833
14
Adaptation
Needs and Options
Coordinating Lead Authors:
Ian R. Noble (Australia), Saleemul Huq (Bangladesh/UK)
Lead Authors:
Yuri A. Anokhin (Russian Federation), JoAnn Carmin (USA), Dieudonne Goudou (Niger),
Felino P. Lansigan (Philippines), Balgis Osman-Elasha (Sudan), Alicia Villamizar (Venezuela)
Contributing Authors:
Jessica Ayers (UK), Frans Berkhout (Netherlands), Kirsten Dow (USA), Hans-Martin Füssel
(Germany), Joel Smith (USA), Kathleen Tierney (USA), Helena Wright (UK)
Review Editors:
Anthony Patt (Austria), Kuniyoshi Takeuchi (Japan)
Volunteer Chapter Scientist:
Eric Chu (USA)
This chapter should be cited as:
Noble
, I.R., S. Huq, Y.A. Anokhin, J. Carmin, D. Goudou, F.P. Lansigan, B. Osman-Elasha, and A. Villamizar, 2014:
Adaptation needs and options. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A:
Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the
Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea,
T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken,
P.R. Mastrandrea, and L.L. White (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New
York, NY, USA, pp. 833-868.
14
834
Executive Summary ........................................................................................................................................................... 836
14.1. Introduction ............................................................................................................................................................ 838
14.2. Adaptation Needs ................................................................................................................................................... 839
14.2.1. Biophysical and Environmental Needs .............................................................................................................................................. 840
14.2.2. Social Needs ..................................................................................................................................................................................... 841
14.2.3. Institutional Needs ........................................................................................................................................................................... 842
14.2.4. Need for Engagement of the Private Sector ...................................................................................................................................... 843
14.2.5. Information, Capacity, and Resource Needs ..................................................................................................................................... 844
14.3. Adaptation Options ................................................................................................................................................ 844
14.3.1. Structural and Physical Options ........................................................................................................................................................ 845
14.3.1.1. Engineering and Built Environment ................................................................................................................................... 846
14.3.1.2. Technological Options ....................................................................................................................................................... 846
14.3.1.3. Ecosystem-Based Adaptation ............................................................................................................................................ 846
14.3.1.4. Service Options ................................................................................................................................................................. 847
14.3.2 Social Options ................................................................................................................................................................................... 847
14.3.3. Institutional Options ......................................................................................................................................................................... 848
14.3.4. Selecting Adaptation Options ........................................................................................................................................................... 849
14.4. Adaptation Assessments ......................................................................................................................................... 850
14.4.1. Purpose of Assessments .................................................................................................................................................................... 850
14.4.2. Trends in Assessments ...................................................................................................................................................................... 850
14.4.3. Issues and Tensions in the Use of Assessments ................................................................................................................................. 851
14.4.4. National Assessments ....................................................................................................................................................................... 852
14.5. Measuring Adaptation ............................................................................................................................................. 853
14.5.1. What Is to Be Measured? ................................................................................................................................................................. 854
14.5.2. Established Metrics ........................................................................................................................................................................... 855
14.5.2.1. Vulnerability Metrics ......................................................................................................................................................... 855
14.5.2.2. Metrics for Resource Allocation ......................................................................................................................................... 855
14.5.2.3. Metrics for Monitoring and Evaluation .............................................................................................................................. 856
14.5.3. Validation of Metrics ........................................................................................................................................................................ 856
14.5.4. Assessment of Existing and Proposed Metrics for Adaptation .......................................................................................................... 857
Table of Contents
835
Adaptation Needs and Options Chapter 14
14
14.6. Addressing Maladaptation ...................................................................................................................................... 857
14.6.1. Causes of Maladaptation .................................................................................................................................................................. 858
14.6.2. Screening for Maladaptation ............................................................................................................................................................ 858
14.6.3. Experiences with Maladaptation ...................................................................................................................................................... 858
14.7. Research and Data Gaps ......................................................................................................................................... 859
References ......................................................................................................................................................................... 860
Frequently Asked Question
14.1: Why do the precise definitions about adaptation activities matter?.................................................................................................. 853
836
Chapter 14 Adaptation Needs and Options
14
Executive Summary
Since the Fourth Assessment Report (AR4), the framing of adaptation has moved further from a focus on biophysical vulnerability
to the wider social and economic drivers of vulnerability and people’s ability to respond (robust evidence, high agreement). These
drivers include the gender, age, health, social status, and ethnicity of individuals and groups, and the institutions in place locally, nationally,
regionally, and internationally. Adaptation goals are often expressed in a framework of increasing resilience, which encourages consideration
of broad development goals, multiple objectives, and scales of operation, and often better captures the complex interactions between human
societies and their environment. The convergence between adaptation and disaster risk management has been further strengthened since AR4,
building on the IPCC Special Report on Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation (SREX).
{14.1-3}
Adaptation needs arise when the anticipated risks or experienced impacts of climate change require action to ensure the safety
of populations and the security of assets, including ecosystems and their services (medium evidence, medium agreement).
Adaptation needs are the gap between what might happen as the climate changes and what we would desire to happen. The use of the term
needs has also shifted with the framing of adaptation. In the National Adaptation Programmes of Action (NAPAs) “needs” were usually
discussed in terms of major vulnerabilities and priority adaptation activities, and, in both developing and developed countries, this hazard-
based approach with a focus on drivers of impacts and options to moderate them is still used often for urban or regional programs. But more
recently, the focus has been on tackling the underlying causes of vulnerability such as informational, capacity, financial, institutional, and
technological needs. {14.2}
Engineered and technological adaptation options are still the most common adaptive responses, although there is growing
experience of the value for ecosystem-based, institutional, and social measures, including the provision of climate-linked safety
nets for those who are most vulnerable (robust evidence, high agreement).
Adaptation measures are increasing and becoming more
integrated within wider policy frameworks. Integration, though it remains a challenge, streamlines the adaptation planning and decision-making
process and embeds climate-sensitive thinking in existing and new institutions and organizations. This can help avoid mismatches with the
objectives of development planning, facilitate the blending of multiple funding streams, and reduce the possibility of maladaptive actions. The
increasing complexity of adaptation practice means that institutional learning is an important component of effective adaptation. {14.3}
Approaches to selecting adaptation options continue to emphasize incremental change to reduce impacts while achieving co-
benefits, but there is increasing evidence that transformative changes may be necessary in order to prepare for climate impacts
(medium evidence, medium agreement).
While no-regret, low-regret, and win-win strategies have attracted the most attention in the past
and continue to be applied, there is increasing recognition that an adequate adaptive response will mean acting in the face of continuing
uncertainty about the extent of climate change and the nature of its impacts, and that in some cases there are limits to the effectiveness of
incremental approaches. While attention to flexibility and safety margins is becoming more common in selecting adaptation options, many
see the need for more transformative changes in our perception and paradigms about the nature of climate change, adaptation, and their
relationship to other natural and human systems. {14.1, 14.3.4}
Among the many actors and roles associated with successful adaptation, the evidence increasingly suggests two to be critical to
progress: those associated with local government and those with the private sector (medium evidence, high agreement).
These
two groups will bear increasing responsibility for translating the top-down flow of risk information and financing and in scaling up the bottom-
up efforts of communities and households in planning and implementing their selected adaptation actions. Local institutions, including local
governments, non-government organizations (NGOs), and civil society organizations, are among the key actors in adaptation but are often
limited by lack of resources and capacity and by continuing difficulties in gaining national government or international support, especially in
developing countries. {14.2.3} Private entities, from individual farmers and small to medium enterprises (SMEs) to large corporations, will seek
to protect and enhance their production systems, supply lines, and markets by pursuing adaptation-related opportunities. These goals will help
expand adaptation activities but they may not align with government or community objectives and priorities without coordination and incentives.
{14.2.4}
837
14
Adaptation Needs and Options Chapter 14
Adaptation assessments, which have evolved in substance and style since AR4, have demonstrably led to a general awareness
among decision makers and stakeholders of climate risks and adaptation needs and options. However, such awareness has often
not translated into adaptation action (medium evidence, high agreement).
Most of the assessments of adaptation done so far have
been restricted to impacts, vulnerability, and adaptation planning, with very few assessing the processes of implementation and evaluation of
actual adaptation actions. {14.4.1} Assessments that include both top-down assessments of biophysical climate changes and bottom-up
assessments of what makes people and natural systems vulnerable to those changes will help to deliver local solutions to globally derived
risks. Also, assessments that are linked more directly to particular decisions and that provide information tailored to facilitate the decision-
making process appear to have most consistently led to effective adaptation measures. {14.4.3}
The evidence to support the most valuable metrics of adaptation needs and effectiveness is limited, but increasing (medium
evidence, high agreement). {14.5.2-3} At present, there are conflicting views concerning the choice of metrics, as governments, institutions,
communities, and individuals value needs and outcomes differently and many of those values cannot be captured in a comparable way by
metrics. {14.5} The demand for metrics to measure adaptation needs and effectiveness is increasing as more resources are directed to adaptation.
These indicators that are proving most useful for policy learning are those that track not just process and implementation, but also the extent
to which targeted outcomes are occurring. {14.5.2.3}
Maladaptation is a cause of increasing concern to adaptation planners, where intervention in one location or sector could
increase the vulnerability of another location or sector, or increase the vulnerability of the target group to future climate change
(medium evidence, high agreement). {14.7.3}
The definition of maladaptation used in AR5 has changed subtly to recognize that
maladaptation arises not only from inadvertent badly planned adaptation actions, but also from deliberate decisions where wider considerations
place greater emphasis on short-term outcomes ahead of longer-term threats, or that discount, or fail to consider, the full range of interactions
arising from the planned actions. {14.6.1}
838
Chapter 14 Adaptation Needs and Options
14
14.1. Introduction
This chapter establishes a foundation for understanding adaptation by
reviewing core concepts related to adaptation, with a focus on mapping
o
ut broad categories of needs and options. Here we use adaptation
needs to refer to circumstances requiring information, resources, and
action to ensure safety of populations and security of assets in response
to climate impacts. Adaptation options are the array of strategies and
measures available and appropriate to address needs. Because identifying
needs and selecting and implementing options require the engagement
of individuals, organizations, and governments at all levels, this chapter
also briefly considers the range of actors involved in these processes
and summarizes the risks of maladaptation.
Other chapters in this report, namely Chapter 4 and in particular Section
4.4, and supported by Chapters 3, 5, 6, and 7, deal with the threats of
climate change on ecosystems and other predominately natural systems
and their prospects and options for adaptation. For the sake of space
and clarity this chapter focuses on the socioeconomic systems that
support human livelihoods, although it also touches on the services
provided by ecosystems (including ecosystem-based adaptation).
This chapter also highlights some important tools for implementing
adaptation, namely approaches to assessing needs at national,
subnational, and sectoral levels, and the challenges of applying metrics
to determine adaptation needs and the effectiveness of adaptation
actions. In the course of these discussions, this chapter establishes a
foundation for the three adaptation chapters that follow. The existence
of adaptation options does not necessarily mean that these options can
be implemented when the need arises. Therefore, Chapter 15 examines
adaptation planning and implementation, including the challenges faced
and how these can be addressed. Chapter 16 focuses on adaptation
opportunities and constraints, while Chapter 17 assesses the economics
o
f adaptation to climate change, including the costs and benefits of
adaptation and of inaction. This chapter also draws on, and seeks not
to repeat, the detailed discussions of human health, well-being, security,
livelihoods, and poverty found in Chapters 11, 12, and 13 that are
so important to the wider discussion of adaptation. These and other
interactions among the adaptation chapters are illustrated in Figure
14-1.
Human and natural systems have a capacity to cope with adverse
circumstances but, with continuing climate change, adaptation will
be needed to maintain this capacity (IPCC, 2012; see also Section 1.4.1
and Box 2.1). The AR5 definition of adaptation (i.e., “The process of
adjustment to actual or expected climate and its effects. In human
systems, adaptation seeks to moderate harm or exploit beneficial
opportunities. In natural systems, human intervention may facilitate
adjustment to expected climate and its effects...”) follows the lead of
the IPCC Special Report on Managing the Risks of Extreme Events and
Disasters to Advance Climate Change Adaptation (SREX) in introducing
a degree of purposefulness by adding the phrase “which seeks to
moderate” rather than simply “which moderates” as in AR4.
1
Human
ability to cope with climate impacts can also be increased by actions
that are not anticipatory or purposefully undertaken in response to
observed or anticipated climate change, sometimes called unplanned
actions. For example, diversifying livelihoods in response to immediate
economic factors can increase long-term ability to cope with a changing
climate. Such actions were often referred to as autonomous adaptation.
However, the use of the term in the literature, including the IPCC reports,
has been inconsistent. The term is often used to refer to purposeful
adaptation actions carried out by agents without external inputs such
as policies, information, or resources (see Chapters 17, 22; Skoufias,
2012), and sometimes to refer to purposeful actions that are reactive to
experienced climate impacts, rather than being proactive or anticipatory
of them (see Glossary and WGII AR3 Section 18.2.3).
Sectors
Needs and
options
Regions
Climate-
resilient
pathways
Adaptation in
natural systems
Emerging
risks
Decision
making
Human
well-being
13
Needs and
options
14
15
16 17
11 12
19
20
4
2
Figure 14-1 | The relationship between the four adaptation chapters (14 to 17) and other closely related chapters. Chapter 14 (Adaptation Needs and Options) draws on and
cross-references many of the issues of human well-being, including health, security, and poverty; the treatment of adaptation of natural ecosystems is dealt with mainly in
Chapter 4 and is not repeated in Chapter 14. Similarly the needs and options synthesized in Chapter 14 are drawn largely from the sectoral (3 to 10) and regional chapters (21 to
30). Chapter 2 provides input to decision-making approaches relevant to Chapter 15 (Adaptation Planning and Implementation). All the adaptation chapters feed into the
synthesis of Chapters 19 (Emerging Risks and Key Vulnerabilities) and 20 (Climate-Resilient Pathways: Adaptation, Mitigation, and Sustainable Development).
839
Adaptation Needs and Options Chapter 14
14
T
he SREX and AR5 definitions of adaptation also clarify the distinction
between adaptation in human and natural systems. Natural systems
have the potential to adapt through multiple autonomous processes
(e.g., phenology changes, migration, compositional changes, phenotypic
acclimation, and/or genetic changes), and humans may intervene to
promote particular adjustments such as reducing non-climate stresses or
through managed migration (see Section 4.5). But successful adaptation
will depend on our ability to allow and facilitate natural systems to
adjust to a changing climate, thus maintaining the ecosystem services
on which all life depends.
Adaptation is becoming increasingly important in climate negotiations
and implementation, and integral to AR5 are the terms incremental and
transformational adaptation—sometimes referred to as a “paradigm
shift as in the Green Climate Fund Governing Instrument (Green Climate
Fund, 2013a). Incremental adaptation refers to actions where the central
aim is to maintain the essence and integrity of the existing technological,
institutional, governance, and value systems, such as through adjustments
to cropping systems via new varieties, changing planting times, or using
more efficient irrigation. In contrast, transformational adaptation seeks
to change the fundamental attributes of systems in response to actual
or expected climate and its effects, often at a scale and ambition greater
than incremental activities. It includes changes in activities, such as
changing livelihoods from cropping to livestock or by migrating to take
up a livelihood elsewhere, and also changes in our perceptions and
paradigms about the nature of climate change, adaptation, and their
relationship to other natural and human systems (Sections 3.3, 8.6.2.3,
20.5 and FAQ 8.2; IPCC, 2012; Kates et al., 2012; Park et al., 2012; Green
Climate Fund, 2013b). Transformational change may be driven by the
pursuit of better opportunities or by the realization of the imminent or
inevitable limits within existing paradigms (Dow et al., 2013; Section
16.4.2). Transformative change may threaten the status quo for many
and require leadership and sometimes triggering events to initiate it
(Kates et al., 2012). However, transformational change is not called for
in all responses to climate change (Pelling, 2010) and ill-prepared
transformative change may bring with it social inequities (O’Brien,
2012). The triggers for transformational change and its implementation
are dealt with in more detail in Sections 16.4 and 20.5. Differentiation
between incremental and transformative adaptation, although indistinct,
is important because it affects how we approach adaptation, how we
integrate it into planning and policy, and how we allocate adaptation
funding in both developed and developing countries (IPCC, 2012; see
also Chapter 17).
Another concept is the adaptation deficit, which is the gap between the
current state of a system and a state that would minimize adverse
impacts from existing climate conditions and variability (see Glossary);
that is, it is essentially inadequate adaptation to the current climate
conditions (Burton et al., 2002; Burton, 2004; Burton and May, 2004;
Parry et al., 2009; see also Sections 17.2.2.2, 17.6.1). Some have
suggested that it is often part of a larger “development deficit” (World
Bank, 2010). Delay in action in both mitigation and adaptation will
increase the adaptation deficit in many parts of the world (IPCC, 2012).
I
n the process of building future adaptive capacity it is important to
reduce the current adaptation deficit along with designing effective risk
management and climate change adaptation measures (Hallegatte,
2011). Failure to close the adaptation deficit, both current and in the
future, will result in residual damages from climate change. There have
been calls for such residual damages to be evaluated and reported
(Parry et al., 2009).
Summary of Key Findings from AR4
In the Working Group II (WGII) Fourth Assessment Report (AR4), the
main chapter on adaptation (Chapter 18) refined the basic terminology
of adaptation and concluded that adaptation to climate change was
already taking place, but on a limited basis. Societies have a long record
of adapting to the impacts of weather and climate through a range of
practices that include crop diversification, irrigation, water management,
disaster risk management, and insurance, but climate change, along
with other drivers of change, poses novel risks often outside the range
of experience.
WGII AR4 found that deliberate adaptation measures in response to
anticipated climate change were being implemented by a range of
public and private actors, on a limited basis, in both developed and
developing countries. These measures are undertaken through policies,
investments in infrastructure and technologies, and behavioral change,
and they are seldom undertaken in response to climate change alone.
Many actions that facilitate adaptation to climate change are undertaken
to deal with current extreme events, such as heat waves and cyclones,
and are often embedded within broader sectoral initiatives such as
water resource planning, coastal defense, and disaster management
planning.
WGII AR4 concluded that there are individuals and groups within all
societies that have insufficient capacity to adapt to climate change. The
capacity to adapt is dynamic and influenced by economic and natural
resources, social networks, entitlements, institutions and governance,
human resources, and technology. However, high adaptive capacity does
not necessarily translate into actions that reduce vulnerability. New
planning processes are being implemented to attempt to overcome
these barriers at local, regional, and national levels in both developing
and developed countries. WGII AR4 noted the establishment of the
National Adaptation Programmes of Action (NAPAs) and that some
developed countries had established national adaptation policy
frameworks. Other conclusions from the WGII AR4 relating to the
implementation of adaptation policies and measures, barriers to
adaptation, and the economic costs of adaptation are summarized in
Chapters 15, 16, and 17 of this report.
14.2. Adaptation Needs
Adaptation involves reducing risk and vulnerability; seeking opportunities;
and building the capacity of nations, regions, cities, the private sector,
communities, individuals, and natural systems to cope with climate
impacts, as well as mobilizing that capacity by implementing decisions
and actions (Tompkins et al., 2010). Vulnerability is the “propensity or
1
Purposefulness was introduced in the SREX definition, which introduced the phrase
“in order to moderate”.
840
Chapter 14 Adaptation Needs and Options
14
p
redisposition [of a system] to be adversely affected” (see Glossary)
and, until AR4, was viewed as comprising three elements: exposure,
sensitivity, and adaptive capacity (IPCC, 2007a). However, in IPCC (2012)
and in this report, vulnerability focuses only on sensitivity and capacity,
with exposure more appropriately incorporated into the concept of risk
(see Glossary; IPCC, 2012, Section 2.2).
Adaptation requires adequate information on risks and vulnerabilities
in order to identify needs and appropriate adaptation options to reduce
risks and build capacity. In framing an approach to adaptation, it is
important to engage people with different knowledge, experience, and
backgrounds in tackling and reaching a shared approach to addressing
the challenges (Preston and Stafford Smith, 2009; Tompkins et al., 2010;
Fünfgeld and McEnvoy, 2011; Eakin et al., 2012) Initially, identifying
needs was most often based on impact assessments (or risk-hazard
approaches), but social vulnerability or resilience assessments are
increasingly being used (Fünfgeld and McEnvoy, 2011; Preston et al.,
2011b). The risk-hazard framework, drawn primarily from risk and
disaster management, focuses on the adverse effects that natural
hazards and other climate impacts can have on a given location (Füssel
and Klein, 2006). The emphasis in this approach is on the physical and
biological aspects of impacts and adaptation (Burton et al., 2002). The
social vulnerability framework focuses on the reasons and ways in
which individuals, groups, and communities are vulnerable to climate
impacts. Here, the focus is on how different factors, such as institutions,
shape the socioeconomic conditions that place human populations at
risk (Adger and Kelly, 1999; Preston et al., 2011b). There are overlaps
and complementarities between these frameworks. Approaches to
identifying needs and options are discussed further in the section on
assessments (Section 14.4).
Comprehensive assessments typically provide insight into the risks and
vulnerabilities that will result from climate change in communities,
cities, nations, and ecosystems and, in turn, offer a means to identify
the presence of adaptation needs and options for addressing those
needs. The term adaptation needs is often used but rarely defined in
the adaptation literature. In the wider literature, a need can be seen as a
problem that can be solved (McKillip, 1987) or as a gap between current
outcomes and desired outcomes (Kaufman and English, 1979). Thus, in
this context, adaptation needs are the gap between what might happen
as the climate changes and what we would desire to happen. Also, the
term adaptation needs is used in several ways in the adaptation literature.
A common use is in the sense of the “urgent and immediate needs”
relating to the adverse effects of climate change, as in the rationale for
the NAPAs, although in this case “needs were usually discussed in terms
of major vulnerabilities and priority adaptation activities (by UNFCCC).
2
The most effective descriptions of these needs combined discussions of
climate and non-climate drivers of impacts, and the resources, capacity,
information, finance, etc., needed to implement options to moderate
those impacts (e.g., GEF, 2002). Assessments of adaptation needs, both
in developing and developed countries, have often taken a hazard-
based approach with a focus on drivers of impacts and options to
moderate them (Moser, 2009; Finzi Hart et al., 2012). But more recently,
the focus has been on tackling the underlying causes of vulnerability
(
Füssel, 2007). One of the few categorizations of needs is that of Burton
et al. (2006), who recognized information, capacity, financial, institutional,
and technological needs. A similar structure is followed in this chapter.
We first discuss biophysical and environmental needs on which all lives
ultimately depend. Then we discuss social needs and capacities and
how they vary throughout society. Third, we discuss our response to
climate risks and impacts and how they are modified by the multitude
of institutions through which humans work, ranging from international
organizations to community-based efforts. Finally, we touch on resources,
including societal needs for information and knowledge and financial
resources.
Although needs are specific to particular groups and places, they fit into
a set of more general categories as summarized in the sections below.
For instance, vulnerability at the national and subnational levels is
affected by geographic location, biophysical conditions, institutional and
governance arrangements, and resource availability, including access
to technology and economic stability (Brooks et al., 2005). At the macro-
level, two broad classes of determinants of vulnerability are recognized:
biophysical determinants and socioeconomic determinants (Preston et
al., 2011a). However, adaptation needs are highly diverse and context
specific, for instance, varying between islands even within nations such
as the Solomon Islands (Section 29.6.1). Different stakeholder groups
and individuals have differential adaptation needs and vulnerabilities.
Adaptation needs are also dynamic, and future adaptation needs are
highly dependent on the mitigation pathway that is taken. Furthermore,
the constraints and limits to adaptation (see Chapter 16) are likely to
mean that not all needs will be met, thereby emphasizing the need for
monitoring to avoid crossing critical thresholds (Section 19.7.3).
14.2.1. Biophysical and Environmental Needs
Climate change is altering ecological systems, biodiversity, genetic
resources, and the benefits derived with ecosystem services (Convention
on Biological Diversity, 2009; Mooney et al., 2009; Hoegh-Guldberg,
2011). Climate change is inducing shifts in habitats that often cannot
be followed by species (Section 4.3.4.1), leading to changed ecosystems,
to local and global extinctions, and to the permanent loss of unique
combinations of genes. For instance, González et al. (2010) used observed
and modeled changes of global patterns of biome shifts under climate
change to conclude that up to half of the terrestrial ecosystems were
vulnerable as a result of changes from secondary stressors, such as
wildfire and disease, and suggested significant changes to natural resource
management plans. In addition to the responses of ecosystems to climatic
change, a number of studies have identified impacts on ecosystem services,
particularly the effects of climate change on agricultural productivity
(Coles and Scott, 2009), freshwater ecosystems (Ormerod et al., 2010),
and downstream industries and enterprises (Preston and Stafford Smith,
2009). Ecosystem services that are already under threat from the impacts
of climate change include pollination, pest, and disease regulation
(Section 4.3.4.4); climate regulation services; and potable water supply
(Section 4.3.4.5). Further stressors will limit our options to respond to
climate change (Section 14.3.2).
2
https://unfccc.int/files/cooperation_support/least_developed_countries_portal/napa_project_database/application/pdf/napa_index_by_country.pdf.
841
Adaptation Needs and Options Chapter 14
14
N
atural systems underpin human livelihoods, health, welfare, food
security, and prosperity. Vital ecosystem services that need to be
maintained include provisioning services such as food, fiber, and potable
water supply; regulating services such as climate regulation, pollination,
disease control, and flood control; and supporting services such as
primary production and nutrient cycling (Section 4.3.4). Much of the
water for human consumption originates on forested lands and the
quality of the water is heavily dependent on the conditions of the
ecosystems through which it flows. Ocean systems also provide climate
regulation services, while coral reefs act as ecological buffers (Section
6.4.1.4). For instance, healthy coastal wetlands and coral reefs can help
to protect against storm surges and rising sea levels (Hoegh-Guldberg,
2011), while the maintenance of wetlands and green spaces can control
runoff and flooding associated with increases in precipitation (Jentsch
and Beierkuhnlein, 2008; Mooney et al., 2009). Meanwhile, fisheries and
aquaculture contribute more than 20% to the dietary animal protein of
nearly 1.5 billion people (Section 5.4.3.3).
Consequently, there is a need to protect these systems and resources
within the changing climate. Goldman et al. (2008) found that research
projects focusing on delivering ecosystem services, rather than on
biodiversity goals, attracted a wider set of funders and better encompassed
the landscapes and the people within them. However, many practices
to intervene to improve and maintain ecosystem services are based on
limited experience and thus still untested assumptions and limited
information (Carpenter et al., 2009). Hence, there is a need to improve
understanding and valuation of ecosystem services provided by different
adaptation options. There is also an urgent need for appropriate ecosystem
monitoring to avoid crossing critical thresholds (see Section 19.7.4).
14.2.2. Social Needs
From a social perspective, vulnerability varies as a consequence of the
capacity of groups and individuals to reduce and manage the impacts
of climate change. Among the key factors determining vulnerability are
gender, age, health, social status, ethnicity, and class (Smit et al., 2001;
Adger et al., 2009a). For instance, the vulnerability to health-related
impacts of climate change varies as a consequence of geographical location
(Section 11.3.1), gender and age (Section 11.3.3), and socioeconomic
status (Section 11.3.4). Poverty and persistent inequality may be the
most salient of the conditions that shape climate-related vulnerability
(Section 13.1.4). Climate change is expected to have a relatively greater
impact on the poor as a consequence of their lack of financial resources,
poor quality of shelter, reliance on local ecosystem services, exposure
to the elements, and limited provision of basic services and their limited
resources to recover from an increasing frequency of losses through
climate events (Tol et al., 2004; Huq et al., 2007; Kovats and Akhtar,
2008; Patz et al., 2008; Revi, 2008; Allison et al., 2009; Shikanga et al.,
2009; Gething et al., 2010; Moser and Satterthwaite, 2010; Rosenzweig
et al., 2010; Skoufias et al., 2012). Owing to limited financial resources
and often compromised health and nutritional status, the poor, along
with the sick and elderly, are at increased risk from trauma, physical
and mental illness, and death from climate impacts such as increased
pollution, higher indoor temperatures, exposure to toxins and pathogens
from floods, and the emergence of new disease vectors (Kasperson and
Kasperson, 2001; Haines et al., 2006; Costello et al., 2009, 2011; O’Neill
a
nd Ebi, 2009; Tonnang et al., 2010; Ebi, 2011; Harlan and Ruddell, 2011;
Huang et al., 2011; McMichael and Lindgren, 2011; Semenza et al., 2012).
Climate change, climate variability, and extreme events can erode natural,
physical, financial, human, and social and cultural assets (Section
13.2.1.1), and poverty traps arise when climate change, variability, and
extreme events make the poor even poorer (Section 13.2.1.4).
Social needs under climate change include understanding emotional and
psychological needs. In Australia, it has been found that extreme events
such as floods, drought, and bushfire can lead to mental suffering,
including post-traumatic stress disorder, resulting in the need for
psychological support and counseling (The Climate Institute, 2011). For
example, drought can increase suicide rates by 8% (Nicholls et al.,
2006). Social psychological adaptation processes powerfully mediate
public risk perceptions and understanding, psychological and social
impacts, and coping responses, as well as behavioral adaptation (Reser
and Swim, 2011). Yet little collaborative work or research has so far
focused on the nature and dynamics of individual-level coping and
adaptation processes and how they influence responses (Reser et al.,
2012).
These individual factors also are often associated with and compounded
by community-level conditions. Women often have unequal access to
and control over resources, including land titles and water rights (UNDP,
2010; CGIAR, 2012; Verner 2012). Many poor and ethnic minorities live
in substandard housing; lack access to basic services, savings, and
insurance; have compromised health; and are at threat due to excessive
densities, poor access roads, and inadequate access to safe water,
sanitation, and drainage (Huq et al., 2007; Kovats and Akhtar, 2008; Revi,
2008; Shikanga et al., 2009; Moser and Satterthwaite, 2010). In rural
areas, adaptation needs also are linked to the viability of agricultural
activity (Bosello et al., 2009). Climate change will lead to higher prices
and increased volatility in agricultural markets, which might undermine
global food supply security (Section 9.3.3.3). Geographically, highly
vulnerable regions are those exposed to sea level rise and extreme
events, overlaid with high concentrations of multidimensional poverty
(Section 13.2.2.1). There will be disproportionate impacts on developing
countries that are dependent on climate-sensitive activities such as
agriculture (Cline, 2007). However, middle-income populations can also
be adversely impacted by climate change as a stressor adding to other
effects.
The causes and solutions of vulnerability take place at different social,
geographic, temporal, and political scales (Ribot, 2010). Therefore, to
identify critical needs of populations, and the underlying conditions
giving rise to these needs, some social assessments can benefit by
looking across institutional domains and by spanning from the local to
the national. Local assessments provide a means to identify existing
vulnerabilities; the policies, plans, and natural hazards contributing to
these vulnerabilities; as well as identifying adaptation actions. Social
needs include the range of needs for human security (see Section
12.1.2), which include the universal and culturally specific, material, and
non-material elements necessary to people to act on behalf of their
interests. More specifically, at this level, social needs can be evaluated
in terms of availability of natural, physical, human, political, and financial
assets; stability of livelihood; and livelihood strategies (Moser, 2006;
Heltberg et al., 2009). Alternatively, regional and national assessments
842
Chapter 14 Adaptation Needs and Options
14
c
an provide a basis for ascertaining institutional conditions associated
with long-standing inequities and development paths that may need
to be addressed in order to generate robust options.
Although different stakeholder groups have specific needs, an overarching
adaptation need for communities, households, private sector, and
institutions is the need for shared learning on adaptation. Adaptation
has itself been referred to as a social learning process (Sections 15.6,
22.4.5.3). In particular, there is the need for human capacity and social
capital to implement adaptation actions, including education and access
to information (Brooks et al., 2005; Adger, 2006; Smit and Wandel,
2006). Improved information for adaptation can benefit from efforts to
combine indigenous and scientific knowledge (Section 12.3).
14.2.3. Institutional Needs
Institutions, informal and formal, are enduring regularities of human
action in situations structured by rules, norms, and shared strategies,
as well as by the physical world (Crawford and Ostrom, 1995) and as such
they provide the enabling environment for implementing adaptation
actions (Bryan et al., 2009; Chuku, 2009; Aakre and Rübbelke, 2010;
Compston, 2010; Moser and Ekstrom, 2011). These institutions provide
the guides, incentives, or constraints that shape the distribution of climate
risks, establish incentive structures that can promote adaptation, foster
the development of adaptive capacity, and establish protocols for both
making and acting on decisions (see Section 14.2.3.2; Chuku, 2009;
Agrawal, 2010; Compston, 2010). In many instances, international and
national-level policies and programs can facilitate localized strategies
through the creation of legal frameworks and the allocation of resources
(Adger, 2001; Bulkeley and Betsill, 2005; Corfee-Morlot et al., 2011).
Overall, there is a need for effective institutions to identify, develop, and
pursue climate-resilient pathways for sustainable development (Sections
20.2, 20.4.2), including strengthening the ability to develop new options
through social, institutional, and technological innovation (Section 20.4.3).
Chapter 15 further considers the institutional needs to mainstream
adaptation into government planning.
Governments at all levels play important roles in advancing adaptation
and in enhancing the adaptive capacity and resilience of diverse
stakeholder groups. National governments are integral to advancing an
adaptation agenda as they decide many of the funding priorities and
tradeoffs, develop regulations, promote institutional structures, and
provide policy direction to district, state, and local governments. In
developing countries, national governments are usually the contact
point and initial recipient of international adaptation financing. In some
countries, both developed and developing, state governments lead the
national government in promoting and implementing adaptation (Mertz
et al., 2009). The engagement of national government actors can help
mobilize political will, support the creation and maintenance of climate
research institutions, establish horizontal networks that promote
information sharing (Westerhoff et al., 2011), and, in some cases,
facilitate the coordination of budgets and financing mechanisms (Alam
e
t al., 2011; Kalame et al., 2011). Governments have the potential to
directly reduce the risk and enhance the adaptive capacity of vulnerable
areas and populations by developing and implementing locally appropriate
regulations including those related to zoning, storm water management
and building codes, and attending to the needs of vulnerable populations
through measures such as basic service provision and the promotion of
equitable policies and plans (Adger et al., 2003b; Brooks et al., 2005;
Nelson et al., 2007; Agrawal and Perrin, 2008; Agrawal, 2010).
Among the important institutions in both developed and developing
countries are those associated with local governments
3
as they have a
major role in translating goals, policies, actions, and investments between
higher levels of international and national government to the many
institutions associated with local communities, civil society organizations,
and non-government organizations (NGOs). SREX Chapter 5 (IPCC,
2012) extensively assesses the role and importance of the local scale
institutions when adapting to extreme weather and climate events,
highlighting that extreme weather and climate events are acutely
experienced at local levels, and that local knowledge is important for
managing impacts (Cutter et al., 2009). As institutional actors, local
governments and community institutions influence the distribution of
climate risks, mediate between levels of government as well as between
social and political processes, and establish incentive structures that affect
both individual and collective action at all levels (Agrawal and Perrin,
2008). They are in a pivotal position to promote widespread support for
adaptation initiatives, foster intergovernmental coordination, and
facilitate implementation, both directly and through mainstreaming into
ongoing planning and work activities (Anguelovski and Carmin, 2011;
Carmin et al., 2012).
There are a number of ongoing political issues that shape the relationships
national and local governments have in managing climate risks (Corfee-
Morlot et al., 2011). Governance failure has a significant influence on
institutional vulnerability (see Section 19.6.1.3.3). For instance, short-
term interests, when dealing with long-term issues, can limit incentives
to make investments. Similarly, the proximity that authorities have to
interest groups can sway their decisions toward other issues, while the
drive to engage the public in planning and other activities can orient
priorities in ways that do not support adaptation (Corfee-Morlot et al.,
2011). Local governments also may lack institutional capacity or have
difficulty gaining coordination among departments as conflicts emerge
to obtain scarce resources (Satterthwaite and Dodman, 2009; Hardoy
and Romero Lankao, 2011). In Bangladesh, the limited access of local
governments to resources has been cited as a barrier to local adaptation
(Christensen et al., 2012).
Tompkins et al. (2010) found from a survey of 300 projects identified as
adaptive at local government level in the UK that more than half were
driven by concerns not directly related to climate change. Nevertheless,
there are a number of indicators that demonstrate whether local
government has institutionalized and mainstreamed adaptation. These
include the presence of an identifiable champion from within government,
climate change being an explicit issue in municipal plans, resources
3
Here local government is used to refer to second or third tiers or lower of government, below national and state or provincial government levels; it includes county, district,
council, municipal, and similar levels of government.
843
Adaptation Needs and Options Chapter 14
14
d
edicated to adaptation, and adaptation incorporated into local political
and administrative decision making (Roberts, 2008, 2010).
Overall, it is important to match the appropriate institutional scale with
the scale of implementation. For example, the Murray-Darling Basin in
Australia includes significant water resources across four states requiring
management institutions involving federal, state, and local governments
to manage and allocate water use (Hussey and Dovers, 2006; see also
Box 25-2). While governments have the potential to influence adaptive
capacity, local governments often lack the human and technological
capacity or mandate to develop and enforce regulations. Local
governments, particularly those in developing countries, are faced with
numerous challenges that limit their ability to identify needs and pursue
adaptation options. Often, these governments must attend to backlogs
of basic and critical services such as housing and water supply or focus
their attention on addressing outmoded and outdated infrastructure.
They also may lack institutional capacity or have difficulty gaining
coordination among departments as conflicts emerge to obtain scarce
resources (Hardoy and Romero Lankao, 2011, Villamizar, 2011). Adaptation
will require an approach that devolves relevant decision making to the
levels where the knowledge and capacity for effective adaptations reside
(see Box 25-5). Sowers et al. (2011) maintain that, in the Middle East
and North Africa, the largely centralized systems of planning, taxation,
and revenue distribution lead to a focus on supply-side issues with little
consideration of changing climates and demand management, which
renders their populations vulnerable to climate-induced impacts on
water resources due to weak integration with local constituencies.
There are critical institutional design issues that can be evaluated in
order to understand institutional needs (Agrawal, 2010; Gupta et al.,
2010). The first is the extent to which institutions are flexible to handle
uncertainty. This includes flexibility across and within institutions to
evaluate and reorganize delivery where necessary. The uncertainty
associated with climate change, presence of rapidly changing information
and conditions, and emerging ideas on how best to foster adaptation
requires continual evaluation, learning, and refinement (Agrawal, 2010;
Gupta et al., 2010). Second is the extent to which adaptation is or has
the potential to be integrated into short- and long-term policy making,
planning, and program development (Conway and Schipper, 2011). Third
is the potential for effective coordination, communication, and cooperation
within and across levels of government and sectors (Schipper, 2009;
Agrawal, 2010; Conway and Schipper, 2011). Finally, to promote adaptive
capacity, it is important to identify the extent to which institutions are
robust enough to attend to the needs of diverse stakeholders and foster
their engagement in adaptation decisions and actions (Urwin and Jordan,
2008; Gupta et al., 2010).
14.2.4. Need for Engagement of the Private Sector
The role of the private sector is important in delivering adaptation. Often,
the focus falls on the role of the private financial sector in providing risk
management options including insurance and finance for large projects
(see Sections 15.4.4, 17.5.1). However, the delivery of adaptation actions
ranges more widely and spans different types of private enterprise, from
small farmers, to small to medium enterprises (SMEs), to multinational
companies. KPMG International (2008) used published reports and
i
nterviews to identify the sectors where businesses considered they face
the greatest climate-related risks. In order of perceived importance, the
core risks were regulatory, physical, reputational, and litigation risks.
The sectors identified as most at risk included an expected cluster
around oil and gas and aviation, and also a group less commonly
perceived to be at risk, including health care, the financial sector,
tourism, and transport.
Khattri et al. (2010) have described three general ways in which the
private sector can become involved in adaptation. The first, internal risk
management, is critical to firms and enterprises protecting their own
interests and ensuring continuity of supply and markets. The second form
of involvement recognizes that business is a stakeholder and therefore
needs to participate in public sector and civil society initiatives, such as
The New York City Panel on Climate Change, which consists of diverse
stakeholders, including representatives from the private sector (Rosenzweig
et al., 2011). Third, climate adaptation also provides a wide range of
new opportunities to the business community. Even in developing
countries, where regulations and markets are often underdeveloped and
business risks are high, Khattri et al., (2010) identified opportunities for
working in the health care, waste and water management, sanitation,
housing, energy, and information sectors through fostering cooperation
across government departments and NGOs and promoting public-private
partnerships.
Despite broad-scale recognition of the need to adapt, such as the World
Economic Forum’s (2012) ranking of the failure to adapt as one of the
highest global risks and on a par with terrorism, and despite some
examples of private sector engagement in adaptation, most assessments
conclude that action in each of the potential arenas has been slow to
emerge and that sharing of knowledge and experience has been limited
(Khattri et al., 2010; Agrawala et al., 2011). KPMG International (2008)
concluded that, while companies are well used to managing business
risk, they have yet to integrate the long-term risks of climate change
into these systems. Nor are they preparing to grasp the competitive
advantages that will accrue to those taking early action. Most of the
businesses interviewed appeared to be unsure of the scale of the threat
and opportunities for their businesses or are awaiting further guidance
and action by governments. They have trouble in accessing and applying
information on the extent of the threats and impacts from climate
change and have yet to engage in the detailed cost-benefit analysis of
adaptive actions or inaction. The European Commission (2009), using
case studies of both the public and private sectors, in eight countries,
came to similar conclusions. A survey by West and Brereton (2013) of
Australian businesses also concluded that most were only vaguely
aware of the breadth of adaptation actions that may be required and
concerned about information sharing and disclosure. The authors
suggest a framework for disclosures of relevant business activities to
both improve practice and cater for the needs of company boards,
investors, and stakeholders. A survey commissioned by the Carbon
Disclosure Project (Gardiner et al., 2007) found that among Standard
and Poor’s (S&P) 500 companies many more (about two-thirds of
respondents) were prepared to report and share information on managing
climate risks and adaptation plans than they were on mitigation.
Also, there are still questions of whether and how adaptation finance
should be made available to the private sector in developing countries
844
Chapter 14 Adaptation Needs and Options
14
a
nd under what circumstances (Persson et al., 2009; IFC, 2010; Agrawala
et al., 2011), although this is being piloted through the Pilot Program
for Climate Resilience (World Bank, 2008; IFC, 2010). Private sector
engagement and investment in adaptation is expected to make a
substantial contribution to reducing climate risk, but the distribution of
its investments will be selective and will be unlikely to match government
and civil priorities (Atteridge, 2011).
14.2.5. Information, Capacity, and Resource Needs
Successful implementation of adaptation actions depends on the
availability of information, access to technology and funding (Yohe
and Tol, 2001; Adger, 2006; Eakin and Lemos, 2006; Smit and Wandel,
2006; World Bank, 2010). In some cases a supposed lack of relevant
and legitimate information has been used as a rationale for inaction
(Moser and Ekstrom, 2011). To address this concern, the Nairobi
Work Program—established at COP-12 in 2006, with a goal of helping
developing countries make better informed decisions based on sound
scientific, technical, and socioeconomic data—has included repeated
calls for better observation systems, information sharing, and modeling
capacity (UNFCCC/SBSTA/2008/3). Developed and developing countries
have acted on this priority by establishing institutions to provide
information services at national, regional, and global scales (CCCCC,
2011; UKCIP, 2011; NCCARF, 2012), and there is an ongoing need to
promote information acquisition and dissemination (OECD, 2009).
For example, information-related adaptation needs in Africa include
additional vulnerability and impact assessments with greater continuity,
country-specific socioeconomic scenarios, and greater knowledge on
costs and benefits of different adaptation measures (Section 22.4.2).
Research and development, knowledge, and technology transfer are
also important for promoting adaptive capacity. However, providing
information does not mean that users will be able to make effective
use of it, and this information will often have to be tailored or translated
to the individual context (Webb and Beh, 2013). Efficacy of scientific
knowledge can be improved by calibration with indigenous knowledge
(Section 20.4.2). There are also opportunities for technology transfer
and innovation to be enhanced through information technologies
(Section 20.4.3).
Financial resources for adaptation have been slower to become available
for adaptation than for mitigation in both developed and developing
countries (see Chapter 17). Adaptation finance made up probably only
a fifth of initial allocations of fast-start funding (Ciplet et al., 2012); and
much of this funding has been directed toward capacity-building,
standalone projects, or pilot programs. This not only has left financial
needs, but has also meant that there is less expertise in adaptation
assessment and implementation, which is further confounded by the
complex relationship between adaptation and more common sustainable
development and/or poverty reduction planning (McGray et al., 2007).
Adaptation cost estimates have been used to estimate the financial
needs for adaptation, and these may well have been underestimated
(see Section 17.4).
Overall, at both international and national levels there is a need to
develop financial instruments that are equitable in both their delivery
of resources and in sharing the burden of supporting the instruments
(
Levina, 2007; World Bank, 2010; see also Chapters 16, 17). In this
regard, the Green Climate Fund (GCF) was established in 2010, based
on the commitment by developed country parties to mobilize jointly
US$100 billion per year by 2020 to address the needs of developing
countries (UNFCCC, 2007). Deliberation over how adaptation finance
needs will be met has become central to the UNFCCC policy agenda
(Section 16.3.4). Also, financial mechanisms for disaster risk management
are also inextricably linked with those for adaptation (Mechler et al.,
2010). Lessons from recent recovery operations have emphasized the
need for disaster preparedness along with longer term goals directed
to building resilience, including maximizing the employment-creation
benefits of adaptation measures (Harsdorff et al., 2011.)
Finances required in the future for climate change are estimated to
approach levels on the order of current development expenditure, and
there is a large gap in funding available for climate change responses
in developing countries (Peskett et al., 2009). Therefore, there is a related
need to design delivery channels so that funding benefits the poor, as
they often are most vulnerable to the impacts of climate change and
climate-related disasters. As well as channeling adaptation finance to
governments, there is a need for finance to reach the most vulnerable
people and for approaches to enable stakeholder participation (Section
15.2.3). For example, for adaptation financing, working at the
subnational level will be important and mechanisms such as microfinance
may be effective (Agrawala and Carraro, 2010). Another important
concern is that, with new money being made available for climate
change research, policy development, and practice, people may place
too much emphasis on addressing climate change as an isolated priority
to the detriment of other equally pressing social, economic, and
environmental issues (Ziervogel and Taylor, 2008). For example, in small
islands, there are concerns that placing adaptation above the critical
development needs of the present could inadvertently reduce resilience
(see Section 29.6).
14.3. Adaptation Options
Identifying needs stemming from climate risks and vulnerabilities
provides a foundation for selecting adaptation options. Over the years,
a number of categories of options have been identified. These options
include a wide range of actions that, as summarized in Table 14-1, are
organized into three general categories: structural/physical, social, and
institutional.
There are many different ways that the range of adaptation options
available could be categorized (Burton, 1996), thus any categorization
is unlikely to be universally agreed on; but this aims to take into account
the diversity of adaptation options for different sectors and stakeholders.
Some options cut across several categories. National, sectoral, or local
adaptation plans are likely to include a number of measures that are
implemented jointly from across various categories including structural,
institutional, and social options. Furthermore, some adaptation options are
interrelated. For instance, institutions and information are prerequisites
for effective early warning systems.
Adaptation constraints and limits mean not all adaptation needs will
be met, and not all adaptation options will be possible (see Chapter 16,
845
Adaptation Needs and Options Chapter 14
14
particularly Section 16.7.1). Moreover, adaptation options are not
available to meet all adaptation needs. For instance, adaptation options
are poorly developed for the broader set of impacts on ocean systems
(see Section 30.6). There is also often going to be a gap between
adaptation needs and the effectiveness of the options to meet these
needs even when well resourced and well implemented. Some of this
gap may be met by procedures to deal with loss and damage (Section
19.7) and some adaptation deficit will remain with us. Many of the
adaptation options intersect with vulnerability reduction and development
options that build adaptive capacity and address the “adaptation deficit”
which may be seen as part of a wider “development deficit” (McGray et
al., 2007; see also Section 2.4.2).
14.3.1. Structural and Physical Options
This category highlights adaptation options that are discrete, with clear
outputs and outcomes that are well defined in scope, space, and time.
They include structural and engineering options; the application of
discrete technologies; the use of ecosystems and their services to serve
adaptation needs; and the delivery of specific services at the national,
regional, and local levels. This category includes much of the notion of
“concrete activities that reflect the priority of the Adaptation Fund,
where the focus is on “discrete activities with a collective objective(s)
and concrete outcomes and outputs that are more narrowly defined in
scope, space, and time” (Adaptation Fund Board, 2013).
Category Examples of options*
S t r u c t u r a l /
physical
E
ngineered
and built
e
nvironment
S
ea walls and coastal protection structures (5.5.2 and 24.4.3.5; Figure 5-5); fl ood levees and culverts (26.3.3); water storage and pump storage (Section
23.3.4); sewage works (3.5.2.3); improved drainage (24.4.5.5); beach nourishment (5.4.2.1); ood and cyclone shelters (11.7); building codes (Section 8.1.5);
s
torm and waste water management (8.2.4.1); transport and road infrastructure adaptation (8.3.3.6); oating houses (8.3.3.4); adjusting power plants and
electricity grids (10.2.2; Table 10-2)
T
echnological New crop and animal varieties (7.5.1.1.1, 7.5.1.1.3, 7.5.1.3; Box 9-3; Table 9-7); genetic techniques (27.3.4.2); traditional technologies and methods (7.5.2,
27.3.4.2, 28.2.6.1, and 29.6.2.1); effi cient irrigation (10.3.6 and 22.4.5.7; Box 20-4); water saving technologies (24.4.1.5 and 26.3.3) including rainwater
h
arvesting (8.3.3.4); conservation agriculture (9.4.3.1 and 22.4.5.7); food storage and preservation facilities (22.4.5.7); hazard mapping and monitoring
t
echnology (15.3.2.3 and 28.4.1); early warning systems (7.5.1.1, 8.1.4.2, 8.3.3.3, 11.7.3, 15.4.3.2, 18.6.4, 22.2.2.1, 22.3.5.3, and 22.4.5.2); building
insulation (8.3.3.3); mechanical and passive cooling (8.3.3.3); renewable energy technologies (29.7.2); second-generation biofuels (27.3.6.2)
E
cosystem-
based
a
C
ross Chapter Box CC-EA, Ecological restoration (5.5.2, 5.5.7, 9.4.3.3, and 27.3.2.2; Box 15-1) including wetland and fl oodplain conservation and
restoration; increasing biological diversity (26.4.3); afforestation and reforestation (Box 22-2); conservation and replanting mangrove forest (15.3.4 and
2
9.7.2); bushfi re reduction and prescribed fi re (Section 24.4.2.5; Box 26-2); green infrastructure (e.g., shade trees, green roofs) (8.2.4.5, 8.3.3, 11.7.4,
and 23.7.4); controlling overfi shing (28.2.5.1 and 30.6.1); sheries co-management (9.4.3.4 and 27.3.3.1); assisted migration or managed translocation
(4.4.2.4, 24.4.2.5, 24.4.3.5, and 25.6.2.3); ecological corridors (4.4.2.4); ex situ conservation and seed banks (4.4.2.5); community-based natural resource
m
anagement (CBNRM) (22.4.5.6); adaptive land use management (Section 23.6.2)
Services Social safety nets and social protection (Box 13-2; 8.3, 17.5.1, and 22.4.5.2); food banks and distribution of food surplus (29.6.2.1); municipal services
i
ncluding water and sanitation (3.5.2.3 and 8.3.3.4); vaccination programs (11.7.1), essential public health services (11.7.2) including reproductive health
services (11.9.2) and enhanced emergency medical services (8.3.3.8); international trade (9.3, 9.4, and 23.9.2)
Social
E
ducational Awareness raising and integrating into education (11.7, 15.2, and 22.4.5.5); gender equity in education (Box 9-2); extension services (9.4.4); sharing
local and traditional knowledge (12.3.4 and 28.4.1) including integrating into adaptation planning (29.6.2.1); participatory action research and social
learning (22.4.5.3); community surveys (Section 8.4.2.2); knowledge-sharing and learning platforms (8.3.2.2, 8.4.2.4, 15.2.4.2, and 22.4.5.4); international
c
onferences and research networks (8.4.2.5); communication through media (22.4.5.5)
Informational Hazard and vulnerability mapping (11.7.2, 8.4.1.5); early warning and response systems (15.4.2.3 and 22.4.5.2) including health early warning systems
(
11.7.3, 23.5.1, 24.4.6.5, and 26.6.3); systematic monitoring and remote sensing (15.4.2.1 and 28.6); climate services (2.3.3) including improved forecasts
(27.3.4.2); downscaling climate scenarios (8.4.1.5); longitudinal data sets (26.6.2); integrating indigenous climate observations (22.4.5.4, 25.8.2.1, and
2
8.2.6.1); community-based adaptation plans (5.5.1.4 and 24.4.6.5) including community-driven slum upgrading (8.3.2.2) and participatory scenario
development (22.4.4.5)
Behavioral Accommodation (5.5.2); household preparation and evacuation planning (23.7.3); retreat (5.5.2) and migration (29.6.2.4), which has its own implications
for human health (11.7.4) and human security (12.4.2); soil and water conservation (23.6.2 and 27.3.4.2); livelihood diversifi cation (7.5.1.1, 7.5.2, and
22.4.5.2); changing livestock and aquaculture practices (7.5.1.1); crop-switching (22.3.4.1); changing cropping practices, patterns, and planting dates
(7.5.1.1.1, 23.4.1, 26.5.4, and 27.3.4.2; Table 24-2); silvicultural options (25.7.1.2); reliance on social networks (Section 29.6.2.2)
Institutional
Economic Financial incentives including taxes and subsidies (Box 8-4; 8.4.3 and 17.5.6); insurance (8.4.2.3, 13.3.2.2, 15.2.4.6, 17.5.1, 26.7.4.3, and 29.6.2.2; Box 25-7)
including index-based weather insurance schemes (9.4.2 and 22.4.5.2); catastrophe bonds (8.4.2.3 and 10.7.5.1); revolving funds (8.4.3.1); payments for
ecosystem services (9.4.3.3 and 27.6.2; Table 27-7); water tariffs (8.3.3.4.1 and 17.5.3); savings groups (8.4.2.3 and 11.7.4; Box 9-4); microfi nance (Box 8-3;
22.4.5.2); disaster contingency funds (22.4.5.2 and 26.7.4.3); cash transfers (Box 13-2)
Laws and
regulations
Land zoning laws (22.4.4.2 and 23.7.4); building standards (8.3.2.2, 10.7.5, and 22.4.5.7); easements (27.3.3.2); water regulations and agreements (26.3.4
and 27.3.1.2); laws to support disaster risk reduction (8.3.2.2); laws to encourage insurance purchasing (10.7.6.2); defi ning property rights and land tenure
security (22.4.6 and 24.4.6.5); protected areas (4.4.2.2); marine protected areas (Box CC-CR Chapter 6; 23.6.5 and 27.3.3.2); fi shing quotas (23.9.2); patent
pools and technology transfer (15.4.3 and 17.5.5)
Government
policies and
programs
National and regional adaptation plans (15.2 and 22.4.4.2; Box 23-3) including mainstreaming climate change; sub-national and local adaptation plans
(15.2.1.3 and 22.4.4.4; Box 23-3); urban upgrading programs (8.3.2.2); municipal water management programs (8.3.3.4; Box 25-2); disaster planning and
preparedness (11.7); city-level plans (8.3.3.3 and 27.3.5.2; Boxes 26-3 and 27-1), district-level plans (26.3.3), sector plans (26.5.4), which may include
integrated water resource management (3.6.1 and 23.7.2), landscape and watershed management (4.4.2.3), integrated coastal zone management (2.4.3,
5.5.4.1, and 23.7.1), adaptive management (2.2.1.3 and 5.5.1.4; Box 5-2), ecosystem-based management (6.4.2.1), sustainable forest management (2.3.4),
sheries management (7.5.1.1.3 and 30.6.2.1), and community-based adaptation (5.5.4.1, 8.4, 15.2.2, 21.3.2, 22.4.4.5, 24.5.2, 29.6.2.2, and 29.6.2.3; Tables
5-4 and 8-4; FAQ 15.1)
Table 14-1 | Categories and examples of adaptation options.
Notes: These adaptation options should be considered overlapping rather than discrete, and are often pursued simultaneously as part of adaptation plans. Examples given can be
relevant to more than one category.
a
A number of these would fall under the term “green infrastructure” in some European Commission documents (European Commission, 2009).
*WGII AR5 sections containing representative sample of adaptation options.
846
Chapter 14 Adaptation Needs and Options
14
14.3.1.1. Engineering and Built Environment
Engineering, and the multidisciplinary teams engineers work with
(architects, planners, legal experts, etc.), is often at the forefront of
delivering adaptation technologies and strategies (Dawson, 2007). Most
engineering options are expert driven, capital-intensive, large-scale, and
highly complex (McEvoy et al., 2006; Morecroft and Cowan, 2010;
Sovacool, 2011). While many of the engineering options—including
management of storm and waste water flow (both inland and coastal),
flood levees, seawalls, upgrading existing infrastructures to improve
wind and flooding resilience, beach nourishment, and retrofitting
buildings (Blanco et al., 2009; Koetse and Rietveld, 2012; Ranger and
Garbett-Shiels, 2012)—are extensions and improvements of existing
practices, plans, and structures; some newer projects are now integrating
changed climate risk into the initial design. For example, during the
engineering design of the Qinghai-Tibet Railway, various measures were
proposed to ensure the stability of the railway embankment in
permafrost regions (Wu et al., 2008). Section 5.5.4.1 describes how new
coastal protection structures in Japan are being upgraded to take into
account future sea level rise.
Engineered adaptation options typically have two general limitations.
First, they often must cope with uncertainties associated with projecting
climate impacts arising from assumptions about future weather,
population growth, and human behavior (Dawson, 2007; Furlow et al.,
2011). Second, the longevity and cost of engineered infrastructure affect
the feasibility at the outset (Koetse and Rietveld, 2012). They also are
subject to consequences that were not anticipated. For example, after
coastal eastern England was devastated by the North Sea storm surge
in 1953, hard-engineered sea walls were put in place to protect the
coast from erosion and inundation. However, the engineered alterations
resulted in a new array of coastal instabilities, including disturbances
in sediment supply and damages to coastal ecosystems (Adger et al.,
2009b; Turner et al., 2010). As a result, many have promoted a “phased
capacity expansion strategy, which allows engineered projects to
undertake design modification as conditions or knowledge change and
facilitate incremental project construction to ease the burden of upfront
financing (Colombo and Byer, 2012). An example is the Thames Estuary
2100 Plan (see Box 5-1) and, in the Netherlands, the Delta Works
(Arnold et al., 2011).
14.3.1.2. Technological Options
Recent advances in technology and information are being combined
with engineering structural adaptation measures in various applications.
In the food and agriculture sector, a suite of adaptation options have
been developed and applied to reduce the adverse impacts of climate
change on production (FAO, 2007; Stokes and Howden, 2010; see also
Chapters 7, 9). Technologies range from more efficient irrigation and
fertilization methods, plant breeding for greater drought tolerance, and
adjusting planting based on projected yields (Semenov, 2006, 2008;
Bannayan and Hoogenboom, 2008) to transfers of traditional technologies
such as floating gardens (Irfanullah et al., 2011a,b). Technology options
for climate change adaptation include both hard” and soft”
technologies, and not only new technologies but also indigenous and
locally made appropriate technology (Glatzel et al., 2012). For example,
t
raditional construction methods have been identified across the Pacific
as a means of adapting to tropical cyclones and floods, including
building low aerodynamic houses and the use of traditional roofing
material such as sago palm leaves to reduce the hazard of iron roofing
being blown away in high winds (see Section 29.6.2.1). Centralized
high-technology systems can increase efficiency under normal conditions,
but also risk cascading malfunctions in emergencies (Section 15.4.3).
With the rapid diffusion of Information and Communication Technologies
(ICT) such as mobile phones and the Internet, the unprecedented speed
at which information is produced and shared is posing a new set of
possibilities for communication. ICT provides opportunities for top-down
dissemination of relevant information such as weather forecasts, hazard
warnings, market information, information sharing, and advisory services.
It can also generate essential information through bottom-up processes
such as “crowd sourcing” of useful information such as local flood levels,
disease outbreaks, and the management of disaster responses. MacLean
(2008) identifies three kinds of effects of the rapid advances in ICT on
adaptation and development in general: direct use for monitoring and
measuring climate change as described earlier, as a medium for raising
awareness, and as an enabler for a “networked governance” based on
networked open organizations. Pant and Heeks (2011) emphasize the
difficulty in foreseeing additional applications arising from planned ICT
applications exploiting local creativity and entrepreneurship, but warn
that ICT itself is not a panacea and that the most effective applications
are embedded in other societal behaviors.
There are repeated calls for technology transfer to and sharing between
developing countries in adaptation to match the programs associated
with mitigation (UNFCCC, 2006). Unlike mitigation, where low-carbon
technologies are often new and protected by patents held in developed
countries, in adaptation the technologies are often familiar and applied
elsewhere. For example, agricultural practices that are well known in a
region some distance away may now be applicable but unfamiliar
within a region of interest (Irfanullah et al., 2011a). Thus, technology
transfer in adaptation may be easier than for mitigation. For example,
to address water scarcity issues in many places, water storage, use, and
water efficiency technologies will all need to be more widely available.
See also Section 15.3.4 on technology transfer and diffusion.
14.3.1.3. Ecosystem-Based Adaptation
Ecosystem-based adaptation (EBA)—which is the use of biodiversity
and ecosystem services as part of an overall adaptation strategy to help
people to adapt to the adverse effects of climate change (Convention
on Biological Diversity, 2009)—is becoming an integral approach to
adaptation (see Box CC-EA). Often, when faced with climate-related
threats, first consideration is given to engineered and technological
approaches to adaptation. However, working with nature’s capacity and
pursing ecological options, such as coastal and wetland maintenance
and restoration, to absorb or control the impact of climate change in
urban and rural areas can be efficient and effective means of adapting
(Huntjens et al., 2010; Jones et al., 2012). The use of mangroves and
salt marshes as a buffer against damage to coastal communities and
infrastructure has been well researched and found to be effective both
physically and financially in appropriate locations (Day et al., 2007;
847
Adaptation Needs and Options Chapter 14
14
M
orris, 2007). They can also provide biodiversity co-benefits, support
fish nurseries, and have carbon sequestration value (Adger et al., 2005;
Reid and Huq, 2005; Convention on Biological Diversity, 2009). Other
EBA activities include integrative adaptive forest management (Bolte
et al., 2009; Guariguata, 2009; Reyer et al., 2009), and the use of agro-
ecosystems in farming systems (Tengö and Belfrage, 2004), ecotourism
activities (Adler et al., 2013), land and water protection and management,
and direct species management (Mawdsley et al., 2009). An analysis of
the 44 submitted NAPAs showed that the value of ecosystem services
was acknowledged in 50% of the national proposals and, in 22% of
the proposals, included the use of ecosystem services mostly in support
of other adaptation activities including infrastructure, soil conservation,
and water regulation (Pramova et al., 2012).
Green infrastructure (including the use of green roofs, porous pavements,
and urban parks) can improve storm water management and reduce
flood risk in cities, and can moderate the heat-island effect, as well as
having co-benefits for mitigation (Section 8.3.3.7). For example, New
York City has a well-established program to enhance its water supply
through watershed protection that is cost-effective compared to
constructing a filtration plant (Section 8.3.3.7). However, there are
trade-offs relating to land use and the availability of space for people
and social, economic, and environmental activities. For example,
providing an effective wetland buffer for coastal protection may require
emphasis on silt accumulation possibly at the expense of wildlife values
and recreation (Convention on Biological Diversity, 2009; Dudley et al.,
2010). Similarly Goldstein et al. (2012) found that in land use decision
making in Hawaii trade-offs existed between carbon storage and water
quality, and between environmental improvement and financial returns.
A further consideration is that ecosystem-based approaches are often
more difficult to implement and assess as they usually require cooperation
across institutions, sectors, and communities, and their benefits are also
spread across a similarly wide set of stakeholders (Jones et al., 2012).
One of the major barriers to EBA is the lack of comparable standards and
methodologies applied to engineering approaches, thus demonstrating the
need for more dialog between engineering and ecological communities.
14.3.1.4. Service Options
Service provision consists of a diverse range of specific and measurable
activities. For instance, one measure to support the most vulnerable
populations is social safety nets. Efforts to address child malnutrition,
which often result from loss of livelihood due to extreme weather events,
particularly floods and droughts (Hoddinott et al., 2008; Alderman et
al., 2009), offer an example of how safety nets can serve as a climate
adaptation measure. While some studies have shown that food programs
can be counterproductive to promoting livelihoods over the longer term
or may not prevent malnutrition in non-emergency situations (e.g.,
Bhutta et al., 2008), programs designed to provide support via food
programs, micro-finance, or insurance at times of extreme events can
provide an important bridge for vulnerable populations (Hoeppe and
Gurenko, 2006; Hochrainer et al., 2007; Alderman et al., 2009; Meze-
Hausken et al., 2009).
Public health services also are important for tackling projected increases
of disease incidences spurred on by climate change (Ebi and Burton, 2008;
G
arg et al., 2009; Edwards et al., 2011; Huang et al., 2011). For example,
in countries where malaria is endemic, frequent adaptation options for
addressing possible outbreaks include increasing use of mosquito nets,
insecticides sprays, and controlling mosquito breeding by reclaiming
land and filling drains (Garg et al., 2009). Governments at all levels are
often also responsible for maintaining adequate access to services that
are projected to be further stressed due to climate change (Laukkonen
et al., 2009). Frequently cited options in this domain include, among
others, clearing drainage systems to prevent floods, diversifying water
supply services to account for changing water supplies (Kiparsky et al.,
2012), and maintaining open public spaces dedicated for disaster recovery
and other emergency purposes (Hamin and Gurran, 2009).
At the local level, infrastructure associated with the provision of basic
services, such as water, sanitation, solid waste disposal, power, storm
water and roadway management, and public transportation are integral
to increasing adaptive capacity (Paavola, 2008; Bambrick et al., 2011;
Barron et al., 2012; see also Section 8.2.4.1). Transport links enable
households to take part in trade, for example, to access agricultural
markets (Section 9.3.3.3.2) although supply chains can be vulnerable
to climate disruption. Housing services are particularly critical because
new patterns in temperature and precipitation will alter the habitability
and stability of residences while increased frequency and intensity of
natural disasters will place settlements and homes on both stable and
unstable land at greater risk (Satterthwaite and Dodman, 2009; see also
Section 8.3.3.3). Although one option is to relocate people inhabiting
vulnerable areas, some argue that in situ upgrading may be more cost-
effective, especially for addressing informal settlements in developing
countries (Revi, 2008).
14.3.2. Social Options
There are various adaptation options that target the specific vulnerability
of disadvantaged groups, including targeting vulnerability reduction
and social inequities. Community-based adaptation (CBA) refers to the
generation and implementation of locally driven adaptation strategies,
operating on a learning-by-doing, bottom-up, empowerment paradigm
that cuts across sectors and technological, social, and institutional
processes. Social protection schemes (see also Section 14.3.1 on services)
include public and private initiatives that transfer income or assets to
the poor, protect against livelihood risks, and raise the social status and
rights of those who are marginalized (see Box 13-2). An example of a
social protection scheme aimed at moving beyond repeated relief
interventions is Ethiopia’s Productive Safety Net Program (PSNP) (Section
22.4.5.2).
The complexity of climate adaptation means that adaptation options
are heavily influenced by forms of learning and knowledge sharing
(Collins and Ison, 2009). Many scholars have noted that education is a
key indicator for how people select adaptation options (Chinowsky et
al., 2011; Sovacool et al., 2012), while a lack of education is a constraint
that contributes to vulnerability (Paavola, 2008). For example, in a study
of how farmers in the Nile Basin of Ethiopia select adaptation options,
the researchers found a positive relationship between the education level
of the household head and the adoption of improved technologies and
adaptation to climate change (Deressa et al., 2009a,b). In Bangladesh,
848
Chapter 14 Adaptation Needs and Options
14
e
ducation about disaster responses was greatly assisted by rising literacy
rates, especially among women (Section 11.7).
Awareness raising, extension, outreach, community meetings, and other
educational programs are important for disseminating knowledge about
adaptation options (Aakre and Rübbelke, 2010; Birkmann and Teichman,
2010) as well as for helping to build social capital that is critical for
social resilience (Adger, 2003; Krasny et al., 2010; Wolf et al., 2010). In
this regard, education can be seen as a public good that promotes
dialog and networks (Boyd and Osbahr, 2010), and therefore allows the
development of resilience at both the level of the individual learner and
at the level of socio-ecological systems (Krasny et al., 2010). Research
partnerships and networks can facilitate knowledge-sharing and
awareness raising at all levels from small groups of individuals to large
institutions (Section 8.4.2.5). Communication and dialog on adaptation
can be a two-way flow of information. Adaptation has itself been
described as a social learning process (Section 15.3.1.2), and a number
of initiatives in Africa emphasized the importance of iterative and
experiential learning for flexible adaptation planning (Suarez et al.,
2009; see also Section 22.4.5.3). In Maryland a half-day role-playing
process has been designed to both help local people, working with key
local and state experts and planners, to plan and prepare for sea level
rise and other coastal impacts. It allows them to experience first hand
the diversity of stakeholders, the conflicting decisions to be made, and
the need to communicate throughout their community to adapt to new
risks (Anon, 2009). A similar role-playing game has been developed for
the Chesapeake Bay of the eastern United States (Learmonth et al.,
2011).
Informational strategies are integral to adaptation. Early warning
systems are critical to ensuring awareness of natural hazards and to
promoting timely response, including evacuation. A number of approaches
are being employed around the world, including tone alert radio,
emergency alert system, presentations, and briefings (Van Aalst et al.,
2008; Ferrara de Giner et al., 2012). Heat wave and health warning
systems (HHWS) can be designed to prevent negative health impacts,
by predicting possible health outcomes, identifying triggers, and
communicating prevention responses (Section 11.7.3).
Climate services initially emerged as an expansion of the tasks provided
by weather services, and can act as “knowledge brokers” that establish
a dialog between science and the public, to facilitate decision support
(Section 2.4.1.2). Linking indigenous and conventional climate
observations can add value, for example, in western Kenya, where
scientists have worked with local rainmakers to develop consensus
forecasts (Section 22.4.5.4). Awareness raising through scenario
development, computer modeling, and role playing is effective in
preparing both responsible authorities and the public. As previously
noted, ICT is facilitating rapid dissemination of information. However,
low-tech measures such as brochures, public service announcements,
and direct contact with local residents also are important to fostering
awareness and response especially where access to ICT is limited or
expensive (National Research Council, 2011).
Behavioral measures are among the suite of options that are integral
to advancing climate adaptation. Government incentives can spark
behavioral change. For example, to slow runoff into storm sewers and
r
educe flooding, a number of cities in the USA run “Disconnect your
Downspout programs to urge homeowners to redirect water from their
roof into a storage tank or small wetlands. These programs will provide
information to households and some offer rebates on supplies. Many
poor and vulnerable communities have taken steps to adapt to changes
in climate, particularly those in flood-prone areas. For instance, some
local communities in Manila are increasing the number of floors in homes
and building makeshift bridges (Porio, 2011). Behavioral adaptation can
include livelihood diversification, which has long been used by African
households to cope with climate shocks and spread risk (Section 22.4.5.2).
Labor migration can be an important strategy for reducing vulnerability
to different sources of stress as it helps households diversify their
livelihoods (Banerjee et al., 2013). However, migration and relocation
do have implications for family relations, health, and human security
(see Sections 11.7.4, 12.4.2).
14.3.3. Institutional Options
Numerous institutional measures can be used to foster adaptation.
These range from economic instruments such as taxes, subsidies, and
insurance arrangements to social policies and regulations (de Bruin et
al., 2009; Hallegatte, 2009; Heltberg et al., 2009). For instance, in the
USA, post-disaster funds for loss reduction are added to funds provided
for disaster recovery and can be used to buy out properties that have
experienced repetitive flood losses and to relocate residents to safer
locations, to elevate structures, and to assist communities with purchasing
property and altering land-use patterns in flood-prone areas, as well as
undertaking other activities designed to lessen the impacts of future
disasters not only on habitation but also on more effective food
production and other livelihoods (FEMA, 2010). Uptake of climate risk
insurance is hindered by expensive premiums. The Caribbean Catastrophe
Risk Insurance Facility (CCRIF) pools together country-level risks into a
more diversified portfolio to offer lower premiums for immediate post-
disaster responses (Section 29.6.2.2).
Laws, regulations, and planning measures such as protected areas,
building codes, and re-zoning are institutional measures that can improve
the safety of hazard-prone communities by designating land use to support
resilience (Biderman et al., 2008; Bartlett et al., 2009). For example,
marine protected areas (MPAs) have the potential to increase ecosystem
resilience and increase recovery of coral reefs after mass coral bleaching
(see Chapter 6, Box CC-CR). While zoning can be used to procure sites
for low-income populations (Biderman et al., 2008; Bartlett et al., 2009;
Satterthwaite and Dodman, 2009), if it increases property and housing
values it also has the potential to exclude the poor from these areas.
Legal rights can also determine adaptive capacity as well as access to
resources. Land tenure security in Africa is widely accepted as being
critical for enabling people to make longer-term decisions, such as
changing farming practices (Section 22.4.6).
A number of funding and financial issues are linked to institutions. At
the international level, agreements and donors have a critical role to
play in promoting and supporting the allocation and flow of financial
resources (OECD, 2011). For instance, the Adaptation Fund, which is set
up under the Kyoto Protocol and funded through a levy on most Clean
Development Mechanism (CDM) projects, is of particular importance to
849
Adaptation Needs and Options Chapter 14
14
d
eveloping countries as it is pioneering the direct access mechanism
which allows countries to access funds without having to work through
a multilateral development agency. The direct access mechanism
highlights the role of institutions in building and maintaining capacity,
not just in the technical aspects of adaptation assessment and project
design, but also in financial management and due diligence (Brown et
al., 2011).
Effective governance is important for efficient operations of institutions.
In general, governance rests on the promotion of democratic and
participatory principles as well as on ensuring access to information,
knowledge, and networks. Institutional strengthening and capacity
building has been highlighted as a priority need in developing countries
(Kumamoto and Mills, 2012). In assessment of river basin planning in
Brazil, Engle and Lemos (2010) found that improving governance
mechanisms appears to enhance adaptive capacity. Similarly water-
trading schemes facilitated by new government measures reduced the
impact of a major drought on the economy in Australia (Mallawaarachchi
and Foster, 2009). The effectiveness of such approaches depends on
both government will and capacity building among those affected.
In terms of national or local adaptation planning and policy making,
Chapter 15 emphasizes that it can be challenging for governments to
move beyond adaptation planning to implementation (Section 15.2.2).
In an evaluation of one of the earliest national adaptation strategies
for Finland, it was found that few measures had been implemented
except in the water sector (see Box 23-2). Adaptation planning can
occur at a number of spatial scales including at the national, regional,
city, district, or local community level. Action plans can include a range
of adaptation options including structural, social, and regulatory
measures. For example, the city of Quito has proposed developing dams,
encouraging a culture of rational water use, reducing water losses, and
developing mechanisms to reduce water conflicts (Section 8.3.3.4).
Table 25-5 lists various urban climate change adaptation options and
their barriers to adoption. See also Section 15.3 on local adaptation
plans.
Institutional adaptation options include the use of various decision-
making and adaptation planning tools (Chapter 15) including iterative
risk management (Chapter 2). There are various decision-making
paradigms that can guide adaptation actions. For example, prominent
institutional frameworks used for management of coastal areas include
Integrated Coastal Zone Management (ICZM) and Adaptive Management
(see Section 5.5.1.4). At the local scale, community-based adaptation
refers to the generation and implementation of locally driven adaptation
strategies through a learning-by-doing, bottom-up approach (Section
5.5.1.4). Community-based approaches to adaptation can also be
mainstreamed into local or regional plans. Refer to Table 5-4 for a
description of community-based adaptation options for coastal areas.
14.3.4. Selecting Adaptation Options
Selecting specific adaptation options can be challenging partly due to
the rate, uncertainty, and cumulative impacts of climate change. How
adaptation is framed will have an impact on how adaptation options
are selected (Fünfgeld and McEvoy, 2011). Policy and market conditions
m
ay be a stronger driver of behavior than the observed climate itself
(Berkhout et al., 2006). Also, rarely will adaptation options be designed
to address climate risks or opportunities alone (IPCC, 2007b); instead
actions will often be undertaken with other goals (such as profit or
poverty reduction) in mind, while also achieving climate-related co-
benefits. Gains in reduced vulnerability, enhanced resilience, or greater
welfare will often be co-benefits generated as a result of changes and
innovations driven by other factors (Khan et al., 2013). Rather than
focusing on adaptation options addressing specific dimensions of climate
change, more attention is being paid to mainstreaming climate change
into wider government policy and private sector activities (see Sections
15.2.1, 15.5.1; Sietz et al., 2011a).
Selection and prioritization of adaptation options is important because
not all adaptation options will be possible owing to constraints such as
insufficient local resources, capacities, and authority (see Section 16.7).
Furthermore, some adaptation options can be maladaptive if they foreclose
other options (see Section 14.7). The viability of adaptation options is
dependent on the time scale and climate scenario, emphasizing that
selecting adaptation options is an iterative process.
A variety of systematic techniques have been developed for selecting
options (e.g., see Section 15.4; De Bruin et al., 2009;ssel, 2009; Ogden
and Innes, 2009). Quantification and other systematic approaches to
selecting options have many virtues. However, they also have limitations.
For instance, most of these methodologies do not account for a range
of critical factors such as leadership, institutions, resources, and barriers
(Smith et al., 2009). For example, cost-benefit analysis of adaptation
options requires valuation of non-market costs and benefits, which
can be impractical (Section 17.3.2). Strategies dominating the early
adaptation literature emphasized maintaining the current system and
minimizing costs while achieving some form of benefit. For instance,
no-regrets measures both reduce climate risk and provide other social,
economic, or environmental benefits (Hallegatte, 2009). Risk management
approaches often lead to no-regrets, low-regrets, or win-win options,
while multi-criteria analysis (MCA) allows assessment of options against
different criteria, as was used in the preparation of NAPAs (UNFCCC,
2011).
As ideas about adaptation have evolved, there has been a shift in
ambition from traditional approaches that emphasize maintaining the
status quo to more dynamic and integrative strategies (see also Sections
2.4.3, 14.1, 16.4.2, 20.5). Adaptive management places an emphasis on
taking action and then using the lessons learned to inform future actions
in order to make better informed, and often incremental, decisions in
the face of uncertainty (Sections 2.2.1.3, 14.4). Lempert and Schlesinger
(2000) have proposed that adaptation options should be robust
against a wide range of plausible climate and societal change futures.
Emerging trends in adaptation place an emphasis on the need for more
transformational changes, which has a distinct logic that differs from
traditional strategies (see Section 14.1).
As research and experience in the practice of adaptation grows, an ever
increasing number of considerations have been advanced as being
important in guiding the selection and sequencing of adaptation options.
It is unlikely that every adaptation program can ever fully meet each of
these considerations, especially as they are increasingly integrated with
850
Chapter 14 Adaptation Needs and Options
14
wider social or development goals, but Table 14-2 seeks to outline the
most common considerations and point to sources in this volume and
the literature for a discussion of some of the core issues.
14.4. Adaptation Assessments
14.4.1. Purpose of Assessments
Identifying adaptation needs requires an assessment of the factors that
determine the nature of, and vulnerability to, climate risks (climate change
assessments, climate impacts and risk assessments, and vulnerability
assessments) and an assessment of adaptation options to reduce risks
(adaptation assessment). The various types of climate change assessments
differ in that they pursue different goals, are underpinned by different
theoretical frameworks, and rely on different forms of data and ultimately
may lead to different adaptation responses (Fünfgeld and McEvoy, 2011).
Assessments help decision makers understand the impacts, vulnerability,
and adaptation options in a region, country, community, or sector. They
are often characterized into “top-down and “bottom-up” assessments.
Top-down assessments are used to measure the potential impacts of
climate change using a scenario and modeling driven approach. Bottom-
up assessments begin at the local scale, address socioeconomic responses
to climate, and tend to be location specific (Dessai and Hulme, 2004).
They are often used to determine the vulnerability of different groups
to current and/or future climate change and their adaptation options,
using stakeholder intervention and analyzing socioeconomic conditions
and livelihoods (UNFCCC, 2010). There are also policy-based assessments,
which assess current policy and plans for their effectiveness under climate
change within a risk-management framework (UNDP, 2004, 2005). The
evolution of assessments has led to a more thorough assessment of
society’s ability to respond to risks through various adaptations, which
can help guide allocation of adaptation resources (Füssel and Klein,
2006). In practice assessments have become increasingly complex, often
combining elements of top-down and bottom-up approaches (e.g.,
Dessai et al., 2005). Decision makers use both in the policy process
(Kates and Wilbanks, 2003; McKenzie Hedger et al., 2006).
14.4.2. Trends in Assessments
A variety of frameworks have been developed for the assessment of
climate impacts, vulnerability, and adaptation (Fünfgeld and McEvoy,
2011). “Impacts-based” approaches focus primarily on the biophysical
climate change impacts to which people and systems need to adapt.
“Vulnerability-basedapproaches focus on the risks themselves by
concentrating on the propensity to be harmed, then seeking to maximize
potential benefits and minimize or reverse potential losses (Adger, 2006;
IPCC, 2007b).Adaptation-based” approaches examine the adaptive
capacity and adaptation measures required to improve the resilience
or robustness of a system exposed to climate change (Smit and Wandel,
2006). In practice these approaches are interrelated, especially with
regard to adaptive capacity (O’Brien et al., 2007). An evolution in the
conceptualization of risk and vulnerability in the past decade has led
to more holistic and integrated approaches to assessment, aiming
toward a more systemic understanding of the complexity of human-
environment interactions (Preston et al., 2011b).
The “standard approach” to assessment has been the climate scenario-
driven impacts-based approach, which developed from the seven-step
assessment framework of the IPCC (Carter et al., 1994; Parry and Carter,
1998): (1) define the problem (including study area and sectors to be
examined), (2) select method of problem assessment, (3) test methods/
conduct sensitivity analyses, (4) select and apply climate change
scenarios, (5) assess biophysical and socioeconomic impacts, (6) assess
autonomous adjustments, and (7) evaluate adaptation strategies. This
approach dominated the assessment sections of the first three IPCC
reports, and aims to evaluate the impacts of climate change under a
given scenario and to assess the need for adaptation and/or mitigation
851
Adaptation Needs and Options Chapter 14
14
t
o reduce any resulting vulnerability to climate risks (IPCC, 2007a). These
frameworks are described as “first generation” or “type 1” assessment
studies (Burton et al., 2002). The standard impact approach is often
described as top-down because it combines scenarios downscaled from
global climate models to the local scale with a sequence of analytical
steps that begin with the climate system and move through biophysical
impacts toward socioeconomic assessment (IPCC, 2007b). The process
of downscaling of global climate models leads to issues of uncertainty
and limited statistical confidence (Fünfgeld and McEvoy, 2011).
A new generation of scenario-based impact assessments has also
emerged linking biophysical, economic, and social analysis tools. Refer
to Section 2.3.2 for examples of large-scale and regional-scale scenario-
based vulnerability assessments that have taken place linking biophysical
and socioeconomic futures. In Europe, a study by Ciscar et al. (2011)
estimated economic welfare losses over four sectors of 0.2 to 1% by the
2080s (Section 2.3.2.1). In Australia, socioeconomic considerations are
beginning to be used to inform assessments of adaptive capacity and
vulnerability (Section 25.3.2). A risk-based framework, based on the risk
management approach, can also be used for assessing adaptation
opportunities, constraints, and limits (Section 16.2). Economic assessments
are also used to estimate the impacts of climate change, including
distributional impacts and adaptation costs and benefits (Chapter 17).
The second generation” vulnerability and adaptation assessments
(Burton et al., 2002) pay greater attention to information around
vulnerability to inform decisions on adaptation. They are characterized
by the intensive involvement of stakeholders and the participation of
vulnerable groups in decision making around adaptation options (Füssel
and Klein, 2006; LDC Expert Group, 2012). Local projects often use
participatory vulnerability assessment (PVA) methods. In Bangladesh,
c
ommunity-based adaptation has combined consensus-building and
participatory rural appraisal (PRA) to assess needs of the communities
and propose adaptation actions (Section 15.2.1). In activities by CARE,
vulnerability assessments were undertaken with men and womens
groups separately to ensure activities were gender sensitive (see Section
7.5.2). Participatory vulnerability assessments offer an opportunity to
avoid maladaptation by involving stakeholders, for example, in a
vulnerability assessment of tourism in the Mamanuca Islands in Fiji,
where stakeholders were explicitly integrated into each step of the
process (Section 29.8).
Adaptation assessments continue to evolve, but most syntheses now
include “top-down” and “bottom-up approaches, and include the
assessment of both biophysical climate change risks and the factors
that make people vulnerable to those risks. There is a shift toward
integrating community-based planning into national adaptation plans.
The Government of Nepal proposes “LAPA assessments” (Local Adaptation
Plans of Action) that seek to integrate top-down and bottom-up models
(Government of Nepal, 2011). There is also increasing attention to
institutional capacity assessments and policy environments as key factors
that can both drive vulnerability and also determine the type and success
of different adaptation options. The generic elements of adaptation and
vulnerability assessment are reflected in the the UK Climate Impacts
Program (UKCIP) guidelines presented in Figure 14-2.
14.4.3. Issues and Tensions in the Use of Assessments
Adaptation and risk assessments give rise to various tensions, three of
which are discussed here. The first is the adaptation paradox, which
recognizes that climate change is a global problem while vulnerability
R t it d i k t d i t
NO
NO
852
Chapter 14 Adaptation Needs and Options
14
i
s locally experienced (Ayers, 2011). Top-down assessments of climate
scenarios are deemed necessary in order to understand the climate
change scenarios that render climate risk. However, the factors that
make people vulnerable to climate risks are often locally generated, so
require locally driven bottom-up analysis, while factors at the national
and regional levels also determine vulnerabilities. Bottom-up analysis
tends to prioritize groups based on factors related to poverty and
development that drive vulnerability. Top-down assessments tend to
prioritize those most exposed to climate risks. Analysis in Nepal that
assessed both under-development and climate change impacts showed
that, at the household scale, there was a strong correlation between
local measures of poverty and vulnerability to climate change (Ghimire
et al., 2010). However, when indicators were aggregated at district scale,
the correlation was weaker—even when the vulnerability index used
included poverty as a proxy for adaptive capacity alongside climate
hazard risk and exposure (Ghimire et al., 2010).
There are also tensions around ownership and participation. Assessments
managed under global climate change governance structure of the UN
Framework Convention on Climate Change are developed under an
“impacts-based” paradigm (Burton et al., 2002). This impacts-based
approach requires external scientific and technological expertise for
defining climate change problems, and formulating technological
adaptation solutions, based on specific knowledge of future climate
conditions. Such assessments are necessarily “top-down” because this
expertise exists at the global and national level. At the local level, the
capacity to adapt is based on the underlying securities that determine
vulnerability to these impacts in the first place (Adger et al., 2003a).
Accessing this information requires “bottom-upand participatory
assessments that engage local vulnerable people. These vulnerable
groups and institutions often do not have access to the climate impacts
science necessary to fully apply top-down impacts-based assessments.
Some places also do not have accurate historical weather data, making
it difficult to validate climate trends and models and hence develop
evidence-based scenarios of what will happen with any degree of
accuracy (Conway, 2009).
The numerous assessments that have been carried out have led to
increased awareness among decision makers and stakeholders of climate
risks and adaptation needs and options. But this awareness is often not
translated into the implementation of even simple adaptation measures
within ongoing activities or within risk management planning. There is
a bottleneck in adaptation assessments, which may need to be overcome
by linking more directly to particular decisions and tailoring the
information to local contexts to facilitate the decision-making process
(Preston and Stafford Smith, 2009; Brown et al., 2011). Specific techniques
such as decision scaling, which seeks to understand which climate
conditions would result in hazardous conditions of concern for particular
stakeholder groups, are a step in this direction (Brown et al., 2012;
Moody and Brown, 2012). Decision support must recognize that human
psychological dimensions play a crucial role in the way people perceive
risks and make decisions (Section 2.2.1.2). Impacts and adaptation
options will also have to be successfully communicated to the local
scale. One example of this is local-scale visualization of impacts and
adaptation measures, as has taken place in British Columbia, Canada
(see Section 2.2.1.3). Use of ICT tools can foster new ways to assimilate
or translate information (see Box 15-1). Vulnerability mapping, including
t
he use of geographic information systems (GIS), can help stakeholders
to visualize the impacts of climate change on the landscape, while
integration with participatory processes can facilitate learning and
deliberation (Preston et al., 2011a).
14.4.4. National Assessments
Under the UNFCCC, all parties are encouraged (Annex 1 countries are
required) to report on activities in relation to “vulnerability assessment,
climate change impacts and adaptation measures” (FCCC/CP/1999/7).
Parties are encouraged to use the IPCC Technical Guidelines for Assessing
Climate Change Impacts and Adaptations (Carter et al., 1994) and the
UNEP Handbook on Methods for Climate Change Impacts Assessment
and Adaptation Strategies, which focuses on the impacts of sea level rise
and uses the seven-step assessment framework (described previously).
Annex 1 countries are due to submit their sixth Communications by
2014 and most non-Annex I countries are due to have submitted at
least one Communication; some are on their fifth. As such, National
Communications have formed the first avenue for assessing and
reporting on climate risk and vulnerability assessments at the national
level. Most initial National Communications to the UNFCCC produced
by developing countries were first-generation vulnerability assessments,
which did not seek to assess the feasibility of implementing adaptations
(Füssel and Klein, 2006). Undertaking such assessment is resource
intensive, underscoring the need for further resources, training, and
expertise.
There is a range of emerging national experiences on adaptation and
vulnerability assessments. For coastal areas under sea level rise, a
summary of the results from coastal vulnerability assessments is shown
in Table 5-5. Such assessments show that vulnerability is highly dependent
on the greenhouse gas (GHG) mitigation scenario. In Kenya, a study by
the Stockholm Environment Institute (SEI) estimated the economics of
climate change under a range of scenarios (see Figure 22-6) and
estimated that, by 2050, more than 300,000 people could be flooded
per year under a high-emissions scenario. In 2012, the UK’s first Climate
Change Risk Assessment (CCRA) was undertaken based on a similar
framework to that shown in Figure 14-2, to assess risks in and across
eleven sectors to inform priorities for action and appropriate adaptation
measures (DEFRA, 2012).
National Adaptation Programmes of Action are designed as a vehicle for
Least Developed Countries (LDCs) to communicate their most “urgent
and immediate adaptation needs” to the UNFCCC for funding from the
Least Developed Countries Fund (LDCF). “Urgent and immediate needs
are defined as those for which further delay in implementation would
increase vulnerability or increase adaptation costs at a later stage (LDC
Expert Group, 2009). The approaches adopted for vulnerability assessment
under NAPAs vary. Although the guidelines call for more participatory and
“bottom-up” mechanisms to be adopted, time and funding limitations
have meant that often the NAPA process remains largely top-down,
focused on impacts and consulting the communities only to verify this
information (Huq and Khan, 2006; Ayers, 2011). Moreover, available
financial resources were too limited to fully assess and address the
needs of all sectors and all vulnerable regions of the country (LDC Expert
Group, 2012; see also Section 15.2.1.2).
853
Adaptation Needs and Options Chapter 14
14
Under the Cancun Adaptation Framework (CAF), a process was
established to enable LDC parties to formulate and implement National
Adaptation Plans (NAPs). NAPs are intended to build on NAPAs but
shift the focus toward identifying medium- and long-term adaptation
needs and developing and implementing strategies and programs to
address those needs. NAPs are intended to facilitate the integration of
climate change adaptation into relevant national and subnational
development and sectoral planning (LDC Expert Group, 2012). Other
developing country parties are also invited to employ the modalities
formulated to support the national adaptation plans in the elaboration
of their planning efforts. Early guidelines (LDC Expert Group, 2009)
propose a country-specific approach tailored to national circumstances,
mixing top-down policy-first assessments with bottom-up approaches.
Recent guidelines propose that this should be non-prescriptive and
should facilitate country-driven, gender-sensitive, participatory action,
taking into consideration vulnerable groups, communities, and ecosystems
(LDC Expert Group, 2012). Refer also to Sections 2.4.3 and 15.2.1.2 for
further details of national and subnational adaptation planning including
NAPAs and NAPs.
14.5. Measuring Adaptation
Adaptation has tended to lag behind mitigation efforts both in research
and in the climate negotiations. In part this is because adaptation and
development specialists, governments, NGOs, and international agencies
have found it difficult to clearly define and identify precisely what
constitutes adaptation, how to track its implementation and effectiveness,
and how to distinguish it from effective development (Burton et al.,
2002; Arnell, 2009; Doria et al., 2009). A contributing reason is that
adaptation has no common reference metrics in the same way that
tonnes of GHGs or radiative forcing values are for mitigation. This
section seeks to explore the feasibility of finding metrics for measuring
adaptation effectiveness.
The search for metrics
6
for adaptation will remain contentious, with
many alternative uses competing for attention. This is inevitable as there
are multiple purposes and viewpoints in approaching the measurement
of adaptation (Hulme, 2009). Brooks et al. (2011) asked “what constitutes
successful adaptation” and suggested that the criteria by which success
Frequently Asked Questions
FAQ 14.1 | Why do the precise definitions about adaptation activities matter?
Humans have always adapted to changing conditions: personal, social, economic, and climatic. The rapid rate of
climate change now means that many groups, ranging from communities to parliaments, now have to factor climate
change into their deliberations and decision making more than ever before. Having a term and working definition
is always useful in discussing how to tackle a challenge as it helps define scope. Is adaptation all about minimizing
damage or are there opportunities as well? Can adaptation proceed only through deliberately planned actions
focused specifically on adaptation to climate change? How much must be known about future climates to make
decisions about adaptation? How does the adaptation of humans systems differ from adaptation in natural systems?
Can adaptation to climate change be distinguished from normal development and planning processes? Need it be?
Are we adequately adapted to current climates, or do we have an “adaptation deficit”? The phrase “maladaptation
immediately turns thoughts to how could plans go wrong and possibly cause greater suffering. A definition does
not answer all these questions but it provides a framework for discussing them.
There is also a political reason for needing a precise definition of adaptation. Developed countries have agreed to
bear the adaptation costs of developing countries to human-induced climate change and that these funds should
represent “new and additional resources,”
4
and the Cancun Agreement and subsequent discussions suggest that
for adaptation these funds could amount to tens of billions US$ per year.
5
In most cases adaptation is best carried
out when integrated with wider planning goals such as improved water allocation, more reliable transport systems,
and so forth. How much of the cost of upgrading a coastal road that is already subject to frequent damage from
bad weather should be attributed to normal development and how much to adaptation to climate change? A
precise answer may never be possible but the closer we agree as to what constitutes adaptation, the easier it will
be to come to workable agreements.
4
Bali Action Plan, 2007; FCCC/CP/2007/6/Add.1.
5
Cancun Agreements 2010, FCCC/CP/2010/7/Add.1, paras 98 & 102.
6
There is no consistent use of the terms metric, measure, and indicator in the literature. Here we try to stay as close as possible to the dictionary meanings (although they overlap).
A measure is the amount or degree of something, that is, a description of its (presumably current) state. A metric is often a group of values (measures) that taken together give
a broader indication of the state or the degree of progress to some desired state. An indicator is a sign, or estimate of the state of something and often of the future state of
something. Most often in seeking to understand the state of vulnerability or adaptation, etc., we need a metric (i.e., a group of measures) and we use the term in that way. In
describing the components of a metric we will give preference to the term indicator over measure.
854
Chapter 14 Adaptation Needs and Options
14
m
ight be assessed include feasibility, efficacy/effectiveness, efficiency,
acceptability/legitimacy, and equity (derived from Yohe and Tol, 2001;
Adger et al., 2005; Stern, 2006), to which they added sustainability
(Fankhauser and Burton, 2011). Effective integration and coherence
with wider national policies and development goals is another often
sought criterion (World Bank, 2010). Also institutions, communities, and
individuals value things differently and many of those values cannot be
captured in a comparable way within metrics (Adger and Barnett, 2009).
At least three uses of metrics for adaptation are relevant, each requiring
different characteristics of the indicators used. The first use seeks metrics
to help determine the need or determinates of that need for adaptation.
These metrics usually focus on measuring vulnerability, but that term
is not well defined. For example, Hinkel (2011) identifies six uses that
vulnerability indicators are sometimes expected to serve and concludes
that they can truly serve only their core purpose, that is, to identify
vulnerable people, communities, and regions. Further, even with metrics
focusing on vulnerability the goal often is not to produce a score or rating
to identify vulnerable groups but to elucidate information on the nature
of vulnerability and to better identify adaptation options (Smit and
Wandel, 2006; Sietz et al., 2011b). The second use of metrics relates to
measuring and tracking the process of implementing adaptive actions,
such as spending on coastal protection, the number of early warning
plans implemented as part of a program, or the number of agricultural
specialists with appropriate training in climate risks. Here the selection of
appropriate metrics is usually less contentious but there is disagreement
as to how much they capture adaptation rather than normal development.
The third use of metrics relates to measuring the effectiveness of
adaptation such as in monitoring and evaluation. This set is essential
to help measure progress and provide feedback on the effectiveness of
actions, but is among the most difficult to identify as adaption outcomes
take time to become identifiable and are often subject to evolving
conditions and objectives.
14.5.1. What Is to Be Measured?
The measurement of vulnerability is central to many adaptation metrics
and initially it was approached from an impacts point of view. Here
vulnerability is usually defined as a function of (1) exposure to specific
hazards or stressors, (2) sensitivity to their impacts, and (3) the target
population’s capacity to adapt (IPCC, 2001, Chapter 17). This approach
continues to be used as the basis of many assessments and adaptation
prioritization efforts. Recently the emphasis has moved from better
defining exposure and potential impacts to a better understanding of
the factors that affect societies’ sensitivity to those impacts and their
capacity to adapt. This reflects the increasing recognition of the
importance of considering social vulnerability alongside biophysical
vulnerability. Various terms have been used to describe these different
emphases including biophysical versus social vulnerability, outcome
versus contextual vulnerability (Section 14.2.1.1; Eakin and Luers, 2006;
Füssel and Klein, 2006; Eriksen and Kelly, 2007; Füssel, 2010), and
scientific framing versus a human-security framing of vulnerability
(O’Brien, 2007). O’Brien et al. (2007) argue that scientific and human-
security frameworks affect the way we approach adaptation, with the
scientific framework leading to building local and sectoral capacity to
make changes rather than address the fundamental causes of vulnerability,
o
r climate change itself, within their broader geopolitical and economic
contexts.
Other questions also arise even within a given conceptual framework
for considering vulnerability. A system of measurement is usually
developed to allow comparisons between different places, social groups,
or sectors of activity, although experience repeatedly cautions us to be
careful in doing so (Schröter et al., 2005). The challenge is as much of
integration across widely differing research domains and traditions (Polksy
et al., 2007). Also, a system’s vulnerability is not static but can respond
rapidly to changes in economic, social, political, and institutional
conditions over time (Smit and Pilifosova, 2003; Smit and Wandel, 2006).
Much of the effort in relation to estimating social vulnerability is
reviewed in Cutter et al. (2009).
It has also been suggested that a framework based on the concept of
resilience is more appropriate than a vulnerability framework in many
contexts (see IPCC, 2012, Chapter 2 and Section 8.3.3 for more details).
For example, in a development context resilience “evokes positive and
broad development goals (e.g., education, livelihood improvements,
food security), includes multiple scales (temporal and spatial) and
objectives, better captures the complex interactions between human
societies and their environments, and emphasizes learning and feedbacks
(Berkes, 2007; Moss et al., 2012, p. 6). A resilience approach also leads
to more focus on interactions between social and biophysical systems
(Nelson et al., 2007). However, others feel that resilience promotes too
great a focus on the return of the overall system to pre-impact conditions
and not enough on the human agents and their need to adapt to changing
conditions (Nelson et al., 2007; IPCC, 2012, Section 8.3.3). The concept
of resilience has been difficult to apply in practice and is particularly
resistant to attempts to establish commonly accepted sets of indicators.
Some (e.g., Klein et al., 2003) have suggested that resilience has become
an umbrella concept that has not been able to support effectively
planning or management.
Recently Brooks et al. (2011) have outlined a framework tracking
adaptation that combines the establishment of upstream metrics to assess
how well risks are being managed by institutions, and downstream
metrics to track whether the interventions are reducing the vulnerability
of affected groups. The upstream metrics would focus on assessments
of institutional capacity, managerial performance, and integration of
climate risk management into planning processes and tracking and
feedback processes. The downstream metrics would focus on indicators
to track development performance and changes in vulnerability.
Attribution of these changes to particular interventions would be
desirable but not essential to track progress.
But understanding vulnerability does not necessarily translate to effective
adaptation. Smit et al. (2001), Osman-Elasha et al. (2009), and others
have suggested that the focus should be on increasing adaptive capacity
within the context of the full range of biophysical and socioeconomic
stressors. However, as the scope of the metrics is widened to include
aspects of development and sustainability they often become less suitable
for other purposes such as helping to identify “the full and additional
costs of adaptation” (McGray et al., 2007). In deriving indices of
vulnerability there are again several broadly different approaches. One
is to deductively identify indicators that theoretically should be strongly
855
Adaptation Needs and Options Chapter 14
14
related to vulnerability (e.g., Dolan and Walker, 2003; Polsky et al.,
2007), while the other is inductive and uses observed data to seek
correlations between indicators and observed consequences of
vulnerability, such as the number of people killed or affected by climate-
related events in recent history. There is some commonality in identifying
the desirable criteria for selecting indicators, and though no list can ever
be complete, Table 14-3, based initially on Perch-Nielsen (2010), seeks
to bring together some of the most common criteria.
14.5.2. Established Metrics
Numerous metrics continue to be prepared for a variety of purposes
and at scales ranging from estimating the vulnerability of individuals
and communities to comparing countries. Several reviews, including
Moss (2001, 2012), Srinivasan and Prabhakar (2009), and Prabhakar
and Srinivasan (2011), discuss both the design and effectiveness of
many of the existing proposals for adaptation metrics.
14.5.2.1. Vulnerability Metrics
Eriksen and Kelly (2007) found strong divergence among five metrics
(or indices) for comparing national vulnerability published over the
period 1995–2003: the Dimensions of Vulnerability of Downing et al.
(1995); the Index of Human Insecurity (IHI) of Lonergan et al. (1999); the
Vulnerability-Resilience Indicators of Moss et al. (2001); the Environmental
Sustainability Index of the World Economic Forum (2002); and the
Country-Level Risk Measures of Brooks and Adger (2003). Between them,
29 indicators were used, with only five indicators appearing in more
than one study. They were able to compare the 20 countries ranked as
most vulnerable from three of the studies and found little overlap, with
only five countries ranked in the top 20 in more than one study.
However, it must be noted that the metrics were developed at different
times and for different purposes. They concluded that the indices
focused on measuring a snapshot of aggregate conditions rather than
on delivering guidance on societal processes that can be targeted to
reduce vulnerability.
There are a series of disaster-related indices designed to assess relative
risks across countries and regions, and to provide benchmarks on which
to assess progress. Among them are the Disaster Risk Index (UNDP,
2004); Hotspots Index (Dilley et al., 2005); the Americas Index (Cardona,
2005); and an index for South Asia (Moench et al., 2009). Again, there
has been little effort to further analyze, validate, or compare these
metrics.
14.5.2.2. Metrics for Resource Allocation
Vulnerability indices have usually been selected to better understand
the drivers of vulnerability or to compare countries, regions, communities,
and so forth in terms of the risks they face from climate change and
their capacity to deal with them. This is not necessarily the same as
designing an allocation index or rule to be used to allocate limited
resources equitably and efficiently among entities (countries, regions
or other administrative groups, or different proponents of adaptation).
For allocation, vulnerability and coping/adaptive capacity might be
856
Chapter 14 Adaptation Needs and Options
14
e
xpected to remain a core consideration, but so also should the relative
costs of implementation in relation to the potential benefits and the
ability of the recipients to absorb the funding and implement policies
and projects to actually achieve the projected benefits (UNFCCC, 2007;
Parry et al., 2009; Wheeler, 2011).
One of the longest running and prominent uses of metrics in funding is
the World Bank’s process of allocating IDA concessional funds to
developing countries, which faces many issues analogous to the same
process for adaptation. The World Bank uses the Country Policy and
Institutional Assessment (CPIA) based on 16 criteria, many qualitative,
to estimate the extent to which a country’s policy and institutional
framework supports sustainable growth and poverty reduction, and
consequently the effective use of development assistance. These criteria
are the main components used to calculate a Country Performance Rating,
which in turn is a major component, along with population and recent
performance measures, in calculating allocations to the poorest developing
countries with long-term, no interest (IDA) loans. The CPIA and the ultimate
IDA allocation formulae are controversial, much debated (Alexander,
2010), and often fine-tuned (IEG, 2009) but still commonly used as a
reference point for this type of procedure (GTZ, 2008).
An explicit example of the use, and non-use, of adaptation metrics was
in establishment of the Pilot Program for Climate Resilience (PPCR). The
governing body, made up of contributors, recipients, and other stakeholders,
set up an independent expert group to make recommendations as to
which countries might be included as pilots within the approximately
US$1 billion program (Climate Investment Funds, 2009). The expert
group refrained from using a simple index, but instead country selection
was done across nine regions and each based on a suite of indicators
appropriate for the region using expert judgment. It is interesting to note
that on moving to the next step of deciding on allocation of financial
resources to the selected pilot countries the governing body of the PPCR
chose not to use an approach based on indicators, but to provide guidance
to the countries of the possible range of funding and to base allocations
on the quality of the proposals brought forward (Climate Investment Funds,
2009). Similarly, none of the other governing bodies of international
adaptation funding mechanisms (e.g., the Global Environment Facility,
the Adaptation Fund) has chosen to use a defined set of metrics within
their decision making.
Wheeler (2011) has developed an index of vulnerability based on
weather-related disasters, sea level rise, and agricultural productivity.
The index can be adjusted according to user preferences to develop
allocation formulas based only on biophysical vulnerability, further
adjusted for economic development and governance, and finally for
project costs and probability of success. Klein and Möhner (2011) have
discussed the options for the Green Climate Fund based on experience
to date and conclude that science cannot be relied on for a single
objective ranking of vulnerability.
14.5.2.3. Metrics for Monitoring and Evaluation
The IPCCs Fourth Assessment Report provided little discussion of the role
of evaluation and monitoring of adaptation responses as a component
of building adaptive capacity (Adger et al., 2007). Preston et al. (2011a)
i
dentify three specific roles of evaluation: (1) ensuring reduction in
societal and ecological vulnerability, (2) facilitating learning and
adaptive management, and (3) providing accountability for adaptation
investments (see also GIZ, 2011). A central challenge in developing
robust monitoring and evaluation frameworks for adaptation is the
existence of multiple, valid points of view that can be used to evaluate
adaptation actions and their continuing effectiveness (Gagnon-Lebrun
and Agrawala, 2006; Perkins et al., 2007; Füssel, 2008; Smith et al., 2009;
Ford et al., 2011; Preston et al., 2011b). This challenges the selection of
appropriate metrics for the monitoring and evaluation of adaptation
and its contribution to vulnerability reduction (Burton and May, 2004;
Gagnon-Lebrun and Agrawala, 2007; McKenzie Hedger et al., 2008; Ford
et al., 2011).
One of the central unresolved tensions in progressing evaluation is the
relative merit of comparatively easy and objective targeting of the
completion of the processes and outputs needed to implement an
adaptation program versus the outcomes, such as changes in livelihoods
or reduction in risks. Assessment of outcomes is less objective, subject
to whether appropriate circumstances occur (e.g., that floods occur so
that risk reduction can be demonstrated) and usually take much longer
to establish. Preston et al. (2011b) suggest the evaluation of adaptation
processes may be a more robust approach to evaluation, owing to the
difficulties in attributing future outcomes to adaptation strategies and
the long-time lags that may be needed to assess the performance of a
particular strategy (Berkhout, 2005; Dovers and Hezri, 2010; Ford et al.,
2011). The OECD analyzed the monitoring and evaluation processes across
106 adaptation projects across six development agencies and found
that Results Based Management and Logical Framework approaches
dominated, as they do in normal development projects (Lamhauge et
al., 2011). They also drew attention to the need for appropriate baselines
and complementary sets of indicators that track not just process and
implementation, but also the extent to which targeted changes are
occurring. Monitoring programs themselves will need careful design to
ensure that they remain in place over the long time frames needed
for the outcomes to be identified; that they contain incentives for
beneficiaries to comply with conditions; and that compliance itself does
not impose undue burdens.
A number of national and international organizations have guides to
monitoring and evaluating adaptation activities (McKenzie Hedger et
al., 2008; UNDP, 2008; WRI, 2009; World Bank, 2010; GIZ, 2011). These
guides tend to focus on the wider framework of identifying and
managing adaptation-related activities and within that the criteria for
the selection of metrics for monitoring and evaluating those activities.
These issues are dealt with in Chapters 15 and 16.
14.5.3. Validation of Metrics
The practice of developing and applying metrics in adaptation has been
subject to much scrutiny. Eakin and Luers (2006) express serious concerns
about national-scale vulnerability assessments ranging from the quality
of the available data, the selection and creation of indicators, the
assumptions used in weighting of variables, and the mathematics of
aggregation. Nevertheless metrics will continue to be used and the
challenge is to identify and maintain basic standards of best practice.
857
Adaptation Needs and Options Chapter 14
14
O
ne of the most comprehensive attempts to validate a system for
measuring important components of adaptation is that of Brooks et al.
(2005). They used the probability of national climate-related mortality
from the CRED database of climate-related disasters
7
as a proxy for risk
and selected a set of 46 social, governance, economic, and biophysical
measures as indicators of social vulnerability. They found that 11 were
effective indicators of mortality rates and these were confirmed as useful
by a small focus group of seven adaptation experts. These experts also
ranked the variables in terms of their perception of their usefulness
leading to a total of 12 different rankings to which was added an equal
ranked set to give 13 measures of vulnerability. Countries were then
scored against these 13 rankings, and the number of times a country
appeared in the top quintile of countries in a particular ranking was
used as an indicator of its overall vulnerability.
Progress continues to be made in the methodologies of deriving
vulnerability metrics. For example, Rygel et al. (2006) have demonstrated
the value of using a Pareto method for combining scores from a
collection of indices without having to apply either implicit or explicit
weightings. Alcamo et al. (2008) sought to increase the consistency of
incorporating expert opinion on different disciplinary approaches
(sociology, environmental psychology, economics, and political science) to
the estimate of vulnerability, in this case of drought events, in three regions.
Based on inference models about what contributes to vulnerability to
drought and using fuzzy set theory (Eierdanz et al., 2008) to compute
susceptibilities, they were able to show that high combined susceptibilities
were associated with water stress crises.
Perch-Nielsen (2010) developed an index to estimate the vulnerability of
beach tourism using a systematic approach by establishing a framework
to identify the types of indicators needed and a systematic approach to
identify indicators that covered the range of countries and time scales.
The derivation of the index from the separate indicators was also
subjected to robustness (sensitivity) testing to determine the most
appropriate methods of scaling and combining the measures.
14.5.4. Assessment of Existing and Proposed Metrics
for Adaptation
Srinivasan and Prabhakar (2009) conducted a wide-ranging stakeholder
survey to assess the attitudes to, and requirements for, indicators of
adaptation. Stakeholders agreed that no single metric can capture the
multiple dimensions of adaptation and that refinements of methodologies
(e.g., rationale for index selection, aggregation methods, and data
checking) are badly needed. Preston et al. (2009) have suggested that,
rather than seeking particular metrics, researchers should focus on
developing rigorous processes for selecting metrics that can be applied
in a range of contexts. But metrics for adaptation remain a necessity.
Their derivation challenges the adaptation community to clarify its goals,
conceptual models, definitions, and applications. But both theory and
practice have shown indices alone are not sufficient to guide decisions
on which adaptation actions to take, on how to modify sustainable
d
evelopment activities, or on resource allocation. Downing (2003) noted
that the climate change community was far from adopting common
standards, paradigms, or analytic language. This still appears to be true,
making the search for commonly accepted metrics, even within well-
specified contexts, a challenging task.
14.6. Addressing Maladaptation
The adaptation literature is replete with advice to avoid maladaptation,
but it is less clear precisely what is included as “maladaptation. In a
general sense maladaptation refers to actions, or inaction that may lead
to increased risk of adverse climate-related outcomes, increased
vulnerability to climate change, or diminished welfare, now or in the
future (see Glossary). For example, the construction of well-engineered
climate-resilient roads designed to withstand current and future climate
extremes may foster new settlement into areas highly exposed to the
impacts of future climates; or increased water harvesting upstream to
cope with erratic rainfall may harm and reduce the opportunities for
communities downstream to manage their own risks. Actions that are
potentially maladaptive need not be inadvertent (as in the IPCC AR3
and AR4 definition), nor be taken ostensibly to avoid or reduce
vulnerability to climate change” (Barnett and O’Neill, 2010) as the
actions may be assessed as appropriate in the context of the full range
of climate and non-climate considerations and pressures that apply to
the decision. There should be clarity as to what is maladaptive action,
or lack of action, lest the avoidance of potential maladaptation becomes
a barrier to effective implementation of adaptation. In the road example
above, the immediate and multiple benefits to the community of a reliable
road system (including as evacuation route in floods, etc.) might be
judged as outweighing the longer term risk of inappropriate settlement
patterns (Lamhauge et al., 2011). This may be seen as an example of
an “unavoidable” ex post maladaptation (see Section 16.3.6.1) as it is
an appropriate decision based on the information and circumstances
at the time. The true maladaptation in this case would be the failure to
implement appropriate incentives or regulations to avoid vulnerable
settlements in the highly exposed areas.
The wide range of actions and circumstances that have been described
as maladaptive demonstrates the complexity of the concept and
terminology. Thomsen et al. (2012) describe actions that are not
respectful of the intrinsic integrity and internal self-regulation of social-
ecological systems as manipulative and likely to prove maladaptive.
Their example is the management of Noosa Beach in northern Australia,
where the coastline is characterized by cycles of erosion and depletion
of beach sands. Rather than enhance the self-regulatory processes and
adapting by managed retreat and expansion according to the cycle,
management has sought to maintain a static beach profile through hard
constructions and beach nourishment. Niemeyer et al. (2005) also
describe the state of individual beliefs about climate change that might
change from adaptive to inaction and possibly maladaptive behaviors
as the perceived magnitude of climate change increases, while Eriksen
et al. (2011) and Brown (2011) discuss avoiding outcomes that are
7
CRED, the Centre for Research on the Epidemiology of Disasters, has maintained a database of disasters, including those that are climate related. Rationale, methodologies, and
data are available at http://www.cred.be/.
858
Chapter 14 Adaptation Needs and Options
14
e
ssentially maladaptive as they run counter to sustainable development
goals.
14.6.1. Causes of Maladaptation
Maladaptation arises in many forms but several broad causes can be
identified. Actions that may benefit a particular group, or sector, at a
particular time may prove to be maladaptive to those same groups or
sectors in future climates or to other groups or sectors in existing
climates. For example, some development policies and measures that
deliver short-term benefits or economic gains but lead to greater
vulnerability in the medium to long term, such as in cases where the
construction of hardinfrastructure reduces the flexibility and the
range of future adaptation options (Adger et al., 2003b; Eriksen and
Kelly, 2007; OECD, 2009), or the failure to encompass the full range of
risks in the design of new structures, such as the effects of increasing
storm surge in the design of a coastal defense system (UNFCCC, 2007).
Adaptation efforts aimed at armoring the coastline may result in coastal
erosion elsewhere while building levees along a flood-prone area provides
protection to coastal population and infrastructure but might encourage
unwanted development within that area, often accentuated by an
exaggerated sense of safety (Grothmann and Patt, 2005; Repetto, 2008;
National Research Council, 2011) and the levees may increase damage
when they fail, as in Bangladesh in 1999 and New Orleans in 2003 (Huq
and Khan, 2006; Masozera et al., 2007; Pouliotte et al., 2009). Similarly,
agricultural policies that promote the growing of high-yielding crop
varieties through subsidies with the objective of boosting production
and increasing revenues may achieve these objectives in the short term,
but will also reduce agro-biodiversity and increase exposure and
vulnerability of mono-crops to climate change and finally undermine
the adaptive capacity of farmers in the long term (World Bank, 2010).
Another cause of maladaptation is the failure to account for multiple
interactions and feedbacks between systems and sectors leading to
inadequate or inaccurate information for developing adaptive responses
a
nd strategies that are maladaptive (Scheraga et al., 2003; Satterthwaite
et al., 2009; Pittock, 2011). An assessment of the downstream impacts
of upstream rainwater harvesting in a semiarid basin in southern India
showed that, once the full range of externalities were accounted for,
the net benefits were insufficient to pay back investment costs (Bouma
et al., 2011). Similarly, the conversion of coastal mangroves into shrimp
farms may lead to increased economic productivity and improved
livelihoods, but could also lead to increased vulnerability to flooding
and storm surges (Klein, 2010). Maladaptation may also occur if the
true potential of an option or a technology is unduly over-emphasized,
making it over-rated. Floating gardening has been suggested as an
example in this connection (Irfanullah, 2009, 2013). Further examples
of the range of maladaptive actions across a range of sectors and
regions in this report are outlined in Table 14-4.
14.6.2. Screening for Maladaptation
Five dimensions of maladaptation were identified by Barnett and O’Neill
(2010), including actions that, relative to alternatives: (1) increase
emissions of GHGs, (2) disproportionately burden the most vulnerable,
(3) have high opportunity costs, (4) reduce incentives and capacity to
adapt, and (5) set paths that limit future choices. These dimensions are
useful pointers to the potential for maladaptation but their application
depends on subjective assessments. The first suggests that any action
that increases GHG emissions is maladaptive, whereas a judgment on
the relative benefits and dis-benefits will need to be made in such cases;
the second turns on the interpretation of “disproportionately;” and the
third on “high” and on how opportunity costs are compared with current
benefits. The dimensions were used by Barnett and O’Neill (2010) to
describe maladaptive potential of the Wonthaggi desalinization plant
to improve water supply to Melbourne, Australia. The plant was included
as part of a wider water management plan for Melbourne that includes
both demand- and supply-side management and incentives (Heffernan,
2012; Porter, 2013). Barnett and O’Neill (2010) argue that the plant
(1) will increase GHG emissions (even if the planned wind power
859
Adaptation Needs and Options Chapter 14
14
e
nergy source is completed); (2) may lead to higher water costs that
will disproportionally affect the poorer households; (3) may divert
money and attention from more cost-effective recycling and rainwater
harvesting; (4) may reduce incentives to adapt through water conservation
approaches; and (5) as a large sunk cost has locked out other options.
The plant also affected significant cultural sites of the Bunurong
Aboriginal community (Lee and Chung, 2007).
14.6.3. Experiences with Maladaptation
Maladaptation is a cause of increasing concern to adaptation planners,
where intervention in one sector could increase vulnerability of another
sector or increase the vulnerability of a group to future climate change.
An example is the situation experienced by subsistence and smallholder
agriculturalists in Palca, Bolivia, who in the face of stressors relating to
land access, small holdings, and so forth moved away from their long
established practices of diversification of crop varieties and planting
locations to more intensive farming practices and cash cropping. They
are now seeing evidence of climate change, and the new practices make
them more vulnerable to these changes, leading to a risk of insufficient
adaptation and maladaptation (McDowell and Hess, 2012). But there
can also be tensions between development goals and climate change
goals, where people may be aware of a climate related risk but are
willing to take that risk (or they may have limited choice) given their
current circumstances (IPCC, 2012, Section 4.2.2).
Some studies warn against the simplistic use of maladaptation to
communicate the state of high exposure to risks resulting from certain
type of livelihoods. For example, the periodic movement of the nomadic
pastoralists following the grass and water is a traditional and effective
way of dealing with climate variability (Agrawal and Perrin, 2008), but
is increasingly being described by some as maladaptive. More focused
studies such as Young et al. (2009) put the breakdown of traditional
pastoralism in the Sudan into the wider social and political context that
led to restrictions on movement, asset stripping, and escalating violence
and was undermined by policies not conducive to mobility.
14.7. Research and Data Gaps
A long list of research questions could be identified and prioritized to
address gaps and assist the practice of adaptation, and many of these
are found in the subsequent adaptation chapters. In this chapter research
priorities would range from metrics for adaptation to the psychology of
communication about livelihood and life-threatening events. But, the
preparation of this report has shown that the practice of adaptation
has outstripped the rate at which relevant peer-reviewed research can
be produced and disseminated.
Many dedicated researches have become engaged in smaller, often
community-based or urban activities where results can be gathered in
relatively short time frames and direct interactions between the
researchers and the implementers are common. Here research and
action can, and are, serving each other and these interactions can be
encouraged with support for further cross-community, cross-cultural,
and cross-sectoral comparisons.
E
ffective and timely interaction is more difficult at larger scales. National
or multinational programs are often longer and complex and it is
difficult to identify the “adaptation” effort within a wider set of policy
objectives. Research inputs into decision making too often centers only
on better projections of future conditions or post hoc assessments of
completed projects. The task is made more difficult by relatively short-
term research grants, often starting late in the process, or after the
process is finished, and by the often rapid turnover of planning and
implementation staff, making a close working relationship difficult. But
there are models that work. Models based on established and ongoing
research teams with a close link to policy such as the EC programs and
its formation of targeted research teams across the European Union,
CSIR in South Africa, Commonwealth Scientific and Industrial Research
Organisation (CSIRO) and the National Climate Change Research Facility
(NCCARF) in Australia, the Corps of Engineers and the Regional Integrated
Sciences and Assessments (RISAs) in the USA, and UKCIP and its
successors in the UK do appear to be more effective in maintaining a
dialog with those “on the ground,” and this shows in the number of
well designed, insightful, and reviewed documents arising from these
collaborations.
Unfortunately this model has not been replicated at scale in most
developing countries. One might ask why is there only one reference in
this volume to any lesson learned from the PPCR a billion dollar
program set up to better understand the challenges of integrating, or
mainstreaming, adaptation into development planning with on the
ground implementation of many larger than normal adaptation activities?
The planning for the PPCR started only in 2008, but the planning process
itself is of research relevance, and over the past 2 to 3 years 18 countries
have been working through how to bring adaptation into their national
planning programs; it is surely a core research interest and opportunity
and one whose lifespan already exceeds that of many research projects.
Similarly the Adaptation Fund is mentioned only descriptively in these
chapters. So where were the groups of independent researchers observing
from their point of view, comparing and contrasting countries, and
simply conducting the process of independent and collaborative
research? The benefit would flow not just from the research itself but
also from the interactions with those charged with implementing
adaptation and from the challenge to interpret that research so that its
implications are relevant to the users, be they government officials or
smallholder farmers.
There are models in developing countries. The Consultative Group on
International Agricultural Research (CGIAR) network is already making
contributions, albeit in the broad domain of agriculture which may be
another model. The Coordinated Regional Climate Downscaling
Experiment (CORDEX) project will make high-quality high-resolution
climate projections available to all countries. The NEPAD Framework for
African Agricultural Productivity is another, and there are numerous
smaller and effective research efforts too numerous to list here, but few
can claim even regional coverage. The Cancun Agreement has already
raised the prospect of establishing in a developing country “an
international centre to enhance adaptation research and coordination.
This may provide the vehicle to tackle some of the problems described
above. The UN Agencies, the MDBs, and many bilateral agencies, which
are heavy users and sometimes producers of “research, could be major
beneficiaries and supporters.
860
Chapter 14 Adaptation Needs and Options
14
T
wo points in a review of a decade of experience in the RISA process in
the USA stand out. One was an insistence that research team members
should primarily be residents in their region of study, and to paraphrase
another insight, “knowing what one ought to do is not the same as
knowing how to do it” (Pulwarty et al., 2009). In arguing for the
establishment of the skills to establish an Australian film industry,
Phillip Adams advised the Prime Minister,
8
“It’s time to see our own
landscapes, hear our own voices and dream our own dreams. Those
words could just as well apply to tackling adaptation.
References
Aakre, S. and D.T.G. Rübbelke, 2010: Objectives of public economic policy and
the adaptation to climate change. Journal of Environmental Planning and
Management, 53, 767-791.
Adaptation Fund Board, 2013: Operational Policies and Guidelines for Parties to
Access Resources from the Adaptation Fund. The Adaptation Fund, Washington,
DC, USA, 52 pp.
Adger, W.N., 2001: Scales of governance and environmental justice for adaptation
and mitigation of climate change. Journal of International Development, 13(7),
921-931.
Adger, W.N., 2003: Social capital, collective action, and adaptation to climate change.
Economic Geography, 79, 387-404.
Adger, W.N., 2006: Vulnerability. Global Environmental Change, 16, 268-281.
Adger, W.N. and J. Barnett, 2009: Four reasons for concern about adaptation to
climate change. Environment and Planning A, 41(12), 2800-2805.
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, G. Bentham, M. Agnew, and S. Eriksen, 2003a: New Indicators
of Vulnerability and Adaptive Capacity. Technical Report No. 7, Tyndall Centre
for Climate Research, School of Environmental Sciences, University of East
Anglia, Norwich, UK, 128 pp.
Adger, W.N., S. Huq, K. Brown, D. Conway, and M. Hulme, 2003b: Adaptation to
climate change in the developing world. Progress in Development Studies, 3(3),
179-195.
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.
Adger, W.N., S. Agrawala, M. 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 and New York, NY, USA, 717-743.
Adger, W.N., S. Dessai, M. Goulden, M. Hulme, I. Lorenzoni, D. Nelson, L.-O. Naess, J.
Wolf, and A. Wreford, 2009a: Are there social limits to adaptation to climate
change? Climatic Change, 93(3-4), 335-354.
Adger, W.N., I. Lorenzoni, and K. O’Brien (eds.), 2009b: Adapting to Climate Change:
Thresholds, Values, Governance. Cambridge University Press, Cambridge, UK
and New York, NY, USA, 514 pp.
Adler, C.E., D. McEvoy, P. Chhetri, and E. Kruk, 2013: The role of tourism in a changing
climate for conservation and development. A problem-oriented study in the
Kailash Sacred Landscape, Nepal. Policy Sciences, 46, 160-178.
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 Publications, Washington, DC,
USA, pp. 173-197.
Agrawal, A. and N. Perrin, 2008: Climate Adaptation, Local Institutions, and Rural
Livelihood. IFRI Working Paper W08I-6, International Forestry Resources and
Institutions Program (IFRI), School of Natural Resources and Environment,
University of Michigan, Ann Arbor, MI, USA, 17 pp.
Agrawala, S., 2005: Putting climate change in the development mainstream:
introduction and framework. In: Bridge Over Troubled Waters: Linking Climate
Change and Development [Agrawala, S. (ed.)]. Organization for Economic Co-
operation and Development (OECD), OECD Publishing, Paris, France, pp. 23-43.
Agrawala, S. and M. Carraro, 2010: Assessing the Role of Microfinance in Fostering
Adaptation to Climate Change. ENV/WKP(2010)1, OECD Environment Working
Paper No. 15, Organization for Economic Co-operation and Development
(OECD), OECD Publishing, Paris, France, 37 pp.
Agrawala, S. and M. van Aalst, 2008: Adapting development cooperation to adapt
to climate change. Climate Policy, 8(2), 183-193.
Agrawala, S., M. Carraro, N. Kingsmill, E. Lanzi, M. Mullan, and G. Prudent-Richard,
2011: Private Sector Engagement in Adaptation to Climate Change: Approaches
to Managing Climate Risks. OECD Environment Working Paper No. 39,
Organization for Economic Co-operation and Development (OECD), OECD
Publishing, Paris, France, 55 pp.
Alam, K., M. Shamsuddoha, T. Tanner, M. Sultana, M.J. Huq, and S.S. Kabir, 2011: The
political economy of climate resilient development planning in Bangladesh. IDS
Bulletin, 42(3), 52-61.
Alderman, H., H. Hoogeveen, and M. Rossi, 2009: Preschool nutrition and subsequent
schooling attainment: longitudinal evidence from Tanzania. Economic
Development and Cultural Change, 57, 239-260.
Alcamo, J., L. Acosta-Michlik, A. Carius, D. Krömker, F. Eierdanz, R. Klein, D., Kromker, and
D. Tänzler, 2008: A new approach to quantifying and comparing vulnerability
to drought. Regional Environmental Change, 8, 137-149.
Alexander, C., N. Bynum, E. Johnson, U. King, T. Mustonen, P. Neofotis, N. Oettle, C.
Rosenzweig, C. Sakakibara, V. Shadrin, M. Vicarelli, J. Waterhouse, and B. Weeks,
2011: Linking indigenous and scientific knowledge of climate change.
BioScience, 61(6), 477-484.
Alexander, N., 2010: The Country Policy and Institutional Assessment (CPIA) and
Allocation of IDA Resources: Suggestions for Improvements to Benefit African
Countries. Economic Governance Program, Heinrich Bell Foundation, Washington,
DC, USA, 43 pp.
Allison, E.H., A.L. Perry, M.-C. Badjeck, W.N. Adger, K. Brown, D. Conway, A.S. Halls,
G.M. Pilling, J.D. Reynolds, N.L. Andrew, and N.K. Dulvy, 2009: Vulnerability of
national economies to the impacts of climate change on fisheries. Fish and
Fisheries, 10, 173-196.
Anguelovski, I. and J. Carmin, 2011: Climate action planning in Quito, Ecuador. In:
Climate Change and Cities: First Assessment Report of the Urban Climate
Change Research Network [Rosenzweig, C., W.D. Solecki, S.A. Hammer, and S.
Mehotra (eds.)]. Cambridge University Press, Cambridge, UK and New York, NY,
USA, 262 pp.
Anon, 2009: Building Coast Smart Communities: HowWill MarylandAdapt to
ClimateChange? Consensus Building Institute and MIT-USGS Science Impact
Collaborative, Cambridge, MA, USA and the Maryland Department of Natural
Resources, Annapolis, MD, USA, 21 pp.
Arnell, N.W., 2009: Adapting to climate change: an evolving research programme.
Climatic Change, 100(1), 107-111.
Arnold, G., H. Bos, R. Doef, R. Goud, N. Kielen, and F. van Lujin, 2011: Water
Management in the Netherlands. This report is a joint publication of the Ministry
of Infrastructure and Environment, Directorate-General Water and Rijkswaterstaat,
Centre for Water Management, The Hague, Netherlands, 80 pp.
Atteridge, A., 2011: Will Private Finance Support Climate Change Adaptation in
Developing Countries? Historical Investment Patterns as a Window on Future
Private Climate Finance. SEI Working Paper No. 2011-5, Stockholm Environment
Institute (SEI), Stockholm, Sweden, 35 pp.
Ayers, J., 2011: Resolving the adaptation paradox: exploring the potential for
deliberative adaptation policy-making in Bangladesh. Global Environmental
Politics, 11(1), 62-88.
Ayers, J. and D. Dodman, 2010: Climate change adaptation and development I: the
state of the debate. Progress in Development Studies, 10(2), 161-168.
Bambrick, H.J., A.G. Capon, G.B. Barnett, R.M. Beaty, and A.J. Burton, 2011: Climate
change and health in the urban environment: adaptation opportunities in
Australian cities. Asia Pacific Journal of Public Health, 23(2), 67-79.
Banerjee, S., J.-Y. Gerlitz, and D. Kniveton, 2013: A methodology for assessing
patterns of labour migration in mountain communities exposed to water hazards.
In: Disentangling Migration and Climate Change: Methodologies, Political
8
http://www.abc.net.au/dimensions/dimensions_in_time/Transcripts/s796788.htm, accessed 3rd Oct 2013.
861
Adaptation Needs and Options Chapter 14
14
Discourses and Human Rights [Faist, T. and J. Schade (eds.)]. Springer
International, Heidelberg, Germany and New York, NY, USA, pp. 81-100.
Bannayan, M. and G. Hoogenboom, 2008: Weather analogue: a tool for real-time
prediction of weather data realizations based on a modified k-nearest neighbor
approach. Environmental Modeling and Software, 23(6), 703-713.
Barnett, J. and S. O’Neill, 2010: Maladaptation. Global Environmental Change, 20(2),
211-213.
Barron, S., G. Canete, J. Carmichael, D. Flanders, E. Pond, S. Sheppard, and K. Tatebe,
2012: A climate change adaptation planning process for low-lying, communities
vulnerable to sea level rise. Sustainability, 4(12), 2176-2208.
Bartlett, S., D. Dodman, J. Hardoy, D. Satterthwaite, and C. Tacoli, 2009: Social Aspects
of Climate Change in Urban Areas in Low- and Middle-Income Nations. Paper
presented at the World Bank Fifth Urban Research Symposium, “Cities and Climate
Change: Responding to an Urgent Agenda,” 28-30 June 2009, Marseilles,
France, Research Cluster 5, 44 pp.
Berkes, F., 2007: Understanding uncertainty and reducing vulnerability: lessons from
resilience thinking. Natural Hazards, 41(2), 283-295.
Berkhout, F., 2005: Rationales for adaptation in EU climate change policies. Climate
Policy, 5(3), 377-391.
Berkhout, F., J. Hertin, and D.M. Gann, 2006: Learning to adaptation: organizational
adaptation to climate change impacts. Climatic Change, 78(1), 135-156.
Bhutta, Z.A., T. Ahmed, R.E. Black, S. Cousens, K. Dewey, E. Giugliani, B.A. Haider, B.
Kirkwood, S.S. Morris, H.P.S. Sachdev, and M. Shekar, 2008: What works?
Interventions for maternal and child undernutrition and survival. Lancet,
371(9610), 417-440.
Biderman, C., M. Smolka, and A. Sant’Anna, 2008: Urban Housing Informality: Does
Building and Land Use Regulation Matter? Lincoln Institute of Land Policy, Land
Lines, July 2008, Cambridge, MA, USA, 6 pp.
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.
Blanco, H., M. Alberti, A. Forsyth, K.J. Krizek, D.A. Rodríguez, E. Talen, and C. Ellis,
2009: Hot, congested, crowded and diverse: emerging research agendas in
planning. Progress in Planning, 71(4), 153-205.
Bolte, A., C. Ammer, M. Löf, P. Madsen, G. Nabuurs, P. Schall, P. Spathelf, and J. Rock,
2009: Adaptive forest management in central Europe: climate change impacts,
strategies and integrative concept. Scandinavian Journal of Forest Research,
24(6), 473-482.
Bosello, F., C. Carraro, and E. De Cian, 2009: An Analysis of Adaptation as a Response
to Climate Change. Report prepared for the Copenhagen Consensus 2009,
Copenhagen Consensus Center, Copenhagen Business School, Frederiksberg,
Denmark, 76 pp.
Bouma, J.A., W. Trent, and M. Laurens, 2011: The downstream externalities of
harvesting rainwater in semi-arid watersheds: an Indian case study. Agricultural
Water Management, 98(7), 1162-1170.
Boyd, E. and H. Osbahr, 2010: Responses to climate change: exploring organisational
learning across internationally networked organisations for development.
Environmental Education Research, 16(5-6), 629-643.
Brooks, N. and W.N. Adger, 2003: Country Level Risk Measures of Climate-related
Natural Disasters and Implications for Adaptation to Climate Change. Tyndall
Centre for Climate Research Working Paper No. 26, University of East Anglia,
Norwich, UK, 25 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 Part A, 15(2), 151-163.
Brooks, N., S. Anderson, J. Ayers, I. Burton, and I. Tellam, 2011:Tracking Adaptation
and Measuring Development.IIED Working Paper No. 1, Climate Change Group,
International Institute for Environment and Development (IIED), London, UK,
35 pp.
Brown, A., M. Gawith, K. Lonsdale, and P. Pringle, 2011: Managing Adaptation:
Linking Theory and Practice. UK Climate Impacts Programme (UKCIP), School
of Geography and the Environment, Oxford University Centre for the Environment
(OUCE), Oxford, UK, 47 pp.
Brown, C., Y. Ghile, M. Laverty, and K. Li, 2012: Decision scaling: linking bottom-up
vulnerability analysis with climate projections in the water sector. Water
Resources Research, 48(9), W09537, doi:10.1029/2011WR011212.
Brown, K., 2011: Sustainable adaptation: an oxymoron? Climate and Development,
3(1), 21-31.
Bryan, E., T.T. Deressa, G.A. Gbetibouo, and C. Ringler, 2009: Adaptation to climate
change in Ethiopia and South Africa: options and constraints. Environmental
Science and Policy, 12(4), 413-426.
Bulkeley, H. and M. Betsill, 2005: Rethinking sustainable cities: multilevel governance
and the ‘urban’ politics of climate change. Environmental Politics 14, 42-63.
Burton, I., 1996: The growth of adaptation capacity: practice and policy. In: Adapting
to Climate Change: An International Perspective [Smith, J., N. Bhatti, G.
Menzhulin, R. Benioff, M.I. Budyko, M. Campos, B. Jallow, and F. Rijsberman
(eds.)]. Springer-Verlag, Berlin Heidelberg, Germany, pp. 55-67.
Burton, I., 2004: Climate Change and the Adaptation Deficit. Occasional Paper 1,
The Adaptation and Impacts Research Group (AIRG), Meteorological Service
of Canada, Environment Canada, Toronto, ON, Canada, 6 pp.
Burton, I. and E. May, 2004: The adaptation deficit in water resources management.
IDS Bulletin, 35(3), 31-37.
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.
Burton, I., E. Diringer, and J. Smith, 2006: Adaptation to Climate Change: International
Policy Options. The Pew Center on Global Climate Change, Arlington, VA, USA,
28 pp.
Cardona, O.D., 2005. Indicators of Disaster Risk and Risk Management: Program for
Latin America and the Caribbean Summary Report. Environment Division,
Sustainable Development Department, Inter-American Development Bank,
Washington, DC, USA, 43 pp.
Carmin, J. and D. Dodman, 2013: Engaging science and managing scientific uncertainty
in urban climate adaptation planning. In: Successful Adaptation to Climate
Change: Linking Science and Policy in a Rapidly Changing World [Moser, S.C.
and M.T. Boykoff (eds.)]. Routledge, Abingdon UK and New York, NY, USA, pp.
220-234.
Carmin, J., I. Anguelovski, and D. Roberts, 2012: Urban climate adaptation in the
global south: planning in an emerging policy domain. Journal of Planning
Education and Research, 32(1), 18-32.
Carpenter, S.R., H.A. Mooney, J. Agard, D. Capistrano, R.S. DeFries, S. Diaz, T. Dietz,
A.K. Duraiappah, A. Oteng-Yeboah, H.M. Pereira, C. Perrings, W.V. Reid, J.
Sarukhan, R.J. Scholes, and A. Whyte, 2009: Science for managing ecosystem
services: beyond the millennium ecosystem assessment. Proceedings of the
National Academy of Sciences of the United States of America, 106(5), 1305-1312.
Carter, T.R., M.L. Parry, H. Harawasa, and S. Nishioka, 1994: IPCC Technical Guidelines
for Assessing Climate Change Impacts and Adaptations. CGER-I015-’94, Climate
Change Impacts and Adaptations, Intergovernmental Panel on Climate Change,
World Meteorological Organization (WMO)/United Nations Environment
Programme (UNEP), Published by the Department of Geography, University
College London, London, UK and the Center for Global Environmental Research,
National Institute for Environmental Studies, Tsukuba, Japan, 59 pp.
CCCCC, 2011: Developing an Implementation Plan for the CARICOM “Regional
Framework for Achieving Development Resilient to Climate Change”. Feedback
report from in-country dialogues, produced by Climate Risk Management, Ltd.,
(Acclimatise), for the Caribbean Community Climate Change Centre (CCCCC),
Acclimatise, Newark, UK, 138 pp.
CGIAR, 2012: Gender and Climate Change Research in Agriculture and Food Security
for Development. Training Guide, The CGIAR Program on Climate Change,
Agriculture and Food Security (CCAFS), Food and Agriculture Organization of
the United Nations (FAO), Rome, Italy, 127 pp.
Chinowsky, P., C. Hayles, A. Schweikert, N. Strzepek, K. Strzepek, and C.A. Schlosser,
2011: Climate change: comparative impact on developing and developed
countries. Engineering Project Organization Journal, 1(1), 67-80.
Christensen, K., S. Raihan, R. Ahsan, A.M.N. Uddin, C.S. Ahmed, and H. Wright, 2012:
Ensuring Access for the Climate Vulnerable in Bangladesh: Financing Local
Adaptation. ActionAid Bangladesh (AAB), Action Research for Community
Adaptation in Bangladesh (ARCAB), Bangladesh Centre for Advanced Studies
(BCAS), and International Centre for Climate Change and Development
(ICCCAD), AAB, Dhaka, Bangladesh, 126 pp.
Chuku, C.A., 2009: Pursuing an integrated development and climate policy framework
in Africa: options for mainstreaming. Mitigation and Adaptation Strategies for
Global Change, 15(1), 41-52.
Ciplet, D., S. Fields, K. Madden, M. Khan, and J.T. Roberts, 2012: The Eight Unmet
Promises of Fast Start Climate Finance. IIED Briefing, International Institute for
Environment and Development (IIED), London, UK, 8 pp.
862
Chapter 14 Adaptation Needs and Options
14
Ciscar, J.-C., A. Iglesias, L. Feyen, L. Szabó, D. Van Regemorter, B. Amelung, R. Nicholls,
P. Watkiss, O.B. Christensen, and R. Dankers, 2011: Physical and economic
consequences of climate change in Europe. Proceedings of the National
Academy of Sciences of the United States of America, 108, 2678-2683.
Climate Investment Funds, 2009: The Selection of Countries to Participate in the
Pilot Program for Climate Resilience (PPCR). Report of the Expert Group
(established by the Pilot Programme on Climate Resilience Subcommittee,
PPCR-SC), to the PPCR Subcommittee, Climate Investment Funds, The World
Bank Group, Washington, DC, USA, 61 pp.
Cline, W.R., 2007: Global Warming and Agriculture: Impact Estimates by Country.
The Center for Global Development (CGD) and the Peterson Institute for
International Economics, Washington, DC, USA, 186 pp.
Coles, A.R. and C.A. Scott, 2009: Vulnerability and adaptation to climate change and
variability in semi-arid rural southeastern Arizona, USA. Natural Resources
Forum, 33(4), 297-309.
Collins, K. and R. Ison, 2009: Jumping off Arnstein’s ladder: social learning as a new
policy paradigm for climate change adaptation. Environmental Policy and
Governance, 19(6), 358-373.
Colombo, A.F. and P.H. Byer, 2012: Adaptation, flexibility and project decision-making
with climate change uncertainties. Impact Assessment and Project Appraisal,
30(4), 229-241.
Compston, H., 2010: The politics of climate policy: strategic options for national
governments. Political Quarterly, 81(1), 107-115.
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.
Conway, G., 2009: The Science of Climate Change in Africa: Impacts and Adaptation.
Grantham Institute for Climate Change Discussion Paper No. 1, Imperial
College, London, UK, 24 pp.
Convention on Biological Diversity, 2009: Connecting Biodiversity and Climate
Change Mitigation and Adaptation: Report of the Second Ad Hoc Technical
Expert Group on Biodiversity and Climate Change. CBD Technical Series No. 41,
Secretariat of the CBD, Montreal, QC, Canada, 126 pp.
Corfee-Morlot, J., I. Cochran, S. Hallegatte, and P.-J. Teasdale, 2011: Multilevel risk
governance and urban adaptation policy. Climatic Change, 104(1), 169-197.
Costello, A., M. Abbas, A. Allen, S. Ball, S. Bell, R. Bellamy, S. Frield, 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.A. Puppim 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. Lancet, 373, 1693-1733.
Costello, A., M. Maslin, H. Montgomery, A.M. Johnson, and P. Ekins, 2011: Global
health and climate change: moving from denial and catastrophic fatalism to
positive action. Philosophical Transactions of the Royal Society A, 369(1942),
1866-1882.
Crawford, S.E.S. and E. Ostrom, 1995: A grammar of institutions. American Political
Science Review, 89, 582-600.
Cutter, S.L., C.T. Emrich, J.J. Web, and D. Morath, 2009: Social Vulnerability to Climate
Variability Hazards: A Review of the Literature. Final Report to Oxfam America,
Hazards Vulnerability Research Institute, University South Carolina, Columbia,
SC, USA, 44 pp.
Dawson, R., 2007: Re-engineering cities: a framework for adaptation to global
change. Philosophical Transactions of the Royal Society A, 365(1861), 3085-3098.
Day, J.W., D.F. Boesch, E.J. Clairain, G.P. Kemp, S.B. Laska, W.J. Mitsch, K. Orth, H.
Mashriqui, D.J. Reed, L. Shabman, C.A. Simenstad, B.J. Streever, R.R. Twilley,
C.C. Watson, J.T. Wells, and D.F. Whigham, 2007: Restoration of the Mississippi
Delta: lessons from Hurricanes Katrina and Rita. Science, 315(5819), 1679-1684.
De Bruin, K., R.B. Dellink, A. Ruijs, L. Boldwidt, 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, 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.
DEFRA, 2012: UK Climate Change Risk Assessment: Government Report. Presented to
Parliament pursuant to Section 56 of the Climate Change Act 2008, Department
for Environment, Food and Rural Affairs (DEFRA), The Stationery Office (TSO),
Norwich, UK, 43 pp.
Deressa, T., R. Hassan, and C. Ringler, 2009b: Assessing Household Vulnerability to
Climate Change: The Case of Farmers in the Nile Basin of Ethiopia. IFPRI
Discussion Paper 00935, International Food Policy Research Institute (IFPRI),
Washington, DC, USA, 18 pp.
Deressa, T.T., R.M. Hassan, C. Ringler, T. Alemu, and M. Yesuf, 2009a: Determinants
of farmers’ choice of adaptation methods to climate change in the Nile Basin
of Ethiopia. Global Environmental Change, 19(2), 248-255.
Dessai, S. and M. Hulme, 2004: Does climate adaptation policy need probabilities?
Climate Policy, 4(2), 107-128.
Dessai, S., X. Lu, and J.S. Risbey, 2005: On the role of climate scenarios for adaptation
planning. Global Environmental Change, 15(2), 87-97.
Dilley, M., R.S. Chen, U. Deichmann, A.L. Lerner-Lam, and M. Arnold, with J. Agwe, P.
Buys, O. Kjekstad, B. Lyon, and G. Yetman, 2005: Natural Disaster Hotspots: A
Global Risk Analysis. Disaster Risk Management Series No. 5, The International
Bank for Reconstruction and Development / The World Bank and Columbia
University, World Bank, Washington, DC, USA, 132 pp.
Dolan, A.H. and I.J. Walker, 2003: Understanding vulnerability of coastal communities
to climate change related risks. Journal of Coastal Research, 39(SI), 1317-1324.
Doria, M.D., E. Boyd, E.L. Tompkins, and W.N. Adger, 2009: Using expert elicitation
to define successful adaptation to climate change. Environmental Science and
Policy, 12(7), 810-819.
Dovers, S.R. and A.A. Hezri, 2010: Institutions and policy processes: the means to
the ends of adaptation. Climate Change, 1(2), 212-231.
Dow, K., F. Berkhout, B.L. Preston, R.J.T. Klein, G. Midgley, and M.R. Shaw, 2013:
Limits to adaptation. Nature Climate Change, 3, 305-307.
Dowlatabadi, H., 2007: On integration of policies for climate change global change.
Mitigation and Adaptation Strategies for Global Change, 12, 651-663.
Downing, T., 2003: Toward a vulnerability/adaptation science: lessons from famine
early warning and food security. In: Climate Change Adaptive Capacity and
Development [Smith, J., R. Klein, and S. Huq (eds.)]. Imperial College Press,
London, UK, pp. 77-100.
Downing, T.E., M.J. Watts, and H.G. Bohle, 1995: Climate change and food insecurity:
towards a sociology and geography of vulnerability. In: Climate Change and
World Food Security [Downing, T.E. (ed.)]. Springer, Berlin, Germany, pp. 183-206.
Dudley, N., S. Stolton, A. Belokurov, L. Krueger, N. Lopoukhine, K. MacKinnon, T. Sandwith,
and N. Sekhran (eds.), 2010: Natural Solutions: Protected Areas Helping People
Cope with Climate Change. The International Union for Conservation of Nature-
World Commission on Protected Areas (IUCN-WCPA), the Nature Conservancy
(TNC), the United Nations Development Programme (UNDP), the Wildlife Con-
servation Society (WCS), The World Bank, and the World Wildlife Fund (WWF),
IUCN-WCPA and WWF International, Gland, Switzerland, TNC, Arlington, VA,
USA, UNDP, New York, NY, USA, WCS, Bronx, NY, USA, and The World Bank,
Washington, DC, USA, 126 pp.
Eakin, H. and M.C. Lemos, 2006: Adaptation and the state: Latin America and the
challenge of capacity-building under globalization. Global Environmental
Change, 16(1), 7-18.
Eakin, H. and A. Luers, 2006: Assessing the vulnerability of social-environmental
systems. Annual Review of Environment and Resources, 31(1), 365-394.
Eakin, H., E.L. Tompkins, D.R. Nelson, and J.M. Anderies, 2012: Hidden costs and
disparate uncertainties: trade-offs in approaches to climate policy. In: Adapting
to Climate Change: Thresholds, Values, Governance [Adger, W.N., I. Lorenzoni,
and K.L. O’Brien (eds.)]. Cambridge University Press, Cambridge, UK and New
York, NY, USA, pp. 212-226.
Ebi, K., 2011: Climate change and health risks: assessing and responding to them
through “adaptive management”. Health Affairs, 30(5), 924-930.
Ebi, K.L. and I. Burton, 2008: Identifying practical adaptation options: an approach
to address climate change-related health risks. Environmental Science and
Policy, 11(4), 359-369.
ECA, 2009: Shaping Climate-Resilient Development: A Framework for Decision-
Making. A Report of the Economics of Climate Adaptation (ECA) Working
Group, a partnership of the Climate Works Foundation, Global Environment
Facility, European Commission, McKinsey and Company, The Rockefeller
Foundation, Standard Chartered Bank, and Swiss Re, 159 pp.
Edwards, F., J. Dixon, S. Friel, G. Hall, K. Larsen, S. Lockie, B. Wood, M. Lawrence, I.
Hanigan, A. Hogan, and L. Hattersley, 2011: Climate change adaptation at the
intersection of food and health. Asia Pacific Journal of Public Health, 23(2),
91-104.
Eierdanz, F., J. Alcamo, L. Acosta-Milchlik, D. Krömker, and D. Tänzler, 2008: Using
fuzzy set theory to address the uncertainty of susceptibility of drought. Regional
Environmental Change, 8(4), 197-205.
Engle, N.L. and M.C. Lemos, 2010: Unpacking governance: building adaptive capacity
to climate change of river basins in Brazil. Global Environmental Change, 20(1),
4-13.
863
Adaptation Needs and Options Chapter 14
14
Eriksen, S.H. and P.M. Kelly, 2007: Developing credible vulnerability indicators for
climate adaptation policy assessment. Mitigation and Adaptation Strategies
for Global Change, 12, 495-524.
Eriksen, S., P. Aldunce, C.S. Bahinipati, R. D’Almeida 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.
European Commission, 2009: Adapting to Climate Change: Towards a European
Framework for Action. White Paper, Brussels 1.4.2009 COM(2009) 147 final,
Publications Office of the European Union, Luxembourg, Luxembourg, 16 pp.
Fankhauser, S. and I. Burton, 2011: Spending adaptation money wisely. Climate
Policy, 11(3), 1037-1049.
FAO, 2007: Adaptation to Climate Change in Agriculture, Forestry and Fisheries:
Perspective, Framework and Priorities. Interdepartmental Working Group on
Climate Change, Food and Agriculture Organization of the United Nations
(FAO), Rome, Italy, 24 pp.
FEMA, 2010: Hazard Mitigation Assistance Unified Guidance. Federal Emergency
Management Agency (FEMA), Department of Homeland Security, FEMA,
Washington, DC, USA, 168 pp.
Ferrara de Giner, G., M.T. Martelo, R. Lairet, A. Villamizar, and J.C. Sánchez, 2012: A
proposal for community adaptive capacity on adaptation to climate change:
San Pedro community, Caracas, Venezuela. In: E-Annals from ABES-AIDIS. XXXIII
Congreso Interamericano de Engenharia Sanitária e Ambiental (Associação
Brasileira de Engenharia Sanitária e Ambiental), Theme VIII-024, pp. 1-7.
Finzi Hart, J.A., P.M. Grifman, S.C. Moser, A. Abeles, M.R. Myers, S.C. Schlosser, and J.A.
Ekstrom, 2012: Rising to the Challenge: Results of the 2011 Coastal California
Adaptation Needs Assessment. Technical Report: USCSG-TR-01-2012, University
of Southern California (USC) Sea Grant, USC, Los Angeles, CA, 76 pp.
Ford, J.D., L. Berrang-Ford, and J. Paterson, 2011: A systematic review of observed
climate change adaptation in developed nations. Climatic Change, 106(2), 327-
336.
Fünfgeld, H. and D. McEvoy, 2011: Framing Climate Change Adaptation in Policy
and Practice. VCCCAR Working Paper 1, VCCCAR Project: Framing Adaptation
in the Victorian Context, Victorian Centre for Climate Change Adaptation
Research (VCCCAR), Melbourne, Australia, 65 pp.
Furlow, J., J.B. Smith, G. Anderson, W. Breed, and J. Padgham, 2011: Building resilience
to climate change through development assistance: USAID’s Climate Adaptation
Program. Climatic Change, 108(3), 411-421.
Füssel, H.-M., 2007: Vulnerability: a generally applicable conceptual framework for
climate change research. Global Environmental Change, 17(2), 155-167.
Füssel, H.-M., 2009: An updated assessment of the risks from climate change based
on research published since the IPCC Fourth Assessment Report. Climatic
Change, 97(3-4), 469-482.
Füssel, H.-M., 2010: How inequitable is the global distribution of responsibility,
capability, and vulnerability to climate change: a comprehensive indicator-
based assessment. Global Environmental Change, 20(4), 597-611.
Füssel, H.-M. and R.J.T. Klein, 2006: Climate change vulnerability assessments: an
evolution of conceptual thinking. Climatic Change, 75(3), 301-329.
Gagnon-Lebrun, F. and S. Agrawala, 2006: Progress on Adaptation to Climate Change
in Developed Countries: An Analysis of Broad Trends. Organisation for Economic
Co-operation and Development (OECD), OECD Publishing, Paris, France, 61 pp.
Gardiner, D., M. Anderson, R. Schlesinger, J. Fox Gorte, and D. Zeller, 2007: Climate
Risk Disclosure by the S&P 500. David Gardiner & Associates report for Ceres
and Calvert analyzing the S&P 500 CDP4 responses to the Carbon Disclosure
Project information request regarding the risks and opportunities of climate
change, Ceres, Boston, MA, USA, 72 pp., www.calvert.com/documents/ceres_
calvert_sandp_500.pdf.
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, 43(5), 779-789.
GEF, 2002: Note on GEF Support for National Adaptation Plans of Action (NAPA).
GEF/C.19/Inf.7, May 8, 2002, Global Environment Facility (GEF), GEF Secretariat,
Washington, DC, USA, 28 pp.
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(7296), 342-345.
Ghimire, Y.N., G.P. Shivakoti, and S.R. Perret, 2010: Household-level vulnerability to
drought in hill agriculture of Nepal: implications for adaptation planning.
International Journal of Sustainable Development and World Ecology, 17(3),
225-230.
GIZ, 2011: Making Adaptation Count: Concepts and Options for Monitoring and
Evaluation of Climate Change Adaptation. Deutsche Gesellschaft für Internationale
Zusammenarbeit (GIZ) GmbH, on behalf of Bundesministerium für
wirtschaftliche Zusammenarbeit und Entwicklung (BMZ), and The World
Resources Institute (WRI), GIZ, Eschborn, Germany, 92 pp.
Glatzel, K., H. Wright, and Z. Makuch, 2012: Technology innovation and the law - the
example of climate adaptation technologies. In: Environmental and Energy Law
[Makuch, K.E. and R. Pereira (eds.)]. Wiley-Blackwell, Chichester, UK, pp. 92-116.
Goldman, R.L., H. Tallis, P. Kareiva, and G.C. Daily, 2008: Field evidence that ecosystem
service projects support biodiversity and diversify options. Proceedings of the
National Academy of Sciences of the United States of America, 105(27), 9445-
9448.
Goldstein, J.H., G. Caldarone, T.K. Duarte, D. Ennaanay, N. Hannahs, G. Mendoza, S.
Polasky, S. Wolny, and G.C. Daily, 2012: Integrating ecosystem-service tradeoffs
into land-use decisions. Proceedings of the National Academy of Sciences of
the United States of America, 109(19), 7565-7570.
Gonzalez, P., R.P. Neilson, J.M. Lenihan, and R.J. Drapek, 2010: Global patterns in
the vulnerability of ecosystems to vegetation shifts due to climate change.
Global Ecology and Biogeography, 19(6), 755-768.
Government of Nepal, 2011: National Framework on Local Adaptation Plan for
Action (LAPA), Kathmandu, Nepal, 56 pp.
Green Climate Fund, 2013a: Governing Instrument for the Green Climate Fund,
Green Climate Fund, Interim Secretariat of the Green Climate Fund, Bonn,
Germany, 16 pp., gcfund.net/fileadmin/00_customer/documents/pdf/GCF-
governing_instrument-120521-block-LY.pdf.
Green Climate Fund, 2013b: Business Model Framework: Objectives, Results and
Performance Indicators. GCF/B.04/03, Green Climate Fund, Meeting of the
Board, 26-28 June 2013, Songdo, Republic of Korea, 27 pp., www.gcfund.net/
fileadmin/00_customer/documents/pdf/B-04_03_BMF_Objectives_Results_
PerformanceIndicators_10Jun13.pdf.
Grothmann, T. and A. Patt, 2005: Adaptive capacity and human cognition: the process
of individual adaptation to climate change. Global Environmental Change,
15(3), 199-213.
GTZ, 2008: Adaptation and Refinement of the World Bank’s “Country Policy and
Institutional Assessment” (CPIA). Deutsche Gesellschaft für Technische
Zusammenarbeit (GTZ) GmbH, on behalf of Bundesministerium für
wirtschaftliche Zusammenarbeit und Entwicklung (BMZ), and the Global Public
Policy Institute (GPPi), GTZ, Eschborn, Germany, 90 pp.
Guariguata, M., 2009: El manejo forestal en el contexto de la adaptación al cambio
climático. Revista de Estudios Sociales, 32, 98-112.
Gupta, J., C. Termeer, J. Klostermann, S. Meijerink, M. van den Brink, P. Jong, S.
Nooteboom, and E. Bergsma, 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.
Haines, A., R.S. Kovats, D. Campbell-Lendrum, and C. Corvalan, 2006: Climate change
and human health: impacts, vulnerability, and mitigation. Lancet, 367(9528),
2101-2109.
Hallegatte, S., 2009: Strategies to adapt to an uncertain climate change. Global
Environmental Change, 19(2), 240-247.
Hallegatte, S., 2011: How economic growth and rational decisions can make disaster
losses grow faster than wealth. Policy Research Working Paper 5617, Office of
the Chief Economist, Sustainable Development Network, The International Bank
for Reconstruction and Development / The World Bank, Washington, DC, 21 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.
Hardoy, J. and P. Romero Lankao, 2011: Latin American cities and climate change:
challenges and options to mitigation and adaptation responses. Current Opinion
in Environmental Sustainability, 3(3), 158-163.
Harlan, S.L. and D.M. Ruddel, 2011: Climate change and health in cities: impacts of
heat and air pollution and potential co-benefits from mitigation and adaptation.
Current Opinion in Environmental Sustainability, 3(3), 126-134.
Harsdorff, M., M. Lieuw-Kie-Song, and M. Tsukamoto, 2011: Towards an ILO
Approach to Climate Change Adaptation. ILO Employment Working Paper No.
104, International Labour Organization (ILO), Geneva, Switzerland, 85 pp.
Heffernan, O., 2012: No going back. Nature, 491, 659-661.
Heltberg, R., P.B. Siegel, and S.L. Jorgensen, 2009: Addressing human vulnerability
to climate change: toward ano-regrets” approach. Global Environmental
Change, 19(1), 89-99.
864
Chapter 14 Adaptation Needs and Options
14
Hinkel, J., 2011: Indicators of vulnerability and adaptive capacity: towards a
clarification of the science-policy interface. Global Environmental Change,
21(1), 198-208.
Hochrainer, S., R. Mechler, and G. Pflug, 2007: Climate change and weather insurance
in Malawi: assessing the impact. In: Feasibility of Risk Financing Schemes for
Climate Adaptation: The Case of Malawi [Suarez, P., J. Linnerooth-Bayer, and
R. Mechler (eds.)]. International Institute for Applied Systems Analysis (IIASA)
Report to the World Bank Development Economics Research Group, IIASA,
Laxenburg, Austria, pp. 53-86.
Hoddinott, J., J.A. Maluccio, J.R. Behrman, R. Flores, and R. Martorell, 2008: Effect of
a nutrition intervention during early childhood on economic productivity in
Guatemalan adults. Lancet, 371(9610), 411-416.
Hoegh-Guldberg, O., 2011: Coral reef ecosystems and anthropogenic climate
change. Regional Environmental Change, 11(Suppl. 1), 215-227.
Hoeppe, P. and E.N. Gurenko, 2006: Scientific and economic rationales for innovative
climate insurance solutions. Climate Policy, 6(6), 607-620.
Huang, C., P. Vaneckova, X. Wang, G. Fitzgerald, Y. Gro, and S. Tong, 2011: Constraints
and barriers to public health adaptation to climate change: a review of the
literature. American Journal of Preventive Medicine, 40(2), 183-190.
Hulme, M., 2009: Why We Disagree about Climate Change: Understanding Controversy,
Inaction and Opportunity. Cambridge University Press, Cambridge, UK and New
York, NY, USA, 392 pp.
Huntjens, P., C. Pahl-Wostl, and J. Grin, 2010: Climate change adaptation in European
river basins. Regional Environmental Change, 10(4), 263-284.
Huq, S. and M.R. Khan, 2006: Equity in National Adaptation Programs of Action
(NAPAs): the case of Bangladesh.In: Fairness in Adaptation to Climate Change
[Adger, W.N., J. Paavola, S. Huq, and M.J. Mace (eds.)]. MIT Press, Cambridge,
MA, USA, pp. 131-153.
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.
Hussey, K. and S. Dovers, 2006: Trajectories in Australian water policy. Journal of
Contemporary Water Research and Education, 135(1), 36-60.
IEG, 2009: The World Bank’s Country Policy and Institutional Assessment: An
Evaluation. Independent Evaluation Group (IEG), The World Bank Group,
Washington, DC, USA, 142 pp.
IFC, 2010: Climate Risk and Financial Institutions: Challenges and Opportunities.
International Finance Corporation (IFC), Washington, DC, USA, 120 pp.
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 and New
York, NY, USA, 1032 pp.
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, Pachauri, R.K. and A. Reisinger
(eds.)]. IPCC, Geneva, Switzerland, 104 pp.
IPCC, 2007b: 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, USA, 976 pp.
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.
Irfanullah, H.M., 2009: Floating gardening in Bangladesh: already affected by climate
variability? In: Biodiversity Conservation and Response to Climate Variability
at Community Level [Irfanulla, H.M., R. Amin, and A. Nishat (eds.)]. International
Union for Conservation of Nature (IUCN), United Nations Environment
Programme (UNEP), and United Nations University (UNU), IUCN, Dhaka,
Bangladesh, pp. 7-14.
Irfanullah, H.M., 2013. Floating gardening: a local lad becoming a climate celebrity?
Clean Slate, 88, 26-27.
Irfanullah, H.M., A. Adrika, A. Ghani, Z.A. Khan, and M.A. Rashid, 2011a: Introduction
of floating gardening in the north-eastern wetlands of Bangladesh for nutritional
security and sustainable livelihoods. Renewable Agriculture and Flood Systems,
23(2), 89-96.
Irfranullah, H.M., M.A.K. Azad, A.K.M. Kamruzzaman, and M.A. Wahed, 2011b:
Floating gardening in Bangladesh: a means to rebuild lives after devastating
flood. Indian Journal of Traditional Knowledge, 10(1), 31-38.
Jentsch, A. and C. Beierkuhnlein, 2008: Research frontiers in climate change: effects
of extreme meteorological events on ecosystems. Comptes Rendus Geoscience,
340(9-10), 621-628.
Jones, H.P., D.G. Hole, and E.S. Zavaleta, 2012: Harnessing nature to help people
adapt to climate change. Nature Climate Change, 2, 504-509.
Kalame, F.B., D. Kudejira, and J. Nkem, 2011: Assessing the process and options for
implementing National Adaptation Programmes of Action (NAPA): a case study
from Burkina Faso. Mitigation and Adaptation Strategies for Global Change,
16(5), 535-553.
Kasperson, J.X. and R.E. Kasperson, 2001: Global Environmental Risk. United
National University Press, Tokyo, Japan and New York, NY, USA, 574 pp.
Kates, R.W. and T.J. Wilbanks, 2003: Making the global local: responding to climate
change concerns from the ground. Environment, 45(3), 12-23.
Kates, R.W., W.R. Travis, and T.J. Wilbanks, 2012: Transformational adaptation when
incremental adaptation climate change insufficient. Proceedings National
Academy Sciences of the United States of America, 109(19), 7156-7161.
Kaufman, R. and F.W. English, 1979: Needs Assessment: Concept and Application.
Educational Technology Publications, Englewood Cliffs, NJ, USA, 355 pp.
Khan, M.M.R., S.H. Miah, and H.M. Irfanullah, 2013: Small-scale silage-making
technology for the extreme poor on floodplains. International Journal of
Environmental Studies, 70(2), 192-202.
Khattri, A., D. Parameshwar, and S. Pellech, 2010: Opportunities for Private Sector
Engagement in Urban Climate Change Resilience Building. Report by Intellectual
Capital Advisory Services Pvt. Ltd. (Intellecap) with the support of the Asian
Cities Climate Change Resilience Network (ACCCRN) and the Rockefeller
Foundation, Intellecap, Hyderabad, India and the Rockefeller Foundation,
Bangkok, Thailand, 90 pp.
Kiparsky, M., A. Milman, and S. Vicuña, 2012: Climate and water: knowledge of
impacts to action on adaptation. Annual Review of Environment and Resources,
37(1), 163-194.
Klein, R.J.T., 2010: Mainstreaming climate adaptation in development: a policy
dilemma. In: Climate Governance and Development [Ansohn, A. and B. Pleskovic
(eds.)]. Berlin Workshop Series 2010, The International Bank for Reconstruction
and Development / The World Bank, Washington, DC, USA, pp. 35-52.
Klein, R.J.T. and A. Möhner, 2011: The political dimension of vulnerability: implications
for the Green Climate Fund. IDS Bulletin, 42(3), 15-22.
Klein, R.J.T., R.J. Nicholls, and F. Thomalla, 2003: Resilience to natural hazards: how
useful is this concept? Global Environmental Change Part B: Environmental
Hazards, 5(1-2), 35-45.
Klein, R.J.T., S. Huq, F. Denton, T.E. Downing, R.G. Richels, J.B. Robinson, and F.L. Toth,
2007: Inter-relationships between adaptation and mitigation. In: Climate
Change 2007: Impacts, Adaptation and Vulnerability. Contributions of Working
Group II to the Fourth Assessment Report of the Intergovernmental Panel on
Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikot, P.J. van der Linden, and
C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK and New York,
NY, USA, pp. 745-777.
Koetse, M.J. and P. Rietveld, 2012: Adaptation to climate change in the transport
sector. Transport Reviews, 32(3), 267-286.
Kovats, S. and R. Akhtar, 2008: Climate, climate change and human health in Asian
cities. Environment and Urbanization, 20(1), 165-175.
KPMG International, 2008: Climate Changes Your Business: KPMG’s Review of the
Business Risks and Economic Impacts at Sector Level. KPMG International,
Amstelveen, Netherlands, 76 pp.
Krasny, M.E., C. Lundholm, and R. Plummer, 2010: Resilience in social-ecological
systems: the roles of learning and education. Environmental Education
Research, 16(5-6), 463-474.
Kumamoto, M. and A. Mills, 2012: What African countries perceive to be adaptation
priorities: results from 20 countries in the Africa Adaptation Programme.
Climate and Development, 4(4), 265-274.
Lamhauge, N., E. Lanzi, and S. Agrawala, 2011: Monitoring and Evaluation for
Adaptation: Lessons from Development Co-operation Agencies. OECD
Environment Working Paper No. 38, Organization for Economic Co-operation
and Development (OECD) Publishing, Paris, France, 49 pp.
Laukkonen, J., P.K. Blanco, J. Lenhart, M. Keiner, B. Cavric, and C. Kinuthia-Njenga,
2009: Combining climate change adaptation and mitigation measures at the
local level. Habitat International, 33(3), 287-292.
865
Adaptation Needs and Options Chapter 14
14
LDC Expert Group, 2009: The Least Developed Countries National Adaptation
Programmes of Action: Overview of Preparation, Design of Implementation
Strategies and Submission of Revised Project Lists and Profiles. United Nations
Framework Convention on Climate Change (UNFCCC), UNFCCC Secretariat,
Bonn, Germany, 32 pp.
LDC Expert Group, 2012: National Adaptation Plans: Technical Guidelines for the
National Adaptation Plan Process. LDC Expert Group, United Nations Framework
Convention on Climate Change (UNFCCC), UNFCCC Secretariat, Bonn, Germany,
148 pp.
Learmonth, G., D.E. Smith, W.H. Sherman, M.A. White, and J. Plank, 2011: A practical
approach to the complex problem of environmental sustainability: the UVa Bay
Game.The Innovation Journal, 16(1), 2-8.
Lee, K.S. and E.S. Chung, 2007: Hydrological effects of climate change, groundwater
withdrawal, and land use in a small Korean watershed. Hydrological Processes,
21(22), 3046-3056.
Lempert, R.J. and M.E. Schlesinger, 2000: Robust strategies for abating climate
change. Climatic Change, 45(3-4), 387-401.
Levina, E., 2007: Adaptation to Climate Change: International Agreements for Local
Needs. COM/ENV/EPOC/IEA/SLT(2007)6, Organization for Economic Co-Operation
and Development (OECD) and the International Energy Agency (IEA), OECD
and IEA, Paris, France, 60 pp.
Lonergan, S., K. Gustavson, and M. Harrower, 1999: Mapping human insecurity. In:
Environmental Change, Adaptation, and Security [Lonergen, S.C. (ed.)].
NATO/Kluwer, Dordrecht, Netherlands, pp. 397-413.
MacLean, D., 2008: ICTs, Adaptation to Climate Change, and Sustainable Development
at the Edges. IISD Commentary, International Institute for Sustainable
Development (IISD), Winnipeg, MB, Canada, 5 pp.
Mallawaarachchi, T. and A. Foster, 2009: Dealing with Irrigation Drought: The Role
of Water Trading in Adapting to Water Shortages in 2007-08 in the Southern
Murray-Darling Basin. ABARE Research Report 09.6 to the Department of the
Environment, Water, Heritage, and the Arts, Australian Bureau of Agriculture
and Resource Economics (ABARE), Canberra, Australia, 32 pp.
Martens, P., D. McEvoy, and C. Chang, 2009: The climate change challenge: linking
vulnerability, adaptation, and mitigation. Current Opinion in Environmental
Sustainability. 1(1), 14-18.
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
Mawdsley, J.R., R. O’Malley, and D.S. Ojima, 2009: A review of climate-change
adaptation strategies for wildlife management and biodiversity conservation.
Conservation Biology, 23(5), 1080-1089.
McDowell, J.Z. and J.J. Hess, 2012: Accessing adaptation: multiple stressors on
livelihoods in the Bolivian Highlands under a changing climate. Global
Environmental Change, 22(2), 342-352.
McEvoy, D., S. Lindley, and J. Handley, 2006: Adaptation and mitigation in urban
areas: synergies and conflicts. Proceedings of the Institution of Civil Engineers:
Municipal Engineer, 159(4), 185-191.
McGray, H., R. Bradley, A. Hammill, E.L. Schipper, and J. Parry, 2007: Weathering the
Storm: Options for Framing Adaptation and Development. World Resources
Institute, Washington, DC, USA, 57 pp.
McKenzie Hedger, M., R. Connell, and P. Bramwell, 2006: Bridging the gap:
empowering adaptation decision-making through the UK Climate Impacts
Programme. Climate Policy, 6, 201-215.
McKenzie Hedger, M., T. Mitchell, J. Leavy, M. Greenly, A. Downie, and L. Horrocks,
2008: Desk Review: Evaluation of Adaptation to Climate Change from a
Development Perspective. Report commissioned by the GEF Evaluation Office
(EO) and financed by DFID, as part of the GEF Evaluation Office International
Conference on Evaluating Climate Change and Development (Alexandria, May
10th to 13th, 2008), Institute of Development Studies (IDS), London, UK, 60 pp.
McKillip, J., 1987: Need Analysis: Tools for the Human Services and Education. Sage
Publications, Thousand Oaks, CA, USA, 143 pp.
McMichael, A.J. and E. Lindgren, 2011: Climate change: present and future risks to
health, and necessary responses. Journal of Internal Medicine, 270(5), 401-
413.
Mechler, R., S. Hochrainer, G. Pflug, A. Lotsch, and K. Williges, 2010: Assessing the
Financial Vulnerability to Climate-Related Natural Hazards. World Bank Policy
Research Paper No. 5232, Background Paper to the 2010 World Development
Report, The International Bank for Reconstruction and Development / The World
Bank, Washington, DC, USA, 30 pp.
Mertz, O., K. Halsnaes, J.E. Olsen, and K. Rasmussen, 2009: Adaptation to climate
change in developing countries. Environmental Management, 43(5), 743-752.
Meze-Hausken, E., A. Patt, and S. Fritz, 2009: Reducing climate risk for micro-
insurance providers in Africa: a case study of Ethiopia. Global Environmental
Change, 19(1), 66-73.
Moench, M., E. Fajber, A. Dixit, E. Caspari, and A. Pokhrel, 2009: Catalyzing Climate
and Disaster Resilience. Processes for Identifying Tangible and Economically
Robust Strategies. Final Report of the Risk to Resilience Study, Institute for Social
and Environmental Transition (ISET), ISET-Nepal, Kathmandu, Nepal, 324 pp.
Moody, P. and C. Brown, 2012: Modeling stakeholder-defined climate risk on the
Upper Great Lakes. Water Resources Research, 48(10), doi:10.1029/
2012WR012497.
Mooney, H., A. Larigauderie, M. Cesario, T. Elmquist, O. Hoegh-Guldberg, S. Lavorel,
G. Mace, M. Palmer, R. Scholes, and T. Yahara, 2009: Biodiversity, climate change,
and ecosystem services. Current Opinion in Environmental Sustainability, 1, 46-
54.
Morecroft, M.D. and C. E. Cowan, 2010: Responding to climate change: an essential
component of sustainable development in the 21st century. Local Economy,
25(3), 170-175.
Morris, J., 2007: Ecological engineering in intertidial saltmarshes. Hydrobiologia,
192, 161-168.
Moser, S.C., 2006: Asset-Based Approaches to Poverty Reduction in a Globalized
Context: An Introduction to Asset Accumulation Policy and Summary of
Workshop Findings. Global Economy and Development Working Paper,
Brookings Institution, Washington, DC, USA, 39 pp.
Moser, S.C., 2009: Governance and the art of overcoming barriers to adaptation.
Magazine of the International Human Dimensions Programme on Global
Environmental Change, 3, 31-36.
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 D. Satterthwaite, 2010: Toward pro-poor adaptation to climate
change in the urban centers of low- and middle-income countries. In: Social
Dimensions of Climate Change: Equity and Vulnerability in a Warming World
[Mearns, R. and A. Norton (eds.)]. The International Bank for Reconstruction
and Development / The World Bank, Washington, DC, USA, pp. 231-258.
Moss, R.H., A.L. Brenkert, and E.L. Malone, 2001: Vulnerability to Climate Change: A
Quantitative Approach. Technical Report PNNL-SA-33642, Prepared by the
Pacific Northwest National Laboratory for the Department of Energy, Office of
Scientific and Technical Information, Oak Ridge, TN, USA and U.S. Department
of Commerce, Springfield, VA, USA, 70 pp.
Moss, R.H., E.L. Malone, N.L. Engle, A. de Bremond, and A. Delgado, 2012: Ready or
Not: Towards a Resilience Framework for Making Climate-Change Adaptation
Decisions. Report prepared for the World Bank by the Joint Global Change
Research Institute.
National Research Council, 2011: Adapting to the Impacts of Climate Change.
Americas Climate Choices: Panel on Adapting to the Impacts of Climate
Change, National Research Council, The National Academies Press, Washington,
DC, USA, 292 pp.
NCCARF, 2012: Climate Change Adaptation Research in Australia: An Overview of
Research Funded by the National Climate Change Adaptation Research Facility.
National Climate Change Adaptation Research Facility (NCCARF), Griffith
University Gold Coast, Queensland, Australia, 27 pp.
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.
Nicholls, N., C.D. Butler, and I. Hanigan, 2006: Inter-annual rainfall variations and
suicide in New South Wales, Australia, 1964-2001. International Journal of
Biometeorology, 50(3), 139-143.
Niemeyer, S., J. Petts, and K. Hobson, 2005: Rapid climate change and society:
assessing responses and thresholds. Risk Analysis, 25(6), 1443-1456.
O’Brien, K.L., 2012: Global environmental change II: from adaptation to deliberate
transformation. Progress in Human Geography, 36(5), 667-676.
O’Brien, K.L., 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’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.
866
Chapter 14 Adaptation Needs and Options
14
OECD, 2009: Integrating Climate Change Adaptation into Development Co-operation:
Policy Guidance. Organization for Economic Co-operation and Development
(OECD), OECD Publishing, Paris, France, 193 pp.
OECD, 2011: Financing Climate Change Action and Boosting Technology Change:
Key Messages and Recommendations from Current OECD Work. Organisation
for Economic Co-operation and Development (OECD), OECD Publishing, Paris,
France, 15 pp.
Ogden, A.E. and J. L. Innes, 2009: Adapting to climate change in the southwest Yukon:
locally identified research and monitoring needs to support decision making
on sustainable forest management. Arctic, 62(2), 159-174.
Ormerod, S.J., M. Dobson, A.G. Hildrew, and C.R. Townsend, 2010: Multiple stressors
in freshwater ecosystems. Freshwater Biology, 55(Suppl. 1), 1-4.
Osman-Elasha, B., N. Goutbi, E. Spanger-Siegfried, B. Dougherty, A. Hanafi, S.
Zakieldeen, E-A. Sanjak, H.A. Atti, and H.M. Elhassan, 2009: Community
development and coping with drought in rural Sudan. In: Climate Change and
Adaptation [Leary, N., J. Adejuwon, V. Barros, I. Burton, J. Kulkarni, and R. Lasco
(eds.)]. Earthscan, London, UK, pp. 90-108.
Paavola, J., 2008: Livelihoods, vulnerability and adaptation to climate change in
Morogoro, Tanzania. Environmental Science and Policy, 11(7), 642-654.
Pant, L.P. and R. Heeks, 2011: ICT-Enabled Development of Capacity for Climate Change
Adaptation Centre for Development Informatics Institute for Development
Policy and Management. Centre for Development Informatics, Institute for
Development Policy and Management, SED, University of Manchester,
Manchester, UK, 26 pp.
Park, S.E., N.A. Marshall, E. Jakku, A.M. Dowd, S.M. Howden, E. Mendham, and A.
Fleming, 2012: Informing adaptation responses to climate change through
theories of transformation. Global Environmental Change, 22(1), 115-
126.
Parry, M. and T. Carter, 1998: Climate Impact and Adaptation Assessment: A Guide
to the IPCC Approach. Earthscan, London, UK, 166 pp.
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 (IIED) and Grantham
Institute for Climate Change, London, UK, 113 pp.
Patz, J.A., S.H. Olson, C.K. Uejio, and H.K. Gibbs, 2008: Disease emergence from global
climate and land use change. Medical Clinics of North America, 92(6), 1473-
1491.
Pelling, M., 2010: Adaptation to Climate Change: From Resilience to Transformation.
Routledge, Abingdon, UK and New York, NY, USA, 224 pp.
Perch-Nielsen, S., 2010: The vulnerability of beach tourism to climate change - an
index approach. Climatic Change, 100(3-4), 579-606.
Perkins, W., D., Ojima, and R., Corell, 2007: A survey of climate change adaptation
planning. The Heinz Center, Washington, DC, USA, 52 pp.
Persson, A., R.J.T Klein, C.K. Siebert, A. Atteridge, B. Mueller, J. Hoffmaister, M. Lazarus,
and T. Takama, 2009: Adaptation Finance under a Copenhagen Agreed
Outcome. Stockholm Environment Institute (SEI), Stockholm, Sweden, 187 pp.
Peskett, L., N. Grist, M. Hedger, T. Lennartz-Walker, and I. Scholz, 2009: Climate
Change Challenges for EU Development Co-operation: Emerging Issues.
European Development Co-operation to 2020 Working Paper No. 3, European
Association of Development Research and Training Institutes (EADI), Bonn,
Germany, 23 pp.
Pittock, J., 2011: National climate change policies and sustainable water management:
conflicts and synergies. Ecology and Society, 16(2), 25, www.ecologyand
society.org/vol16/iss2/art25/.
Polsky, C., R. Neff, and B. Yarnal, 2007: Building comparable global change vulnerability
assessments: the vulnerability scoping diagram. Global Environmental Change,
17(3-4), 472-485.
Porio, E., 2011: Vulnerability, adaptation, and resilience to floods and climate change-
related risks among marginal, riverine communities in Metro Manila. Asian
Journal of Social Science, 39(4), 425-445.
Porter, M.G., 2013: A Tale of Two Cities: Desalination and Drought in Perth and
Melbourne. A report prepared for NCEDA under: ‘Desalination for Australian
Economic Development,’ National Centre of Excellence in Desalination Australia
(NCEDA), Alfred Deakin Research Institute, Deakin University, Melbourne,
Australia, 43 pp.
Pouliotte, J., B. Smit, and L. Westerhoff, 2009: Adaptation and development: livelihoods
and climate change in Subarnabad, Bangladesh. Climate and Development,
1(1), 31-46.
Prabhakar, S.V.R.K. and A. Srinivasan, 2011: Metrics for mainstreaming adaptation
in agriculture sector. Climate Change and Food Security in South Asia, 8, 551-567.
Pramova E., B. Locatelli, M. Brockhaus, and S. Fohlmeister, 2012: Ecosystem services
in the National Adaptation Programmes of Action. Climate Policy, 12(4), 393-409.
Preston, B. and M. Stafford-Smith, 2009: Framing Vulnerability and Adaptive Capacity
Assessment: Discussion Paper. CSIRO Climate Adaptation National Research
Flagship Working Paper No. 2, CSIRO Marine and Atmospheric Research,
Aspendale, Victoria, Australia, 52 pp.
Preston, B.L., R.M. Westaway, S. Dessai, and T. Smith, 2009: Are We Adapting to
Climate Change? Research and Methods for Evaluating Progress. In: 89
th
American Meteorological Society Annual Meeting: Fourth Symposium on Policy
and Socio-Economic Research, Phoenix, AZ, 15 pp.
Preston, B.L., R.M. Westaway, and E.J. Yuen, 2011a: 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.
Preston, B.L., E.J. Yuan, and R.M. Westaway, 2011b: Putting vulnerability to climate
change on the map: a review of approaches, benefits and risks. Sustainability
Science, 6, 177-202.
Pulwarty, R.S., C. Simpson, and C.R. Nierenberg, 2009: The Regional Integrated
Sciences and Assessments (RISA) Program: Crafting effective assessments for
the long haul. In: Integrated Regional Assessment of Global Climate Change
[Knight, C.G. and J. Jäger (eds.)]. Cambridge University Press, Cambridge, UK,
pp. 367-393.
Ranger, N. and S.-L. Garbett-Shiels, 2012: Accounting for a changing and uncertain
climate in planning and policymaking today: lessons for developing countries.
Climate and Development, 4(4), 288-300.
Reid, H. and S. Huq, 2005: Climate change- biodiversity and livelihood impacts. In:
Tropical Forests and Adaptation to Climate Change: In Search of Synergies
[Robledo, C., M. Kanninen, and L. Pedroni (eds.)]. Center for International
Forestry Research, Bogor Barat, Indonesia, pp. 57-70.
Repetto, R., 2008: The Climate Crisis and the Adaptation Myth. Yale School of
Forestry and Environmental Studies Working Paper 13, Yale School of Forestry
and Environmental Studies, New Haven, CT, USA, 21 pp.
Reser, J.P. and J.K. Swim, 2011: Adapting to and coping with the threat and impacts
of climate change. American Psychologist, 66(4), 277-289.
Reser, J.P., G. Bradley, and M. Ellul, 2012: Coping with climate change: bringing
psychological adaptation in from the cold. In: Handbook of the Psychology of
Coping: New Research [Molinelli B. and V. Grimaldo (eds.)]. Nova Science
Publishers, Hauppauge, NY, USA, pp. 1-34.
Revi, A., 2008: Climate change risk: an adaptation and mitigation agenda for Indian
cities. Environment and Urbanization, 20(1), 207-229.
Reyer, C., M. Guericke, and P.L. Ibisch, 2009: Climate change mitigation via afforestation,
reforestation and deforestation avoidance: and what about adaptation to
environmental change? New Forests, 38(1), 15-34.
Ribot, J., 2010: Vulnerability does not fall from the sky: toward multiscale, pro-poor
climate policy. In: Social Dimensions of Climate Change: Equity and Vulnerability
in a Warming World [Mearns, R. and A. Norton (eds.)]. The International Bank
for Reconstruction and Development / The World Bank, Washington, DC, USA,
pp. 47-74.
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.
Rosenzweig, C., W. Solecki, S.A. Hammer, and S. Mehrotra, 2010: Cities lead the way
in climate-change action. Nature, 467, 909-911.
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 the New York City infrastructure-shed: process, approach,
tools, and strategies. Climatic Change, 106(1), 93-127.
Rygel, L., D. O’Sullivan, and B. Yarnal, 2006: A method for constructing a social
vulnerability index: an application to hurricane storm surges in a developed
country. Mitigation and Adaptation Strategies for Global Change, 11, 741-764.
Satterthwaite, D. and D. Dodman, 2009: The costs of adapting infrastructure to
climate change. In: Assessing the Costs of Adaptation to Climate Change: A
Review of the UNFCCC and Other Recent Estimates [Parry, M., N. Arnell, P. Berry,
D. Dodman, S. Fankhauser, C. Hope, S. Kovats, R. Nicholls, D. Satterthwaite, R.
Tiffin, and T. Wheeler (eds.)]. International Institute for Environment and
867
Adaptation Needs and Options Chapter 14
14
Development (IIED) and the Grantham Institute for Climate Change, IIED,
London, UK, pp. 73-89.
Satterthwaite, D., S. Huq, H. Reid, M. Pelling, and P. Romero Lankao, 2009: Adapting
to climate change in urban areas: the possibilities and constraints in low- and
middle-income nations. In: Adapting Cities to Climate Change [Bicknell, J., D.
Dodman, and D. Satterthwaite (eds.)]. Earthscan, London, UK, pp. 3-47.
Scheraga, J.D., K.L. Ebi, J. Furlow, and A.R. Moreno, 2003: From science to policy:
developing responses to climate change. In:Climate Change and Human
Health: Risks and Responses [McMichael, A.J., D. Lendrum, C.F. Corvalan, K.L.
Ebi, A. Githeko, and J.D. Scheraga (eds.)]. World Health Organization (WHO),
World Meteorological Organization (WMO), United Nations Environment
Programme (UNDP), World Health Organization, Geneva, Switzerland, pp.
237-266.
Schipper, L., 2009: Meeting the crossroads? Exploring the linkages between climate
change adaptation and disaster risk reduction. Climate and Development, 1(1),
16-30.
Schröter, D., C. Polsky, and A. Patt, 2005: Assessing vulnerabilities to the effects of
global change: an eight step approach. Mitigation and Adaptation Strategies
for Global Change, 10(4), 573-595.
Semenov, M.A., 2006: Using weather generators in crop modeling. Acta Horticulturae,
707, 93-100.
Semenov, M.A., 2008: Simulation of extreme weather events by a stochastic weather
generator. Climate Research, 35(3), 203-212.
Semenza, J.C., S. Herbst, A. Rechenburg, J.E. Suk, C. Hoser, C. Schreiber, and T.
Kistemann, 2012: Climate change impact assessment of food- and waterborne
diseases. Critical Reviews in Environmental Science and Technology, 42(8),
857-890.
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 cholera outbreak
in western Kenya after post-election violence in 2008. The American Journal
of Tropical Medicine and Hygiene, 81(6), 1085-1090.
Sietz, D., M. Boschutz, and R.J.T. Klein, 2011a: Mainstreaming climate adaptation
into development assistance: rationale, institutional barriers and opportunities
in Mozambique. Environmental Science and Policy, 14(4), 493-502.
Sietz, D., M.K.B. Lüdeke, and C. Walther, 2011b: Categorisation of typical vulnerability
patterns in global drylands. Global Environmental Change, 21(2), 431-440.
Skoufias, E. (ed.), 2012: The Poverty and Welfare Impacts of Climate Change:
Quantifying the Effects, Identifying the Adaptation Strategies. World Bank
Publications, Washington, DC, USA, 184 pp.
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, pp. 9-28.
Smit, B. and J. Wandel, 2006: Adaptation, adaptive capacity and vulnerability. Global
Environmental Change, 16(3), 282-292.
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 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 and New York, NY, USA, pp. 877-912.
Smith, J.B., J.M. Vogel, and J.E. Cromwell III, 2009: An architecture for government
action on adaptation to climate change. Climatic Change, 95 (1-2), 53-61.
Sovacool, B.K., 2011: Hard and soft paths for climate change adaptation. Climate
Policy, 11(4), 1177-1183.
Sovacool, B.K., A.L. D’Agostino, H. Meenawat, and A. Rawlani, 2012: Expert views
of climate change adaptation in least developed Asia. Journal of Environmental
Management, 97, 78-88.
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.
Srinivasan, A. and S.V.R.K. Prabhakar, 2009: Measures of Adaptation to Climatic
Change and Variability (Adaptation Metrics). Final Report to the World Bank,
Contract No. 7145543, Institute for Global Environmental Strategies (IGES)
Research Report, IGES, Hayama, Japan and The World Bank, Washington, DC,
USA, 108 pp., http://pub.iges.or.jp/modules/envirolib/view.php?docid=2940.
Stafford Smith, M., L. Horrocks, A. Harvey, and C. Hamilton, 2010: Rethinking
adaptation for a 4°C world. Philosophical Transactions of the Royal Society A,
369(1934), 196-216.
Stern, N., 2006: The Economics of Climate Change. Cambridge University Press,
Cambridge, UK, 712 pp.
Stokes, C. and M. Howden, 2010: Adapting Agriculture to Climate Change: Preparing
Australian Agriculture, Forestry and Fisheries for the Future. CSIRO Publishing,
Canberra, Australia, 296 pp.
Suarez, P., F. Ching, G. Ziervogel, I. Lemaire, D. Turnquest, J. Mendler de Suarez, and
B. Wisner, 2009: Video-mediated approaches for community-level climate
adaptation. IDS Bulletin, 39(4), 96-104.
Swart, R. and F. Raes, 2007: Making integration of adaptation and mitigation work
mainstreaming into sustainable development policies? Climate Policy, 7, 288-
303.
Tengö, M. and K. Belfrage, 2004: Local management practices for dealing with
change and uncertainty: a cross-scale comparison of cases in Sweden and
Tanzania. Ecology and Society, 9(3), 4, www.ecologyandsociety.org/vol9/iss3/
art4/.
The Climate Institute, 2011: A Climate of Suffering: The Real Cost of Living with
Inaction on Climate Change. The Climate Institute, Melbourne and Sydney,
Australia, 28 pp.
Thomsen, D.C., T.F. Smith, and N. Keys, 2012: Adaptation or manipulation? Unpacking
climate change response strategies. Ecology and Society, 17(3), 20,
www.ecologyandsociety.org/vol17/iss3/art20/.
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.
Tonnang, H.E.Z., R.Y.M. Kangalawe, and P.Z. Yanda, 2010: Predicting and mapping
malaria under climate change scenarios: the potential redistribution of malaria
vectors in Africa. Malaria Journal, 9, 111-120.
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(4),
627-635.
Turner, W.R., B.A. Bradley, L.D. Estes, D.G. Hole, M. Oppenheimer, and D.S. Wilcove,
2010: Climate change: helping nature survive the human response. Conservation
Letters, 3(5), 304-312.
UKCIP, 2011: Making Progress: UKCIP & Adaptation in the UK. UK Climate Impacts
Programme (UKCIP), School of Geography and the Environment, at the
University of Oxford, Oxford, UK, 99 pp.
UNDP, 2004: Reducing Disaster Risk: A Challenge for Development. United Nation
Development Programme (UNDP), Bureau for Crisis Prevention and Recovery,
New York, NY, USA, 146 pp.
UNDP, 2005: Adaptation Policy Frameworks for Climate Change: Developing
Strategies, Policies and Measures. Cambridge University Press, New York, NY,
USA, 258 pp.
UNDP, 2008: Human Development Report 2007/8. Fighting Climate Change: Human
Solidarity in a Divided World. United Nations Development Programme (UNDP),
Palgrave Macmillan, Houndmills, Basingstoke, Hampshire, UK and New York,
NY, USA, 384 pp.
UNDP, 2010: Gender, Climate Change and Community-Based Adaptation. A
Guidebook for Designing and Implementing Gender-Sensitive Community-
Based Adaptation Programmes and Projects, United Nations Development
Programme (UNDP), New York, NY, USA, 69 pp.
UNFCCC, 2006: Technologies for Adaptation to Climate Change. The United Nations
Framework Convention on Climate Change (UNFCCC), UNFCCC Secretariat,
Bonn, Germany, 38 pp.
UNFCCC, 2007: Investment and Financial Flows to Address Climate Change. The
United Nations Framework Convention on Climate Change (UNFCCC), UNFCCC
Secretariat, Bonn, Germany, 272 pp.
UNFCCC, 2010: Synthesis Report on Efforts Undertaken to Assess the Costs and
Benefit of Adaptation Options, and Views on Lessons Learned, Good Practices,
Gaps and Needs. UNFCCC Secretariat, Bonn, Germany, 13 pp.
UNFCCC, 2011: Assessing the Costs and Benefits of Adaptation Options: An Overview
of Approaches. The Nairobi Work Programme and United Nations Framework
Convention on Climate Change (UNFCCC), UNFCCC Secretariat, Bonn,
Germany, 48 pp.
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 Aalst, M.K., T. Cannon, and I. Burton, 2008: Community level adaptation to
climate change: the potential role of participatory risk assessment. Global
Environmental Change, 18(1), 165-179.
868
Chapter 14 Adaptation Needs and Options
14
Verner, D., 2012: Gender-responsive climate change adaptation: ensuring effectiveness
and sustainability. In: Adaptation to a Changing Climate in the Arab Countries:
A Case for Adaptation Governance and Leadership in Building Climate
Resilience [Verner, D. (ed.)]. The International Bank for Reconstruction and
Development / The World Bank, Washington, DC, USA, pp. 277-308.
Villamizar, A., 2011: Adaptación al Cambio Climático ypolíticas públicas: una deuda
de alto riesgo para América Latina?Mundo Nuevo,1(6), 301-338.
Webb, R. and J. Beh, 2013:Leading Adaptation Practices and Support Strategies
forAustralia: An International and Australian Review of Products and Tools.
Synthesis and Integrative Research: Final Report, NationalClimate Change
Adaptation Research Facility (NCCARF) and Australian National University,
NCCARF, Griffith University Gold Coast Campus, Southport, Australia, 105 pp.
West, J.M. and D. Brereton, 2013: Climate Change Adaptation in Industry and
Business. Synthesis and Integrative Research Final Report, National Climate
Change Adaptation Research Facility (NCCARF), Griffith University Gold Coast
Campus, Southport, Australia, 144 pp.
Westerhoff, L., E.C.H. Keskitalo, and S. Juhola, 2011: Capacities across scales: local
to national adaptation policy in four European countries. Climate Policy, 11(4),
1071-1085.
Wheeler, D., 2011: Quantifying Vulnerability to Climate Change: Implications for
Adaptation Assistance. CGD Working Paper 240, Center for Global Development
(CGD), Washington, DC, USA, 49 pp.
Wolf, J., W.N. Adger, I. Lorenzoni, V. Abrahamson, and R. Raine, 2010: Social capital,
individual responses to heat waves and climate change adaptation: an empirical
study of two UK cities. Global Environmental Change, 20(1), 44-52.
World Bank, 2008: The Pilot Program for Climate Resilience under the Strategic
Climate Fund. CTF-SCF/7 November 10, 2008, The World Bank, Washington,
DC, USA, 7 pp.
World Bank, 2010: World Development Report 2010: Development in a Changing
Climate Concept Note. The International Bank for Reconstruction and
Development / The World Bank, Washington, DC, USA, 43 pp.
World Economic Forum, 2002: 2002 Environmental Sustainability Index. An Initiative
of the Global Leaders for Tomorrow Environment Task Force, World Economic
Forum (WEF), Annual Meeting 2002, in collaboration with Yale Center for
Environmental Law and Policy (YCELP), Yale University and Center for
International Earth Science Information Network (CIESIN), Columbia University,
YCELP, New Haven, CT, USA, 297 pp.
World Economic Forum, 2012: Global Risks 2012.7
th
edn., An initiative of the Risk
Response Network, World Economic Forum (WEF), Geneva, Switzerland, 62 pp.
WRI, 2009: Bellagio Framework for Adaptation Assessment and Prioritization. WRI
Working Paper, World Resources Institute (WRI), Washington, DC, USA, 6 pp.
Wu, Q., G. Cheng, W. Ma, and Y. Liu, 2008: Railway construction techniques adapting
to climate warming in permafrost regions. Advances in Climate Change
Research, 4, 60-66.
Yohe, G. and R.S.J. Tol, 2001: Indicators for social and economic coping capacity –
moving toward a working definition of adaptive capacity. Global Environmental
Change, 12(1), 25-40.
Young, H., A.M. Osman, A.M. Abusin, M. Asher, and O. Egemi, 2009: Livelihoods,
Power and Choice: The Vulnerability of the Northern Rizaygat, Darfur, Sudan.
Feinstein International Center, Tufts University, Medford, MA, USA, 115 pp.
Ziervogel, G. and A. Taylor, 2008: Feeling stressed: integrating climate adaptation
with other priorities in South Africa. Environment, 50(2), 32-41.