 436(7051), 696-700.
Kondo, H., N. Seo, T. Yasuda, M. Hasizume, Y. Koido, N. Ninomiya, and Y. Yamamoto,
2002: Post-flood infectious diseases in Mozambique. Prehospital and Disaster
Medicine, 17(3), 126-133, pdm.medicine.wisc.edu/Volume_17/issue_3/kondo.pdf.
Kosatsky, T., 2005: The 2003 European heat waves. Euro Surveillance, 10, 148-149.
Koster, R.D., M.J. Suarez, P. Liu, U. Jambor, A. Berg, M. Kistler, R. Reichle, M. Rodell,
and J. Famiglietti, 2004: Realistic initialization of land surface states: impacts on
subseasonal forecast skill. Journal of Hydrometeorology, 5, 1049-1063.
Koster, R.D., S.P.P. Mahanama, T.J. Yamada, G. Balsamo, A.A. Berg, M. Boisserie, P.A.
Dirmeyer, F.J. Doblas-Reyes, G. Drewitt, C.T. Gordon, Z. Guo, J.-H. Jeong, D.M.
Lawrence, W.-S. Lee, Z. Li, L. Luo, S. Malyshev, W.J. Merryfield, S.I. Seneviratne, T.
Stanelle, B.J.J.M. van den Hurk, F. Vitart, and E.F. Wood, 2010: Contribution of
land surface initialization to subseasonal forecast skill: first results from a
multi-model experiment. Geophysical Research Letters, 37, L02402.
Kovats, R.S. and S. Hajat, 2008: Heat stress and public health: a critical review.
Annual Review of Public Health, 29, 41-55.
Kovats, S. and R. Akhtar, 2008: Climate, climate change and human health in Asian
cities. Environment and Urbanization, 20(1), 165-175.
Kreimer, A. and M. Arnold, 2000: World Bank’s role in reducing impacts of disasters.
Natural Hazards Review, 1(1), 37-42.
Kulling, P., M. Birnbaum, V. Murray, and G. Rockenschaub, 2010: Guidelines for
reports on health crises and critical health events. Prehospital and Disaster
Medicine, 25(4), 377-383.
Kumar, S., 1998: India’s heat wave and rains result in massive death toll. Lancet,
351, 1869.
Kundzewicz, Z.W., L.J. Mata, N.W. Arnell, P. Döll, P. Kabat, B. Jiménez, K.A. Miller, T.
Oki, Z. Sen, and I.A. Shiklomanov, 2007: Freshwater resources and their
management. In: Climate Change 2007. Impacts, Adaptation and Vulnerability.
Contribution of Working Group II to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P.
Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, 173-210.
Kunreuther, H. and E. Michel-Kerjan, 2009: At War with the Weather: Managing
Large-Scale Risks in a New Era of Catastrophes. MIT Press, Cambridge, MA.
Kunreuther, H. and R.J. Roth, 1998: Paying the Price: The Status and Role of
Insurance Against Natural Disasters in the United States
. Joseph Henry Press,
Washington, DC.
Laaidi, K., M. Pascal, M. Ledrans, A. Le Tertre, S. Medina, C. Caserio, J.C. Cohen, J.
Manach, P. Beaudeau, and P. Empereue-Bissonet, 2004: Le système français
d’alerte canicule et santé (SACS 2004): Un dispositif intégéré au Plan National
Canicule (The French Heatwave Warning System and Health: A System
Integrated into the National Heatwave Plan). Institut de Veille Sanitaire, Saint-
Maurice, France.
Lagadec, P., 2004: Understanding the French 2003 heat wave experience: beyond
the heat, a multi-layered challenge. Journal of Contingencies and Crisis
Management, 12, 160-169.
Lalani, B., 2011: Drought mitigation: Assessing technological trade-offs and
challenges posed by policy solutions; A case study of Salamieh District, Syria.
World Applied Sciences Journal, 13(4), 936-946.
Larsen, P.H., S. Goldsmith, O. Smith, M. Wilon, K. Strzepek, P. Chinowsky, and B.
Saylor, 2008: Estimating future costs for Alaska public infrastructure at risk from
climate change. Global Environmental Change, 18, 442-457.
Lateef, F., 2009: Cyclone Nargis and Myanmar: a wake up call. Journal of
Emergencies, Trauma, and Shock, 2(2), 106-113.
Lawrimore, J., R. Heim, T. Owen, M. Svoboda, V. Davydova, D. Chobanik, R. Rippey,
and D. Lecomte, 2007: The North American drought monitor and a global
drought early warning system. In: The Full Picture. Group on Earth
Observations, Tudor Rose, London, UK, pp. 142-144.
Lazo, J.K., R.E. Morss, and J.L. Demuth, 2009: 300 billion served: Sources, perceptions,
uses, and values of weather forecasts. Bulletin of the American Meteorological
Society, 90(6),785-798.
Lemmen, D.S., F.J. Warren, J. Lacroix, and E. Bush (eds.), 2008: From Impacts to
Adaptation: Canada in a Changing Climate 2007. Government of Canada,
Ottawa, Canada, 448 pp.
Létard, V., H. Flandre, and S. Lepeltier, 2004: La France et les Français face à la
canicule: les leçons d’une crise (How France and French people faced a
heatwave: lessons from a crisis). Report No. 195 (2003-2004) to the Sénat,
Government of France, Paris, France, 391 pp.
Levick J.J., 1859: Remarks on sunstroke. American Journal of the Medical Sciences,
73, 40-42.
Liniger, M.A., H. Mathis, C. Appenzeller, and F.J. Doblas-Reyes, 2007: Realistic
greenhouse gas forcing and seasonal forecasts. Geophysical Research Letters,
34, L04705.
Linnerooth-Bayer, J. and R. Mechler, 2009: Insurance against Losses from Natural
Disasters in Developing Countries. DESA Working Paper No. 85, United Nations
Department of Economic and Social Affairs, New York, NY, USA, 37 pp.
Linnerooth-Bayer, J., R. Mechler, and G. Pflug, 2005: Refocusing disaster aid.
Science, 309, 1044-1046.
Linnerooth-Bayer, J., R. Mechler, and M.J. Mace, 2007: Insurance for assisting
adaptation to climate change in developing countries. A proposed strategy.
Climate Policy, 6, 621-636.
Chapter 9 Case Studies
536
Linnerooth-Bayer, J., P. Suarez, M. Victor and R. Mechler, 2009: Drought insurance
for subsistence farmers in Malawi. Natural Hazards Observer, 33(5), 5-9.
Lipp, E.K., A. Huq, and R.R. Colwell, 2002: Effects of global climate on infectious
disease: the cholera model. Clinical Microbiology Reviews, 15, 757-770.
Lise, W., S. Hess, and B. Purev, 2006: Pastureland degradation and poverty among
herders in Mongolia: data analysis and game estimation. Ecological Economics,
58, 350-364.
Lizarraga-Partida, M.L., E. Mendez-Gomez, A.M. Rivas-Montano, E. Vargas-
Hernandez, A. Portillo-Lopez, A.R. Gonzalez-Ramirez, A. Huq, and R.R. Colwell,
2009: Association of Vibrio cholerae with plankton in coastal areas of Mexico.
Environmental Microbiology, 11(1), 201-208.
Llosa, S. and I. Zodrow, 2011: Disaster risk reduction legislation as a basis for effective
adaptation. In: Global Assessment Report on Disaster Risk Reduction: Revealing
Risk, Redefining Development. United Nations International Strategy for
Disaster Reduction, Geneva, Switzerland.
Lopez, A.L., J.D. Clemens, J. Deen, and L. Jodar, 2008: Cholera vaccines for the
developing world. Human Vaccines, 4(2), 165-169.
Loster, T. and D. Reinhard, 2010: Microinsurance and climate change. In:
Microinsurance- An innovative tool for risk and disaster management [Morelli,
E., G.A. Onnis, W.J. Amman, and C. Sutter (eds.)]. Global Risk Forum, Davos,
Switzerland, pp. 39-42.
Loster, T. and A. Wolf, 2007: IntoAction. Flood warning system in Mozambique.
Completion of Búzi project. Munich Re Foundation, Munich, Germany, 16 pp.
Louka, P., G. Vachon, J.-F. Sini, P.G. Mestayer, and J.-M. Rosant, 2002: Thermal effects
on the airflow in a street canyon – Nantes ‘99 experimental results and model
simulation. Water, Air and Soil Pollution: Focus, 2(5-6), 351-364.
Luber, G. and M. McGeehin, 2008: Climate change and extreme heat events.
American Journal of Preventive Medicine, 35(5), 429-435.
Luna, E.M., M.L.P. Bautista, and M.P. de Guzman, 2008: Mainstreaming Disaster Risk
Reduction in the Education Sector in the Philippines: Integrating Disaster Risk
Reduction in the School Curriculum, Impacts of Disasters on the Education
Sector, School Construction Current Practices and Improvements Needed.
Centre for Disaster Preparedness, Quezon City, the Philippines, 160 pp.
Luque Fernandez, M.A., A. Bauernfeind, J.D. Jimenez, C.L. Gil, N. El Omeiri, and D.H.
Guibert, 2009: Influence of temperature and rainfall on the evolution of cholera
epidemics in Lusaka, Zambia, 2003-2006: analysis of a time series. Transactions
of the Royal Society of Tropical Medicine & Hygiene, 103(2), 137-143.
Mahul, O. and J. Skees, 2005: Managing Agricultural Risk at the Country Level: The
Case of Index-Based Livestock Insurance in Mongolia. World Bank, Washington,
DC, 35 pp.
Marin, A., 2008: Between cash cows and golden calves. Adaptations of Mongolian
pastoralism in the ‘age of the market.Nomadic Peoples, 12(2), 75-101.
Marin, A., 2010: Riders under storms: Contributions of nomadic herders’ observations
to analysing climate change in Mongolia. Global Environmental Change –
Human and Policy Dimensions
, 20(1), 162-176.
Marinacci, C., M. Marino, E. Ferracin, L. Fubini, L. Gilardi, P. Visentin, E. Cadum, and
G. Costa, 2009: Testing of interventions for prevention of heat wave related
deaths: results among frail elderly and methodological problems.
Epidemiologia e Prevenzione, 33(3), 96-103.
Maskey, M., J.S. Shastri, K. Saraswathi, R. Surpam, and N. Vaidya, 2006: Leptospirosis
in Mumbai: post-deluge outbreak 2005. Indian Journal of Medical
Microbiology, 24(S4), 337-338.
Mason, P.R., 2009: Zimbabwe experiences the worst epidemic of cholera in Africa.
Journal of Infection in Developing Countries, 3(2), 148-151.
Matsuda, F., S. Ishimura, Y. Wagatsuma, T. Higashi, T. Hayashi, A.S. Faruque, D.A. Sack,
M. Nishibuchi, and A.S.G. Faruque, 2008: Prediction of epidemic cholera due to
Vibrio cholerae O1 in children younger than 10 years using climate data in
Bangladesh. Epidemiology & Infection, 136(1), 73-79.
Mattingly, S., 2002: Policy, legal and institutional arrangements. Proceedings of
Regional Workshop on Best Practices in Disaster Mitigation, Asian Disaster
Preparedness Centre, Bangkok, Thailand, 51 pp.
McBean, G.A., 2000. Forecasting in the 21st Century. Publication 916, World
Meteorological Organization, Geneva, Switzerland, 18 pp.
McBean, G., G. Alekseev, D. Chen, E. Førland, J. Fyfe, P.Y. Groisman, R. King, H.
Melling, R. Vose, and P.H. Whitfield, 2005: Arctic climate: past and present. In:
Arctic Climate Impact Assessment. Cambridge University Press, Cambridge, UK,
pp. 22-60.
McCormick, S., 2010a: Dying of the heat: diagnostic debates, calculations of risk, and
actions to advance preparedness. Environment and Planning A, 42, 1513-1518.
McCormick, S., 2010b: Hot or not? Obstacles to the emergence of climate-induced
illness movements. In: Social Movements and Health Care in the United States
[Zald, M., J. Banaszak-Holl, and S. Levitsky (eds.)]. Oxford University Press,
Oxford, UK.
McGranahan, G., D. Balk, and B. Anderson, 2007: The rising tide: assessing the risks
of climate change and human settlements in low elevation coastal zones.
Environment and Urbanization, 19(1), 17-37.
McMichael, A.J., R.E. Woodruff, and S. Hales, 2006: Climate change and human
health: present and future risks. Lancet, 367(9513), 859-869.
McWhinney, I.R., 2001: The value of case studies. European Journal of General
Practice, 7(3), 88-89.
Mearns, R., 2004: Sustaining livelihoods on Mongolia’s pastoral commons: Insights
from a participatory poverty assessment. Development and Change, 35(1),
107-139.
Mechler, R., 2004: Natural Disaster Risk Management and Financing Disaster Losses
in Developing Countries. Verlag für Versicherungswirtschaft, Karlsruhe,
Germany.
Mechler, R., J. Linnerooth-Bayer, and D. Peppiatt, 2006: Disaster Insurance for the
Poor? A Review of Microinsurance for Natural Disaster Risks in Developing
Countries. ProVention Consortium and IIASA, Geneva, Switzerland and
Laxenburg, Austria.
Medina-Cetina, Z. and F. Nadim, 2008: Stochastic design of an early warning
system. Georisk: Assessment and Management of Risk for Engineered Systems
and Geohazards, 2(4), 223-236.
Meehl, G.A., T.F. Stocker, W.D. Collins, P. Friedlingstein, A.T. Gaye, J.M. Gregory, A.
Kitoh, R. Knutti, J.M. Murphy, A. Noda, S.C.B. Raper, I.G. Watterson, A.J. Weaver,
and Z.-C. Zhao, 2007: Global climate projections. In: Climate Change 2007: The
Physical Science Basis. Contribution of Working Group I to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change
[Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor
and H.L. Miller (eds.)]. Cambridge University Press, Cambridge, UK, and New
York, NY, pp. 747-845.
Meehl, G.A., L. Goddard, J. Murphy, R.J. Stouffer, G. Boer, G. Danabasoglu, K. Dixon,
M.A. Giorgetta, A.M. Greene, E. Hawkins, G. Hegerl, D. Karoly, N. Keenlyside, M.
Kimoto, B. Kirtman, A. Navarra, R. Pulwarty, D. Smith, D. Stammer, and T.
Stockdale, 2009: Decadal prediction: can it be skillful? Bulletin of the American
Meteorological Society, 90, 1467-1485.
Meinke, H., R. Nelson, P. Kokic, R. Stone, R. Selvaraju, and W. Baethgen, 2006:
Actionable climate knowledge – from analysis to synthesis. Climate Research,
33, 101-110.
Melbourne Water, 2009a: Water Report. Melbourne Water, Melbourne, Australia,
www.melbournewater.com.au/content/water_storages/water_report/water_
report.asp.
Melbourne Water, 2009b: Bushfires in Catchments. Melbourne Water, Melbourne,
Australia, www.melbournewater.com.au/content/water_storages/bushfires_in_
catchments/bushfires_in_catchments.asp.
Mendelsohn, J. and T. Dawson, 2008: Climate and cholera in KwaZulu-Natal,
South Africa: the role of environmental factors and implications for epidemic
preparedness. International Journal of Hygiene & Environmental Health,
211(1-2), 156-162.
Météo France, 2003: Retour sur la canicule d’août 2003 - Bilan de la canicule
d’août 2003 adressé par Météo-France au ministère de l’Equipement, des
Transports, du Logement et de la Mer le 20 août 2003 (Analysis of the August
2003 heatwave presented by Météo-France to the Ministry for Equipement,
Transport, Housing and Sea on 20 August 2003). Météo France, Paris, France,
france.meteofrance.com/france/actu/bilan/archives/2003/canicule?page_id=10
035&document_id=4523&portlet_id=40336.
Chapter 9Case Studies
537
Météo France, 2006: Retour sur la canicule de juillet 2006 (Analysis of the July
2006 heatwave). Météo France, Paris, France, france.meteofrance.com/
france/actu/bilan/archives/2006/canicule?page_id=10043&document_id=451
6&portlet_id=40408.
MGH, 2003: State Policy for Food and Agriculture. Annex 29 of resolution of
Mongolia Great Hural (MGH), Ulaanbaatar, Mongolia, mofa.gov.mn/
livestock/index.php?option=com_content&view=article&id=73%3A2010-09-
06-08-07-14&catid=52%3A2010-07-02-06-32-10&Itemid=59&lang=en.
Michelozzi, P., F.K. de’ Donato, A.M. Bargagli, D. D’Ippoliti, M. de Sario, C. Marino, P.
Schifano, G. Cappai, M. Leone, U. Kirchmayer, M. Ventura, M. di Gennaro, M.
Leonardi, F. Oleari, A. de Martino, and C.A. Perucci, 2010: Surveillance of summer
mortality and preparedness to reduce the health impact of heat waves in Italy.
International Journal of Environmental Research and Public Health, 7(5),
2256-2273.
MIJ, 2009: Informe Nacional del Progreso en la Implementación del Marco de Acción
de Hyogo. Report to UNISDR HFA Monitor, Colombia Ministerio del Interior y de
Justicia, Dirección de Prevención y Atención de Desastres, Bogota, Colombia.
Mijiddorj, R., 2008: Global Warming and Challenge to Desert. Mongolia University
of Science and Technology, Ulaanbaatar, Mongolia, 62 pp.
Mills, E., 2005: Insurance in a climate of change. Science, 309, 1040-1044.
Milly, P.C.D., J. Betancourt, M. Falkenmark, R.M. Hirsch, Z.W. Kundzewicz, D.P.
Lettenmaier, and R.J. Stouffer, 2008: Stationarity is dead: whither water
management? Science, 319, 573-574.
Mimura, N., L. Nurse, R.F. McLean, J. Agard, L. Briguglio, P. Lefale, R. Payet, and G.
Sem, 2007: Small islands. In: Climate Change 2007. Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani,
J.P. Palutikof, P.J. Van Der Linde, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK, pp. 687-716.
Mirza, M.M.Q., 2003: Climate change and extreme weather events: can developing
countries adapt? Climate Policy, 3, 233-248.
Mitchell, T., T. Tanner, and K. Haynes, 2009: Children as Agents of Change for Disaster
Risk Reduction: Lessons from El Salvador and the Philippines. Working Paper 1,
Children in a Changing Climate, IDS, Brighton, UK, 48 pp.
MMS, 2009: National Climate Risk Management Strategy and Action Plan (Draft) [by
expert team headed by Togtokh, N. and D. Tuvdendorj]. National Emergency
Management Authority, Mongolia Meteorological Society (MMS), United
Nations Development Programme, Ulaanbaatar, Mongolia.
Morelli, E., G.A. Onnis, W.J. Amman, and C. Sutter (eds.), 2010: Microinsurance – An
innovative tool for risk and disaster management. Global Risk Forum, Davos,
Switzerland, 360 pp.
Morinaga, Y., S. Tian, and M. Shinoda, 2003: Winter snow anomaly and atmospheric
circulation in Mongolia. International Journal of Climatology, 23, 1627-1636.
Morris, L., 2011: Death and suffering in the land of Genghis Khan. CMAJ, 183(5),
doi:10.1503/cmaj.109-3796.
Mouillot, F., J.P. Ratte, R. Joffre, J.M. Moreno, and S. Rambal, 2003: Some determinants
of the spatio-temporal fire cycle in a Mediterranean landscape (Corsica,
France). Landscape Ecology, 18, 665-674.
Munich Re, 2010: Topics Geo: Natural catastrophes 2009 – Analyses, assessments,
positions. Munich Re, Munich, Germany, www.munichre.com.
Myers, S.S. and J.A. Patz, 2009: Emerging threats to human health from global
environmental change. Annual Review of Environment and Resources, 34(1),
223-252.
NASA, 2008: Heatwave in Northern Africa and Southern Europe. Earth Observatory,
NASA Goddard Space Flight Center, Greenbelt, MD, earthobservatory.nasa.gov/
NaturalHazards/view.php?id=15257.
National Climate Centre, 2009: The exceptional January-February 2009
heatwave in southeastern Australia. Special Climate Statement 17, Bureau of
Meteorology, Melbourne, Australia, www.bom.gov.au/climate/current/statements/
scs17c.pdf.
National Hurricane Center, 2006: Dennis, Katrina, Rita, Stan, and Wilma “Retired”
from List of Storm Names. National Oceanic and Atmospheric Administration,
Washington, DC, www.noaanews.noaa.gov/stories2006/s2607.htm.
Naughton, M.P., A. Henderson, M.C. Mirabelli, R. Kaiser, J.L. Wilhelm, S.M. Kieszak,
C.H. Rubin, and M.A. McGeehin, 2002: Heat-related mortality during a 1999
heat wave in Chicago. American Journal of Preventive Medicine, 22(4), 221-227.
NDMC, 2007: Inaugural Annual Report 2006-2007. National Disaster Management
Centre, Provincial and Local Government Department, Pretoria, South Africa,
172 pp.
NDMC, 2010: National Education, Training and Research Needs and Resources
Analysis (NETaRNRA). Consolidated report compiled by the National Disaster
Management Centre, South African Department of Cooperative Governance
and Traditional Affairs, Pretoria, South Africa, 23 pp.
Nelson, F., O. Anisimov, and N. Shiklomanov, 2002: Climate change and hazard
zonation in the circum-arctic permafrost regions. Natural Hazards, 26, 203-225.
Nicholls, N., 2001: Atmospheric and climatic hazards: improved monitoring and
prediction for disaster mitigation. Natural Hazards, 23, 137-155.
Nicholls, N., C. Skinner, M. Loughnan, and N. Tapper, 2008: A simple heat alert
system for Melbourne, Australia. International Journal of Biometeorology,
52(5), 375-384.
Nicholls, R.J., P.P. Wong, V.R. Burkett, J.O. Codignotto, J.E. Hay, R.F. McLean, S.
Ragoonaden, and C.D. Woodroffe, 2007: Coastal systems and low-lying areas.
In: Climate Change 2007. Impacts, Adaptation and Vulnerability. Contribution of
Working Group II to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der
Linden, and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK,
315-356.
Nielson, D., 2007: The City of Iqaluit’s Climate Change Impacts, Infrastructure Risks
and Adaptive Capacity Project. City of Iqaluit, Canada.
Nixdorf-Miller, A., D.M. Hunsaker, and J.C. Hunsaker III, 2006: Hypothermia and
hyperthermia medicolegal investigation of morbidity and mortality from exposure
to environmental temperature extremes. Archives of Pathology & Laboratory
Medicine, 130, 1297-1304.
Noji, E.K., 2000: The public health consequences of disasters. Prehospital and
Disaster Medicine, 15(4), 147-157.
Noss, R.F., J.F. Franklin, W.L. Baker, T. Schoennagel, and P.B. Moyle, 2006: Managing
fire-prone forests in the western United States. Frontiers in Ecology and
Environment, 4, 481-487.
NRC, 2003: Fair Weather: Effective Partnerships in Weather and Climate Services.
National Research Council, National Academies Press, Washington, DC.
NRCAN
, 2010: Adaptation case studies: Iqaluit’s Sustainable Subdivision. Natural
Resources Canada, Ottawa, Canada, adaptation.nrcan.gc.ca/case/iqaluit_e.php.
NRCAN, 2011a: Geological Survey of Canada. Natural Resources Canada, Ottawa,
Canada, gsc.nrcan.gc.ca/permafrost/whatis_e.php.
NRCAN, 2011b: Regional Adaptation Collaboratives: The Nunavut Mining Sector and
Climate Change: Risks, Vulnerabilities and Good Environmental Practices
Project. Natural Resources Canada, adaptation.nrcan.gc.ca/collab/no_e.php.
NRCAN, 2011c: Nunavut Climate Change Partnership. Natural Resources Canada,
Ottawa, Canada, www.nrcan.gc.ca/com/elements/issues/49/panpcn-eng.php.
NRTEE, 2009: True North: Adapting Infrastructure to Climate Change in Northern
Canada. National Round Table on the Environment and the Economy of
Canada, Ottawa, Canada, 160 pp.
NSO, 2011: Mongolian Statistical Yearbook, 2010. National Statistical Office of
Mongolia (NSO), Ulaanbaatar, Mongolia, 463 pp.
Nuclear Claims Tribunal, Republic of the Marshall Islands, 2007: Agreement
between the Government of the United States and the Government of the
Marshall Islands for the Implementation of Section 177 of the Compact of Free
Association. www.nuclearclaimstribunal.com/177text.htm.
Nunes, M.C.S., M.J. Vasconcelos, J.M.C. Pereira, N. Dasgupta, R.J. Alldredge, and F.C.
Rego, 2005: Land cover type and fire in Portugal: do fires burn land cover
selectively? Landscape Ecology, 20, 661-673.
Nurse, L. and G. Sem, 2001: Small island states. In: Climate Change 2001: Impacts,
Adaptation, and Vulnerability. Contribution of Working Group II to the Third
Assessment Report of the Intergovernmental Panel on Climate Change
[McCarthy, J.J., O.F. Canziani, N.A. Leary, D.J. Dokken, and K.S. White (eds.)].
Cambridge University Press, Cambridge, UK, 834-875.
Chapter 9 Case Studies
538
O’Neill, M.S., R. Carter, J.H. Kish, C.J. Gronlund, J.L. White-Newsome, X. Manarolla, A.
Zanobetti, and J.D. Schwartz, 2009: Preventing heat-related morbidity and
mortality: new approaches in a changing climate. Maturitas, 64(2), 98-103.
Oswald Spring, Ú., 2011: Social vulnerability, discrimination, and resilience-building
in disaster risk reduction. In: Coping with Global Environmental Change,
Disasters and Security – Threats, Challenges, Vulnerabilities and Risks [Brauch,
H.G., U. Oswald Spring, C. Mesjasz, J. Grin, P. Kameri-Mbote, B. Chourou, P.
Dunay, and J. Birkmann (eds.)]. Springer Verlag, Berlin, Germany, pp. 1169-1188.
Oyun, R. (ed.), 2004: Impact of Current Climate Hazards on the Livelihoods of the
Herders’ Households. United Nations Development Programme, Poverty
Research Group, JEMR Consulting, Ulaanbaatar, Mongolia, 87 pp.
Palmer,T.N., F.J. Doblas-Reyes,A. Weisheimer, and M.J. Rodwell, 2008: Toward seamless
prediction: calibration of climate change projections using seasonal forecasts.
Bulletin of the American Meteorological Society, 89, 459-470.
Parker, D.J., S.J. Priest and S.M. Tapsell, 2009: Understanding and enhancing the
public’s behavioural response to flood warning information. Meteorological
Applications, 16, 103-114.
Parliament of Victoria, 2009: 2009 Victorian Bushfires Royal Commission –
Interim Report. No. 225 – Session 2006–09, Government Printer for the State
of Victoria, Melbourne, Australia.
Parliament of Victoria, 2010a: 2009 Victorian Bushfires Royal Commission, Final
Report, Volume 1. Government Printer for the State of Victoria, Melbourne,
Australia, www.royalcommission.vic.gov.au/Commission-Reports/Final-Report/
Volume-1.
Parliament of Victoria, 2010b: 2009 Victorian Bushfires Royal Commission, Final
Report, Volume 2. Government Printer for the State of Victoria, Melbourne,
Australia, www.royalcommission.vic.gov.au/Commission-Reports/Final-Report/
Volume-2.
Parliament of Victoria, 2010c: 2009 Victorian Bushfires Royal Commission,
Summary Report, Volume 2. Government Printer for the State of Victoria,
Melbourne, Australia, www.royalcommission.vic.gov.au/Commission-Reports/
Final-Report/Summary.
Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. van der Linden, and C.E. Hanson (eds.),
2007: Climate Change 2007: Impacts, Adaptation and Vulnerability.
Contribution of Working Group II to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change. Cambridge University Press,
Cambridge, UK, 1000 pp.
Pascual, M., M.J. Bouma, and A.P. Dobson, 2002: Cholera and climate: revisiting the
quantitative evidence. Microbes & Infection, 4(2), 237-245.
Patt, A., P. Suarez, and C. Gwata, 2005: Effects of seasonal climate forecasts and
participatory workshops among subsistence farmers in Zimbabwe. Proceedings
of the National Academy of Sciences, 102(35), 12623-12628.
Patz, J., D. Engelberg, and J. Last, 2000: The effects of changing weather on public
health. Annual Review of Public Health, 21, 271-307.
Paul, A. and M. Rahman, 2006: Cyclone mitigation perspectives in the islands in
Bangladesh: a case study of Swandip and Hatia Islands. Coastal Management,
34(2), 199-215.
Paul, B.K., 2009: Why relatively fewer people died? The case of Bangladesh cyclone
Sidr. Natural Hazards, 50
, 289-304.
Paul, B.K. and S. Dutt, 2010: Hazard warnings and responses to evacuation orders:
the case of Bangladesh’s cyclone Sidr. Geographical Review, 100(3), 336-355.
Pearce, T., J. Ford, G. Laidler, B. Smit, F. Duerden, M. Allarut, M. Andrachuk, S. Baryluk,
A. Dialla, P. Elee, A. Goose, T. Ikummaq, E. Joamie, F. Kataoyak, E. Loring, S. Meakin,
S. Nickels, K. Shappa, J. Shirley, and J. Wandel, 2009: Community collaboration and
climate change research in the Canadian Arctic. Polar Research, 28, 10-27.
Peduzzi, P., B. Chatenoux, H. Dao, C. Herold, and G. Giuliani, 2011: Preview Global
Risk Data Platform. UNEP/GRID and UNISDR, Geneva, Switzerland,
preview.grid.unep.ch/index.php?preview=tools&cat=1&lang=eng.
Peduzzi, P., U. Deichmann, A. Maskrey, F.A. Nadim, H. Dao, B. Chatenoux, C. Herold,
A. Debono, G. Giuliani, and S. Kluser, 2009: Global disaster risk: patterns, trends
and drivers. In: Global Assessment Report on Disaster Risk Reduction. United
Nations International Strategy for Disaster Reduction, Geneva, Switzerland, pp.
17-57.
Pelling, M. and A. Holloway, 2006: Legislation for Mainstreaming Disaster Risk
Reduction. Tearfund, Teddington, UK, 36 pp.
Pelling, M., A. Ozerdem, and S. Barakat, 2002: The macro-economic impact of disasters.
Progress in Development Studies, 2(4), 283-305.
Peng, R.D., J.F. Bobb, C. Tebaldi, L. McDaniel, M.L. Bell, and F. Dominici, 2010: Toward
a quantitative estimate of future heat wave mortality under global climate
change. Environmental Health Perspectives, 119(5), 701-706.
Pernetta, J.C., 1992: Impacts of climate change and sea-level rise on small island
states: National and international responses. Global Environmental Change,
2(1), 19-31.
Pirard, P., S. Vandentorren, M. Pascal, K. Laaidi, A. Le Tertre, S. Cassadou, and M.
Ledrans, 2005: Summary of the mortality impact assessment of the 2003 heat
wave in France. Euro Surveillance, 10(7), pii=554.
Poolman, E., H. Chikoore, and F. Lucio, 2008: Public benefits of the Severe Weather
Forecasting Demonstration Project in southeastern Africa. Meteo World, Dec 2008,
www.wmo.int/pages/publications/meteoworld/archive/dec08/swfdp_en.html.
Poumadere, M., C. Mays, S. LeMer, and R. Blong, 2005: The 2003 heat wave in
France: dangerous climate change here and now. Risk Analysis, 25, 1483-1494.
ProAct Network, 2008: The Role of Environmental Management and eco-engineering
in Disaster Risk Reduction and Climate Change Adaptation. ProAct Network,
Gaia Group, Finnish Ministry of Environment, and the United Nations
International Strategy for Disaster Reduction Secretariat, Switzerland,
www.preventionweb.net/english/professional/publications/v.php?id=4148.
Pruzzo, C., L. Vezzulli, and R.R. Colwell, 2008: Global impact of Vibrio cholerae inter-
actions with chitin. Environmental Microbiology, 10(6), 1400-1410.
PSIDS, 2009: Fiji, Marshall Islands, Micronesia (Federated States of), Nauru, Palau,
Papua New Guinea, Samoa, Solomon Islands, Tonga, Tuvalu, Vanuatu: Views on
the Possible Security Implications of Climate Change to be included in the
report of the Secretary-General to the 64th Session of the United Nations
General Assembly, Pacific Small Island Developing States, Permanent Mission of
the Republic of Nauru to the United Nations, New York, NY, www.un.org/
esa/dsd/resources/res_pdfs/ga-64/cc-inputs/PSIDS_CCIS.pdf.
Quarless, D., 2007: Addressing the vulnerability of SIDS. Natural Resources Forum,
31(2007), 99-101.
Ranger, N., S. Hallegatte, S. Bhattacharya, M. Bachu, S. Priya, K. Dhore, F. Rafique, P.
Mathur, N. Naville, F. Henriet, C. Herweijer, S. Pohit, and J. Corfee-Morlot, 2011:
An assessment of the potential impact of climate change on flood risk in
Mumbai. Climatic Change, 104, 139-167.
Read, R., 2010: Economic Vulnerability and Resilience in Small Island Developing
State. Global Platform on Climate Change, Trade and Sustainable Energy,
International Centre for Trade and Sustainable Development, Geneva,
Switzerland.
Reale, O., W.K. Lau, J. Susskind, R. Rosenberg, E. Brin, E. Liu, L.P. Riishojgaard, M.
Fuentes, and R. Rosenberg, 2009: AIRS impact on the analysis and forecast
track of tropical cyclone Nargis in a global data assimilation and forecasting
system. Geophysical Research Letters, 36, L06812, doi:10.1029/2008GL037122.
Reason, C.J.C. and A. Keibel, 2004: Tropical Cyclone Eline and its unusual penetration
and impacts over the Southern African mainland. Weather and Forecasting, 19,
789-805.
Reid, C.E., M.S. O’Neill, C.J. Gronlund, S.J. Brines, D.G. Brown, A.V. Diez-Roux, and J.
Schwartz, 2009: Mapping community determinants of heat vulnerability,
Environmmental Health Perspectives, 117, 1730-1736.
Republic of Mozambique, 2001: The National Action Plan for the Reduction of
Absolute Poverty, 2001-2005. Government of Mozambique, Maputo,
Mozambique.
Republic of Mozambique, 2006a: Master Plan for Prevention and Mitigation of
Natural Disasters. Council of Ministers, Republic of Mozambique, Maputo,
Mozambique.
Republic of Mozambique, 2006b: Action Plan for the Reduction of Absolute
Poverty, 2006-2009 (PARPA II). Republic of Mozambique, Maputo,
Mozambique.
Republic of South Africa, 2002: Disaster Management Act, 2002. Act No. 57 of
2002. Government Printer, Pretoria, South Africa.
Chapter 9Case Studies
539
Republic of South Africa, 2005: National Disaster Management Policy
Framework. Government Printer, Pretoria, South Africa.
Republic of the Philippines, 2009: Climate Change Act of 2009. Republic Act
9729, Fourteenth Congress of the Philippines, Manila, Philippines.
Republic of the Philippines, 2010: National Disaster Risk Reduction and
Management Act. Republic Act 10121, Fourteenth Congress of the Philippines,
Manila, Philippines.
Reuveny, R., 2007: Climate change-induced migration and violent conflict. Political
Geography, 26(6), 656-673.
Revi, A., 2005: Lessons from the deluge: Priorities for multi-hazard risk mitigation.
Economic and Political Weekly, 40(36), 3911-3916.
Revi, A., 2008: Climate change risk: an adaptation and mitigation agenda for Indian
cities. Environment and Urbanization, 20(1), 207-229.
Richardson, G.R.A., 2010: Adapting to Climate Change: An Introduction for
Canadian Municipalities. Natural Resources Canada, Ottawa, Canada, 40 pp.
Robine, J.M., S.L. Cheung, S. Le Roy, H. Van Oyen, C. Griffiths, J.P. Michel, and F.R.
Herrmann, 2008: Death toll exceeded 70,000 in Europe during the summer of
2003. Comptes Rendus Biologies, 331(2), 171-178.
Rodó, X., M. Pascual, G. Fuchs, and A. Faruque, 2002: ENSO and cholera: a
nonstationary link related to climate change? Proceedings of the National
Academy of Sciences, 99, 12901-12906.
Royal Society, 2008: Ground-Level Ozone in the 21st Century: Future Trends,
Impacts and Policy Implications. RS Policy document 15/08, Royal Society,
London, UK.
Rubio, J.C.M., I.J.M Pérez, J.J. Criado-Álvarez, C. Linares, and J.D. Jiménez, 2010: Heat
health warning systems. Possibilities of improvement. Revista Espanola de
Salud Publica, 84(2), 137-149.
Sack, D.A., R.B. Sack, and C.-L. Chaignat, 2006: Getting serious about cholera. New
England Journal of Medicine, 355(7), 649-651.
SADC, 2000: Revised Protocol on Shared Watercourses. Southern African
Development Community, Gabarone, Botswana.
Saenger, P. and N.A. Siddiqi, 1993: Land from the sea: the mangrove afforestation
program of Bangladesh. Ocean and Coastal Management, 20(1), 23-39.
Sailor, D.J. and A.A. Pavlova, 2003: Air conditioning market saturation and long-term
response of residential cooling energy demand to climate change. Energy,
28(9), 941-951.
Saizen, I., A. Maekawa, and N. Yamamura, 2010: Spatial analysis of time-series
changes in livestock distribution by detection of local spatial associations in
Mongolia. Applied Geography,
30, 639-649.
Santamouris, M., N. Papanikolaou, I. Koronakis, I. Livada, and D. Assimakopoulos,
1999: Thermal and airflow characteristics in a deep pedestrian canyon, under
hot weather conditions. Atmospheric Environment, 33, 4503-4521.
SARPMETT, 2010: Syrian Economy Bulletin. Syrian Arab Republic Prime Ministry
Economic Technical Team, Damascus, Syria, zunia.org/uploads/media/knowledge/
Syrian%20Economy%20Bulletin%20EN1284201253.pdf.
Schubert, S.D., M.J. Suarez, P.J. Pegion, R.D. Koster, and J.T. Bacmeister, 2008:
Potential predictability of long-term drought and pluvial conditions in the
United States Great Plains. Journal of Climate, 21, 802-816.
SCSCDD, 2008a: Percepção de risco, a descoberta de um novo olhar. Santa Catarina
State Civil Defence Department, Florianópolis, Brazil. EIRD Informa, 15, 2 pp.
SCSCDD, 2008b: O Projeto: Percepção de risco, a descoberta de um novo olhar.
Santa Catarina State Civil Defence Department, Florianópolis, Brazil, 2 pp.,
www.percepcaoderisco.sc.gov.br/?ver=projeto.
SCSCDD, 2008c: Finalizada capacitação sobre percepção de risco para comunidades
em Florianópolis. Notícias,18 December, Santa Catarina State Civil Defence
Department, Florianópolis, Brazil, 1 pp., www.percepcaoderisco.sc.gov.br/
?ver=noticia-completa&noticia=49.
SCSCDD, 2008d: Morro da penitenciária tem ação comunitária sobre riscos locais.
Notícias, 9 October, Santa Catarina State Civil Defence Department,
Florianópolis, Brazil, 1 pp.
Semenza, J.C., 2005: Building healthy cities: A focus on interventions. In: Handbook
of Urban Health [Galea, S. and D. Vlahov (eds.)]. Springer Science and Business
Media, New York, NY.
Semenza, J.C., T. March, and B. Bontempo, 2007: Community-initiated urban
development: an ecological intervention. Journal of Urban Health, 84(1), 8-20.
Semenza, J., D.J. Wilson, J. Parra, B.D. Bontempo, M. Hart, D.J. Sailor, and L.A. George,
2008: Public perception and behavior change in relationship to hot weather
and air pollution. Environmental Research, 107(3), 401-411.
Shapiro, M.A., J. Shukla, B. Hoskins, K. Trenberth, M. Beland, G. Brasseur, M. Wallace,
G. McBean, J. Caughey, D. Rogers, G. Brunet, L. Barrie, A. Henderson-Sellers, D.
Burridge, T. Nakazawa, M. Miller, P. Bougeault, R. Anthes, Z. Toth, and T. Palmer,
2007: The socio-economic and environmental benefits of a revolution in weather,
climate and Earth-system analysis and prediction. In: The Full Picture, Group on
Earth Observations, Tudor Rose, London, UK, pp. 137-139.
Shapiro, M.A., J. Shukla, G. Brunet, C. Nobre, M. Béland, R. Dole, K. Trenberth, R.
Anthes, G. Asrar, L. Barrie, P. Bougeault, G. Brasseur, D. Burridge, A. Busalacchi,
J. Caughey, D. Chen, J. Church, T. Enomoto, B. Hoskins, Ø. Hov, A. Laing, H. Le
Treut, J. Marotzke, G. McBean, G. Meehl, M. Miller, B. Mills, J. Mitchell, M.
Moncrieff, T. Nakazawa, H. Olafsson, T. Palmer, D. Parsons, D. Rogers, A.
Simmons, A. Troccoli, Z. Toth, L. Uccellini, C. Velden, and J.M. Wallace, 2010: An
earth-system prediction initiative for the 21st century. Bulletin of the American
Meteorological Society, 91, 1377-1388.
Sharma, D. and S. Tomar, 2010: Mainstreaming climate change adaptation in Indian
cities. Environment and Urbanization, 22(2), 451-465.
Sharma, U., A. Patwardhan, and D. Parthasarathy, 2009: Assessing adaptive capacity
to tropical cyclones in the east coast of India: a pilot study of public response
to cyclone warning information. Climatic Change, 94(1-2), 189-209.
Shaw, R., K. Shiwaku, H. Kobayashi, and M. Kobayashi, 2004: Linking experience,
education, perception and earthquake preparedness. Disaster Prevention and
Management
, 13(1), 39-49.
Sheridan, S.C., 2003: Heat, mortality, and level of urbanisation. Climate Research,
24, 255-265.
Sheridan, S.C., 2006: A survey of public perception and response to heat warnings
across four North American cities: an evaluation of municipal effectiveness.
International Journal of Biometeorology, 52, 3-15.
Shikanga, O.T., D. Mutonga, M. Abade, S. Amwayi, M. Ope, H. Limo, E.D. Mintz, R.E.
Quick, R.F. Breiman, and D.R. Feikin, 2009: High mortality in a cholera outbreak
in western Kenya after post-election violence in 2008. American Journal of
Tropical Medicine & Hygiene, 81(6), 1085-1090.
Shimoda, Y., 2003: Adaptation measures for climate change and the urban heat island
in Japan’s built environment. Building Research & Information, 31(3-4), 222-230.
Shiwaku, K., R. Shaw, C. Kandel, S.N. Shrestha, and A.M. Dixit, 2006: Promotion of
disaster education in Nepal: role of teachers as change agents. International
Journal of Mass Emergency and Disaster, 24(3), 403-420.
Shukla, J., R. Hagedorn, B. Hoskins, J. Kinter, J. Marotzke, M. Miller, T. Palmer, and J.
Slingo, 2009: Revolution in climate prediction is both necessary and possible: a
declaration at the world modelling summit for climate prediction. Bulletin of
the American Meteorological Society, 90, 16-19.
Shukla, J., T.N. Palmer, R. Hagedorn, B. Hoskins, J. Kinter, J. Marotzke, M. Miller, and J.
Slingo, 2010: Towards a new generation of world climate research and computing
facilities. Bulletin of the American Meteorological Society, 91, 1407-1412.
Shultz, A., J.O. Omollo, H. Burke, M. Qassim, J.B. Ochieng, M. Weinberg, D.R. Feikin,
and R.F. Breiman, 2009: Cholera outbreak in Kenyan refugee camp: risk factors
for illness and importance of sanitation. American Journal of Tropical Medicine
& Hygiene, 80(4), 640-645.
Sidley, P., 2000; Malaria epidemic expected in Mozambique. BMJ, 320, 669.
Sietz, D., M. Boschütz, R.J.T. Klein, and A. Lotsch, 2008: Mainstreaming Climate
Adaptation into Development Assistance in Mozambique: Institutional Barriers
and Opportunities. Policy Research Working Paper 4711, World Bank,
Washington, DC.
Silva, H.R., P.E. Phelan, and J.S. Golden, 2010: Modeling effects of urban heat island
mitigation strategies on heat-related morbidity: a case study for Phoenix,
Arizona, USA. International Journal of Biometeorology, 54, 13-22.
Simmons, A.J. and A. Hollingsworth, 2002: Some aspects of the improvement in skill
of numerical weather prediction. Quarterly Journal of the Royal Meteorological
Society, 128, 647-677.
Chapter 9 Case Studies
540
Skaff, M. and Sh. Masbate, 2010: Precipitation change, and its potential effects on
vegetation and crop productivity in Syrian Al Jazerah Region. The Arab Journal
for Arid Environments, 3(2), 71-78.
Skees, R. and A. Enkh-Amgalan, 2002: Examining the Feasibility of Livestock
Insurance in Mongolia. Policy Research Working Paper 2886, World Bank,
Washington, DC.
Smith, S.K. and C. McCarty, 2009: Fleeing the storm(s): an examination of evacuation
behavior during Florida’s 2004 hurricane season. Demography, 46(1), 127-145.
Smithers, J.C., R.E. Schulze, A. Pike, and G.P.W. Jewitt, 2001: A hydrological perspective
of the February 2000 floods: a case study in the Sabie River Catchment. Water
SA, 27(3), 25-31.
Smoyer, K.E., D.G. Rainham, and J.N. Hewko, 2000: Heat stress-related mortality in
five cities in Southern Ontario: 1980-1996. International Journal of
Biometeorology, 44(4), 190-197.
Smoyer-Tomic, K.E. and D.G.C. Rainham, 2001: Beating the heat: Development and
evaluation of a Canadian hot weather health-response plan. Environmental
Health Perspectives, 109(12), 1241-1248.
Solh, M., 2010: Tackling the drought in Syria. Nature Middle East, doi:10.1038/
nmiddleeast.2010.206.
Sowers, J. and E. Weinthal, 2010: Climate Change Adaptation in the Middle East and
North Africa: Challenges and Opportunities. Working Paper No. 2, Dubai
Initiative, Dubai School of Government, Dubai, United Arab Emirates,
belfercenter.ksg.harvard.edu/files/Sowers-Weinthal_DI_Working-Paper-2.PDF.
Sowers, J., A. Vengosh, and E. Weinthal, 2011: Climate change, water resources, and
the politics of adaptation in the Middle East and North Africa. Climatic Change,
104(3-4), 599-627.
Sphere Project, 2004: Humanitarian Charter and Minimum Standards in Disaster
Response. The Sphere Project, Geneva, Switzerland.
Srivastava, A.K, M.M. Dandekar, S.R. Kshirsagar, and S.K. Dikshit, 2007: Is summer
becoming more uncomfortable at Indian Cities? Mausam, 58(3), 335-344.
Stal, M., 2011. Flooding and relocation: The Zambezi River Valley in Mozambique.
International Migration, 49(S1), e125-e145.
Stedman, J.R., 2004: The predicted number of air pollution related deaths in the UK
during the August 2003 heatwave. Atmospheric Environment, 38, 1087-1090.
Stein, R.M., L. Dueñas-Osorio, and D. Subramanian, 2010: Who evacuates when
hurricanes approach? The role of risk, information, and location. Social Science
Quarterly, 91(3), 816-834.
Sternberg, T., 2010: Unravelling Mongolia’s extreme winter disaster of 2010.
Nomadic Peoples, 14(1), 72-86.
Stewart, B.I., A.F. Asfary, A. Belloum, K. Steiner, and T. Friedrich, 2008: Conservation
Agriculture for Sustainable Land Management to Improve the Livelihood of
People in Dry Areas. Arab Center for the Studies of Arid Zones and Dry Lands
and the German Agency for Technical Cooperation, Damascus, Syria and
Eschborn, Germany, www.fao.org/ag/ca/doc/CA%20Workshop%20procedding
%2008-08-08.pdf.
Stott, P.A., D.A. Stone, and M.R. Allen, 2004: Human contribution to the European
heat wave of 2003. Nature, 432, 610-614.
Suarez, P. and J. Linnerooth-Bayer, 2010: Micro-insurance for local adaptation. Wiley
Interdisciplinary Reviews: Climate Change, 1(2), 271-278.
Subbarao, I., N.A. Bostick, and J.J. James, 2008: Applying yesterday’s lessons to today’s
crisis: Improving the utilization of recovery services following catastrophic
flooding. Disaster Medicine and Public Health Preparedness, 2(3), 132-133.
Sur, D., A.L. Lopez, S. Kanungo, A. Paisley, B. Manna, M. Ali, S.K. Niyogi, J.K. Park, B.
Sarkar, M.K. Puri, D.R. Kim, J.L. Deen, J. Holmgren, R. Carbis, R. Rao, T.V. Nguyen,
A. Donner, N.K. Ganguly, G.B. Nair, S.K. Bhattacharya, and J.D. Clemens, 2009:
Efficacy and safety of a modified killed-whole-cell oral cholera vaccine in India:
an interim analysis of a cluster-randomised, double-blind, placebo-controlled
trial. Lancet, 374(9702), 1694-1702.
Surminski, S., 2010: Adapting to the Extreme Weather Impacts of Climate Change –
How Can the Insurance Industry Help? Climate Wise, London, UK.
Syphard, A.D., V.C. Radeloff, T.J. Hawbaker, and S.I. Stewart, 2009: Conservation
threats due to human-caused increases in fire frequency in Mediterranean-
climate ecosystems. Conservation Biology, 23(3), 758-769.
Talavera, A. and E.M. Perez, 2009: Is cholera disease associated with poverty.
Journal of Infection in Developing Countries, 3(6), 408-411.
TDB, 2007: “Structurally weak, vulnerable and small economies”: Who are they?
What can UNCTAD do for them? Trade and Development Board, United Nations
Conference on Trade and Development, Geneva, Switzerland, www.unctad.org/
en/docs/tdb54crp4_en.pdf.
Thomson, M.C., F.J. Doblas-Reyes, S.J. Mason, R. Hagedorn, S.J. Connor, T. Phindela,
A.P. Morse, and T.N. Palmer, 2006: Malaria early warnings based on seasonal
climate forecasts from multi-model ensembles. Nature, 439, 576-579.
Tolhurst, K., 2010: Report on Fire Danger Ratings and Public Warning. The University
of Melbourne, Melbourne, Australia.
Twigg, J. and H. Bottomley, 2011: Disaster Risk Reduction NGO Inter-Agency Group
Learning Review. Action Aid, Christian Aid, Plan International, Practical Action,
and Tearfund, London, UK, 40 pp.
UN, 2005: Report of International Meeting to Review the Implementation of the
Programme of Action for the Sustainable Development of Small Island
Developing States, Port Louis, Mauritius, 10-14 January 2005, United Nations,
New York, NY, www.sidsnet.org/docshare/other/20050622163242_English.pdf.
UN, 2006: Global Survey of Early Warning Systems. Prepared by UN International
Strategy for Disaster Reduction for the United Nations, United Nations, Geneva,
Switzerland, 46 pp.
UN, 2008: World Urbanization Prospects – The 2007 Revision. United Nations
Department of Economic and Social Affairs, Population Division, New York, NY,
www.un.org/esa/population/publications/wup2007/2007WUP_Highlights_
web.pdf.
UN, 2009: Syria Drought Response Plan. United Nations, New York, NY,
reliefweb.int/sites/reliefweb.int/files/resources/2A1DC3EA365E87FB8525760F0
051E91A-Full_Report.pdf.
UN, 2011: General Assembly Report of the Special Rapporteur on the right to
food, Olivier De Schutter. Mission to the Syrian Arab Republic. United Nations,
New York, NY, www2.ohchr.org/english/bodies/hrcouncil/docs/16session/
A.HRC.16.49.Add.2_en.pdf.
UNCTAD, 2004: Is a Special Treatment of Small Island Developing States Possible?
United Nations Conference on Trade and Development, New York, NY,
www.unctad.org/en/docs/ldc20041_en.pdf.
UNDESA, 2010: Trends in Sustainable Development: Small Island Developing States.
Division for Sustainable Development, United Nations Department of Economic
and Social Affairs, New York, NY, www.cbd.int/islands/doc/sids-trends-report-
v4-en.pdf.
UNDESA, 2011: Case Study Detail Record, Marshall Islands: Groundwater Supply on
the Marshall Islands. Division for Sustainable Development, United Nations
Department of Economic and Social Affairs, New York, NY, webapps01.un.org/
dsd/caseStudy/public/displayDetailsAction.do?code=326.
UNDP, 2005: A Global Review: Support to Institutional and Legislative Systems for
Disaster Risk Management. United Nations Development Programme, New
York, NY.
UNDP, 2006: Country Programme Action Plan. 2006 – 2009. United Nations
Development Programme, New York, NY, www.bb.undp.org/uploads/file/pdfs/
general/UNDP-GRN%20CPAP%202006-2009.pdf.
UNDP, 2007: A Global Review: UNDP Support to Institutional and Legislative Systems
for Disaster Risk Management. United Nations Development Programme,
Bureau for Crisis Prevention and Recovery, New York, NY.
UNDP Mongolia, 2009a: The Millennium Development Goals Implementation, Third
National Report. UNDP, Ulaanbaatar, Mongolia, 142 pp.
UNDP Mongolia, 2009b: Moving Forward to Reduce Climate Change Risks. United
Nations Development Programme, Ulaanbaatar, Mongolia, www.undp.mn/
news-snrm-230409.html.
UNDP Mongolia, 2010: Project Report: Support to the Dzud in Mongolia/early
recovery. United Nations Development Programme, Ulaanbaatar, Mongolia,
www.undp.mn/pprojects.html.
UNEP, 2004: Impacts of Summer 2003 Heat Wave in Europe. Environment Alert
Bulletin 2, United Nations Environment Programme, Geneva, Switzerland,
www.grid.unep.ch/product/publication/download/ew_heat_wave.en.pdf.
Chapter 9Case Studies
541
UNESC, 2000: Assistance to Mozambique following the devastating floods: Report of
the Secretary-General. A/55/123-E/2000/89, United Nations General Assembly
Economic and Social Council, New York, NY, www.un.org/documents/ga/docs/
55/a55123.pdf.
UNESC, 2004: Freshwater management: progress in meeting the goals, targets and
commitments of Agenda 21, the Programme for the Further Implementation of
Agenda 21, and the Johannesburg Plan of Implementation. Report of the
Secretary-General, A/CONF.207/3, United Nations Economic and Social Council,
New York, NY, www.preventionweb.net/files/resolutions/N0423881.pdf.
UNESC, 2011: Integrated Analysis of United Nations System Support to Small Island
Developing States. Report of the Secretary-General,E/2011/110, United Nations
Economic and Social Council, New York, NY, www.un.org/ga/search/view_doc.asp
?symbol=E/2011/110&referer=http://sids-l.iisd.org/news/advance-report-of-
secretary-general-calls-for-improved-msi-delivery/?referrer=linkages-
iisdrs&Lang=E.
UNFCCC, 2007a: Vulnerability and Adaptation To Climate Change In Small Island
Developing States. United Nations Framework Convention on Climate Change,
Bonn, Germany, unfccc.int/files/adaptation/adverse_effects_and_response_
measures_art_48/application/pdf/200702_sids_adaptation_bg.pdf.
UNFCCC, 2007b: Climate Change: Impacts, Vulnerabilities and Adaptation in
Developing Countries, United Nations Framework Convention on Climate
Change, Bonn, Germany, unfccc.int/resource/docs/publications/impacts.pdf.
UNFCCC, 2010: Action On the Ground: A Synthesis of Activities in the Areas of
Education, Training and Awareness-Raising For Adaptation. Nairobi Work
Programme on Impacts, Vulnerability and Adaptation to Climate Change, United
Nations Framework Convention on Climate Change, Bonn, Germany, 116 pp.
Ungern-Sternberg, T., 2003: State Intervention on the Market for Natural Damage
Insurance in Europe. CESifo Working Paper 1067, CESifo Group, Munich, Germany.
UN-HABITAT, 2006: State of the World’s Cities 2006/2007. United Nations Human
Settlements Programme, Nairobi, Kenya.
UNISDR, 2004: Living with Risk: A Global Review of Disaster Reduction Initiatives.
United Nations International Strategy for Disaster Reduction, Geneva,
Switzerland, 429 pp.
UNISDR, 2005a: Review of the Yokohama Strategy and Plan of Action for a Safer
World. Conference document A/CONF.206/L.1. In: Proceedings of the World
Conference on Disaster Reduction, Kobe, Hyogo, Japan, 18-22 January 2005,
United Nations International Strategy for Disaster Reduction, Geneva,
Switzerland, 23 pp.
UNISDR, 2005b: Hyogo Framework for Action: Building the Resilience of Nations and
Communities to Disasters, 2005-2015. United Nations International Strategy
for Disaster Reduction, Geneva, Switzerland.
UNISDR, 2006: Disaster Risk Reduction Begins at School, 2006-2007 International
Disaster Risk Reduction Campaign. United Nations International Strategy for
Disaster Reduction, Geneva, Switzerland, www.unisdr.org/eng/public_aware/
world_camp/2006-2007/wdrc-2006-2007.htm.
UNISDR, 2007a: Drought Risk Reduction Framework and Practices: Contributing to
the Implementation of the Hyogo Framework for Action. United Nations
International Strategy for Disaster Reduction, Geneva, Switzerland, 98 pp.
UNISDR, 2007b: Towards a Culture of Prevention: Disaster Risk Reduction Begins at
School; Good Practices and Lessons Learned. United Nations International
Strategy for Disaster Reduction, Geneva, Switzerland, 143 pp.
UNISDR, 2008a: Climate Change and Disaster Risk Reduction. United Nations
International Strategy for Disaster Reduction, Geneva, Switzerland.
UNISDR, 2008b: Annual Report 2007: The Secretariat of the International Strategy
for Disaster Reduction. United Nations International Strategy for Disaster
Reduction, Geneva, Switzerland, 68 pp.
UNISDR, 2009a: Outcome Document: Chair’s Summary of the Second Session of the
Global Platform for Disaster Risk Reduction. Proceedings of a conference,
Global Platform for Disaster Risk Reduction, 16-19 June 2009, United Nations
International Strategy for Disaster Reduction, Geneva, Switzerland, 4 pp.
UNISDR, 2009b: Terminology on Disaster Risk Reduction, 2009. United Nations
International Strategy for Disaster Reduction, Geneva, Switzerland,
www.unisdr.org/we/inform/terminology.
UNISDR, 2009c: Global Assessment Report on Disaster Risk Reduction. International
Strategy for Disaster Reduction, Geneva, Switzerland, 207 pp.
UNISDR, 2010: Early Warning Practices can Save Many Lives: Good Practices and
Lessons Learned. United Nations Secretariat of the International Strategy for
Disaster Reduction, UN International Strategy on Disaster Risk Reduction,
Bonn, Germany, 67 pp., www.unisdr.org/files/15254_EWSBBLLfinalweb.pdf.
UNISDR, 2011a: Mid-Term Review 2010-2011: Hyogo Framework for Action 2005-2015:
Building the Resilience of Nations and Communities to Disasters. United Nations
International Strategy for Disaster Reduction, Geneva, Switzerland, 107 pp.
UNISDR, 2011b: Global Assessment Report on Disaster Risk Reduction: Revealing
Risk, Redefining Development. United Nations International Strategy for
Disaster Reduction, Geneva, Switzerland, www.preventionweb.net/english/
hyogo/gar/2011/en/home/index.html.
United Kingdom, 2008: Climate Change Act 2008. The Stationary Office Limited,
London, UK.
UNOCHA, 2007: Mozambique 2007 Flash Appeal, Executive Summary. United
Nations Office for the Coordination of Human Affairs, New York, NY,
ochaonline.un.org/humanitarianappeal/webpage.asp?Page=1558.
UN-OHRLLS, 2011: List of Small Island Developing States (UN Members). United
Nations Office of the High Representative for the Least Developed Countries,
Landlocked Developing Countries and the Small Island Developing States, New
York, NY, www.un.org/special-rep/ohrlls/sid/list.htm.
UNRCS and SARPCMSPC, 2005. Second National Report on the Millennium
Development Goals (MDGs) in the Syrian Arab Republic 2005. United Nations
Resident Coordinator System and Syrian Arab Republic Presidency of the
Council of Ministers State Planning Commission, Damascus, Syria,
planipolis.iiep.unesco.org/upload/Syrian%20AR/SyriaMDGSecondReport
English.pdf.
USAID, 2007: Mozambique – Floods and Cyclone. US Agency for International
Development, Washington, DC, www.usaid.gov/our_work/humanitarian_
assistance/disaster_assistance/countries/mozambique/template/fs_sr/
mozambique_fl_fs01_03-22-2007.pdf.
USARC, 2003: Climate Change, Permafrost and Impacts on Civil Infrastructure:
Permafrost Task Force Report. United States Arctic Research Commission,
Washington, DC.
USDA, 2008a: Middle East & Central Asia: Continued Drought in 2009/10 Threatens
Greater Food Grain Shortages. US Department of Agriculture
http://www.pecad.fas.usda.gov/highlights/2008/09/mideast_cenasia_drought/
USDA, 2008b: Syria: Wheat Production in 2008/09 Declines Owing to Season-Long
Drought. US Department of Agriculture, Washington, DC, www.pecad.fas.usda.gov/
highlights/2008/05/Syria_may2008.htm.
Van Niekerk, D., 2006: Disaster risk management in South Africa: The function and
the activity – towards an integrated approach. Politeia, 25(2), 95-115.
Van Niekerk, D., 2011: Concept Paper: The South African Disaster Risk Management
Policy and Legislation – A critique. African Centre for Disaster Studies,
Potchefstroom, South Africa, acds.co.za/uploads/research_reports/
SA_law_2011.pdf.
Van Riet, G. and M. Diedericks, 2010: The placement of the disaster management
function within district, metropolitan and provincial government structures in
South Africa. Administratio Publica, 18(4), 155-173.
Varangis, P., J.R. Skees, and B.J. Barnett, 2002: Weather indexes for developing
countries. In: Climate Risk and the Weather Market [Dischel, B. (ed.)]. Risk
Books, London, UK, pp. 279-294.
Vardoulakis, S., B.E.A. Fisher, K. Pericleous, and N. Gonzalez-Flesca, 2003: Modelling
air quality in street canyons: a review. Atmospheric Environment, 37, 155-182.
Verschuren, P.J.M., 2003: Case study as a research strategy: Some ambiguities and
opportunities.
International Journal of Social Research Methodology: Theory
and Practice, 6(2), 121-139.
Victorian Government, 2009: January 2009 Heatwave in Victoria: an Assessment
of Health Impacts. Victorian Government, Department of Human Services,
Melbourne, Australia.
Vingarzan, R., 2004: A review of surface ozone background levels and trends.
Atmospheric Environment, 38, 3431-3442.
Chapter 9 Case Studies
542
Visser, R. and D. Van Niekerk, 2009: A Funding Model for the Disaster Risk
Management Function of Municipalities. National Disaster Management
Centre, South African Department of Cooperative Governance and Traditional
Affairs, Pretoria, South Africa.
Vogt, F. and D.G. Sapir, 2009: Cyclone Nargis in Myanmar: lessons for public health
preparedness for cyclones. American Journal of Disaster Medicine, 4(5), 273-278.
Volz, A. and D. Kley, 1988: Evaluation of the Montsouris series of ozone measurements
made in the nineteenth century. Nature, 332, 240-242.
Vos, F., J. Rodriguez, R. Below, and D. Guha-Sapir, 2010: Annual Disaster Statistical
Review 2009: The Numbers and Trends. Centre for Research on the
Epidemiology of Disasters, Brussels, Belgium, www.cred.be/sites/default/
files/ADSR_2009.pdf.
Waechter, B., 2005: Working with Nature: Wind Deflection Fins in Action. Rowan
Williams Davies & Irwin Inc., Guelph, Canada, www.rwdi.com/cms/publications/
16/t05.pdf.
Walsh, R., S. Orban, R. Walker, J. Coates, J. Croteau, D. Stone, and T. Strynadks, 2009:
Front Street Paving Project, Dawson City, Yukon: Adapting to Climate Change in
a National Historic District. Yukon Highways and Public Works, Whitehorse,
Canada, www.colascanada.ca/uploads/colascanada/File/expertise/Environmental
AchievementYukonPaper.pdf.
Wang, W.-G., X.-R.Wang, Y.-L. Xu, Y.-H. Duan, and J.-Y. Li, 2008: Characteristics of
super-typhoon “SaoMai” and benefits of forecast service. Journal of Natural
Disasters, 17(3), 106-111.
Warner, K., N. Ranger, S. Surminski, M. Arnold, J. Linnnerooth-Bayer, E. Michel-Kerjan,
P. Kovacs, and C. Herweijer, 2009: Adaptation to Climate Change: Linking
Disaster Risk Reduction and Insurance. United Nations International Strategy
for Disaster Reduction, Geneva, Switzerland, 30pp.
Webster, P.J., 2008: Myanmar’s deadly daffodil. Nature Geoscience, 1, 488-490.
Webster, P.J., J. Jian. T.M. Hopson, C.D. Hoyos
, P.A. Agudelo, H.-R. Chang, J.A Curry,
R.I. Grossman, T.N. Palmer, and A.R. Subbiah
, 2010: Extended-range probabilistic
forecasts of Ganges and Brahmaputra floods in Bangladesh. Bulletin of the
American Meteorological Society, 91, 1493-1514.
Weisskopf, M.A., H.A. Anderson, S. Foldy, L.P. Hanrahan, K. Blair, T.J. Torok, and P.D.
Rumm, 2002: Heat wave morbidity and mortality, Milwaukee, Wis, 1999 vs 1995:
An improved response? American Journal of Public Health, 92(5), 830-833.
WHO, 2006a: Air Quality Guidelines: Global Update 2005, Particulate Matter, Ozone,
Nitrogen Dioxide And Sulphur Dioxide. World Health Organization Regional
Office for Europe, Copenhagen, Denmark.
WHO, 2006b: First meeting of the project: Improving Public Health Responses to
Extreme Weather/Heat-waves. EuroHEAT Report on a WHOMeeting in Rome,
Italy, 20–22 June 2005, World Health Organization Regional Office for Europe,
Copenhagen, Denmark.
WHO, 2006c: Cholera 2005. Weekly Epidemiologic Record, 81, 297-308.
WHO, 2007: Improving Public Health Responses to Extreme Weather/Heat-Waves –
EuroHEAT,
Technical Summary. World Health Organization Regional Office for
Europe, Copenhagen, Denmark.
WHO, 2008a: Cholera in Zimbabwe: Epidemiological Bulletin Number 1. World
Health Organization, Harare, Zimbabwe.
WHO, 2008b: Health System Problems Aggravate Cholera Outbreak in Zimbabwe.
News Release, World Health Organization, Geneva, Switzerland,
www.who.int/mediacentre/news/releases/2008/pr49/en/index.html.
WHO, 2010: Cholera vaccines: WHO position paper. Weekly Epidemiologic Record,
85(13), 117-128.
Wisner, B., 2006: Let our Children Teach Us! A Review of the Role of Education and
Knowledge in Disaster Risk Reduction. UNISDR System Thematic Cluster/
Platform on Knowledge and Education, Geneva, Switzerland, 135 pp.
WMO, 2011a: Climate Knowledge for Action: A Global Framework for Climate
Services – Empowering the Most Vulnerable. The Report of the High-Level
Taskforce for the Global Framework for Climate Services, WMO No. 1065, World
Meteorological Organization, Geneva, Switzerland.
WMO, 2011b: Tropical Cyclone Programme. Report to Plenary on item 4.3. World
Meteorological Organization, Geneva, Switzerland, www.hydrometeoindustry.org/
Reports2011/CONGRESS2011/d04-3_TROPICAL_CYCLONE.pdf.
Woodroffe, C., 2008: Reef-island topography and the vulnerability of atolls to sea-
level rise. Global and Planetary Change, 247(1-2), 159-177.
World Bank, 2005a: Learning Lessons from Disaster Recovery: The Case of
Mozambique. Disaster Risk Management Working Paper Series No. 12, World
Bank, Washington, DC.
World Bank, 2005b: Regional Engagement Framework FY 2006-2009 for Pacific Islands.
Report No. 32261-EAP, World Bank, Washington, DC, siteresources.worldbank.org/
INTPACIFICISLANDS/Resources/PI-Strategy-05.pdf.
World Bank, 2005c: Republic of Maldives, Tsunami Impact and Recovery. Joint
assessment, World Bank, Asian Development Bank, and UN System,
Washington, DC, www.adb.org/Documents/Reports/Tsunami/joint-needs-
assessment.pdf .
World Bank, 2005d: Managing Agricultural Production Risk. World Bank,
Washington, DC, siteresources.worldbank.org/INTARD/Resources/Managing_Ag_
Risk_FINAL.pdf.
World Bank, 2006a: Opportunities to Improve Social Services in the Republic of the
Marshall Islands. Summary Report, World Bank, Washington, DC, www-
wds.worldbank.org/external/default/WDSContentServer/WDSP/IB/2007/05/17/
000310607_20070517100419/Rendered/PDF/397780EAP0P0791Development
01PUBLIC1.pdf.
World Bank, 2006b: Ethiopia – Second Productive Safety Net APL Project. The World
Bank, Washington, DC, www-wds.worldbank.org/external/default/
main?pagePK=64193027&piPK=64187937&theSitePK=523679&menuPK=64
187510&searchMenuPK=64187283&siteName=WDS&entityID=000020953_2
0070117142245.
World Bank, 2007: The Caribbean Catastrophe Risk Insurance Initiative. Results of
Preparation Work on the Design of a Caribbean Catastrophe Risk Insurance
Facility. World Bank, Washington, DC.
World Bank, 2009: Reducing the Risk of Disasters and Climate Variability in the
Pacific Islands. Republic of the Marshall Islands Country Assessment. World
Bank, Washington, DC, siteresources.worldbank.org/INTPACIFICISLANDS/
Resources/MARSHALL_ASSESSMENT_f.pdf .
World Bank, 2011a: World Development Indicators Database. World Bank,
Washington, DC, data.worldbank.org/data-catalog/world-development-
indicators.
World Bank, 2011b: Disaster Risk Management Programs for Priority Countries.
Global Facility for Disaster Risk Reduction, Washington, DC.
World Bank, Global Facility for Disaster Reduction and Recovery, Inter-Agency
Network for Education in Emergencies (INEE), and UNISDR, 2009:
Guidance
Notes on Safer School Construction. Global Facility for Disaster Reduction and
Recovery and INEE Secretariat, Washington, DC and New York, NY, 133 pp.
Yip, F.Y., W.D. Flanders, A. Wolkin, D. Engelthaler, W. Humble, A. Neri, L. Lewis, L.
Backer, and C. Rubin, 2008: The impact of excess heat events in Maricopa
County, Arizona: 2000–2005. International Journal of Biometeorology, 52(8),
765-772.
Ykhanbai, H., E.U. Beket, R. Vernooy, and J. Graham, 2004: Reversing grassland
degradation and improving herders’ livelihoods in the Altai Mountains of
Mongolia. Mountain Research and Development, 24(2), 96-100.
Yohe, G.W., R.D. Lasco, Q.K. Ahmad, N.W. Arnell, S.J. Cohen, C. Hope, A.C. Janetos. and
R.T. Perez, 2007: Perspectives on climate change and sustainability. In: Climate
Change 2007. Impacts, Adaptation and Vulnerability. Contribution of Working
Group II to the Fourth Assessment Report of the Intergovernmental Panel on
Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. Van Der Linde, and
C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK, 811-841.
Yokoi, S. and Y.N. Takayabu, 2010: Environmental and external factors in the genesis
of tropical cyclone Nargis in April 2008 over the Bay of Bengal. Journal of the
Meteorological Society of Japan, 88, 425-435.
Yoshida, Y., Y. Takeuchi, and R. Shaw, 2009: Town watching as a useful tool in urban
risk reduction in Saijo. Community, Environment and Disaster Risk
Management, 1, 189-205.
Zuckerman, J., L. Rombo, and A. Fisch, 2007: The true burden and risk of cholera:
implications for prevention and control. Lancet Infectious Diseases, 7, 521-530.
Chapter 9Case Studies
IV Annexes I to IV
545
This annex should be cited as:
IPCC, 2012: Authors and expert reviewers annex. In: Managing the Risks of Extreme Events and Disasters to Advance Climate
Change Adaptation [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner,
S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups I and II of the Intergovernmental Panel on
Climate Change (IPCC). Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 545-553.
I
Authors and Expert Reviewers
ANNEX
546
Argentina
Vicente Barros, CIMA/Universidad de Buenos Aires
Ines Camilloni, CIMA/Universidad de Buenos Aires
Hernan Carlino, Universidad Torcuato Di Tella
Mario Nunez, CIMA/Universidad de Buenos Aires
Matilde Rusticucci, Universidad de Buenos Aires
Haris Eduardo Sanahuja, Senior Consultant
Pablo Suarez, Boston University, Red Cross/Red Crescent Climate Centre
Carolina Vera, CIMA/Universidad de Buenos Aires
Australia
Jonathan Abrahams, World Health Organization
Lisa Alexander, The University of New South Wales
Julie Arblaster, National Center for Atmospheric Research, Australian Bureau of
Meteorology
Jon Barnett, Melbourne University
Ian Carruthers, National Climate Change Adaptation Research Facility
Bob Cechet, Geoscience Australia
Lynda Chambers, Australian Bureau of Meteorology
John Church, Commonwealth Scientific and Industrial Research Organization
Paul Della-Marta, Partner Reinsurance Company
Amy Dumbrell, Australian Government Department of Climate Change and Energy
Efficiency
Jill Edwards, Australasian Fire and Emergency Service Authorities Council
Ailie Gallant, University of Melbourne
John Handmer, Centre for Risk and Community Safety, RMIT University
Mark Hemer, Commonwealth Scientific and Industrial Research Organization
Adriana Keating, RMIT University
Monique Ladds, RMIT University
Padma Narsey Lal, International Union for Conservation of Nature-Oceania
Yun Li, Commonwealth Scientific and Industrial Research Organisation
Kathleen McInnes, Commonwealth Scientific and Industrial Research Organization,
Marine and Atmospheric Research
Neville Nicholls, Monash University, School of Geography and Environmental Science
Lauren Amy Rickards, University of Melbourne
Anthony Swirepik, Australia Government Department of Climate Change and
Energy Efficiency
Frank Thomalla, Macquarie University
Richard Thornton, Bushfire Cooperative Research Centre
Blair Trewin, Australian Bureau of Meteorology
Anya M. Waite, University of Western Australia
Xiaoming Wang, Commonwealth Scientifc and Industrial Research Organisation
Penny Whetton, Commonwealth Scientific and Industrial Research Organisation
Joshua Whittaker, RMIT University
Austria
Stefan Hochrainer, International Institute for Applied System Analysis
Helmut Hojesky, Bundesministerium fur Land-und Forstwirtschaft, Umwelt und
Wasserwirtschaft
Klaus Radunsky, Umweltbundesamt GmbH
Petra Tschakert, Pennsylvania State University
Bangladesh
Md. Siarjul Islam, North South University
Tarik ul Islam, United Nations Development Programme-Bangladesh
Alimullah Miyan, South Asian Disaster Management Centre, International University
of Business, Agriculture, and Technology
Ainun Nishat, BRAC University
Ataur Rahman, Centre for Global Environmental Culture, International University of
Business, Agriculture, and Technology
Belgium
Johan Bogaert, Flemish Government
Lieven Bydekerke, VITO Flemish Institute for Technological Research
Cathy Clerbaux, Universite Libre de Bruxelles and CNRS France
Luc Feyen, Joint Research Centre, European Commission
Leen Gorissen, VITO – Flemish Institute for Technological Research
Julien Hoyaux, Agence Wallonne de l’Air et du Climat
Philippe Marbaix, Université Catholique de Louvain
Anne Mouchet, Universite de Liege
Andrea Tilche, European Commission Directorate
Hans van de Vyvere, Royal Meteorological Institute of Belgium
Jean-Pascal van Ypersele, Université Catholique de Louvain
Martine Vanderstraeten, Belgian Federal Science Policy
Patrick Willems, Katholieke Universiteit Leuven
Botswana
Pauline Dube, University of Botswana
Brazil
Andre Odenbreit Carvalho, Environmental Policy and Sustainable Development
Jose Marengo, National Institute for Space Research, Earth System Science Centre
Jose Domingos Gonzales Miguez, Tecnologia Commission for Global Climate Change
Carlos Nobre, Ministry of Science and Technology
Vinicius Rocha, Operador Nacional do Sistema Elétrico
Maria Assuncao Silva Dias, University of Sao Paulo
Canada
Heather Auld, Environment Canada
Anik Beaudoin, Foreign Affairs and International Trade Canada
Peter Berry, Climate Change and Health Office, Health Canada
Marie Boehm, Agriculture and Agri-Food Canada
Roy Brooke, United Nations
Ross Brown, Environment Canada at Ouranos
Stephen Burridge, Foreign Affairs and International Trade Canada
Ian Burton, University of Toronto
Elizabeth Bush, Environment Canada
John Clague, Simon Fraser University
J. Graham Cogley, Trent University
Neil Comer, Environment Canada
John Cooper, Health Canada
Thea Dickinson, Burton Dickinson Consulting
Karen Dodds, Environment Canada
Jimena Eyzaguirre, National Round Table on the Environment and the Economy
Donald Forbes, Natural Resources Canada
Jim Frehs, Health Canada
Nathan P. Gillett, Environment Canada
Sunling Gong, Environment Canada
Christian Gour, Foreign Affairs and International Trade Canada
Patrick Hebert, Foreign Affairs and International Trade Canada
Ole Hendrickson, Environment Canada
Matt Jones, Environment Canada
Dan Jutzi, Environment Canada
Viatcheslav V. Kharin, Environment Canada
Grace Koshida, Environment Canada
Paul Kovacs, Institute for Catastrophic Loss Reduction
Beth Lavender, Foreign Affairs and International Trade Canada
Donald Lemmen, Natural Resources Canada
Guilong Li, Environment Canada
Brad Little, Environment Canada
John Loder, Fisheries and Oceans Canada
Heather Low, Foreign Affairs and International Trade Canada
Katie Lundy, Environment Canada
Gordon McBean, Institute for Catastrophic Loss Reduction
Brian Mills, Environment Canada
Seung-Ki Min, Environment Canada
Authors and Expert Reviewers Annex I
547
Monirul Mirza, Environment Canada
Niall O’Dea, Natural Resources Canada
Michael Ott, Fisheries and Oceans Canada
William Perrie, Fisheries and Oceans Canada
Paul Pestieau, Environment Canada
Caroline Rodgers, Clean Air Partnership, Toronto
Slobodan Simonovic, University of Western Ontario
Sharon Smith, Natural Resources Canada
Ronald Stewart, University of Manitoba
John Stone, Carleton University
Roger B. Street, UK Climate Impacts Programme
Kit Szeto, Environment Canada
Richard Tarasofsky, Foreign Affairs and International Trade Canada
Martin Tremblay, Indian and Northern Affairs Canada
Liette Vasseur, Brock University
Elizabeth Walsh, Natural Resources Canada
Xiaolan Wang, Environment Canada
Bin Yu, Environment Canada
Xuebin Zhang, Environment Canada
Francis Zwiers, Pacific Climate Impacts Consortium, University of Victoria
Chile
Paulina Aldunce, University of Chile
Daniel Barrera, Ministry of Agriculture
Gonzalo Leon, Ministry of Environment
Alejandro León, Universidad de Chile
Sebastian Vicuna, Pontificia Universidad Catolica de Chile
China
Xing Chen, Nanjing University
Ying Chen, Chinese Academy of Social Sciences
Qin Dahe, China Meteorological Administration
Shibo Fang, Chinese Academy of Meteorological Sciences
Qiang Feng, Chinese Academy of Sciences
Ge Gao, National Climate Center, China Meteorological Administration
Daoyi Gong, State Key Laboratory of Earth Surface Processes and Resource Ecology
Anhong Guo, Agrometeorological Center of National Meteorological Centre
Tong Jiang, China Meteorological Administration
Jianping Li, Institute of Atmospheric Physics, Chinese Academy of Sciences
Jing Li, Beijing Normal University
Ning Li, Beijing Normal University
Erda Lin, Chinese Academy of Agricultural Sciences
Hongbin Liu, China Meteorological Administration
Xinhua Liu, Severe Weather Prediction Center of National Meteorological Center
Houquan Lu, National Meteorological Center
Xianfu Lu, United Nations Framework Convention on Climate Change
Yali Luo, Chinese Academy of Sciences
Wenjun Ma, Shanghai Jiao Tong University
Huixin Meng, Institute for Urban and Environment Studies
Jiahua Pan, Chinese Academy of Social Sciences
Gubo Qi, College of Humanities and Development Studies
Fumin Ren, China Meteorological Administration
Guoyu Ren, Chinese Academy of Social Sciences
WU Shaohong, Institute of Geographical Sciences, Chinese Academy of Sciences
Shangbai Shi, Chinese Academy of Social Sciences
Ying Sun, China Meteorological Administration
Changke Wang, China Meteorological Administration
Dongming Wang, Policy Study
Jianwu Wang, Chinese Academy of Social Sciences
Ming Wang, Academy of Disaster Reduction and Emergency Management
Xiaoyi Wang, Institute of Sociology
Yongguang Wang, China Meteorological Administration
Fengying Wei, China Meteorological Administration
Jet-Chau Wen, National Yunlin University of Science and Technology
Xiangyang Wu, Research Centre for Sustainable Development
Liyong Xie, Shenyang Agricultural University
Wei Xu, State Key Laboratory of Earth Surface Processes and Resource Ecology
Yinlong Xu, Chinese Academy of Agricultural Sciences
Zheng Yan, Institute of Urban and Environmental Studies
Saini Yang, Beijing Normal University
Tao Ye, State Key Lab of Earth Surface Processes and Resource Ecology
Yi Yuan, National Disaster Reduction Center of China
Yin Yunhe, Institute of Geographical Sciences, Chinese Academy of Sciences
Panmao Zhai, Chinese Academy of Meteorological Sciences
Chenyi Zhang, China Meteorological Administration
Zhao Zhang, State Key Laboratory of Earth Surface Processes and Resource Ecology
Zong-Ci Zhao, National Climate Center
Guangsheng Zhou, China Meteorological Administration
Hongjian Zhou, Nationals Disaster Reduction Center of China
Tao Zhou, State Key Laboratory of Earth Surface Processes and Resource Ecology
Tianjun Zhou, Chinese Academy of Sciences, Institute of Atmospheric Physics
Furong Zhu, Ningxia Economic Research Center
Xukai Zou, China Meteorological Administration
Colombia
Ana Campos-Garcia, Consultant
Omar-Dario Cardona, Universidad Nacional de Colombia
Carmen Lacambra, Cambridge Coastal Research Unit
Pedro Simon Lamprea Quiroga, Colombian Institute of Hydrology, Meteorology, and
Environmental Studies
Walter Vergara, Inter-American Development Bank
Gustavo Wilches-Chaux, Universidad Andina Simon Bolivar
Cook Islands
Arona Ngari, Cook Islands Meteorological Service
Costa Rica
Allan Lavell, Programme for the Social Study of Risk and Disaster
Roberto Villalobos Flores, Instituto Meteorologico Nacional
Cuba
Raul J. Garrido Vazquez, Ministry Science, Technology and Environment
Tomas Gutierrez Perez, Instituto de Meteorologia
Avelino G. Suarez, Institute of Ecology and Systematic, Cuban Environmental Agency
Cyprus
Silas Michaelides, Ministry of Agriculture, Natural Resources, and Environment
Denmark
Kirsten Halsnaes, Risø DTU
Torkil Jonch Clausen, DHI
Anne Mette Jorgensen, Centre Danish Meteorological
Karen G. Villholth, GEUS, Geological Survey of Denmark and Greenland
Egypt
Fatma El Mallah, League of Arab States
Wadid Fawzy Erian, Arab Center for the Studies of Arid Zones and Dry Lands
Amal Saad-Hussein, National Research Centre
Adel Yasseen, Ain Shams University, Institute of Environmental Research and Studies
El Salvador
Luis Ernesto Romano, Centro Humboldt Nicaragua
Finland
Timothy Carter, Finnish Environment Institute
Hilppa Gregow, Finnish Meteorological Institute
Annex I Authors and Expert Reviewers
548
Simo Haanpää, Aalto University
Pirkko Heikinheimo, Prime Minister’s Office of Finland
Susanna Kankaanpää, HSY Helsinki Region Environmental Services Authority
Sanna Luhtala, Ministry of Agriculture and Forestry
Markku Niinioja , Ministry for Foreign Affairs
Hannu Raitio, Finnish Forest Research Institute
Karoliina Saarniharo, United Nations Framework Convention on Climate Change
Kristiina Säntti, Finnish Meteorological Institute
Heikki Tuomenvirta, Finnish Meteorological Institute
Elina Vapaavuori, Finnish Forest Research Institute
Hanna Virta, Finnish Meteorological Institute
France
Franck Arnaud, Ministry of Ecology, Sustainable Development, Transport, and Housing
Slimane Bekki, Institut Pierre Simon Laplace, Laboratoire Atmospheres, Milieux,
Observations Spatiales
Nicolas Beriot, Ministry of Ecology, Sustainable Development, Transport, and Housing
Olivier Bommelaer, Ministry of Ecology, Sustainable Development, Transport, and
Housing
Olivier Boucher, Met Office Hadley Centre
Paul-Henri Bourrelier, French Association for Disaster Risk Reduction
Jean-Marie Carriere, Meteo-France
Fabrice Chauvin, Meteo-France
Chantal Claud, Institute Pierre-Simon, Laboratoire de Meteorologie Dynamique
Fabio D’Andrea, Institute Pierre-Simon, Laboratoire de Meteorologie Dynamique
Sylvie de Smedt, Ministry of Ecology, Sustainable Development, Transport and Housing
Henri Decamps, Centre National de la Recherche Scientifique
Pascale Delecluse, Meteo-France
Michel Deque, Meteo-France
Pierre-Yves Dupuy, Service Hydrographique et Oceanographique de la Marine
Nicolas Eckert, Cemagref
Rene Feunteun, French Association for Disaster Risk Reduction
JC Gaillard, The University of Auckland
Francois Gerard, French Association for Disaster Risk Reduction
Marc Gillet, Meteo-France
Pascal Girot, IUCN
Frederic Grelot, Cemagref
Delphine Grynszpan, UK Health Protection Agency
Eric Guilyardi, Institut Pierre Simon Laplace, Laboratoire D’Oceanographie et du Climat
Stephane Hallegatte, CIRED and Meteo-France
Sylvie Joussaume, Paris Consortium on Climate
Reza Lahidji, Groupement de Recherches en Gestion a HEC
Michel Lang, Cemagref
Goneri Le Cozannet, Bureau de Recherches Geologiques et Minieres
Jean-Michel Le Quentrec, French Association for Disaster Risk Reduction
Antoine Leblois, CIRED
Alexandre Magnan, Institute fior Sustainable Development and International Relations
Eric Martin, Meteo-France
Olivier Mestre, Meteo-France
David Meunier, Ministry of Ecology, Sustainable Development, Transport, and Housing
Hormoz Modaressi, Bureau of Geological and Mining Research
Roland Nussbaum, Mission Risques Naturels
Sylvie Parey, EDF-France
Cedric Peinturier, Ministry of Ecology, Sustainable Development, Transport, and Housing
Michel Petit, Conseil General de L’industrie, de L’energie et des Technologies
Pierre Picard, Ecole Polytechnique
Serge Planton, Méto-France
Valentin Przyluski, Centre International de Recherche sur l’Environnement et le
Developpement
Jean-Luc Salagnac, Centre Scientifique et Technique de Batiment
Bernard Seguin, INRA
Karen Sudmeier-Reiux, University of Lausanne, IUCN Commission on Ecosystems
Management
Nicolas Taillefer, Centre Scientifique et Technique de Batiment
Jean-Philippe Torterotot, Cemagref
Robert Vautard, Institut Pierre Simon Laplace, Laboratoire des Sciences du Climat et
de l’Environnement
Jean-Philippe Vidal, Cemagref
Vincent Viguié, Centre International de Recherche sur l’Environnement et le
Développement
Pascal Yiou, Laboratoire des Sciences du Climat et de l’Environnement
Germany
Gotelind Alber, Women for Climate Justice
Christoph Bals, Germanwatch
Hubertus Bardt, Cologne Institute for Economic Research
Andreas Baumgärtner, Project Management Agency of DLR, German IPCC
Coordination Office
Paul Becker, German Weather Service
Joern Birkmann, UN University Institute for Environment and Human Security
Hans-Georg Bohle, University of Bonn
Hans Gunter Brauch, Freie Universitaet Berlin
Michael Bründl, WSL Institute for Snow and Avalanche Research SLF
Achim Daschkeit, German Federal Environment Agency
Carmen de Jong, University of Savoy
Thomas Deutschländer, German Meteorological Service
Paul Dostal, Project Management Agency of DLR
Dirk Engelbart, Federal Ministry of Transport, Building and Urban Development
Eberhard Faust, Munich Reinsurance Company
Roland Fendler, German Federal Environment Agency
Tobias Fuchs, German Weather Service
Hans-Martin Fuessel, European Environment Agency
Stefan Goessling-Reisemann, University of Bremen
Robert Grassmann, Deutsche Welthungerhilfe e.V.
Edeltraud Guenther, Technische Universität Dresden
Josef Haider, KfW Development Bank
Angela Michiko Hama, United Nations International Strategy for Disaster Reduction
Sven Harmeling, Germanwatch
Fred Fokko Hattermann, Potsdam Institute for Climate Impact Research
Gabriele Hegerl, University of Edinburgh
Hans-Joachim Herrmann, German Federal Environment Agency
Anne Holsten, Potsdam Institute of Climate Impact Research
Anke Jentsch, University of Koblenz-Landau
Marcus Kaplan, German Development Institute
Christina Koppe-Schaller, Deutscher Wetterdienst
Christoph Kottmeier, Karlsruhe Institute of Technology
Frank Kreienkamp, Climate and Environment Consulting Potsdam GmbH
Christian Kuhlicke, Helmhotz Centre for Environmental Research
Birgit Kuna, Project Management Agency of DLR
Nana Künkel, German Agency for International Development
Michael Kunz, Karlsruhe Institute of Technology
Ole Langniss, Fichtner GmbH & Co KG
Rocio Lichte, United Nations Framework Convention on Climate Change
Petra Mahrenholz, German Federal Environment Agency
Reinhard Mechler, International Institute for Applied Systems Analysis, Vienna
University of Economics
Bettina Menne, World Health Organization, Regional Office for Europe
Annette Mohner, United Nations Framework Convention on Climate Change
Guido Mücke, German Federal Environment Agency
Christian L.C. Müller, Federal Ministry for the Environment, Nature Conservation,
and Nuclear Safety
Claudia Pahl-Wostl, Institute of Environmental Systems Research, University of
Osnabruck
Gertrude Penn-Bressel, German Federal Environment Agency
Jurgen Pohl, University of Bonn
Joerg Rapp, Deutscher Wetterdienst
Authors and Expert Reviewers Annex I
549
Markus Reichstein, Max-Planck Institute for Biogeochemistry
Joachim Rock, Johann Heinrich von Thuenen-Institute
Benno Rothstein, University of Applied Forest Sciences Rottenburg
Peter Rottach, Diakonie Katastrophenhilfe Consultant
Julia Rufin, Federal Minitry for the Environment, Nature Conservation, and Nuclear
Safety
Evelina Santa, Federal Ministry of Education and Research
Philipp Schmidt-Thome, Geological Survey of Finland
Gudrun Schütze, German Federal Environment Agency
Reimund Schwarze, Helmholtz Center for Environmental Research
Joachim H. Spangenberg, Sustainable Europe Research Institute
Frank Sperling, World Wildlife Fund, Norway
Jochen Stuck, Project Management Agency of DLR
Swenja Surminski, Association of British Insurers
Christiane Textor, Project Management Agency of DLR, German IPCC Coordination
Office
Annegret Thieken, University of Potsdam
Uwe Ulbrich, Freie Universitat Berlin
Martine Vatterodt, Federal Ministry for Economic Cooperation and Development
Monika Vees, German Federal Environment Agency
Gottfried von Gemminingen, Federal Ministry for Economic Cooperation and
Development
Hans von Storch, GKSS Research Center
Martin Voss, Katastrophenforschungsstelle Berlin
Koko Warner, United Nations University, Institute for Environment and Human Security
Juergen Weichselgartner, GKSS Research Center
Johanna Wolf, Memorial University of Newfoundland
Sabine Wurzler, North Rhine Westphalia State Environment Agency
Karl-Otto Zentel, Deutsches Komitee Katastrophenvorsorge e.V.
Ghana
Seth Vordzorgbe, United Nations Development Programme
Greece
Christina Anagnostopoulou, Aristotle University of Thessaloniki
Helena Flocas, University of Athens
Panagiota Galiatsatou, Aristotle University of Thessaloniki
Antonis Koussis, National Observatory of Athens
Aristeidis Koutrouli, Technical University of Crete
Athanasios Louka, University of Thessaly
Petroula Louka, Hellenic National Meteorological Service
Dimitrios Melas, Aristotle University of Thessaloniki
Panayotis Prinos, Aristotle University of Thessaloniki
Ioannis Tsanis, Technical University of Crete
Adonis Velegrakis, University of the Aegean
Christos Zerefos, Academy of Athens
Guatemala
Edwin Castellanos, Universidad del Valle de Guatemala
Hungary
Joseph Feiler, Ministry of National Development
Ferenc L. Tóth, International Atomic Energy Agency
Iceland
Halldor Bjornsson, Icelandic Meteorological Office
HalldorSigrun Karlsdottir, Icelandic Meteorological Office
Arni Snorrason, Icelandic Meteorological Office
India
Unnikrishnan Alakkat, National Institute of Oceanography
Subbiah Arjunapermal, Asian Disaster Preparedness Center
Suruchi Bhadwal, The Energy and Resources Institute
Mihir Bhatt, India Disaster Mitigation Institute
Amit Garg, Indian Institute of Management Ahmedabad
B.N. Goswami, Indian Institute for Tropical Meteorology
Manu Gupta, SEEDS
Umesh Haritashya, University of Dayton
Ritesh Kumar, Wetlands International - South Asia
Pradeep Mujumdar, Indian Institute of Science
Anand Patwardhan, Indian Institute of Technology Bombay
Apurva Sanghi, The World Bank
Akhilesh Surjan, United Nations University
Indonesia
Edvin Aldrian, Badan Meteorologi Klimatologi dan Geofisika
Iran
Rahman Davtalab, Ministry of Energy
Saeid Eslamian, Isfahan University of Technology
Mahnaz Khazaee, Atmospheric Science and Meteorological Research Center
Mohammad Rahimi, Semnan University
Fatemeh Rahimzadeh, Atmospheric Science and Meteorological Research Center
Saviz Sehat Kashani, Atmospheric Sciences and Meteorological Research Center
Ireland
Ian Bryceson, Norwegian University of Life Sciences
Noel Casserly, Department of the Environment
Italy
Marina Baldi, National Research Council, Institute of Biometeorology
Roberto Bertollini, World Health Organization
Francesco Bosello, Fondazione Eni Enrico Mattei, Milan University
Stefano Bovo, ARPA Piemonte
Carlo Giupponi, University Ca’ Foscari of Venice and Euro-Mediterranean Centre for
Climate Change
Georg Kaser, University of Innsbruck
Valentina Pavan, ARPA Emilia-Romagna
Roberto Ranzi, University of Brescia
Carlo Scaramella, World Food Programme
Rodica Tomozeiu, ARPA Emilia-Romagna
Japan
Shiho Asano, Forestry and Forest Products Research Institute
Fumiaki Fujibe, Meteorological Research Institute
Koji Fujita, Nagoya University
Masahiro Hashizume, Institute of Tropical Medicine, Nagasaki University
Yasushi Honda, University of Tsukuba
Shinjiro Kanae, Tokyo Institute of Technology
Takehiro Kano, Division Ministry of Foreign Affairs
Miwa Kato, UNFCCC Secretariat
Hiroyasu Kawai, Port and Airport Research Institute
So Kazama, Tohoku University
Akio Kitoh, Meteorological Research Institute
Masahide Kondo, University of Tsukuba
Kazuo Kurihara, Meteorological Research Institute
Shoji Kusunoki, Meteorological Research Institute
Takao Masumoto, National Institute for Rural Engineering, National Agriculture and
Food Research Organization
Nobuo Mimura, Ibaraki University
Hisayoshi Morisugi, Nihon University
Toshiyuki Nakaegawa, Meteorological Research Institute
Elichi Nakakita, Kyoto University
Motoki Nishimori, National Institute for Agro-Environmental Sciences
Taikan Oki, University of Tokyo
Rajib Shaw, Kyoto University
Annex I Authors and Expert Reviewers
550
Hideo Shiogama, National Institute for Environmental Studies
Yasuto Tachikawa, Kyoto University
Kiyoshi Takahashi, National Institute for Environmental Studies
Izuru Takayabu, Meteorological Research Institute
Kuniyoshi Takeuchi, International Centre for Water Hazard and Risk Management
Tadashi Tanaka, University of Tsukuba
Makoto Tani, Kyoto University
Tsugihiro Watanabe, Research Institute for Humanity and Nature
Hiroya Yamano, National Institute for Environmental Studies
Kenya
Peter Ambenje, Kenya Meteorological Department
Samwel Marigi, Kenya Meteorological Department
Charles Mutai, Ministry of Environment and Mineral Resources
Christopher Oludhe, University of Nairobi, Department of Meteorology
Latvia
Olga Vilima, United Nations International Strategy for Disaster Reduction
Malaysia
Joy Jacqueline Pereira, Universiti Kebangsaan Malaysia
Salmah Zakaria, United Nations Economic and Social Commission
Mauritania
Gueladio Cisse, Swiss Tropical and Public Health Institute
Mexico
Victor Cardenas, Climate Change and Natural Disaster Risk Management
Tereza Cavazos, El Centro de Investigación Científica y de Educación Superior de
Ensenada
Carolina Neri, Universidad Nacional Autonoma de Mexico
Ursula Oswald-Spring, Universidad Nacional Autónoma de México
Ricardo Zapata-Marti, United Nations Economic Commission for Latin America and
the Caribbean
Mongolia
Ravsal Oyun, JEMR Consulting Company
Morocco
Abdalah Mokssit, Direction de la Météorologie Nationale
Mozambique
Felipe Lucio, Global Framework for Climate Services Office, WMO
New Zealand
Reid Basher, Secretariat of the High-Level Taskforce on the Global Framework for
Climate Services
Leonard Brown, Ministry for the Environment - Manatu Mo Te Taiao
John Campbell, University of Waikato
John Hay, University of the South Pacific
Glenn McGregor, University of Auckland
Matthew McKinnon, DARA
Helen Plume, Ministry for the Environment - Manatu Mo Te Taiao
David Wratt, National Institute of Water and Atmospheric Research
Niger
Abdelkrim Ben Mohamed, University of Niamey
Norway
Torgrim Asphjell, Climate and Pollution Agency
Rasmus Benestad, The Norwegian Meteorological Institute
Tor A. Benjaminsen, Norwegian University of Life Sciences
Elzbieta Maria Bitner-Gregersen, Det Norske Veritas AS
Oyvind Christophersen, Climate and Pollution Agency
Solveig Crompton, Ministry of the Environment
Linda Dalen, Norwegian Directorate for Nature Management
Lars Ingolf Eide, Det Norske Veritas
Siri Eriksen, Norwegian University of Life Sciences
Christoffer Grønstad, Climate and Pollution Agency
Hege Haugland, Climate and Pollution Agency
Hege Hisdal, Norwegian Water Resources and Energy Directorate
Dag O. Høgvold, Directorate for Civil Protection and Emergency Planning
Linn Bryhn Jacobsen, Climate and Pollution Agency
Vikram Kolmannskog, Norwegian Refugee Council
Ole-Kristian Kvissel, Climate and Pollution Agency
Farrokh Nadim, International Centre for Geohazards
Lars Otto Naess, Institute of Development Studies
Karen O’Brien, University of Oslo
Ellen Øseth, Norwegian Polar Institute
Marit Viktoria Pettersen, Ministry of Foreign Affairs
Asgeir Sorteberg, University of Bergen
Linda Sygna, University of Oslo
Kirsten Ulsrud, University of Oslo
Vigdis Vestreng, Climate and Pollution Agency
Pakistan
Muhammad Mohsin Iqbal, Global Change Impact Studies Centre
Jawed Ali Khan, Ministry of Environment
Maira Zahur, Women for Climate Justice
Palestinian National Authority
Nedal Katbeh-Bader, Environment Quality Authority
Peru
Eduardo Calvo, Universidad Nacional Mayor de San Marcos
Encinas Carla, Intercooperation
Silvia Llosa, United Nations International Strategy for Disaster Reduction
Philippines
Imelda Abarquez, Oxfam Hong Kong
Sanny Jegillos, United Nations Development Programme
Rosa Perez, Manila Observatory
Poland
Janusz Filipiak, Institute of Meteorology and Water Management
Zbigniew Kundzewicz, Polish Academy of Sciences
Zbigniew Ustrnul, Institute of Meteorology and Water Management, Jagiellonian
University
Joanna Wibig, University of Lodz
Republic of Korea
So-Min Cheong, University of Kansas
Tae Sung Cheong, National Emergency Management Agency
Soojeong Myeong, Korea Environment Institute
Republic of Maldives
Amjad Abdulla, IPCC Vice Chair WG II, Climate Change Energy Department,
Ministry of Housing and Environment
Romania
Roxana Bojariu, National Meteorological Administration
Sorin Cheval, National Meteorological Administration
Russian Federation
E. M. Akentyeva, Main Geophysical Observatory
Sergey Borsch, Hydromet Center of Russia
Authors and Expert Reviewers Annex I
551
N. V. Kobysheva, Main Geophysical Observatory
Boris Porfiriev, Institute for Economic Forecasting, Russian Academy of Sciences
Vladimir Semenov, A.M. Obukhov Institute of Atmospheric Physics
Boris Sherstyukov, All Russian Research Institute of Hydrometeorological
Information World Data Center
Senegal
Cherif Diop, Senegalese Meteorological Agency
South Africa
Reinette (Oonsie) Biggs, Stockholm Resilience Centre, Stockholm University
Bruce Glavovic, Massey University
Bettina Koelle, Indigo Development and Change
Noel Oettle, Environmental Monitoring Group
Coleen Vogel, University of Witwatersrand
Gina Ziervogel, University of Cape Town
Spain
Enric Aguilar, Universitat Rovira i Virgili
Gerardo Benito, Spanish Council for Scientific Research
Jorge Bonnet Fernandez Trujillo, Government of the Canary Islands
Francisco Garcia Novo, University of Seville
José Manuel Gutiérrez, Consejo Superior de Investigaciones Científicas
Ana Iglesias, Universidad Politecnica de Madrid
José Antonio López-Díaz, Agencia Estatal de Meteorología
Concepcion Martinez-Lope, Spanish Bureau for Climate Change
José Moreno, University of Castilla-La Mancha
Francisco Pascual, Spanish Bureau for Climate Change
Jose Ramon Picatoste-Ruggeroni, Spanish Bureau for Climate Change
Ernesto Rodriguez-Camino, Spanish Meteorological Agency
Sabater Sergi, University Girona
Sudan
Ismail Fadl El Moula Mohamed, Sudan Meteorological Authority
Balgis Osman-Elasha, African Development Bank
Sweden
Cecilia Alfredsson, Swedish Civil Contingencies Agency
Lars Barring, Swedish Meteorological and Hydrological Institute
Sten Bergstrom, Swedish Meteorological and Hydrological Institute
Pelle Boberg, Swedish Meteorological and Hydrological Institute
Henrik Carlsen, Swedish Defence Research Agency
Carl Folke, The Beijer Institute, Stockholm University
Clarisse Kehler Siebert, Stockholm Environment Institute
Carina Keskitalo, Umea University
Richard Klein, Stockholm Environment Institute
Georg Lindgren, Lund University
Elin Lovendahl, Swedish Meteorological and Hydrological Institute
Barbro Naslund-Landenmark, Swedish Civil Contingencies Agency
Carin Nilsson, Swedish Meteorological and Hydrological Institute
Ulrika Postgard, Swedish Civil Contingencies Agency
Markku Rummukainen, Swedish Meteorological and Hydrological Institute
Johan Schaar, Ministry of Foreign Affairs
Lisa Schipper, Stockholm Environment Institute
Ake Svensson, Swedish Civil Contingencies Agency
Switzerland
Simon Allen, IPCC WGI Technical Support Unit
Walter J. Ammann, Global Risk Forum GRF Davos
Neville Ash, United Nations Environment Programme
Stefan Brönnimann, University of Bern
Carlo Casty, Partner Reinsurance Company
Nicole Clot, Intercooperation
Paul Della-Marta, Partner Reinsurance Company
Andreas Fischlin, Swiss Federal Institute of Technology, Systems Ecology
Markus Gerber, University of Bern, Climate and Environmental Physics
Peter Greminger, Federal Office for Environment
Christian Huggel, University of Zurich
Matthias Huss, University of Fribourg
Daniel Kull, International Federation of Red Cross and Red Crescent Societies
Juerg Luterbacher, Justus Liebig University
Joy Muller, International Federation of Red Cross and Red Crescent Societies
Urs Neu, Swiss Academy of Sciences
Boris Orlowsky, Swiss Federal Institute of Technology Zurich
Pascal Peduzzi, United Nations Environment Programme
Gian-Kasper Plattner, IPCC WGI Technical Support Unit
Dieter Rickenmann, Swiss Federal Research Institute WSL
Stephan Rist, Centre for Development and Environment, University of Bern
Jose Romero, Federal Office for the Environment
Sonia Seneviratne, Swiss Federal Institute of Technology Zurich
Andreas Spiegel, Swiss Re
Thomas Stocker, University of Bern
Philippe Thalmann, EPFL Swiss Federal Institute of Technology Lausanne
Heinz Wanner, University of Bern
Andre Wehrli, European Environment Agency
Heini Wernli, Swiss Federal Institute of Technology Zurich
Irina Zodrow, United Nations International Strategy for Disaster Reduction
Tanzania
Emmanuel Mpeta, Tanzania Meteorological Agency
Khamaldin Daud Mutabazi, Sokoine University of Agriculture
Pius Zebhe Yanda, University of Dar es Salaam
Thailand
Seree Supratid, Rangsit University
The Netherlands
Frans Berkhout, Vrije University
Laurens Bouwer, Institute for Environmental Studies, Vrije University
Hein W. Haak, The Royal Netherlands Meteorological Institute
Albert Klein Tank, The Royal Netherlands Meteorological Institute
Irene Kreis, Health Protection Agency
Adriaan Perrels, Finnish Meteorological Institute
Maarten van Aalst, Red Cross Red Crescent Climate Centre
Bart van den Hurk, The Royal Netherlands Meteorological Institute
Henny A.J. van Lanen, Wageningen University
Geert Jan van Oldenborgh, The Royal Netherlands Meteorological Institute
Jeroen Warner, Wageningen University
Trinidad and Tobago
Veronica Belgrave, Ministry of Planning, Housing, and the Environment
Turkey
Salahattin Incecik, Istanbul Technical University
United Kingdom
Neil Adger, Tyndall Centre, University of East Anglia
Alex Arnall, University of Reading
Nigel Arnell, University of Reading
Victoria Bell, Centre for Ecology and Hydrology
Enrico Biffis, Imperial College, London
Katrina Brown, University of East Anglia
Simon Brown, Met Office Hadley Centre
Sal Burgess, UK Department for Environment, Food, and Rural Affairs
Harriet Caldin, Health Protection Agency
Diarmid Campbell-Lendrum, World Health Organization
Annex I Authors and Expert Reviewers
552
Catriona Carmichael, Health Protection Agency
Amanda Charles, UK Government Office for Science
Declan Conway, University of East Anglia
Tim Conway, UK Department for International Development
Anita Cooper, Health Protection Agency
Geoff Darch, Atkins Consultants and University of East Anglia
Ian Davis, Cranfield University
Ken De Souza, UK Department for International Development
Andrew Dlugolecki, Climatic Research UnitUniversity of East Anglia
Maureen Fordham, Northumbria University
Tim Forsyth, London School of Economics and Political Science
Clare Goodess, University of East Anglia Climatic Research Unit
Jim Hall, Newcastle University
Lucy Hayes, UK Department of Energy and Climate Change
Clare Heaviside, Centre for Radiation, Chemical and Environmental Hazards, Health
Protection Agency
Debbie Hillier, Oxfam International
Robert Hodgson, University of Exeter
Sari Kovats, London School of Hygiene and Tropical Medicine
Bo Lim, United Nations Development Programme
Andrew Maskrey, UN International Strategy for Disaster Reduction
Michael McCall, Universidad Nacional Autonoma de Mexico
William McGuire, University College London
Thomas Mitchell, Overseas Development Institute
John Morton, Natural Resources Institute, University of Greenwich
Alessandro Moscuzza, UK Department for International Development
Robert Muir-Wood, Risk Management Solutions
Virginia Murray, Health Protection Agency
Katherine Nightingale, Christian Aid
Geoff O’Brien, Northumbria University
Phil O’Keefe, Northumbria University
Jean Palutikof, Griffith University
Mark Pelling, King’s College London
Emily Polack, Institute of Development Studies
Ben Ramalingam, Overseas Development Institute
Nicola Ranger, London School of Economics
John Rees, Natural Environment Research Council, UK
Hannah Rowlatt, Health Protection Agency and University of Sheffield
Sohel Saikat, Health Protection Agency
David Satterthwaite, International Institute for Environment & Development
Chris Sear, UK Department of Energy and Climate Change
A. Simmons, European Centre for Medium-Range Weather Forecasts
Robert Siveter, International Petroleum Industry Environmental Conservation
Association
David Smith, University of the West Indies
Stephen Smith, UK Committee on Climate Change
David Stephenson, University of Exeter
Peter Stott, Met Office Hadley Centre
Robert Sykes, International Petroleum Industry Environmental Conservation
Association
Thomas Tanner, Institute of Development Studies
Addis Taye, Health Protection Services
Emma Tompkins, University of Southampton, Highfield Campus
John Twigg, University College London
Sotiris Vardoulakis, Health Protection Agency
Emma Visman, Humanitarian Futures Programme, King’s College, London
Tim Waites, UK Department for International Development
David Warrilow, UK Department of Energy and Climate Change
Paul Watkiss, Paul Watkiss Associates
Robert Wilby, University of Loughborough
Michelle Winthrop, UK Department for International Development
Philip Woodworth, National Oceanography Centre
Ronald Young, Young International Ltd / Knowledge Associates International Ltd
United States of America
David Allen, US Global Change Research Program
Tom Armstrong, US Global Change Research Program
Jeff Arnold, US Army Corps of Engineers
Margaret Arnold, The World Bank
Bilal Ayyub, University of Maryland
Donald Ballantyne, MMI Engineering
Ko Barrett, National Oceanic and Atmospheric Administration
Stephen Bender, Organization of American States (retired)
Lisa M. Butler-Harrington, The Wharton School, Kansas State University
JoAnn Carmin, Massachusetts Institute of Technology
Edward Carr, US Administration for International Development
DeWayne Cecil, National Oceanic and Atmospheric Administration
Christina Chan, US Department of State
David Cleaves, US Forest Service
Thomas Cronin, US Geological Survey
Susan Cutter, University of South Carolina
Kirstin Dow, University of South Carolina
David Easterling, National Oceanic and Atmospheric Administration, National
Climatic Data Center
Kristie Ebi, IPCC WGII Technical Support Unit
Barbara Ellis, Center for Disease Control
Christopher Emrich, University of South Carolina
Sandy Eslinger, National Oceanic and Atmospheric Administration
Ross Faith, US Subcommittee on Disaster Reduction
Christopher Field, Carnegie Institution for Science
Stephen Gill, National Oceanic and Atmospheric Administration
Justin Ginnetti, United Nations International Strategy for Disaster Reduction
William Gutowski, Iowa State University
D.E. (Ed) Harrison, National Oceanic and Atmospheric Administration, Pacific Marine
Environmental Laboratory
Jerry Hatfield, US Department of Agriculture
Robert Heilmayr, Emmett Interdisciplinary Program for Environment and Resources,
Stanford University
Molly Hellmuth, International Research Institute for Climate and Society
Jeremy Hess, Centers for Disease Control and Prevention
Robert Hirsch, US Geological Survey
Robert Jarrett, US Geological Survey
Terry Jeggle, University of Pittsburgh
Mark Keim, Centers for Disease Control and Prevention
Paul Knappenberger, New Hope Environmental Services
Thomas Knutson, National Oceanic and Atmospheric Administration
James Kossin, National Oceanic and Atmospheric Administration, National Climatic
Data Center
Howard Kunreuther, The Wharton School, University of Pennsylvania
David Lea, US Department of State
Arthur Lee, Chevron Services Company
Robin Leichenko, Rutgers University
Maria Carmen Lemos, University of Michigan
Robert Lempert, RAND Corporation
David Levinson, US Forest Service
Joanne Linnerooth-Bayer, International Institute for Applied Systems Analysis
Peter Liotta, Independent Scholar
Chris Little, Woodrow Wilson School of Public and International Affairs, Princeton
University
David Lobell, Stanford University
Pat Longstaff, Syracuse University
Alexander Lotsch, The World Bank
Michael MacCracken, Climate Institute
Katharine Mach, IPCC WGII Technical Support Unit
Simon Mason, Columbia University
Michael Mastrandrea, IPCC WGII Technical Support Unit
Sabrina McCormick, US Environmental Protection Agency
Authors and Expert Reviewers Annex I
553
Linda Mearns, National Center for Atmospheric Research
Jerry Meehl, National Center for Atmospheric Research
Chris Milly, US Geological Survey
James Mitchell, Rutgers University
Marcus Moench, Institute for Social and Environmental Transition
Susanne Moser, Susanne Moser Research and Consulting
Meredith Muth, National Oceanic and Atmospheric Administration
Robert J Naiman, University of Washington
Robert O’Connor, National Science Foundation
Ian O’Donnell, Asian Development Bank
Michael Oppenheimer, Princeton University
Jacob Park, Green Mountain College
Roger Pielke Jr., University of Colorado
David Pierce, Scripps Institution of Oceanography
Mark Powell, National Oceanic and Atmospheric Administration
Michael Prather, University of California, Irvine
Roger Pulwarty, National Oceanic and Atmospheric Administration
David Reidmiller, US Department of State
Dian Seidel, National Oceanic and Atmospheric Administration
Emil Simiu, National Institute of Standards and Technology
Anthony-Oliver Smith, Emeritus, University of Florida
Joel Smith, Stratus Consulting
Susan Solomon, National Oceanic and Atmospheric Administration
Doreen Stabinsky, College of the Atlantic
Amanda Staudt, National Wildlife Federation
Ronald Stouffer, National Oceanic and Atmospheric Administration
Trigg Talley, US Department of State
Wassila Thiaw, National Oceanic and Atmospheric Administration
John Tiefenbacher, Texas State University
Sezin Tokar, US Agency for International Development
Kevin E. Trenberth, National Center for Atmospheric Research
Thomas Wagner, National Aeronautics and Space Administration
Robert Webb, National Oceanic and Atmospheric Administration
Elke Weber, Columbia University
Michael Wehner, Lawrence Berkeley National Laboratory
Jason Westrich, University of Georgia, Odum School of Ecology
Thomas Wilbanks, Oak Ridge National Laboratory
Benjamin Wisner, Aon-Benfield UCL Hazard Research Centre, University College
London
Richard Wright, American Society of Civil Engineers
Donald Wuebbles, University of Illinois
Tingjun Zhang, University of Colorado National Snow and Ice Data Center
Venezuela
Maria Teresa Abogado, Ministry of People’s Powers for Foreign Affairs
Jose Azuaje, Ministry of People’s Powers for Foreign Affairs
Salvano Briceno, United Nations
Claudia Salerno Caldera, Ministry of People’s Powers for Foreign Affairs
Isabel Di Carlo Quero, Ministry of People’s Powers for Foreign Affairs
Rafael Hernandez, Ministry of People’s Powers for Foreign Affairs
Federico Lagarde, Ministry of People’s Powers for Foreign Affairs
Alejandro Linayo, Research Center on Disaster Risk Reduction
Luis Jose Mata, International Monetary Fund
Yessica Pereira, Ministry of People’s Powers for Foreign Affairs
Reina Perez, Ministry of People’s Powers for Foreign Affairs
Rafael Rebolledo, Ministry of People’s Powers for Foreign Affairs
Dirk Thielen, Ministry of People’s Powers for Foreign Affairs
Vietnam
Mai Trong Nhuan, Vietnam National University
Bach Tan Sinh, National Institute for Science and Technology Policy and Strategy
Studies
Zambia
Raban Chanda, University of Botswana
Annex I Authors and Expert Reviewers
554
Authors and Expert Reviewers Annex I
555
This annex should be cited as:
IPCC, 2012: Glossary of terms. In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation
[Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen,
M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate
Change (IPCC). Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 555-564.
Glossary of Terms
II
ANNEX
556
Abrupt climate change
The nonlinearity of the climate system may lead to abrupt climate
change, sometimes called rapid climate change, abrupt events, or even
surprises. The term abrupt often refers to time scales faster than the
typical time scale of the responsible forcing. However, not all abrupt
climate changes need be externally forced. Some changes may be truly
unexpected, resulting from a strong, rapidly changing forcing of a
nonlinear system.
Adaptation
In human systems, the process of adjustment to actual or expected
climate and its effects, in order to moderate harm or exploit beneficial
opportunities. In natural systems, the process of adjustment to actual
climate and its effects; human intervention may facilitate adjustment to
expected climate.
Adaptation assessment
The practice of identifying options to adapt to climate change and
evaluating them in terms of criteria such as availability, benefits, costs,
effectiveness, efficiency, and feasibility.
Adaptive capacity
The combination of the strengths, attributes, and resources available to
an individual, community, society, or organization that can be used to
prepare for and undertake actions to reduce adverse impacts, moderate
harm, or exploit beneficial opportunities.
Aerosols
A collection of airborne solid or liquid particles, with a typical size
between 0.01 and 10 μm, that reside in the atmosphere for at least
several hours. Aerosols may be of either natural or anthropogenic
origin. Aerosols may influence climate in several ways: directly through
scattering and absorbing radiation, and indirectly by acting as cloud
condensation nuclei or modifying the optical properties and lifetime of
clouds.
Albedo
The fraction of solar radiation reflected by a surface or object, often
expressed as a percentage. Snow-covered surfaces have a high albedo,
the surface albedo of soils ranges from high to low, and vegetation-
covered surfaces and oceans have a low albedo. The Earth’s planetary
albedo varies mainly through varying cloudiness, snow, ice, leaf area,
and land cover changes.
Anthropogenic
Resulting from or produced by human beings.
Anthropogenic emissions
Emissions of greenhouse gases, greenhouse gas precursors, and aerosols
associated with human activities. These activities include the burning of
fossil fuels, deforestation, land use changes, livestock, fertilization, etc.,
that result in a net increase in emissions.
Atlantic Multi-decadal Oscillation (AMO)
A multi-decadal (65- to 75-year) fluctuation in the North Atlantic, in
which sea surface temperatures showed warm phases during roughly
1860 to 1880 and 1930 to 1960 and cool phases during 1905 to 1925
and 1970 to 1990 with a range of the order of 0.4°C.
Atmosphere
The gaseous envelope surrounding the Earth. The dry atmosphere
consists almost entirely of nitrogen (78.1% volume mixing ratio) and
oxygen (20.9% volume mixing ratio), together with a number of trace
gases, such as argon (0.93% volume mixing ratio), helium, and radiatively
active greenhouse gases such as carbon dioxide (0.035% volume mixing
ratio) and ozone. In addition, the atmosphere contains the greenhouse
gas water vapor, whose amounts are highly variable but typically
around 1% volume mixing ratio. The atmosphere also contains clouds
and aerosols.
Available potential energy
That portion of the total potential energy that may be converted to
kinetic energy in an adiabatically enclosed system.
Baseline/reference
The baseline (or reference) is the state against which change is measured.
It might be a ‘current baseline,’ in which case it represents observable,
present-day conditions. It might also be a ‘future baseline,’ which is a
projected future set of conditions excluding the driving factor of interest.
Alternative interpretations of the reference conditions can give rise to
multiple baselines.
Capacity
The combination of all the strengths, attributes, and resources available
to an individual, community, society, or organization, which can be used
to achieve established goals.
Carbon cycle
The term used to describe the flow of carbon (in various forms, e.g., as
carbon dioxide) through the atmosphere, ocean, terrestrial biosphere,
and lithosphere.
Carbon dioxide (CO
2
)
A naturally occurring gas fixed by photosynthesis into organic matter.
A byproduct of fossil fuel combustion and biomass burning, it is also
emitted from land use changes and other industrial processes. It is the
principal anthropogenic greenhouse gas that affects the Earth’s radiative
balance. It is the reference gas against which other greenhouse gases
are measured, thus having a Global Warming Potential of 1.
Catchment
An area that collects and drains precipitation.
Clausius-Clapeyron relationship (or equation)
The differential equation relating the pressure of a substance (usually
Glossary of Terms Annex II
557
water vapor) to temperature in a system in which two phases of the
substance (water) are in equilibrium.
Climate
Climate in a narrow sense is usually defined as the average weather, or
more rigorously, as the statistical description in terms of the mean and
variability of relevant quantities over a period of time ranging from months
to thousands or millions of years. The classical period for averaging
these variables is 30 years, as defined by the World Meteorological
Organization. The relevant quantities are most often surface variables
such as temperature, precipitation, and wind. Climate in a wider sense
is the state, including a statistical description, of the climate system. In
various chapters in this report different averaging periods, such as a
period of 20 years, are also used.
Climate change
A change in the state of the climate that can be identified (e.g., by
using statistical tests) by changes in the mean and/or the variability
of its properties and that persists for an extended period, typically
decades or longer. Climate change may be due to natural internal
processes or external forcings, or to persistent anthropogenic changes
in the composition of the atmosphere or in land use.
1
See also Climate
variability and Detection and attribution.
Climate extreme (extreme weather or climate event)
The occurrence of a value of a weather or climate variable above (or
below) a threshold value near the upper (or lower) ends of the range of
observed values of the variable. For simplicity, both extreme weather
events and extreme climate events are referred to collectively as ‘climate
extremes.The full definition is provided in Section 3.1.2.
Climate feedback
An interaction mechanism between processes in the climate system is
called a climate feedback when the result of an initial process triggers
changes in a second process that in turn influences the initial one. A
positive feedback intensifies the original process, and a negative feedback
reduces it.
Climate model
A numerical representation of the climate system that is based on the
physical, chemical, and biological properties of its components, their
interactions, and feedback processes, and that accounts for all or some of
its known properties. The climate system can be represented by models
of varying complexity, that is, for any one component or combination of
components a spectrum or hierarchy of models can be identified, differing
in such aspects as the number of spatial dimensions, the extent to which
physical, chemical, or biological processes are explicitly represented, or
the level at which empirical parameterizations are involved. Coupled
Atmosphere-Ocean Global Climate Models (AOGCMs), also referred to as
Atmosphere-Ocean General Circulation Models, provide a representation
of the climate system that is near the most comprehensive end of the
spectrum currently available. There is an evolution toward more complex
models with interactive chemistry and biology. Climate models are
applied as a research tool to study and simulate the climate, and for
operational purposes, including monthly, seasonal, and interannual climate
predictions.
Climate projection
A projection of the response of the climate system to emissions or
concentration scenarios of greenhouse gases and aerosols, or radiative
forcing scenarios, often based upon simulations by climate models.
Climate projections are distinguished from climate predictions in order
to emphasize that climate projections depend upon the emission/
concentration/radiative-forcing scenario used, which are based on
assumptions concerning, e.g., future socioeconomic and technological
developments that may or may not be realized and are therefore
subject to substantial uncertainty.
Climate scenario
A plausible and often simplified representation of the future climate,
based on an internally consistent set of climatological relationships that
has been constructed for explicit use in investigating the potential
consequences of anthropogenic climate change, often serving as input
to impact models. Climate projections often serve as the raw material
for constructing climate scenarios, but climate scenarios usually require
additional information such as about the observed current climate.
Climate system
The climate system is the highly complex system consisting of five major
components: the atmosphere, the oceans, the cryosphere, the land
surface, the biosphere, and the interactions between them. The climate
system evolves in time under the influence of its own internal dynamics
and because of external forcings such as volcanic eruptions, solar
variations, and anthropogenic forcings such as the changing composition
of the atmosphere and land use change.
Climate threshold
A critical limit within the climate system that induces a non-linear
response to a given forcing. See also Abrupt climate change.
Climate variability
Climate variability refers to variations in the mean state and other
statistics (such as standard deviations, the occurrence of extremes, etc.) of
the climate at all spatial and temporal scales beyond that of individual
weather events. Variability may be due to natural internal processes
within the climate system (internal variability), or to variations in natural
Annex II Glossary of Terms
____________
1
This definition differs from that in the United Nations Framework Convention on
Climate Change (UNFCCC), where climate change is defined as: “a change of climate
which is attributed directly or indirectly to human activity that alters the composition
of the global atmosphere and which is in addition to natural climate variability
observed over comparable time periods.” The UNFCCC thus makes a distinction
between climate change attributable to human activities altering the atmospheric
composition, and climate variability attributable to natural causes.
558
or anthropogenic external forcing (external variability). See also Climate
change.
Cold days/cold nights
Days where maximum temperature, or nights where minimum
temperature, falls below the 10th percentile, where the respective
temperature distributions are generally defined with respect to the
1961-1990 reference period.
Community-based disaster risk management
See Local disaster risk management.
Confidence
Confidence in the validity of a finding, based on the type, amount,
quality, and consistency of evidence and on the degree of agreement.
Confidence is expressed qualitatively.
Control run
A model run carried out to provide a ‘baseline’ for comparison with
climate change experiments. The control run uses constant values for
the radiative forcing due to greenhouse gases and anthropogenic
aerosols appropriate to pre-industrial conditions.
Convection
Vertical motion driven by buoyancy forces arising from static instability,
usually caused by near-surface cooling or increases in salinity in the case
of the ocean and near-surface warming in the case of the atmosphere.
At the location of convection, the horizontal scale is approximately the
same as the vertical scale, as opposed to the large contrast between
these scales in the general circulation. The net vertical mass transport is
usually much smaller than the upward and downward exchange.
Coping
The use of available skills, resources, and opportunities to address,
manage, and overcome adverse conditions, with the aim of achieving
basic functioning in the short to medium term.
Coping capacity
The ability of people, organizations, and systems, using available skills,
resources, and opportunities, to address, manage, and overcome
adverse conditions.
Detection and attribution
Climate varies continually on all time scales. Detection of climate
change is the process of demonstrating that climate has changed in
some defined statistical sense, without providing a reason for that
change. Attribution of causes of climate change is the process of
establishing the most likely causes for the detected change with some
defined level of confidence.
Diabatic
A process in which external heat is gained or lost by the system.
Disaster
Severe alterations in the normal functioning of a community or a society
due to hazardous physical events interacting with vulnerable social
conditions, leading to widespread adverse human, material, economic,
or environmental effects that require immediate emergency response to
satisfy critical human needs and that may require external support for
recovery.
Disaster management
Social processes for designing, implementing, and evaluating strategies,
policies, and measures thatpromote and improve disaster preparedness,
response, and recovery practices at different organizationaland societal
levels.
Disaster risk
The likelihood over a specified time period of severe alterations in the
normal functioning of a community or a society due to hazardous
physical events interacting with vulnerable social conditions, leading to
widespread adverse human, material, economic, or environmental
effects that require immediate emergency response to satisfy critical
human needs and that may require external support for recovery.
Disaster risk management (DRM)
Processes for designing, implementing, and evaluating strategies,
policies, and measures to improve theunderstanding of disaster risk,
foster disaster risk reduction and transfer, and promote continuous
improvement in disaster preparedness, response, and recovery practices,
with the explicit purpose of increasing human security, well-being,
quality of life,andsustainable development.
Disaster risk reduction (DRR)
Denotes both a policy goal or objective, and the strategic and
instrumental measures employed for anticipating future disaster risk;
reducing existing exposure, hazard, or vulnerability; and improving
resilience.
Diurnal temperature range
The difference between the maximum and minimum temperature
during a 24-hour period.
Downscaling
Downscaling is a method that derives local- to regional-scale (up to
100 km) information from larger-scale models or data analyses. The full
definition is provided in Section 3.2.3.
Drought
A period of abnormally dry weather long enough to cause a serious
hydrological imbalance. Drought is a relative term (see Box 3-3),
therefore any discussion in terms of precipitation deficit must refer to
the particular precipitation-related activity that is under discussion. For
example, shortage of precipitation during the growing season impinges
on crop production or ecosystem function in general (due to soil moisture
Glossary of Terms Annex II
559
drought, also termed agricultural drought), and during the runoff and
percolation season primarily affects water supplies (hydrological drought).
Storage changes in soil moisture and groundwater are also affected by
increases in actual evapotranspiration in addition to reductions in
precipitation. A period with an abnormal precipitation deficit is defined as
a meteorological drought. A megadrought is a very lengthy and pervasive
drought, lasting much longer than normal, usually a decade or more.
Early warning system
The set of capacities needed to generate and disseminate timely and
meaningful warning information to enable individuals, communities, and
organizations threatened by a hazard to prepare and to act appropriately
and in sufficient time to reduce the possibility of harm or loss.
El Niño-Southern Oscillation (ENSO)
The term El Niño was initially used to describe a warm-water current
that periodically flows along the coast of Ecuador and Peru, disrupting the
local fishery. It has since become identified with a basin-wide warming
of the tropical Pacific Ocean east of the dateline. This oceanic event is
associated with a fluctuation of a global-scale tropical and subtropical
surface pressure pattern called the Southern Oscillation. This coupled
atmosphere-ocean phenomenon, with preferred time scales of 2 to
about 7 years, is collectively known as the El Niño-Southern Oscillation.
It is often measured by the surface pressure anomaly difference between
Darwin and Tahiti and the sea surface temperatures in the central and
eastern equatorial Pacific. During an ENSO event, the prevailing trade
winds weaken, reducing upwelling and altering ocean currents such
that the sea surface temperatures warm, further weakening the trade
winds. This event has a great impact on the wind, sea surface temperature,
and precipitation patterns in the tropical Pacific. It has climatic effects
throughout the Pacific region and in many other parts of the world,
through global teleconnections. The cold phase of ENSO is called La Niña.
Emissions scenario
A plausible representation of the future development of emissions of
substances that are potentially radiatively active (e.g., greenhouse gases,
aerosols), based on a coherent and internally consistent set of assumptions
about driving forces (such as technological change, demographic and
socioeconomic development) and their key relationships. Concentration
scenarios, derived from emissions scenarios, are used as input to a climate
model to compute climate projections. In the IPCC 1992 Supplementary
Report, a set of emissions scenarios was presented, which were used as
a basis for the climate projections in the IPCC Second Assessment Report.
These emissions scenarios are referred to as the IS92 scenarios. In the
IPCC Special Report on Emissions Scenarios, new emissions scenarios, the
so-called SRES scenarios, were published. SRES scenarios (e.g., A1B,
A1FI, A2, B1, B2) are used as a basis for some of the climate projections
shown in Chapter 3 of this report.
Ensemble
A group of parallel model simulations used for climate projections.
Variation of the results across the ensemble members gives an estimate
of uncertainty. Ensembles made with the same model but different
initial conditions only characterize the uncertainty associated with
internal climate variability, whereas multi-model ensembles including
simulations by several models also include the impact of model
differences. Perturbed parameter ensembles, in which model parameters
are varied in a systematic manner, aim to produce a more objective
estimate of modeling uncertainty than is possible with traditional multi-
model ensembles.
Evapotranspiration
The combined process of evaporation from the Earth’s surface and
transpiration from vegetation.
Exposure
The presence of people; livelihoods; environmental services and resources;
infrastructure; or economic, social, or cultural assets in places that could
be adversely affected.
External forcing
External forcing refers to a forcing agent outside the climate system
causing a change in the climate system. Volcanic eruptions, solar variations,
and anthropogenic changes in the composition of the atmosphere and
land use change are external forcings.
Extratropical cyclone
Any cyclonic-scale storm that is not a tropical cyclone. Usually refers to
a middle- or high-latitude migratory storm system formed in regions of
large horizontal temperature variations. Sometimes called extratropical
storm or extratropical low.
Extreme coastal high water (also referred to as extreme sea level)
Extreme coastal high water depends on average sea level, tides, and
regional weather systems. Extreme coastal high water events are
usually defined in terms of the higher percentiles (e.g., 90th to 99.9th)
of a distribution of hourly values of observed sea level at a station for a
given reference period.
Extreme weather or climate event
See Climate extreme.
Famine
Scarcity of food over an extended period and over a large geographical
area, such as a country. Famines may be triggered by extreme climate
events such as drought or floods, but can also be caused by disease,
war, or other factors.
Flood
The overflowing of the normal confines of a stream or other body of
water, or the accumulation of water over areas that are not normally
submerged. Floods include river (fluvial) floods, flash floods, urban
floods, pluvial floods, sewer floods, coastal floods, and glacial lake
outburst floods.
Annex II Glossary of Terms
560
Frozen ground
Soil or rock in which part or all of the pore water is frozen. Perennially
frozen ground is called permafrost. Ground that freezes and thaws
annually is called seasonally frozen ground.
Glacial lake outburst flood (GLOF)
Flood associated with outburst of glacial lake. Glacial lake outburst
floods are typically a result of cumulative developments and occur (i)
only once (e.g., full breach failure of moraine-dammed lakes), (ii) for the
first time (e.g., new formation and outburst of glacial lakes), and/or (iii)
repeatedly (e.g., ice-dammed lakes with drainage cycles, or ice fall).
Glacier
A mass of land ice that flows downhill under gravity (through internal
deformation and/or sliding at the base) and is constrained by internal
stress and friction at the base and sides. A glacier is maintained by
accumulation of snow at high altitudes, balanced by melting at low
altitudes or discharge into the sea.
Global climate model (also referred to as
general circulation model, both abbreviated as GCM)
See Climate model.
Global surface temperature
The global surface temperature is an estimate of the global mean surface
air temperature. However, for changes over time, only anomalies, as
departures from a climatology, are used, most commonly based on the
area-weighted global average of the sea surface temperature anomaly
and land surface air temperature anomaly.
Governance
The way government is understood has changed in response to social,
economic, and technological changes over recent decades. There is a
corresponding shift from government defined strictly by the nation-state
to a more inclusive concept of governance, recognizing the contributions
of various levels of government (global, international, regional, local)
and the roles of the private sector, of nongovernmental actors, and of
civil society.
Greenhouse effect
Greenhouse gases effectively absorb thermal infrared radiation, emitted
by the Earth’s surface, by the atmosphere itself due to the same gases,
and by clouds. Atmospheric radiation is emitted to all sides, including
downward to the Earth’s surface. Thus, greenhouse gases trap heat
within the surface-troposphere system. This is called the greenhouse
effect. Thermal infrared radiation in the troposphere is strongly coupled
to the temperature of the atmosphere at the altitude at which it is
emitted. In the troposphere, the temperature generally decreases with
height. Effectively, infrared radiation emitted to space originates from
an altitude with a temperature of, on average, -19°C, in balance with
the net incoming solar radiation, whereas the Earth’s surface is kept at
a much higher temperature of, on average, 14°C. An increase in the
concentration of greenhouse gases leads to an increased infrared
opacity of the atmosphere and therefore to an effective radiation into
space from a higher altitude at a lower temperature. This causes a
radiative forcing that leads to an enhancement of the greenhouse effect,
the so-called enhanced greenhouse effect.
Greenhouse gas
Greenhouse gases are those gaseous constituents of the atmosphere,
both natural and anthropogenic, which absorb and emit radiation at
specific wavelengths within the spectrum of thermal infrared radiation
emitted by the Earth’s surface, by the atmosphere itself, and by clouds.
This property causes the greenhouse effect. Water vapor (H
2
O), carbon
dioxide (CO
2
), nitrous oxide (N
2
O), methane (CH
4
), and ozone (O
3
) are
the primary greenhouse gases in the Earth’s atmosphere. Moreover,
there are a number of entirely human-made greenhouse gases in the
atmosphere, such as the halocarbons and other chlorine- and bromine-
containing substances, dealt with under the Montreal Protocol. Besides
CO
2
, N
2
O, and CH
4
, the Kyoto Protocol deals with the greenhouse gases
sulfur hexafluoride (SF
6
), hydrofluorocarbons (HFCs), and perfluorocarbons
(PFCs).
Hazard
The potential occurrence of a natural or human-induced physical event
that may cause loss of life, injury, or other health impacts, as well as
damage and loss to property, infrastructure, livelihoods, service provision,
and environmental resources.
Heat wave (also referred to as extreme heat event)
A period of abnormally hot weather. Heat waves and warm spells have
various and in some cases overlapping definitions. See also Warm spell.
Holocene
The Holocene geological epoch is the latter of two Quaternary epochs,
extending from about 11.6 thousand years before present to and
including the present.
Human security
Human security can be said to have two main aspects. It means, first,
safety from such chronic threats as hunger, disease, and repression. And
second, it means protection from sudden and hurtful disruptions in the
patterns of daily life – whether in homes, in jobs, or in communities.
Such threats can exist at all levels of national income and development.
Hydrological cycle (also referred to as water cycle)
The cycle in which water evaporates from the oceans and the land
surface, is carried over the Earth in atmospheric circulation as water
vapor, condenses to form clouds, precipitates again as rain or snow, is
intercepted by trees and vegetation, provides runoff on the land surface,
infiltrates into soils, recharges groundwater, and/or discharges into
streams and flows out into the oceans, and ultimately evaporates again
from the oceans or land surface. The various systems involved in the
hydrological cycle are usually referred to as hydrological systems.
Glossary of Terms Annex II
561
Impacts
Effects on natural and human systems. In this report, the term ‘impacts’
is used to refer to the effects on natural and human systems of physical
events, of disasters, and of climate change.
Indian Ocean Dipole (IOD)
Large-scale interannual variability of sea surface temperature in the
Indian Ocean. This pattern manifests through a zonal gradient of tropical
sea surface temperature, which in one extreme phase in boreal autumn
shows cooling off Sumatra and warming off Somalia in the west,
combined with anomalous easterlies along the equator.
Insurance/reinsurance
A family of financial instruments for sharing and transferring risk among
a pool of at-risk households, businesses, and/or governments. See
Risk
transfer.
Landslide
A mass of material that has moved downhill by gravity, often assisted
by water when the material is saturated. The movement of soil, rock,
or debris down a slope can occur rapidly, or may involve slow, gradual
failure.
Land surface air temperature
The air temperature as measured in well-ventilated screens over land at
1.5 to 2 m above the ground.
Land use and land use change
Land use refers to the total of arrangements, activities, and inputs
undertaken in a certain land cover type (a set of human actions). The
term land use is also used in the sense of the social and economic
purposes for which land is managed (e.g., grazing, timber extraction,
and conservation). Land use change refers to a change in the use or
management of land by humans, which may lead to a change in land
cover. Land cover and land use change may have an impact on the
surface albedo, evapotranspiration, sources and sinks of greenhouse
gases, or other properties of the climate system and may thus have
radiative forcing and/or other impacts on climate, locally or globally.
Lapse rate
The rate of change of an atmospheric variable, usually temperature,
with height. The lapse rate is considered positive when the variable
decreases with height.
Latent heat flux
The flux of heat from the Earth’s surface to the atmosphere that is
associated with evaporation or condensation of water vapor at the
surface; a component of the surface energy budget.
Likelihood
A probabilistic estimate of the occurrence of a single event or of an
outcome, for example, a climate parameter, observed trend, or projected
change lying in a given range. Likelihood may be based on statistical or
modeling analyses, elicitation of expert views, or other quantitative
analyses.
Local disaster risk management (LDRM)
The process in which local actors (citizens, communities, government,
non-profit organizations, institutions, and businesses) engage in and
have ownership of the identification, analysis, evaluation, monitoring,
and treatment of disaster risk and disasters, through measures that
reduce or anticipate hazard, exposure, or vulnerability; transfer risk;
improve disaster response and recovery; and promote an overall
increase in capacities. LDRM normally requirescoordination with and
support from external actors at the regional, national, or international
levels. Community-based disaster risk management is a subset of
LDRM where community members and organizations are in the center
of decisionmaking.
Mass movement
Mass movement in the context of mountainous phenomena refers to
different types of mass transport processes including landslides,
avalanches, rock fall, or debris flows.
Mean sea level
Sea level measured by a tide gauge with respect to the land upon which
it is situated. Mean sea level is normally defined as the average relative sea
level over a period, such as a month or a year, long enough to average out
transients such as waves and tides. See
Sea level change.
Meridional overturning circulation (MOC)
Meridional (north-south) overturning circulation in the ocean quantified
by zonal (east-west) sums of mass transports in depth or density layers.
In the North Atlantic, away from the subpolar regions, the MOC (which
is in principle an observable quantity) is often identified with the
thermohaline circulation, which is a conceptual interpretation. However,
it must be borne in mind that MOC can also include shallower, wind-
driven overturning cells such as occur in the upper ocean in the tropics
and subtropics, in which warm (less dense) waters moving poleward are
transformed to slightly denser waters and subducted equatorward at
deeper levels.
Mitigation (of disaster risk and disaster)
The lessening of the potential adverse impacts of physical hazards
(including those that are human-induced) through actions that reduce
hazard, exposure, and vulnerability.
Mitigation (of climate change)
A human intervention to reduce the sources or enhance the sinks of
greenhouse gases.
Modes of climate variability
Natural variability of the climate system, in particular on seasonal and
longer time scales, predominantly occurs with preferred spatial patterns
Annex II Glossary of Terms
562
and time scales, through the dynamical characteristics of the atmospheric
circulation and through interactions with the land and ocean surfaces.
Such patterns are often called regimes, modes, or teleconnections.
Examples are the North Atlantic Oscillation (NAO), the Pacific-North
American pattern (PNA), the El Niño-Southern Oscillation (ENSO), the
Northern Annular Mode (NAM; previously called the Arctic Oscillation,
AO), and the Southern Annular Mode (SAM; previously called the
Antarctic Oscillation, AAO).
Monsoon
A monsoon is a tropical and subtropical seasonal reversal in both the
surface winds and associated precipitation, caused by differential
heating between a continental-scale land mass and the adjacent ocean.
Monsoon rains occur mainly over land in summer.
Nonlinearity
A process is called nonlinear when there is no simple proportional
relation between cause and effect. The climate system contains many
such nonlinear processes, resulting in a system with a potentially very
complex behavior. Such complexity may lead to abrupt climate change.
See also
Predictability.
North Atlantic Oscillation (NAO)
The North Atlantic Oscillation consists of opposing variations in barometric
pressure near Iceland and near the Azores. It therefore corresponds to
fluctuations in the strength of the main westerly winds across the
Atlantic into Europe, and thus to fluctuations in the embedded cyclones
with their associated frontal systems.
Northern Annular Mode (NAM)
A winter fluctuation in the amplitude of a pattern characterized by low
surface pressure in the Arctic and strong mid-latitude westerlies. NAM
has links with the northern polar vortex into the stratosphere. Its
pattern has a bias to the North Atlantic and has a large correlation with
the North Atlantic Oscillation.
Pacific Decadal Oscillation (PDO)
The pattern and time series of the first empirical orthogonal function of
sea surface temperature over the North Pacific north of 20°N. PDO
broadened to cover the whole Pacific Basin is known as the Inter-decadal
Pacific Oscillation (IPO). The PDO and IPO exhibit virtually identical
temporal evolution.
Parameterization
In climate models, this term refers to the technique of representing
processes that cannot be explicitly resolved at the spatial or temporal
resolution of the model (sub-grid scale processes) by relationships
between model-resolved larger-scale flow and the area- or time-averaged
effect of such sub-grid scale processes.
Percentile
A percentile is a value on a scale of 100 that indicates the percentage
of the data set values that is equal to or below it. The percentile is often
used to estimate the extremes of a distribution. For example, the 90th
(10th) percentile may be used to refer to the threshold for the upper
(lower) extremes.
Permafrost
Ground (soil or rock and included ice and organic material) that remains
at or below 0°C for at least 2 consecutive years.
Predictability
The extent to which future states of a system may be predicted based
on knowledge of current and past states of the system.
Probability density function (PDF)
A probability density function is a function that indicates the relative
chances of occurrence of different outcomes of a variable. The function
integrates to unity over the domain for which it is defined and has the
property that the integral over a sub-domain equals the probability that
the outcome of the variable lies within that sub-domain. For example,
the probability that a temperature anomaly defined in a particular way
is greater than zero is obtained from its PDF by integrating the PDF
over all possible temperature anomalies greater than zero. Probability
density functions that describe two or more variables simultaneously
are similarly defined.
Projection
A projection is a potential future evolution of a quantity or set of
quantities, often computed with the aid of a model. Projections are
distinguished from predictions in order to emphasize that projections
involve assumptions concerning, for example, future socioeconomic and
technological developments that may or may not be realized, and are
therefore subject to substantial uncertainty. See also
Climate projection
and Climate prediction.
Proxy climate indicator
A proxy climate indicator is a local record that is interpreted, using
physical and biophysical principles, to represent some combination of
climate-related variations back in time. Climate-related data derived in
this way are referred to as proxy data. Examples of proxies include
pollen analysis, tree ring records, characteristics of corals, and various
data derived from ice cores. The term ‘proxy’ can also be used to refer
to indirect estimates of present-day conditions, for example, in the
absence of observations.
Radiative forcing
Radiative forcing is the change in the net, downward minus upward,
irradiance (expressed in W m
–2
) at the tropopause due to a change in
an external driver of climate change, such as, for example, a change in
the concentration of carbon dioxide or the output of the Sun. Radiative
forcing is computed with all tropospheric properties held fixed at their
unperturbed values, and after allowing for stratospheric temperatures,
if perturbed, to readjust to radiative-dynamical equilibrium. Radiative
Glossary of Terms Annex II
563
forcing is called instantaneous if no change in stratospheric temperature
is accounted for. For the purposes of this report, radiative forcing is
further defined as the change relative to the year 1750 and, unless
otherwise noted, refers to a global and annual average value. Radiative
forcing is not to be confused with cloud radiative forcing, a similar
terminology for describing an unrelated measure of the impact of
clouds on the irradiance at the top of the atmosphere.
Reanalysis
Reanalyses are atmospheric and oceanic analyses of temperature, wind,
current, and other meteorological and oceanographic quantities, created
by processing past meteorological and oceanographic data using fixed
state-of-the-art weather forecasting models and data assimilation
techniques. Using fixed data assimilation avoids effects from the
changing analysis system that occur in operational analyses. Although
continuity is improved, global reanalyses still suffer from changing
coverage and biases in the observing systems.
Relative sea level
See Mean sea level.
Resilience
The ability of a system and its component parts to anticipate, absorb,
accommodate, or recover from the effects of a hazardous event in a
timely and efficient manner, including through ensuring the preservation,
restoration, or improvement of its essential basic structures and functions.
Return period
An estimate of the average time interval between occurrences of an
event (e.g., flood or extreme rainfall) of (or below/above) a defined size
or intensity.
Return value
The highest (or, alternatively, lowest) value of a given variable, on
average occurring once in a given period of time (e.g., in 10 years).
Risk transfer
The process of formally or informally shifting the financial consequences
of particular risks from one party to another whereby a household,
community, enterprise, or state authority will obtain resources from
the other party after a disaster occurs, in exchange for ongoing or
compensatory social or financial benefits provided to that other party.
Runoff
That part of precipitation that does not evaporate and is not transpired,
but flows through the ground or over the ground surface and returns to
bodies of water. See Hydrological cycle.
Scenario
A plausible and often simplified description of how the future may
develop based on a coherent and internally consistent set of assumptions
about driving forces and key relationships. Scenarios may be derived
from projections, but are often based on additional information from
other sources, sometimes combined with a narrative storyline. See also
Climate scenario and Emissions scenario.
Sea level change
Changes in sea level, globally or locally, due to (i) changes in the shape
of the ocean basins, (ii) changes in the total mass and distribution of
water and land ice, (iii) changes in water density, and (iv) changes in
ocean circulation. Sea level changes induced by changes in water
density are called steric. Density changes induced by temperature
changes only are called thermosteric, while density changes induced by
salinity changes are called halosteric. See also
Mean sea level.
Sea surface temperature (SST)
The sea surface temperature is the temperature of the subsurface bulk
temperature in the top few meters of the ocean, measured by ships,
buoys, and drifters. From ships, measurements of water samples in
buckets were mostly switched in the 1940s to samples from engine intake
water. Satellite measurements of skin temperature (uppermost layer; a
fraction of a millimeter thick) in the infrared or the top centimeter or so
in the microwave are also used, but must be adjusted to be compatible
with the bulk temperature.
Sensible heat flux
The flux of heat from the Earth’s surface to the atmosphere that is not
associated with phase changes of water; a component of the surface
energy budget.
Significant wave height
The average height of the highest one-third of the wave heights (trough
to peak) from sea and swell occurring in a particular time period.
Soil moisture
Water stored in or at the land surface and available for evapotranspiration.
Southern Annular Mode (SAM)
The fluctuation of a pattern like the Northern Annular Mode, but in the
Southern Hemisphere.
SRES scenarios
See Emissions scenario.
Storm surge
The temporary increase, at a particular locality, in the height of the sea
due to extreme meteorological conditions (low atmospheric pressure
and/or strong winds). The storm surge is defined as being the excess
above the level expected from the tidal variation alone at that time and
place.
Storm tracks
Originally, a term referring to the tracks of individual cyclonic weather
systems, but now often generalized to refer to the regions where the
Annex II Glossary of Terms
564
main tracks of extratropical disturbances occur as sequences of low
(cyclonic) and high (anticyclonic) pressure systems.
Streamflow
Water flow within a river channel, for example, expressed in m
3
s
-1
. A
synonym for river discharge.
Subsidiarity
The principle that decisions of government (other things being equal) are
best made and implemented, if possible, at the lowest most decentralized
level closest to the citizen. Subsidiarity is designed to strengthen
accountability and reduce the dangers of making decisions in places
remote from their point of application. The principle does not necessarily
limit or constrain the action of higher orders of government, it merely
counsels against the unnecessary assumption of responsibilities at a
higher level.
Surface temperature
See Global surface temperature, Land surface air temperature, and Sea
surface temperature
.
Sustainable development
Development that meets the needs of the present without compromising
the ability of future generations to meet their own needs.
Transpiration
The evaporation of water vapor from the surfaces of leaves through
stomata.
Transformation
The altering of fundamental attributes of a system (including value
systems; regulatory, legislative, or bureaucratic regimes; financial
institutions; and technological or biological systems).
Tropical cyclone
The general term for a strong, cyclonic-scale disturbance that originates
over tropical oceans. Distinguished from weaker systems (often named
tropical disturbances or depressions) by exceeding a threshold wind
speed. A tropical storm is a tropical cyclone with one-minute average
surface winds between 18 and 32 m s
-1
. Beyond 32 m s
-1
, a tropical
cyclone is called a hurricane, typhoon, or cyclone, depending on
geographic location.
Uncertainty
An expression of the degree to which a value or relationship is
unknown. Uncertainty can result from lack of information or from
disagreement about what is known or even knowable. Uncertainty may
originate from many sources, such as quantifiable errors in the data,
ambiguously defined concepts or terminology, or uncertain projections
of human behavior. Uncertainty can therefore be represented by
quantitative measures, for example, a range of values calculated by
various models, or by qualitative statements, for example, reflecting the
judgment of a team of experts. See also
Likelihood and Confidence.
Urban heat island
The relative warmth of a city compared with surrounding rural areas,
associated with changes in runoff, the concrete jungle effects on heat
retention, changes in surface albedo, changes in pollution and aerosols,
and so on.
Vulnerability
The propensity or predisposition to be adversely affected.
Warm days/warm nights
Days where maximum temperature, or nights where minimum
temperature, exceeds the 90th percentile, where the respective
temperature distributions are generally defined with respect to the
1961-1990 reference period.
Warm spell
A period of abnormally warm weather. Heat waves and warm spells
have various and in some cases overlapping definitions. See also Heat
wave.
Glossary of Terms Annex II
565
Acronyms
III
ANNEX
566
AAO Antarctic Oscillation
ADAPT Assessment & Design for Adaptation to Climate
Change: A Prototype Tool
AMO Atlantic Multi-decadal Oscillation
AO Arctic Oscillation
AR5 Fifth Assessment Report
CAPE Convective Available Potential Energy
CAT catastrophic risk
CBA cost-benefit analysis or community-based adaptation
CBD Convention on Biological Diversity
CBDR common but differentiated responsibilities and
respective capabilities
CBO community-based organization
CCA climate change adaptation
CCRIF Caribbean Catastrophe Risk Insurance Facility
CCSP Climate Change Science Program (US)
CDD Consecutive Dry Days
CDM Clean Development Mechanism
CEE Centre for Environment Education
CEI Climate Extremes Index
C-ERA-40 Corrected ERA-40 reanalysis
CFR case fatality rate
CH
4
methane
CMIP3 Coupled Model Intercomparison Project 3
CO
2
carbon dioxide
COP Conference of the Parties
CPP cyclone preparedness program
CRED Centre for Research on the Epidemiology of Disasters
CSA Canadian Standards Association
CSO civil society organization
CSR corporate social responsibility
DDI Disaster Deficit Index
DFID Department for International Development (UK)
DJF December-January-February
DRM disaster risk management
DRR disaster risk reduction
DRRM disaster risk reduction management
EbA ecosystem-based adaptation
EBRD European Bank for Reconstruction and Development
EDI Ethiopia Drought Index
ELF Emergency Liquidity Facility
EM-DAT Emergency Events Database
ENSO El Niño-Southern Oscillation
ERA-40 European Centre for Medium Range Weather Forecasts
40-year reanalysis
EVT extreme value theory
EWS early warning system
FAO Food and Agriculture Organization
FONDEN Fund for Natural Disasters
GAR Global Assessment Report on Disaster Risk Reduction
GCM global climate model
GDP gross domestic product
GEC global environmental change
GEF Global Environment Facility
GFCS Global Framework on Climate Services
GFDRR Global Facility for Disaster Reduction and Recovery
GHG greenhouse gas
GIS geographic information system
GLOF glacial lake outburst flood
GNCSODR Global Network of Civil Society Organisations for
Disaster Reduction
GPS Global Positioning System
GSDI Global Spatial Data Infrastructure
H
2
O water
HARS Heat Action Response System
HDI Human Development Index
HEP hydroelectric power
HFA Hyogo Framework for Action
HFC hydrofluorocarbon
HWDI Heat Wave Duration Index
HWS Heat Warning System
IADB Inter-American Development Bank
IAM integrated assessment model
ICSU International Council for Science
ICT information and communication technology
ICZM integrated coastal zone management
IDMC Internal Displacement Monitoring Centre
IDNDR International Decade for Natural Disaster Reduction
IDP internally displaced person
IDRL International Disaster Response Law
IHL international humanitarian law
IOD Indian Ocean Dipole
IPO Inter-decadal Pacific Oscillation
IRDR Integrated Research on Disaster Risk program
ISSC International Social Science Council
ITCZ Inter-Tropical Convergence Zone
IWRM integrated water resource management
JJA June-July-August
LA RED Red de Estudios Sociales en Prevención de Desastres en
América Latina
LDC least-developed country
LDCF Least Developed Countries Fund
LDRM local disaster risk management
LEED Leadership in Energy and Environmental Design
LIDAR Light Detection and Ranging
MDGs Millennium Development Goals
MFI micro-finance institution
MJO Madden-Julian Oscillation
MLP multi-level perspective
MME Multi-Model Ensemble (CMIP3)
MOC meridional overturning circulation
MPBI Indonesian Society for Disaster Management
MSLP mean sea level pressure
N
2
O nitrous oxide
Acronyms Annex III
567
NAM Northern Annular Mode
NAO North Atlantic Oscillation
NAPA National Adaptation Programme of Action
NaTech Natural Hazard Triggering a Technological Disaster
NDMO National Disaster Management Office
NECJOGHA Network of Climate Journalists of the Greater Horn of
Africa
NGO nongovernmental organization
NHC National Hurricane Committee
NIDM National Disaster Management Institute
NMHS national meteorological and hydrological service
NTR non-tide residuals
NU Nunavut
NWP Nairobi Work Programme
NWT Northwest Territories
O
3
ozone
OCHA United Nations Office for the Coordination of
Humanitarian Affairs
ODA official development assistance
OECD Organisation for Economic Co-operation and
Development
OFDA Office of Foreign Disaster Assistance
OLR outgoing longwave radiation
PAR pressure and release
PDF probability density function
PDO Pacific Decadal Oscillation
PDSI Palmer Drought Severity Index
PESETA Projection of Economic impacts of climate change in
Sectors of the European Union based on boTtom-up
Analysis
PFC perfluorocarbon
PICs Pacific Island Countries and Territories
PNA Pacific North American pattern
POPs persistent organic pollutants
PPEA Precipitation Potential Evaporation Anomaly
PPP public-private partnership
Pr precipitation
PSNP Productive Safety Net Programme
PTSD post-traumatic stress disorder
PVI Prevalent Vulnerability Index
RAC Regional Adaptation Collaborative
RANET RAdio and InterNET
RCM regional climate model
REDD reduced carbon emissions from deforestation and forest
degradation
REDD+ reduced carbon emissions from deforestation and forest
degradation, maintaining/enhancing carbon stocks, and
promoting sustainable forest management
RMI Republic of the Marshall Islands
SAM Southern Annular Mode
SAMS South American Monsoon System
SCCF Special Climate Change Fund
SDLE Prepare, Stay and Defend, or Leave Early
SDMP School Disaster Management Plans
SECO Swiss State Secretariat for Economic Affairs
SF
6
sulfur hexafluoride
SHELDUS Spatial Hazard Events and Losses Database for the
United States
SIDS small island developing states
SIS small island states
SMA soil moisture anomaly
SMEs small- and medium-sized enterprises
SOI Southern Oscillation Index
SPA Strategic Priority ‘Piloting an Operational Approach to
Adaptation’
SPEI Standardized Precipitation-Evapotranspiration Index
SPI Standard Precipitation Index
SRES Special Report on Emissions Scenarios
SST sea surface temperature
SWH significant wave height
UN United Nations
UNCCD United Nations Convention to Combat Desertification
UNDP United Nations Development Programme
UNFCCC United Nations Framework Convention on Climate
Change
UNISDR United Nations International Strategy for Disaster
Reduction
WDSI Warm Spell Duration Index
WFP World Food Programme
WHO World Health Organization
WMO World Meteorological Organization
YT Yukon Territory
Annex III Acronyms
568
Acronyms Annex III
569
List of Major IPCC Reports
IV
ANNEX
570
Climate Change: The IPCC Scientific Assessment
Report of the IPCC Scientific Assessment Working Group
1990
Climate Change: The IPCC Impacts Assessment
Report of the IPCC Impacts Assessment Working Group
1990
Climate Change: The IPCC Response Strategies
Report of the IPCC Response Strategies Working Group
1990
Climate Change 1992: The Supplementary Report
to the IPCC Scientific Assessment
Report of the IPCC Scientific Assessment Working Group
1992
Climate Change 1992: The Supplementary Report
to the IPCC Impacts Assessment
Report of the IPCC Impacts Assessment Working Group
1992
Climate Change: The IPCC 1990 and 1992 Assessments –
IPCC First Assessment Report Overview and Policymaker
Summaries, and 1992 IPCC Supplement
1992
Climate Change 1994: Radiative Forcing of Climate Change
and an Evaluation of the IPCC IS92 Emission Scenarios
IPCC Special Report
1994
Climate Change 1995: The Science of Climate Change
Contribution of Working Group I
to the IPCC Second Assessment Report
1996
Climate Change 1995: Impacts, Adaptations, and Mitigation
of Climate Change: Scientific-Technical Analyses
Contribution of Working Group II
to the IPCC Second Assessment Report
1996
Climate Change 1995: Economic and Social Dimensions of
Climate Change
Contribution of Working Group III
to the IPCC Second Assessment Report
1996
Climate Change 1995: IPCC Second Assessment Synthesis
of Scientific-Technical Information Relevant to Interpreting
Article 2 of the UN Framework Convention on Climate Change
1996
Technologies, Policies, and Measures
for Mitigating Climate Change
IPCC Technical Paper I
1996
An Introduction to Simple Climate Models
used in the IPCC Second Assessment Report
IPCC Technical Paper II
1997
Stabilization of Atmospheric Greenhouse Gases:
Physical, Biological, and Socio-Economic Implications
IPCC Technical Paper III
1997
Implications of Proposed CO
2
Emissions Limitations
IPCC Technical Paper IV
1997
The Regional Impacts of Climate Change
IPCC Special Report
1998
Aviation and the Global Atmosphere
IPCC Special Report
1999
Methodological and Technological Issues in Technology Transfer
IPCC Special Report
2000
Land Use, Land-Use Change, and Forestry
IPCC Special Report
2000
Emissions Scenarios
IPCC Special Report
2000
Climate Change 2001: The Scientific Basis
Contribution of Working Group I
to the IPCC Third Assessment Report
2001
Climate Change 2001: Impacts, Adaptation, and Vulnerability
Contribution of Working Group II
to the IPCC Third Assessment Report
2001
Climate Change 2001: Mitigation
Contribution of Working Group III
to the IPCC Third Assessment Report
2001
List of Major IPCC Reports Annex IV
571
Climate Change 2001: IPCC Third Assessment Synthesis Report
2001
Climate Change and Biodiversity
IPCC Technical Paper V
2002
Safeguarding the Ozone Layer and the Global Climate System:
Issues Related to Hydrofluorocarbons and Perfluorocarbons
IPCC Special Report
2005
Carbon Dioxide Capture and Storage
IPCC Special Report
2005
Climate Change 2007: The Physical Science Basis
Contribution of Working Group I
to the IPCC Fourth Assessment Report
2007
Climate Change 2007: Impacts, Adaptation, and Vulnerability
Contribution of Working Group II
to the IPCC Fourth Assessment Report
2007
Climate Change 2007: Mitigation of Climate Change
Contribution of Working Group III
to the IPCC Fourth Assessment Report
2007
Climate Change 2007: Synthesis Report
2008
Climate Change and Water
IPCC Technical Paper VI
2008
Renewable Energy Sources and Climate Change Mitigation
IPCC Special Report
2011
Managing the Risks of Extreme Events and Disasters
to Advance Climate Change Adaptation
IPCC Special Report
2012
Enquiries: IPCC Secretariat, c/o World Meteorological Organization, 7 bis, Avenue
de la Paix, Case Postale 2300, CH - 1211 Geneva 2, Switzerland
Annex IV List of Major IPCC Reports
572
List of Major IPCC Reports Annex IV
573
Index
A
Abrupt climate change[*], 122, 458
Access
to resources, 454-456
to technology, 447-448
ADAPT (Assessment & Design for Adaptation
Tool), 417
Adaptation[*], 6, 36, 443
barriers to, 451-453
community-based, 295, 300, 321, 322
continuum/process, 323, 324
coping and, 50-56
dimensions of, 51
DRM and, 37-39, 47-48, 398, 408-409, 443-462
DRM as, 443-450
DRR and, 237, 396-397, 423-424, 443-446
ecosystem-based, 343, 352, 357, 370-371, 371,
445-446
effectiveness and tradeoffs, 439-440, 448-450
Adaptation, inequalities, 10, 294, 313-317
Adaptation[*]
integration with DRM, 11, 28, 50, 355-357,
396-397, 425-427, 439, 450-454, 469-471
interactions with DRM and mitigation, 20, 440,
458-462
international level, 396-397, 411
learning and, 27-28, 50, 53-56
local knowledge and, 293, 311-312
local level, 293-295, 319-320, 319
mainstreaming, 348, 355-357
maladaptative actions, 55, 368
management options, 16-20, 18-19
past experience and, 10-11, 38-39
portfolio of approaches, 17, 295
robust, 56
sector-based, 352-354, 357
sustainability and, 444
synergies with DRM, 48-50, 68, 357, 469-471
technology and, 414-415, 415
UNFCCC commitments on, 407
Adaptation costs, 264-269, 412-414
assessment of, 273-274, 273, 274
in developing countries, 412, 412
evaluation of, 236, 266-269
funding for, 406-407, 411
Adaptation deficit, 265
Adaptation Fund, 406-407, 411, 413
Adaptation limits, 319-320, 319
Adaptation options, 352-354
Adaptation planning, 349-357, 443-444
Adaptive capacity[*], 6-7, 33, 72-76, 73, 443
local, 294, 308-313, 315
post-disaster recovery/reconstruction and, 293, 439
Adaptive management, 343, 377-379, 467
Adjustments, 450-454
Afforestation, 352, 355
Africa, 253-254
adaptation costs, 274, 274
agricultural impacts, 246-247, 253
drought, 19, 170, 171, 172, 174, 175, 253
floods, 176, 253-254, 297
food security, 368
monsoons, 153, 154
precipitation, 142, 143, 145-146, 147, 193, 199,
253-254
sand/dust storms, 190, 254
temperature, 134, 135, 139-140, 141, 193,
199
Africa Union, 401
Age/aging, 234-235, 314, 315
Agriculture
crop insurance, 323
DRM/adaptation options, 352
food security and, 368-369
impacts, 235, 246-247, 272-273
regional impacts, 253, 259
temperature impacts, 247, 255
vulnerability/exposure, 235
Alaska, 191, 197
drought, 170
precipitation, 145-146, 191, 197
temperature, 135, 139-140, 191, 197
Alpine regions, 251
Amazon region, 194, 200, 371
Antarctica, 261-263
ice sheet, 179, 188
Anthropogenic influences
changes in extremes, 9, 112, 125-126, 268
changes in precipitation, 143-144, 149
changes in temperature, 135-136, 141
climate change, 40
Aral Sea basin, 239
Arctic region, 261-263
Arid environments, 143, 149, 167, 190
case studies, 498-502
Asia, 254-255
adaptation costs, 273, 274
adaptation technologies, 415, 415
Disaster Reduction Hyperbase, 415
drought, 171, 255
floods, 176, 177, 254-255
monsoons, 154
precipitation, 142-143, 145-146, 147-148,
194-195, 201-202
sand/dust storms, 190
temperature, 135, 135, 139-140, 141, 194-195,
201-202, 255
wildfires, 255
Assets, 315-316
Assistance
humanitarian, 10, 293, 299-300
international, 400
Atlantic Multi-decadal Oscillation (AMO)[*],
153, 157
Atolls, 263-264, 263
community relocation, 300-301
Attribution
of changes in extremes, 112, 119-120, 125-128, 268
of changes in precipitation, 143-144
of changes temperature, 135-136, 141
of extreme events, 40
human-induced changes, 9, 12, 125-126
of impacts, 9, 268-269
multi-step, 127-128
single-step, 127-128
Auckland, 260, 261
Australia, 260-261
case study (Victoria), 496-498
coastal impacts, 183, 185, 186, 261
drought, 170, 171, 172, 174, 261-262
heat waves, 496-498
land use change, 261
precipitation, 142, 143, 145-146, 147, 148-149,
195, 202
, 261
temperature, 134, 135, 135, 139-140, 141, 195, 202
waves, 182
wildfire, 239, 261, 496-498
wind, 150
Avalanches, 187-188, 189
See also Landslides
B
Bali Action Plan, 398, 403, 407, 414
Bangladesh
cyclones, 502-505, 503
disaster preparedness, 305, 465
early warning system, 405
floods, 254, 461
Barriers to adaptation, 451-453
Baseline/reference[*], 117, 364
Belmont Challenge, 425-426
Bilateral/multilateral agencies, 348-349
Biodiversity, 343, 370, 371, 410, 446
Bottom-up mechanisms/approach, 266-267, 346,
350-351, 350, 396, 427
Brazil, 194, 200
dams, 304
drought, 172, 174, 175
precipitation, 145-146, 148, 194, 200
public awareness initiative, 527-528
temperature, 139-140, 194, 200
Building codes, 305, 317, 353
LEED standards, 461
national, 347, 367-368
Building measures, 304-305
wind-resistant building, 416
C
Cambodia, 526
Canada, 258-260
case study, 514-517
coastal impacts, 183
CSA (Infrastructure in Permafrost) guide, 367
drought, 170
floods, 176, 177
forest fires, 252, 259
precipitation, 142, 143, 145-146, 148, 191, 197
temperature, 135, 138, 139-140, 191, 197
wind, 150
Cancun Agreements, 397, 398, 403, 408, 413, 414
Capabilities, 33
Capacity[*], 33, 72-76, 73
availability and limitations, 454-456
Capacity building, 33, 308-313, 415
Capacity needs, 74-76
Carbon dioxide (CO
2
)[*], 244
Carbon sequestration, 307, 370
Carbon sink/source, 244
Caribbean, 184, 263
adaptation costs, 273, 274
adaptation strategies, 371
coastal impacts, 183
hurricanes, 19
insurance/risk pool, 372, 400, 419, 524-525
574
Index
Index | Key
Terms defined in the glossary are
marked with an asterisk(*). Bold page
numbers indicate page spans for
entire chapters. Italicized page
numbers denote tables, figures, and
boxed material.
575
resilience building, 378
tourism impacts, 251, 251
Case studies, 487-542
synthesis of, 491-492, 529-530
Caste/class, 454
Catastrophic risk bonds, 419-420, 465
Catchments[*], 41, 113, 177-178, 242
Cayman Islands, 378
Central America, 255-256
disaster losses, 256
drought, 172, 175
hurricanes, 503, 504-505
precipitation, 142, 145-146, 191, 196-197, 201
temperature, 135, 139-140, 191, 196-197, 201
Central Asia, 195
Children, 314, 454, 455
China
adaptation technologies, 415
drought, 174
floods, 254-255
national disaster management, 374
precipitation, 143
sand/dust storms, 190
temperature, 135
tropical cyclones, 254
vulnerability/exposure, 347-348
wind, 150, 152
Cholera, 316, 410, 507-510
Circulation (Walker/Hadley), 149, 151, 154
Cities, 248, 317, 460-461
DRM/adaptation options, 353
megacities, 294, 317, 510-512
See also Urban areas
Civil society organizations, 348, 404, 409-410
Clausius-Clapeyron relationship[*], 126, 143-144
Clean development mechanisms, 411, 413
Climate[*]
global mean, 121-122
is it becoming more extreme? (FAQ), 124-125
Climate change[*], 25-64, 444
abrupt, 122
attribution of impacts to, 9, 268-269
changes in extremes and, 7, 111, 115, 127
concepts/definitions, 30-37
defined, 5, 29
DRM and, 27-28, 37-39, 375-380, 376-377
DRM challenges, 46-47
Climate change adaptation. See Adaptation
Climate Change Green Fund, 397
Climate change mitigation. See Mitigation
Climate events
categories of, 115
disasters and, 115-118
interactions of, 238-239, 239
Climate extremes[*]
, 65-290
changes in, 8-9, 109-230
context, 4-7, 4
costs of, 264-274
definition and analysis, 5, 111, 116-117, 237
factors and confidence in, 111-112, 120-121
impacts on humans and ecosystems, 231-290
impacts on physical environment, 8-9, 167-190
indices of, 116-117, 125
managing changing risks, 16-20
methods and requirements, 122-133
natural and socioeconomic systems, 237-239
observed changes, 7-9, 111-112, 119-120, 133-152
past experience with, 10-11
phenomena related to, 119, 152-166
projected changes, 11-16, 112-114, 119-120,
133-152
regional and global climate, 121-122
regionally based impacts, 252-264
unprecedented, 7, 111
vulnerability/exposure and, 65-108, 239-264
See also Extreme events
Climate feedback[*], 112, 118-120
Climate information, 421-422
Climate models[*], 13, 112-113, 128-133, 147
ensembles, 131-133
modeling tools, 464
planning approach, 350-351, 350
See also Projections
Climate modes[*], 113, 155-158
observed and projected changes, 15-16, 119
uncertainty in projections of, 113
See also El Niño-Southern Oscillation
Climate projections[*], 119-120
Climate scenarios[*]. See Scenarios
Climate services, 409, 422
Climate variability, 7, 115, 155-158
Coastal erosion, 113, 182-183, 185, 186
Coastal floods, 259-260
Coastal impacts, 182-186
case study, 510-512
DRM/adaptation options, 352
exposure and, 249, 258
extreme high water, 9, 15, 18, 113, 178, 182
observed changes, 120, 183
projected changes, 113, 120, 183-186
regional impacts, 254-255, 256-257, 259-260, 263
sea level and, 182, 183, 184, 185
waves and, 180-182
Coastal inundation, 18, 113, 182-183, 184-185,
185, 248, 249
Coastal settlements, 7-8, 235, 248, 249, 258, 460
megacities (case study), 510-512
Cold climate regions (case study), 514-517
Cold days/cold nights[*], 8, 116, 134, 135,
137,
138, 141
regional projections, 191-202
Cold spell, 134
Collective action, 309-310, 309, 321
Colombia, 519-522
Communication, 302-304
gaps, 425
national systems, 349, 376
of risk, 17, 67, 95, 294, 302-304, 303, 376
technologies, 422-423
Community-based adaptation, 295, 300, 321, 322
Community-based DRM[*], 308-313
Community-based DRR, 310, 321
Community-based organizations, 348
Complex Bayesian framework, 133
Complex systems, 46, 48, 53
Complexity, 27, 42-44, 53
Compound events, 118
Conference of the Parties (COP), 406
Confidence[*], 8, 21, 112, 120-121, 132
Conflict/warfare, 297
Contingent credit, 523
Contingent liabilities, 361, 361
Cooperation, 401, 402
Coordination (across scales and sectors), 342,
356, 358-360, 360, 439
international, 408-409
national, 358-360, 360
Copenhagen Accord, 408, 413
Coping[*], 33, 50-56, 450-451
dimensions of, 51
local, 294, 298-301
Coping capacity[*], 6-7, 51-53, 72-76, 73
Coping range, 52-53
Coral reefs, 185, 263
Corporate social responsibility, 347
Cost-benefit analysis, 267-268
Costa Rica, 252
Costs, 264-274
adaptation costs, 264-265, 266-269, 273-274,
273, 274
assessment of, 269-274
damage costs, 264
direct costs, 264, 266, 317
DRM costs, 317-319
economic costs, 267
evaluating (estimating), 236, 266-269
financial costs, 267
framing of, 264-265
global and regional costs, 269-271, 270, 271, 272
indirect costs, 264, 266, 317
intangibles, 264, 266, 317
investment costs, 267
local DRM costs, 317-319
methods for evaluating, 266-269
observed increase in, 270-271, 271
potential damage costs, 264
regional costs, 256, 270, 270
residual damage costs, 264
top-down vs bottom-up approach, 266-267
uncertainty and, 274
Crop insurance, 524
Cultural dimensions, 7, 84-85
Cultural heritage loss, 317, 318, 319
Cultural norms and values, 84-85, 309, 310-311
Culture of safety, 362-366
Cumulative impacts, 6-7, 38, 67, 69
Customary law, 402
Cyclones, 41, 158-166
case study, 502-505
extratropical[*], 119, 163-166
economic losses, 272, 272
observed changes, 163-166
poleward shift of, 8, 164-166
projected changes, 13, 272, 272
tropical[*], 119, 150, 151, 158-163, 502-505
economic losses, 235, 254, 271, 272
exposure to, 240
forecasting/warnings, 416-417
impacts, 248
observed changes, 159-161, 163
projected changes, 13, 161-163, 271, 272
regional impacts, 254
D
Dams, 176, 304, 415
Databases, 364, 415, 423-424, 464
Debt, 93
Decentralization, 28, 46, 312-313, 360, 464
Decisionmaking, 67, 342
decentralized, 28
local, 305-306, 308, 310
maps of, 325
support tools, 463-466
tradeoffs in, 448
Definitions, 5
See also Glossary
Deforestation, 238, 252, 446
Index
576
Demographic changes, 80, 234-235
Desertification, 190, 402
DesInventar database, 423-424
Developed countries
economic/disaster losses in, 9, 234, 265
mainstreaming, 356-357
urban capacities, 460
vulnerability, 78, 265
Developing countries
adaptation costs, 412, 412
adaptation funding, 406-407, 411, 413
disaster preparedness, 369
economic/disaster losses in, 9, 234, 265, 269-270,
400
infrastructure, 367-368
insurance (case study), 522-525
macroeconomic issues, 344, 412
mainstreaming, 357
urban capacities, 460
vulnerability, 77-78, 265, 269, 441
Development, 10, 11, 27
disaster risk and, 10, 27-28, 265-266
extreme events/impacts and, 265-266
international, 410
land use changes and, 238
planning, 439, 443-444, 460-461
post-disaster, 293, 439
setbacks, 410
skewed, exposure/vulnerability and, 10, 67, 70
sustainability and, 20, 437-486
values and interests in, 440, 446-447
vulnerability and, 67, 70, 78-80, 265
See also Sustainable development
Development pathways
disaster risk and, 27, 293
global, 10, 396
vulnerability/exposure and, 78-80
Dhaka, 252, 461, 511
Diarrhea, 252, 297, 506
Dikes/levees, 11, 52, 55, 68, 305
Direct losses, 264, 266, 317
Disaster[*], 27, 31, 39-44
cycle, 35
defined, 5, 31, 237
humanitarian relief for, 10, 293, 299-300
impacts of, 31, 32, 42
weather/climate related to, 115-122
Disaster costs, 264-274, 270, 271, 272
Disaster databases, 364, 423-424
Disaster Deficit Index (DDI), 93
Disaster losses, 9, 11-16
See also Economic losses/impacts
Disaster management[*], 35
Disaster mitigation, 36
Disaster preparedness, 36, 364-366, 376-377
Disaster prevention, 36, 69
Disaster Reduction Hyperbase, 415
Disaster risk[*], 10, 25-64
concept/relationships, 31-32, 31, 44, 444
continuum, 35, 69
defined, 5
development and, 27-28
future, anticipating and responding to, 302
increase in, 29, 405
risk accumulation, 95-96
Disaster risk management (DRM)[*], 16-20,
291-435
adaptation and, 37-39, 47-48, 295, 408-409,
443-462
allocation of efforts, 44-50
bottom up mechanisms, 396
challenges and opportunities, 46-47, 302, 313
climate change and, 27-28, 46-47, 49, 375-380,
376-377
climate variability and, 297
community participation, 28, 295, 300
context, 4-7, 4
coordination, 342, 356
corrective, 36
costs of, 317-319
cycle, 35
defined, 5, 34
effective strategies, 17, 27, 67, 342, 439-440
effectiveness, assessment of, 375-377
implementation, national level, 341-342
information and, 421-425
integration with adaptation, 11, 28, 50, 355-357,
396-397, 425-427, 439, 450-454, 469-471
international level, 393-435
legislation, 342, 358, 359, 519-522
local knowledge and, 293, 311-312
local level, 291-338
mainstreaming, 348, 355-357, 380
national level, 339-392
options/strategies, 6, 16-20, 18-19, 320-323,
439-440
past experience, 10-11
policies and options, 352-354, 361
prospective, 36
public policies/components, 89-90
sector-based strategies, 352-354, 357
short- and long-term responses, 450-454
synergies with adaptation, 36, 48-50, 68, 357,
469-471
top down mechanisms, 396
tradeoffs in, 20, 439-440, 448-450
See also Risk management
Disaster risk reduction (DRR)[*], 34, 366-371
as adaptation, 237, 443-446
adaptation and, 423-424
community/local knowledge and, 28
ecosystem-based solutions, 343
integration with adaptation, 396-397
international level, 396-397, 403-425
investment in, 529
post-disaster recovery/reconstruction, 293, 301,
315, 439
short-term strategies, 305-306
sustainable land management, 293
technical and operational support, 409-410
Discounting, 268, 319, 376
Disease, 252, 259, 410, 507-510
Disease vectors, 252, 253, 259, 316, 506
Displaced people/refugees, 80-81, 238, 457
international refugee law, 396, 402, 412
Displacement, 80-81, 300-301
Distribution
of climate variables, 40, 41, 121, 130
probability distribution, 40, 41, 116-117, 117, 121
Diversity, cultural, 84-85, 456
See also Biodiversity
Downscaling[*], 129-131, 133, 147, 416
Drivers
of capacity, 76
of disaster risk, 10
of economic losses, 16, 235
underlying, need to address, 20, 425, 440, 470
of vulnerability, 70-72, 379
Drought[*], 167-175
agricultural impacts, 247, 252
attribution of, 171-172, 174-175
case study, 498-500
defined, 167-169
drivers of, 168
ecosystem impacts, 246, 252
impacts, 242
indicators/indices, 168-169, 242, 245
management/adaptation options, 19
megadroughts, paleoclimatic, 170
observed changes, 8, 19, 119, 170-171, 174
projected changes, 19, 113-114, 119, 172-175,
173, 191-195, 242
regional impacts, 253, 255, 256, 259, 260-261, 264
regional projections, 191-195
socioeconomic impacts, 238-239
technological management, 415
Dust storms, 190, 254
Dynamical downscaling, 129-130, 133, 147
Dzud, 500-502
E
Early warning systems[*], 303-304, 342, 364-366,
405, 416-417
case study, 517-519
Heat Warning System, 495
strategy of, 517-519
Economic costs of disasters, 264-274
See also Costs; Economic losses/impacts
Economic efficiency, 399-400, 420
Economic growth, 445
Economic losses/impacts, 9, 234, 264-274
assessment of, 269-273
attribution of, 268-269
developed vs developing countries, 234, 265
direct, indirect, intangible, 264, 266, 317
Disaster Deficit Index (DDI), 93
drivers of, 16, 235
exposure and, 9, 16, 234-235, 269, 273
global and regional, 269-271, 270, 271, 272, 441
percent of GDP, 234, 270, 344, 441
Probable Maximum Loss (PML), 93
projected changes, 16
top-down vs. bottom-up approach, 266-267
types of, 264
See also Costs
Economic vulnerability, 86-87
Ecosystem-based adaptation, 343, 352, 357,
370-371, 371, 445-446
Ecosystem management and restoration, 294,
306-307, 315, 404
Ecosystem services, 370, 370, 445-446
Ecosystems, 231-290
DRM and adaptation, 343, 352, 370-371
impacts, 244-246
local management/protection, 294, 306-307, 315
reproduction, climate extremes and, 238
Education, 364-366, 526-529
case study, 526-529
vulnerability/exposure, 81-82
Efficiency, 399-400, 420
El Niño-Southern Oscillation (ENSO)[*], 155-157,
255, 459
observed changes, 119, 155-156
projected changes, 15-16, 113, 119, 156-157
El Salvador, 504, 520
Elderly, 454
Electrical networks, 257, 307
Index
577
EM-DAT database, 364
Emissions scenarios[*], 11-12, 112-113
precipitation projections, 145-146, 147-149
SRES, 112-113, 136-137, 137-140, 141
temperature projections, 137-138, 139-140, 141
Empowerment, 310, 316
Energy systems, 353
Ensembles, 131-133
Entitlements, 315-316
Environment, physical, 109-230
impacts on, 167-190
Environmental dimensions of vulnerability,
76-80
Environmental justice, 320
Epidemic disease, 507-510
Equity, 320, 401, 454-456
Ethiopia, 365, 420, 524
Ethnicity. See Race/ethnicity
Europe, 256-258
adaptation costs, 274
case study (heat waves), 492-496
coastal impacts, 183, 185, 186, 256-257
drought, 170-171, 172-174, 175, 242, 243, 246, 256
ecosystem-based adaptation, 371
floods, 176, 177, 256-257, 258
heat waves, 19, 133, 192, 256, 257, 492-496
landslides, 258
precipitation, 142, 145-146, 148, 149, 192, 198-199
snow, 258
temperature, 133-135, 135, 139-140, 141, 192,
198-199
waves, 181, 182
wildfires, 256
wind, 150, 151, 152, 257
Evapotranspiration[*], 113-114, 118, 167, 167-169
Ex post vs ex ante actions, 74, 90
Exposure[*], 4-7, 32, 65-108
concept/definition, 5, 32, 69, 237, 444
development and, 10, 67, 70
dimensions and trends, 76-89
economic losses and, 9, 16, 234-235, 269, 273
interactions of, 238-239, 239
management options, 18-19
observations of, 7-9, 18-19
regionally based aspects, 252-264
risk and, 69
scales and factors in, 67, 237
system- and sector-based aspects, 239-252
See also Vulnerability
External forcing, 5, 29
Extratropical cyclones. See Cyclones
Extreme events[*]
, 39-44
anthropogenic influences and, 9, 112, 125-126, 268
case studies, 492-507
climate change and, 40, 127
compound events, 118
comprehensive/integral/holistic focus, 38
context, 4-7
costs, 264-274
defined, 30, 40-41, 111, 116-117
extreme weather vs climate, 117
factors in, 4, 111
impacts on humans and ecosystems, 231-290
impacts on physical environment, 109-230
observed changes, 111-112, 119-120, 133-152
physical aspects, 38, 40-41
projected changes, 112-113, 119-120, 133-152
risk management, 16-20
short- and long-term responses to, 450-454
technologies for, 416-417
vulnerability/exposure and, 65-108, 239-264
See also Climate extremes; Extremes; Impacts
Extreme impacts, 27, 41-44, 237
Extreme indices, 116-117
Extreme value theory, 116-117
Extremes
changing climate and, 40, 124-125, 127
defined, 5, 40
diversity of, 40-41
methods and requirements, 122-133
observed changes, 119-120, 133-152
projected changes, 119-120, 133-152
traditional adjustment to, 43-44
See also Climate extremes; Extreme events
F
Fairness, 320
Faith-based organizations, 296, 346, 348
FAQs (frequently asked questions)
climate: is it becoming more extreme?, 124-125
DRM strategies: in changing climate, 49
government: preparedness measures, 376-377
local context: importance of, 298
local level: adaptation limits, 319
local level: cost estimation, 318
local level: lessons learned, 300
relationship: climate change and individual
extreme events, 127
relationship: extreme events and disasters, 33
resilience: practical steps for, 470
technology: emphasis on, 448
transformational changes, 466
Fatality rates, 9, 234, 344, 400
Feedbacks, 112, 118-120
Finance/budgeting, 523
adaptation in developing countries, 406-407, 411,
413
international, 17, 357, 397, 410-411, 412-414
national, 360-362, 523
technology transfer, 417
See also Costs
Fire. See Forest fires; Wildfire
Fisheries, 352
Flexibility, 343, 355, 380
Flood control, 11, 52, 55, 68, 305
Floods[*], 41, 175-178
case studies, 505-507, 510-512
climate change and, 245
costs, 256-257, 272
ecosystem impacts, 246
exposure and, 241, 245
flash floods, 18, 175
floodplains, 77
fluvial (river) floods, 113, 175, 178, 242-244
glacial lake outburst floods[*], 114, 175, 186, 258
observed changes, 8, 119, 175-177, 178
projected changes, 13, 113, 119, 177-178
regional impacts, 253-254, 256-257, 258,
259-260, 261, 262
FONDEN, 362
Food security, 235, 246-247, 368-370
DRM/adaptation options, 352
Forest fires, 252, 256, 259
See also Wildfire
Forestry sector, 235, 446
costs, 273
deforestation, 238, 252, 446
DRM/adaptation options, 352
Frameworks, 403-411
See also UNFCCC; UNISDR
France, 176, 185
Frequently asked questions. See FAQs
Funds for adaptation, 406-407, 411
G
Garifuna women, 82
GDP, 234, 270, 344, 441
Gender, 91, 313, 315
Geographic information systems (GIS), 416
Glaciers[*], 186-189
glacial lake outburst floods (GLOFs)[*], 114, 175,
186, 258
glacier melting, 15, 114, 188
Global Assessment Reports (GARs), 404-405,
408, 425
Global climate models[*], 128-133
Global costs
of adaptation, 273-274, 273
of disasters, 269-271, 270, 271, 272
Global Environment Facility (GEF), 410-411, 413,
417
Global Facility for Disaster Reduction and
Recovery (GFDRR), 411
Global Framework on Climate Services (GFCS),
409, 464
Global interdependence, 10, 396, 399
Global mean climate, 121-122
Global Network of Civil Society Organisations
for Disaster Reduction (GNDR), 404
Global Platforms, 404, 405
Governance[*]
case study, 519-522
international, 403-411
local, 312-313
national, 341, 346-347
preparedness and, 376-377
risk governance framework, 27, 44
risk-neutral approach, 360-361
sub-national governments, 346-347
vulnerability and, 85-86
Grand Challenges, 426
Green Climate Fund, 408, 414
Greenhouse gases[*], 9, 112, 124, 126
ENSO and, 156
mitigation, 446, 458-462
Greenland, 191, 197
ice sheet, 179, 188
precipitation, 145-146, 191, 197
temperature, 135, 139-140, 191, 197
Grenada, 512-514
Gross domestic product. See GDP
Groundwater, 42, 167-168, 242
salinization of, 250, 256
Growing season, 247
Guatemala, 504
H
Hadley/Walker circulation, 149, 151, 154
Hail, 141-142, 143, 148-149
Hazards[*], 31, 69, 444
‘creeping’, 365
multi-hazard risk management, 17, 439, 450
preparedness and, 364-366
Health and well-being, 251-252
DRM/adaptation options, 354
impacts, 235, 251-252, 259, 316
inequalities in, 316
Index
578
regional impacts, 259
vulnerability, 82-84
Heat waves[*]
case study, 492-496
ecosystem impacts, 244-246
human impacts, 248, 251-252, 316
observed changes, 8, 133, 134
projected changes, 15, 114, 191-202
regional impacts, 253, 256, 258-259
Australia, 239, 496-498
Europe, 19, 256, 257, 492-496
regional projections, 191-202
vulnerability to, 234
High-income countries, 9, 234, 270, 343, 344, 347
High-latitude changes, 189-190
Holistic approaches, 379-380
Honduras, 82, 314
Housing options, 352-354
Human development index, 266
Human health. See Health and well-being
Human rights, 401, 403, 447, 457
Human rights law, 396, 401
Human security[*], 293, 297-298, 440, 457-458
Human settlements, 7-8, 247-251, 316-317
coastal, 7-8, 235, 248, 249, 258, 460, 510-512
informal, 237, 247, 460
vulnerability/exposure, 234-235, 247-251
vulnerability reduction, 368-370
See also Cities; Urban areas
Human systems, impacts on, 231-290
Humanitarian/disaster relief, 10, 293, 299-300, 452
Humanitarian sector, research in, 468
Hurricanes, 19, 181, 260
economic losses, 19, 256, 260
ecosystem impacts, 246
Katrina, 9, 55, 158, 260, 308, 315, 457
local response to, 297
management/adaptation options, 19
Mesoamerican (case study), 504-505
storm surge, 158
Hydroclimatic extremes, 41
Hyogo Framework for Action, 398, 403-406
national systems and, 341-342, 344
Strategic Goals, 404
I
Ice cover, 262
Ice sheets, 179, 188
Iceland, 191
Impacts[*], 8-9, 16, 109-290, 441
cascading, 399
costs/economic losses, 234-236, 264-274
cumulative, 6-7, 38, 67, 69
databases, 364
extreme, 6, 27, 41-44, 237
extreme and non-extreme events, 6, 234
global, 399, 441
human systems and ecosystems, 16, 231-290
international, 399
local, 293
management of, 373-375
physical environment, 8-9, 119-120, 167-190
projected, 11-16
regionally-based, 252-264
resistance to, 38
sector-based, 235, 239-252
social, 42-43
system-based, 239-252
trends in, 271-273, 272
uncertainty and, 239
vulnerability/exposure and, 10, 67, 234-235, 238-
239, 271
Impacts-first approach, 350-351, 350
Incremental change, 20, 439, 442
Index-based contracts, 323
Indexes (indices), 92-93, 116-117, 125, 168-169
India
adaptation technologies, 415
floods, 254, 457
wetlands, 370
Indian Ocean, 181, 184
cyclones, 503-504, 503
tsunami (2004), 307, 457
Indian Ocean Dipole (IOD)[*], 157-158
Indigenous peoples, 80-81, 82, 84-85, 298, 311-312
Indirect losses, 264, 266
Individual scale, 38, 39
Indonesia, 359, 457, 526
Inequalities, 10, 294, 313-317, 405, 447, 454
Informal settlements, 237, 247, 460
flash floods and, 18
Information, 342, 356, 363-364, 365, 529
access to, 82
acquisition, 421-422
international level support, 409, 421-425
on scale/nature of events, 416
sharing and dissemination, 422-425
See also Communication
Information and communication technologies
(ICT), 422-423
Information gap, 409
Infrastructure, 16, 248-250, 366-368
adaptation costs, 412
DRM/adaptation options, 353, 366-368
education, 81-82
permafrost and, 367, 515-517
risk reduction, 304-306
technology and, 447, 448
tipping points, 366
transport, 235, 248-249, 249, 250, 348
vulnerability, 235
Infrastructure thresholds, 366
Innovation, 20, 294, 439, 447, 468, 469
Institutional approaches, 464-465
Institutional/governance dimensions, 85-86,
294, 308-313
Insurance[*], 294-295, 343, 371-373, 465, 523
case study, 522-525
crop insurance, 524
distribution of, 343
index-based/index-linked, 323
international programs, 419-420
local level, 322-323
micro-insurance, 322-323, 420, 523, 524
national systems, 343, 346, 360, 371-373
regional pools, 372, 400, 419, 420, 523, 524-525
Intangibles, 264, 266, 317, 446
Integrated models, 131-133
Integrated systems approach, 27, 38, 50
Integration
across scales, 17, 397, 426-427
of DRM and adaptation, 11, 28, 50, 355-357,
396-397, 425-427, 439, 450-454, 469-471
international level, 396, 425-427
short- and long-term responses, 450-454
Intellectual property rights, 414, 448
Interdependence, 10, 396, 399
International Bill of Rights, 411
International conventions, 402
International Decade for Natural Disaster
Reduction (IDNDR), 397, 403
International development, 410
International Disaster Response Law, 411
International finance, 357, 397, 410-411
International humanitarian law, 411
International institutions, 348
See also UNFCCC; UNISDR
International instruments, 402-403
International law, 396, 401-403, 411-412
hard law, 401
limits and constraints, 411-412
soft law, 401, 402, 403-404
International refugee law, 396, 402, 412
International risk management, 393-435
adaptation and, 396-397
context, 398
current actors, 408-411
current governance and institutions, 403-411
economic efficiency, 399-400
finance, 17, 357, 397, 410-411, 412-414
future policy and research, 425-426
integration across scales, 426-427
knowledge/information, 421-425
options, constraints, and opportunities, 411-425
rationale for, 398-403
risk sharing and transfer, 418-421
shared responsibility, 400-401
subsidiarity, 401
systemic risks, 399
technology transfer and cooperation, 414-418
International technical support, 409-410
Irrigation, 448
Island states. See Small island states
J
Japan
disaster risk reduction, 424
precipitation, 148
public awareness campaign, 528
storm surge, 254
waves, 182
Judgments about risk, 45, 46-47
Justice, 320
See also Equity
K
Katrina (hurricane), 55, 158, 308, 457
economic losses, 9, 260
international impacts, 399
recovery and reconstruction, 315
Kenya, 360, 374
Knowledge acquisition, 421-422
Knowledge/information sharing, 422-425, 526-529
Kyoto Protocol, 402, 403, 406, 411
L
Land use/land use change, 69, 238, 293
Land use, land use change, and forestry
(LULUCF), 370
Land use planning, 306-307
Landslides[*], 41, 114, 120, 187-189, 255
regional impacts, 258
Leadership, 469
Learning, 27-29, 53-56, 439, 467-468
humanitarian sector, 468
‘learning by doing’, 378-379, 380
learning loops, 53-54, 56
Index
579
transformation and, 324
See also Case studies
Least Developed Countries Fund (LDCF), 357,
406, 413, 417
Legislation, 358, 359
case study, 519-522
Lessons learned, 469-470, 529-530
local level, 295, 300
See also Case studies
Levees, 55, 254-255, 305
Liabilities, 361, 361
Light Detection and Ranging (LIDAR) data, 186
Likelihood[*], 21, 112, 120-121
Limpopo River/Basin, 253, 309, 506
Livelihoods, 314-315
agricultural sector, 246
ecosystem management and restoration, 294
tourism sector, 251, 251
vulnerability and, 87
Livestock, 237, 253, 259
Local adaptation, 293-295
limits to, 319-320, 319
Local coping, 294, 298-301
differences/inequalities in, 313-317
Local decisionmaking, 305-306, 308, 310, 325
Local disaster risk management (LDRM)[*],
291-338
anticipating future risk, 302-307
capacity building, 308-313
challenges and opportunities, 295, 313-320
community participation, 28, 295, 300, 321, 322
context and, 298, 298
costs, 295, 317-319, 318
current coping, 298-301
differences/inequalities in, 10, 294, 313-317
disaster relief and assistance, 299-300
gaps in information, 323-325
importance of local level, 296-298
insurance, 294-295, 322-323
lack of data, 295
lessons learned, 295, 300
links to global/national levels, 296
risk communication, 294, 302-304, 303
risk sharing, 321-323
strategies, 293, 320-323
Local government, 312-313
Local knowledge, 17, 293, 311-312
Local-level institutions/planning, 294, 308-313
Local recovery and reconstruction, 301, 315
Localized social networks/norms, 310-311
Loss of life/fatalities, 9, 234, 344, 400
Low-income countries, 9, 234, 343, 344, 347
Low regrets strategies, 16-17, 56, 342, 351, 352-
354, 376
M
Maastricht Treaty, 401
Mainstreaming, 348, 355-357, 380, 411
Maladaptative actions, 54, 55, 368, 448
Malaria, 252, 253, 316
Maldives, 370, 512-514
Mangroves, 343, 370, 370, 450
Mass media, 422-423
Media coverage, 422-423
Mediterranean region, 192, 198-199
drought, 172-174, 175, 256, 498-500
precipitation, 142, 145-146, 192, 198-199
temperature, 133-134, 135, 138, 139-140, 192,
198-199
tourism, 251, 251
waves, 182
wind, 150
Mega-deltas, 254
Megacities, 294, 317, 510-512
Melbourne, 239, 497
Mental health impacts, 252, 316
Meridional overturning circulation (MOC)[*], 122
Mesoamerican hurricanes, 503, 504-505
Mexico, 258-260
drought, 170, 172, 174, 175, 259
fund for disasters (FONDEN), 362
hurricanes, 503, 504-505
insurance, 372
precipitation, 142, 145-146, 191, 197
temperature, 135, 139-140, 191, 197
Micro-finance, 419
Micro-insurance, 322-323, 420, 523, 524
Microsatellites, 416
Migration, 16, 80-81, 293, 300-301, 457
as adaptive, 237, 300, 399
destination area issues, 399
Millennium Declaration, 400, 410
Millennium Development Goals (MDGs), 369,
398, 400, 410, 458
Mitigation (climate change), 36, 446
interactions of, 20, 458-462
Mitigation (disasters and disaster risk)[*], 36
Models. See Climate models
Modes. See Climate modes
Moisture content
atmospheric, 126
soil, 118-119, 167-175, 167-169
Mongolia, 419, 500-502
Monsoons[*], 119, 152-155
Moral economy, 309, 309, 447
Mountain environments, 8, 15, 114, 186-189, 248
Mozambique, 253, 297, 405, 505-507
Mudslides, 158, 255
See also Landslides
Multidisciplinary management, 36, 37
Mumbai, 461, 510-512
Myanmar, 502-505, 503
N
Nairobi floods, 18
Nairobi Work Programme, 407, 409, 529
Namibia, 253-254
National adaptation plans, 349-355, 356, 370
National Adaptation Programme of Actions
(NAPA), 369-370, 406
National building standards, 347, 367-368
National Platforms, 359, 360
National systems, 11, 36, 339-392, 346
adaptation options, 352-354, 361
adaptive management, 377-379
aligning with climate change, 342, 351-355,
375-380
communication, 349, 376
coordination, 358-360, 360
culture of safety, 362-366
disaster risk management, 341-343, 352-354, 361
disaster risk reduction, 366-371
finance and budget, 360-362
flexibility in, 343, 355, 380
holistic approaches, 379-380
impact management, 373-375, 374
implementation of DRM, 341-342
insurance, 343, 346, 371-373
legislation and compliance, 358, 359
planning and policies, 349-357, 352-354
practices, methods, and tools, 362-375
preparedness, 364-366, 376-377
public goods, 341, 363
risk assessments, 380
risk pooling, 343
risk transfer, 355, 371-373, 376
sector-based risk management, 352-354, 357
strategies, 357-362, 376-377
systems and actors, 345-349, 346
top-down vs bottom-up approaches, 350-351, 350
uncertainties, managing, 377-379
Natural physical environment. See Physical
environment
Nepal, 526
Netherlands, 52, 257
flood impacts, 272-273, 272
flood management, 469
Networks, 309-311, 309, 404
New Orleans, 55, 158, 260, 308, 315
New York, 303
New Zealand, 260-261
drought, 172, 261
precipitation, 143, 145-146, 195, 202, 261
temperature, 139-140, 195, 202
wildfire, 261
No regrets/low regrets options, 16-17, 56, 351,
352-354, 376
Nongovernmental organizations (NGOs), 313,
403
Normative dimension, 426
Norms, 310-311
North America, 258-260
adaptation costs, 274
drought, 170, 174, 175, 259
floods, 177, 259-260
forest fires, 252
heat waves, 258-259
hurricanes, 19, 260, 315
monsoons, 153, 154
precipitation, 142, 145-146, 149, 191, 196-197
temperature, 134, 135, 135, 139-140, 141, 191,
196-197
wildfire, 259
North Atlantic Oscillation (NAO)[*], 119, 149,
157-158
Northern Annular Mode (NAM)[*], 150
Northwest Territories, 514-517
Nunavut, 514-517
O
Observed changes, 7-9, 18-19, 111-112, 119-120,
133-152
coastal impacts, 120, 183
drought, 19, 119, 170-171, 174
ENSO, 119, 155-156
extratropical cyclones, 163-166
floods, 119, 175-177, 178
glaciers and mountains, 186-188
heat waves, 19
methods for analyzing, 122-125
monsoons, 152-153
permafrost, 120, 186-188, 189-190
sand/dust storms, 190
sea level, 120, 178-179
temperature, 119, 133-136, 141
tropical cyclones, 159-161, 163
wind, 119, 149-150
Index
580
Ocean acidification, 185, 261
Oceania, 260-261
Open oceans, 261
P
Pacific Decadal Oscillation (PDO)[*], 153, 157
Pacific Island Countries, 183, 263-264
Pacific Ocean, 181, 184, 263
Pakistan, 457
Palmer Drought Severity Index (PDSI), 168-169
Permafrost[*], 189-190
case study, 514-517
infrastructure and, 367
observed changes, 120, 186-188, 189-190
projected changes, 15, 114, 120
regional impacts, 263
Perturbed-physics ensembles, 131-133
Philadelphia, 495
Philippines, 239, 519-522, 526
Physical environment, 109-230
impacts on, 119-120, 167-190
Planning, 10, 462-469
adaptation planning, 349-357, 443-444
community participation in, 28
development, 439
local level, 308-313
national level, 349-357, 380
risk reduction, 69
urban planning, 460-461
Polar regions, 261-263
Politician’s dilemma, 452
Polluter pays principle, 400
Population growth, 237, 238
Population movements. See Migration
Post-disaster credit, 419
Post-disaster recovery and reconstruction, 10,
293, 301, 315, 439, 457
Post-traumatic stress disorder, 252
Poverty, 20, 314, 344
rural, 238
vulnerability/exposure and, 87, 247, 400, 441
Poverty reduction strategies, 410
Poverty traps, 451, 452
Precautionary principle, 402
Precipitation, 8, 141-149
attribution of changes, 143-144, 149
observed changes, 41, 119, 142-144, 149
projected changes, 13, 14, 15, 113, 119, 144-149,
144-146, 191-202
rainfall, 113, 119, 142-144, 148, 149
regional impacts, 253-254, 255
regional projections, 191-202
snowfall, 141, 189
uncertainties, 148-149
See also Drought; Floods
Preparedness, 36, 364-366, 369, 376-377, 517-519
Prevention, 36, 69
See also Disaster risk reduction
PreventionWeb, 404, 405, 424
Private sector organizations, 347-348
Probabilistic risk analysis
, 42, 43, 44-45, 446
Probability distributions, 7, 7, 40, 41, 116-117,
117, 121
Probability of occurrence, 40, 41
See also Return period; Return value
Projected changes, 11-16, 18-19, 112-114, 119-120,
133-152
coastal impacts, 113, 120, 183-186
cyclones, extratropical, 13, 272
cyclones, tropical, 161-163, 271, 272
drought, 19, 113-114, 119, 172-175, 173
ENSO, 113, 119, 156-157
floods, 13, 113, 119, 177-178
glaciers and mountains, 189
heat waves, 15, 19, 114
monsoons, 153-157
permafrost, 114, 120, 190
precipitation, 13, 14, 15, 113, 144-149, 191-202
sand/dust storms, 190
sea level, 15, 120, 179-180
temperature, 12, 13, 112-113, 119, 133-152, 137,
138, 191-202
wind, 13, 113, 119, 151-152, 151, 248
Projections[*], 11-16, 119-120, 462-463
likelihood/confidence in, 112, 120-121
time scale of, 112
uncertainty and, 112, 130-133
See also Climate models; Scenarios
Property, 293
Property rights, 306
Psychological/mental health, 252, 316
Public awareness, 364-366, 526-529
Public goods, 341, 363, 399-400
Public health, 83, 529
heat waves and, 492-493
infectious diseases, 252, 253, 259, 316, 507-510
Public-private partnerships, 343, 347
R
Race/ethnicity, 84, 315
RANET, 423, 424
Rationing, 307
Reconstruction, 10, 301, 457
Recovery and reconstruction, 10, 293, 301, 315,
439, 457
funding for, 417
Red Cross/Red Crescent, 348, 359, 403, 409, 410
Refugees, 238, 412
international refugee law, 396, 402, 412
Regime shifts (ecosystems), 122, 170, 445, 448
Regional climate models, 129-133
Regional climate projections, 121-122, 191-202
cyclones and floods, 240-241
precipitation, 145-146, 147-149, 191-202
temperature, 135, 138, 139-140, 141, 191-202
Regional costs/economic losses, 256, 270, 270
adaptation costs, 273-274, 274
Regional development banks, 411
Regional impacts, 252-264
Regional risk pools, 372, 400, 419, 420, 524-525
Regrets/low regrets, 16-17, 56, 342, 351, 352-354,
376
Reinsurance, 323, 362, 364, 524-525
ReliefWeb, 423
Relocation, 293, 300-301
Remittances, 418-419
Republic of Korea, 201, 246
Republic of the Marshall Islands, 512-514
Resilience[*], 20, 34, 437-486
access to resources, 454-458
adaptation and, 443-450
building, 376, 378
concept/definition, 5, 34, 238
disaster risk management and, 443-450, 458-462
equity and, 454-456
integration of policies, 11, 439, 450-454, 469-471
interactions among processes, 324, 440, 458-462
long-term options, 462-469
multi-hazard risk management, 439, 450
multiple approaches and pathways, 17, 440
national/government actions, 376
pathways to, 442
planning and proactive actions, 462-469
practical steps for, 470
questioning assumptions and, 20, 439
risk and uncertainty and, 442
risk sharing/transfer and, 10-11, 465
short- and long-term responses, 450-454
sustainable development, 454-458
synergies of DRM and adaptation, 469-471
thresholds and tipping points, 458-459
transformation and, 442
vulnerability and, 72
Resilience thinking, 453-454
Resistance
, 38
Resources
access to, 454-456
scarcity of, 297, 316
storage and rationing, 307
Responsibility
common but differentiated, 400
corporate, 347
responsibility to protect, 396, 412
shared, 400-401
Return period[*], 46, 113
precipitation projections, 142, 146, 147, 198
temperature projections, 135, 137, 140
Return value[*], 117, 121
precipitation projections, 141, 145-146, 147, 196
temperature projections, 135, 139-140, 141, 196
Rights, 401, 403, 447
International Bill of Rights, 411
Rio Earth Summit, 400-401
Risk, 65-108, 444
acceptability of, 361, 362
factors in, 69-70
judgments about, 45, 46-47
social construction of, 36-37, 45
systemic, 399
unacceptable, 442
underestimation of, 452
Risk acceptance threshold, 361
Risk accumulation, 95-96
Risk assessment, 27, 67, 90-95
challenges in, 27, 46
local level, 294
methods, 274
uncertainty and, 274
Risk awareness, 364-366
Risk communication, 17, 67, 95, 294, 302-304,
303, 376
Risk governance framework, 27, 44
Risk identification, 90, 369
Risk-linked securities, 419-420
Risk management, 16-20, 34, 376
effective, 17, 67, 439
integrated systems approach, 27
international, 393-435
iterative, 47-48
local, 291-338
multi-hazard approaches, 17, 439, 450
national, 339-392
sector-based, 352-354, 357
short-term strategies, 305-306
See also Disaster risk management
Risk-neutral approach, 360-361
Risk pooling, 11, 343, 372, 400, 419, 523, 524-525
Index
581
Risk sharing, 10-11, 397, 523
international level, 397, 418-421
local level, 321-323
national level, 371-373, 376
Risk transfer[*], 10-11, 35, 294-295, 371-373, 465
case study, 522-525
international level, 397, 419-421, 523
local level, 294-295, 321-323, 523
national and sector-based programs, 352-354
national systems, 341, 355, 371-373, 376, 523
River discharge/runoff, 176-178
River regulation, 176, 238
dams, 176, 304, 415
Robustness, 56
Runoff[*], 177, 238, 262
Rural areas, 79-80, 461-462
continuum with urban areas, 298
Russian Federation, 262
S
Safety, culture of, 362-366
Saffir-Simpson scale, 149, 158, 162, 502
Sahara, 193, 199
drought, 174
precipitation, 145-146, 193, 199
sand/dust storms, 190
temperature, 139-140, 193, 199
Sahel, 171, 190
Saijo City, 528
Saltwater intrusion, 114, 184-185
Samaritan’s dilemma, 452
Sand and dust storms, 190, 254
Satellite-based technologies, 416, 423
Saudi Arabia, 415
Scale, 111-112, 132
coordination/linking across, 358-360, 360
exposure and, 237
individual, 38, 39
integration across scales, 397, 426-427
multiple scales, 448-450
small- and medium-scale, 38
territorial, 39
timescales, 88-89, 112
Scenario development, 462-463
Scenarios, 11-12, 112-113
precipitation projections, 145-146, 147-149
SRES, 112-113, 136-137, 137-140, 141
temperature projections, 137-138, 139-140, 141
Scenarios-impacts approach, 350-351, 350
School curricula, 526
Science and technology, 89
Sea ice, 122, 261, 262
Sea level, 178-180
change[*], 120
coastal impacts, 18, 182, 183, 184, 185
extreme, 18, 178-180
mean[*], 178
observed changes, 120, 178-179
projected changes, 15, 120, 179-180
small island states and, 184, 263
Sea level rise, 113, 178-179
impacts, 18, 184, 248, 263
observed changes, 178-179
projected changes, 15, 113, 179-180
Sea surface temperature[*], 185, 261
Sector-based organizations
, 357
Sectors, 235, 239-252
DRM/adaptation, 352-354, 357
See also specific sectors
Settlement patterns, 7-8, 78-80, 234, 235, 247-251
See also Human settlements
Shared responsibility, 400-401
Shelter-in-place, 308-309
Simulations, 462-463
Skewed development processes, 10, 67, 70
Slums. See Informal settlements
Small island states, 18, 184-185, 263-264
case study, 512-514
coastal impacts, 183
drought, 264
impacts, 15, 183, 248, 251
relocation issues, 300-301, 368
sea level and, 184, 263
small island developing states (SIDS), 18, 512-514
tourism impacts, 251, 251
vulnerability/exposure, 114, 184, 235, 248, 263, 263
Snow load, 514, 515
Snowmelt, 113, 119, 175-178
Snowpack, 176-177
Social capital, 310, 315
Social construction, 36-37, 45, 70
Social groups, vulnerability, 81
Social impacts, 42-43
Social justice, 320
Social networks, 310-311, 315
web-based, 423
Social norms, 310
Socio-ecological systems, 441
Socio-political networks, 309, 309
Socioeconomic change/inequities, 10, 238
Socioeconomic conditions, 16, 234, 235, 238, 410
Socioeconomic systems, 237-239
Soft law, 401, 402, 403-404
Soft vs hard engineering, 448
Soil moisture[*], 118-119, 124, 167-175, 167-169, 173
soil moisture anomalies (SMA), 15, 169, 172, 173
soil moisture drought, 168-169, 171-172
Solidarity, 400-401, 420-421, 523
South Africa, 193,
199
case study (DRM legislation), 519-522
collective action in, 322
precipitation, 145-146, 147, 193, 199
temperature, 139-140, 193, 199
South America, 255-256
adaptation costs, 273, 274
drought, 171, 174
floods, 176
precipitation, 142, 145-146, 148, 194, 200, 255
temperature, 134, 135, 135, 138, 139-140, 141,
194, 200
waves, 181
wildfires, 255-256
South Asia, 195, 201
Southern Annular Mode (SAM)[*], 150, 157-158
Space-based monitoring, 416
Special Climate Change Fund (SCCF), 406
SRES emissions scenarios, 112-113, 136-137,
137-140, 141
Stationarity/non-stationarity, 46, 131
Statistical downscaling, 129-131, 147
Storm surge[*], 158, 179-180
regional impacts, 254, 260, 261
Storm tracks[*], poleward shift in, 8, 164-166
Streamflow[*], 176-177
Structural measures, 293, 304-306, 415
Sub-Saharan Africa
adaptation costs, 273, 274
floods, 176
food security, 368
MDGs, 410
Subsidiarity[*], 401
Surprises, 122
Susceptibility, 72, 183
Sustainability, 20, 437-486
adaptation and, 444
DRM and, 440, 443-450, 458-462
of ecosystem services, 445-446
incremental steps, 439, 442
interactions of mitigation, adaptation, and DRM,
20, 440, 458-462
multiple approaches, 440
practical steps for, 470
prerequisites for, 440
synergies of DRM and adaptation, 469-471
thresholds and tipping points, 20, 458-459
transformational change, 20, 439, 465-469
See also Resilience
Sustainable development[*], 20, 439-440
climate change and, 444
DRM and, 440, 443-446
questioning assumptions in, 20, 439
technology and, 414-418
Sustainable land management, 293
Synergies, 48-50, 68, 357, 397, 469-471
Syria, 498-500
Systemic risks, 399
T
Tampere Convention, 402, 411
Tanzania, 39
Teacher training, 526-527
Technology, 89, 366-368, 447-448
adaptation and, 414-415, 415
emphasis on (FAQ), 448
extreme events and, 416-417
hard vs soft, 448
international technical support, 409-410
Technology choices, 447-448
Technology Mechanism, 408
Technology transfer, 397, 414-418, 448
financing for, 417
Temperature, 133-141
agricultural impacts, 247
attribution of changes, 135-136, 141
case studies, 492-498, 514-517
extremes, 135, 137, 138, 141
extremes, projected, 13, 112-113, 119, 133-152,
137, 138
global surface[*], 126
observed changes, 119, 133-136, 141
polar regions, 262-263
projected changes, 12, 13, 119, 135, 136-141,
137-140, 191-202
regional projections, 191-202
sea surface[*], 185, 261
summary, 141
uncertainties, 141
See also Heat waves
Thailand, 448
Thermohaline circulation, 122
Three Mile Island, 45
Thresholds, 20, 458-459
abrupt climate change, 122
absolute, 116
extreme events, 116
infrastructure, 366
risk acceptance, 361
Index
582
tipping points, 122, 351, 458-459
vulnerability/thresholds approach, 350-351, 350
Thunderstorms, 123, 141, 143
Tibetan plateau, 195, 202
precipitation, 143, 145-146, 195, 202
temperature, 139-140, 195, 202
wind, 150
Timescales, 88-89, 112
Tipping points, 122, 351, 366, 458-459
Top-down mechanisms/approach, 266, 346,
350-351, 350, 396, 427
Tornadoes, 151, 152
Tourism, 235, 250-251, 251
Tradeoffs, 20, 440, 448-450
Traditional behaviors, 84-85
Traditional knowledge, 311-312, 456
Transformation[*], 324, 441, 465-469
context, 441-442
defined, 5
local framework, 323
pathways, 442
Transformational change, 20, 439, 442, 465-469,
466
facilitating, 466-469
Transpiration[*], 113-114, 118, 167, 167-169
Transportation sector, 83, 88, 235, 246, 259, 353
infrastructure, 235, 248-249, 250, 348
Tropical cyclones. See Cyclones
Tsunami (2004), 307, 457
U
Uncertainty[*], 21, 47-48, 112, 130-133, 132
UNFCCC, 344, 396, 406-408, 411-412
adoption of, 403
commitments on adaptation, 407
finance and, 412-413
technology transfer, 397
UNISDR, 396, 397, 403, 408-409
evaluation of, 405-406
United Kingdom
coastal impacts, 185, 186
floods, 176, 177
precipitation, 142, 148
United Nations Development Programme
(UNDP), 404
United Nations Framework Convention on
Climate Change. See UNFCCC
United Nations International Strategy for
Disaster Reduction. See UNISDR
United States, 258-260
drought, 170, 174, 259
floods, 176
heat waves, 258-259
hurricanes, 19, 55, 158, 260, 308, 315, 457
precipitation, 142, 143, 148
temperature, 134, 138
waves, 181, 182
wetlands, 370
wind, 152
Urban areas, 460-461
continuum with rural areas, 298
growth in, 294, 317, 366
infrastructure, 250, 366-367
vulnerability/exposure, 78-79, 238, 247-248, 317
See also Cities
Urban floods, 175
Urban heat islands[*], 235, 248, 493
Urban planning, 460-461
Urbanization, 294, 317, 460
exposure/vulnerability and, 67, 78, 234, 248, 294
migration and, 238
Utilitarian approach, 447
V
Values and perceptions, 443, 446-447
Vanuatu, 309
Variability
climate, 7, 15, 115, 157
climate modes, 15-16, 155-158
Vectors of disease, 252, 253, 259, 316, 506
Victoria, Australia, 496-498
Vietnam, 317, 358, 370, 371
Violent conflict, 297, 457
Volcanoes, 188-189
Vulnerability[*], 4-7, 65-108
age/aging and, 234-235
capacity and, 72-74
concept/definition, 5, 32-33, 69-70, 237-238, 444
conceptual frameworks, 71-72
dimensions and trends, 76-89
drivers of, 70-72, 379
economic dimensions, 86-87
environmental dimensions, 76-80
extreme and nonextreme events and, 67
interactions and integration, 87-88, 238-239, 239
management options, 18-19
observations of, 7-9, 18-19
Prevalent Vulnerability Index (PVI), 92
regionally based aspects, 92-93, 252-264
risk and, 69-70
risk identification and assessment, 89-95
scales and factors in, 67, 69-72
sector-based aspects, 235, 239-252
skewed development and, 10, 67, 70
social dimensions, 80-86
socioeconomic systems and, 237-239, 410
system-based aspects, 239-252
timing and timescales, 88-89
underlying causes, need to address, 20, 425, 440,
470
urbanization and, 67, 78, 234, 248, 294
Vulnerability assessments, 320-321
Vulnerability thresholds approach, 350-351, 350
W
Warm days/warm nights[*], 8, 134-135, 137-138,
137, 138, 141
regional projections, 191-202
Warm spells[*], 134, 141, 191-202
Warning systems, 303-304, 364-366, 416-417
Water management, 16, 469
Water scarcity, 167, 238-239, 255
Water sector, 241-244
DRM/adaptation options, 353
impacts, 235, 241-244
vulnerability/exposure, 235, 241-244
water supply system, 242
Watersheds. See Catchments
Waves/wave height, 180-182
WeAdapt/WikiAdapt, 424
Wealth, 314
Weather derivatives, 523, 524
Weather events
categories of, 115
disasters and, 115-118
Weather extremes
defined, 40, 116-117
impacts on humans and ecosystems, 231-290
impacts on physical environment, 109-230
observed changes, 119-120, 133-152
phenomena related to, 119, 152-166
projected changes, 119-120, 133-152
vs climate extremes, 117
See also Climate extremes; Extreme events;
Impacts
Weather forecasting, 416-417
Weather indices, 116-117, 125, 168-169
Wetland reduction, 238, 444, 461
Wetland services, 248, 255, 370, 370, 445
Wildfire
Australia, 239, 496-498
case study (Victoria), 496-498
regional impacts, 255-256, 259, 261
See also Forest fires
Win-win options, 352-354, 355, 357
Wind, 149-152
impacts, 248
observed changes, 119, 149-150
projected changes, 13, 113, 119, 151-152, 151,
248
regional impacts, 257
technologies for, 416
Winners and losers, 456-457
Women, 81, 313, 454-456
Garifuna women, 82
proactive role (Honduras),
314
World Bank, 361, 404, 409, 411, 417, 419, 420
World Climate Conference-3 (2009), 409
World Conference on Disaster Reduction
(Kobe), 398, 403
World Food Programme, 419, 420, 524
World Trade Organization, 410
Y
Yangtze Basin, 147, 148, 176
Yucatan Peninsula, 503, 504
Yukon, 514-517
Z
Zambezi River/Basin, 253, 506, 507
Zimbabwe, 506, 508-509
Index