1267
23
Europe
Coordinating Lead Authors:
R. Sari Kovats (UK), Riccardo Valentini (Italy)
Lead Authors:
Laurens M. Bouwer (Netherlands), Elena Georgopoulou (Greece), Daniela Jacob (Germany),
Eric Martin (France), Mark Rounsevell (UK), Jean-Francois Soussana (France)
Contributing Authors:
Martin Beniston (Switzerland), Maria Vincenza Chiriacò (Italy), Philippe Cury (France),
Michael Davies (UK), Paula Harrison (UK), Olaf Jonkeren (Netherlands), Mark Koetse
(Netherlands), Markus Lindner (Finland), Andreas Matzarakis (Greece/Germany),
Reinhard Mechler (Germany), Annette Menzel (Germany), Marc Metzger (UK),
Luca Montanarella (Italy), Antonio Navarra (Italy), Juliane Petersen (Germany), Martin Price
(UK), Boris Revich (Russian Federation), Piet Rietveld (Netherlands), Cristina Sabbioni (Italy),
Yannis Sarafidis (Greece), Vegard Skirbekk (Austria), Donatella Spano (Italy), Jan E. Vermaat
(Netherlands), Paul Watkiss (UK), Meriwether Wilson (UK), Thomasz Zylicz (Poland)
Review Editors:
Lucka Kajfez Bogataj (Slovenia), Roman Corobov (Moldova), Ramón Vallejo (Spain)
This chapter should be cited as:
Kovats
, R.S., R. Valentini, L.M. Bouwer, E. Georgopoulou, D. Jacob, E. Martin, M. Rounsevell, and J.-F. Soussana,
2014: Europe. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part B: Regional Aspects.
Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate
Change [Barros, V.R., C.B. Field, D.J. Dokken, M.D. Mastrandrea, K.J. Mach, 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. 1267-1326.
23
1268
Executive Summary.......................................................................................................................................................... 1270
23.1. Introduction .......................................................................................................................................................... 1274
23.1.1. Scope and Route Map of Chapter ................................................................................................................................................... 1274
23.1.2. Policy Frameworks .......................................................................................................................................................................... 1274
23.1.3. Conclusions from Previous Assessments ......................................................................................................................................... 1274
23.2. Current and Future Trends .................................................................................................................................... 1275
23.2.1 Non-Climate Trends ........................................................................................................................................................................ 1275
23.2.2. Observed and Projected Climate Change ........................................................................................................................................ 1275
23.2.2.1. Observed Climate Change ............................................................................................................................................... 1275
23.2.2.2. Projected Climate Changes ............................................................................................................................................. 1276
23.2.2.3. Projected Changes in Climate Extremes .......................................................................................................................... 1276
23.2.3. Observed and Projected Trends in Riverflow and Drought .............................................................................................................. 1279
23.3. Implications of Climate Change for Production Systems and Physical Infrastructure ......................................... 1279
23.3.1. Settlements ..................................................................................................................................................................................... 1279
23.3.1.1. Coastal Flooding ............................................................................................................................................................. 1279
23.3.1.2. River and Pluvial Flooding ............................................................................................................................................... 1280
23.3.1.3. Windstorms ..................................................................................................................................................................... 1281
23.3.1.4. Mass Movements and Avalanches .................................................................................................................................. 1281
23.3.2. Built Environment ........................................................................................................................................................................... 1281
23.3.3. Transport ........................................................................................................................................................................................ 1281
23.3.4. Energy Production, Transmission, and Use ...................................................................................................................................... 1282
23.3.5. Industry and Manufacturing ........................................................................................................................................................... 1283
23.3.6. Tourism ........................................................................................................................................................................................... 1283
23.3.7. Insurance and Banking ................................................................................................................................................................... 1283
23.4. Implications of Climate Change for Agriculture, Fisheries, Forestry, and Bioenergy Production ......................... 1284
23.4.1. Plant (Food) Production .................................................................................................................................................................. 1284
23.4.2. Livestock Production ....................................................................................................................................................................... 1286
23.4.3. Water Resources and Agriculture .................................................................................................................................................... 1286
23.4.4. Forestry ........................................................................................................................................................................................... 1287
23.4.5. Bioenergy Production ..................................................................................................................................................................... 1288
Box 23-1. Assessment of Climate Change Impacts on Ecosystem Services by Sub-region ....................................................... 1288
23.4.6. Fisheries and Aquaculture ............................................................................................................................................................... 1290
23.5. Implications of Climate Change for Health and Social Welfare ............................................................................ 1290
23.5.1. Human Population Health ............................................................................................................................................................... 1290
23.5.2. Critical Infrastructure ...................................................................................................................................................................... 1291
Table of Contents
1269
Europe Chapter 23
23
23.5.3. Social Impacts ................................................................................................................................................................................. 1291
23.5.4. Cultural Heritage and Landscapes .................................................................................................................................................. 1292
Box 23-2. Implications of Climate Change for European Wine and Vineyards .......................................................................... 1292
23.6. Implications of Climate Change for the Protection of Environmental Quality and Biological Conservation ...... 1293
23.6.1. Air Quality ...................................................................................................................................................................................... 1293
23.6.2. Soil Quality and Land Degradation ................................................................................................................................................. 1293
23.6.3. Water Quality ................................................................................................................................................................................. 1294
23.6.4. Terrestrial and Freshwater Ecosystems ........................................................................................................................................... 1294
23.6.5. Coastal and Marine Ecosystems ..................................................................................................................................................... 1294
23.7. Cross-Sectoral Adaptation Decision Making and Risk Management ................................................................... 1295
Box 23-3. National and Local Adaptation Strategies ................................................................................................................ 1295
23.7.1. Coastal Zone Management ............................................................................................................................................................ 1296
23.7.2. Integrated Water Resource Management ....................................................................................................................................... 1296
23.7.3. Disaster Risk Reduction and Risk Management .............................................................................................................................. 1296
23.7.4. Land Use Planning .......................................................................................................................................................................... 1296
23.7.5. Rural Development ......................................................................................................................................................................... 1297
23.7.6. Economic Assessments of Adaptation ............................................................................................................................................. 1297
23.7.7. Barriers and Limits to Adaptation ................................................................................................................................................... 1298
23.8. Co-benefits and Unintended Consequences of Adaptation and Mitigation ......................................................... 1298
23.8.1. Production and Infrastructure ......................................................................................................................................................... 1298
23.8.2. Agriculture, Forestry, and Bioenergy ............................................................................................................................................... 1299
23.8.3. Social and Health Impacts .............................................................................................................................................................. 1299
23.8.4. Environmental Quality and Biological Conservation ....................................................................................................................... 1299
23.9. Synthesis of Key Findings ..................................................................................................................................... 1300
23.9.1. Key Vulnerabilities .......................................................................................................................................................................... 1300
23.9.2. Climate Change Impacts Outside Europe and Inter-regional Implications ...................................................................................... 1303
23.9.3. Effects of Observed Climate Change in Europe ............................................................................................................................... 1303
23.9.4. Key Knowledge Gaps and Research Needs ..................................................................................................................................... 1304
References ....................................................................................................................................................................... 1306
Frequently Asked Questions
23.1: Will I still be able to live on the coast in Europe? ........................................................................................................................... 1305
23.2: Will climate change introduce new infectious diseases into Europe? ............................................................................................. 1305
23.3: Will Europe need to import more food because of climate change? ............................................................................................... 1305
1270
Chapter 23 Europe
23
Executive Summary
Observed climate trends and future climate projections show regionally varying changes in temperature and rainfall in Europe
(high confidence), {23.2.2} in agreement with Fourth Assessment Report (AR4) findings, with projected increases in temperature throughout
Europe and increasing precipitation in Northern Europe and decreasing precipitation in Southern Europe. {23.2.2.2} Climate projections show a
marked increase in high temperature extremes (high confidence), meteorological droughts (medium confidence), {23.2.3} and heavy precipitation
events (high confidence), {23.2.2.3} with variations across Europe, and small or no changes in wind speed extremes (low confidence) except
increases in winter wind speed extremes over Central and Northern Europe (medium confidence). {23.2.2.3}
Observed climate change in Europe has had wide ranging effects throughout the European region including the distribution,
phenology, and abundance of animal, fish, and plant species (high confidence) {23.6.4-5; Table 23-6}; stagnating wheat yields in
some sub-regions (medium confidence, limited evidence) {23.4.1}; and forest decline in some sub-regions (medium confidence).
{23.4.4}
Climate change has affected both human health (from increased heat waves) (medium confidence) {23.5.1} and animal health
(changes in infectious diseases) (high confidence). {23.4.2} There is less evidence of impacts on social systems attributable to observed climate
change, except in pastoralist populations (low confidence). {23.5.3}
Climate change will increase the likelihood of systemic failures across European countries caused by extreme climate events
affecting multiple sectors (medium confidence). {23.2.2.3, 23.2.3, 23.3-6, 23.9.1} Extreme weather events currently have significant
impacts in Europe in multiple economic sectors as well as adverse social and health effects (high confidence). {Table 23-1} There is limited evidence
that resilience to heat waves and fires has improved in Europe (medium confidence), {23.9.1, 23.5} while some countries have improved their
flood protection following major flood events. {23.9.1, 23.7.3} Climate change is very likely to increase the frequency and intensity of heat
waves, particularly in Southern Europe (high confidence), {23.2.2} with mostly adverse implications for health, agriculture, forestry, energy
production and use, transport, tourism, labor productivity, and the built environment. {23.3.2-4, 23.3.6, 23.4.1-4, 23.5.1; Table 23-1}
The provision of ecosystem services is projected to decline across all service categories in response to climate change in Southern
Europe (high confidence). {23.9.1; Box 23-1} Both gains and losses in the provision of ecosystem services are projected for the other
European sub-regions (high confidence), but the provision of cultural services is projected to decline in the Continental, Northern, and Southern
sub-regions (low confidence). {Box 23-1}
Climate change is expected to impede economic activity in Southern Europe more than in other sub-regions (medium confidence)
{23.9.1; Table 23-4}, and may increase future intra-regional disparity (low confidence). {23.9.1} There are also important differences
in vulnerability within sub-regions; for example, plant species and some economic sectors are most vulnerable in high mountain areas due to
lack of adaptation options (medium confidence). {23.9.1} Southern Europe is particularly vulnerable to climate change (high confidence),
as multiple sectors will be adversely affected (tourism, agriculture, forestry, infrastructure, energy, population health) (high confidence).
{23.9; Table 23-4}
The impacts of sea level rise on populations and infrastructure in coastal regions can be reduced by adaptation (medium
confidence). {23.3.1, 23.5.3}
Populations in urban areas are particularly vulnerable to climate change impacts because of the high density of
people and built infrastructure (medium confidence). {23.3, 23.5.1}
Synthesis of evidence across sectors and sub-regions confirm that there are limits to adaptation from physical, social, economic,
and technological factors (high confidence). {23.7; Table 23-3}
Adaptation is further impeded because climate change affects multiple
sectors. {23.7} The majority of published assessments are based on climate projections in the range 1°C to 4°C global mean temperature per
century. Limited evidence exists regarding the potential impacts in Europe under high rates of warming (>4°C global mean temperature per
century). {23.9.1}
1271
Europe Chapter 23
23
Impacts by Sector
Sea level rise and increases in extreme rainfall are projected to further increase coastal and river flood risk in Europe and,
without adaptive measures, will substantially increase flood damages (people affected and economic losses) (high confidence).
{23.3.1, 23.5.1}
Adaptation can prevent most of the projected damages (high confidence, based on medium evidence, high agreement) but
there may be constraints to building flood defenses in some areas. {23.3.1, 23.7.1} Direct economic river flood damages in Europe have
increased over recent decades (high confidence) but this increase is due to development in flood zones and not due to observed climate
change. {23.3.1.2; SREX 4.5} Some areas in Europe show changes in river flood occurrence related to observed changes in extreme river
discharge (medium confidence). {23.2.3}
Climate change is projected to affect the impacts of hot and cold weather extremes on transport leading to economic damage
and/or adaptation costs, as well as some benefits (e.g., reduction of maintenance costs) during winter (medium confidence).
{23.3.3}
Climate change is projected to reduce severe accidents in road transport (medium confidence) and adversely affect inland water
transport in summer in some rivers (e.g., the Rhine) after 2050 (medium confidence). Damages to rail infrastructure from high temperatures
may also increase (medium confidence). Adaptation through maintenance and operational measures can reduce adverse impacts to some
extent.
Climate change is expected to affect future energy production and transmission. {23.3.4} Hydropower production is likely to decrease
in all sub-regions except Scandinavia (high confidence). {23.3.4} Climate change is unlikely to affect wind energy production before 2050
(medium confidence) but will have a negative impact in summer and a varied impact in winter after 2050 (medium confidence). Climate
change is likely to decrease thermal power production during summer (high confidence). {23.3.4} Climate change will increase the problems
associated with overheating in buildings (medium confidence). {23.3.2} Although climate change is very likely to decrease space heating demand
(high confidence), cooling demand will increase (very high confidence) although income growth mostly drives projected cooling demand up to
2050 (medium confidence). {23.3.4} More energy-efficient buildings and cooling systems as well as demand-side management will reduce future
energy demands. {23.3.4}
After 2050, tourism activity is projected to decrease in Southern Europe (low confidence) and increase in Northern and Continental
Europe (medium confidence). No significant impacts on the tourism sector are projected before 2050 in winter or summer tourism except for
ski tourism in low-altitude sites and under limited adaptation (medium confidence). {23.3.6} Artificial snowmaking may prolong the activity of
some ski resorts (medium confidence). {23.3.6}
Climate change is likely to increase cereal yields in Northern Europe (medium confidence, disagreement) but decrease yields in
Southern Europe (high confidence). {23.4.1} In Northern Europe, climate change is very likely to extend the seasonal activity of pests and
plant diseases (high confidence). {23.4.1} Yields of some arable crop species like wheat have been negatively affected by observed warming in
some European countries since the 1980s (medium confidence, limited evidence). {23.4.1} Compared to AR4, new evidence regarding future
yields in Northern Europe is less consistent regarding the magnitude and sign of change. Climate change may adversely affect dairy production in
Southern Europe because of heat stress in lactating cows (medium confidence). {23.4.2} Climate change has contributed to vector-borne disease
in ruminants in Europe (high confidence) {23.4.2} and northward expansion of tick disease vectors (medium confidence). {23.4.2, 23.5.1}
Climate change will increase irrigation needs (high confidence) but future irrigation will be constrained by reduced runoff,
demand from other sectors, and by economic costs. {23.4.1, 23.4.3}
By the 2050s, irrigation will not be sufficient to prevent damage
from heat waves to crops in some sub-regions (medium confidence). System costs will increase under all climate scenarios (high confidence).
{23.4.3} Integrated management of water, also across countries’ boundaries, is needed to address future competing demands among agriculture,
energy, conservation, and human settlements. {23.7.2}
1272
Chapter 23 Europe
23
As a result of increased evaporative demand, climate change is likely to significantly reduce water availability from river
abstraction and from groundwater resources (medium confidence)
, in the context of increased demand (from agriculture, energy and
industry, and domestic use) and cross-sectoral implications that are not fully understood. {23.4.3, 23.9.1} Some adaptation is possible through
uptake of more water-efficient technologies and water-saving strategies. {23.4.3, 23.7.2}
Climate change will change the geographic distribution of wine grape varieties (high confidence) and this will reduce the value
of wine products and the livelihoods of local wine communities in Southern and Continental Europe (medium confidence) and
increase production in Northern Europe (low confidence). {23.4.1, 23.3.5, 23.5.4; Box 23-2}
Some adaptation is possible through
technologies and good practice. {Box 23-2}
Climate warming will increase forest productivity in Northern Europe (medium confidence), {23.4.4} although damage from pests
and diseases in all sub-regions will increase due to climate change (high confidence). {23.4.4} Wildfire risk in Southern Europe (high confidence)
and damages from storms in Central Europe (low confidence) may also increase due to climate change. {23.4.4} Climate change is likely to
cause ecological and socioeconomic damages from shifts in forest tree species range (from southwest to northeast) (medium confidence), and
in pest species distributions (low confidence). {23.4.4} Forest management measures can enhance ecosystem resilience (medium confidence).
{23.4.4}
Observed warming has shifted marine fish species ranges to higher latitudes (high confidence) and reduced body size in species
(medium confidence). {23.4.6} There is limited and diverging evidence on climate change impacts on net fisheries economic turnover. Local
economic impacts attributable to climate change will depend on the market value of (high temperature tolerant) invasive species. {23.4.6}
Climate change is unlikely to entail relocation of fishing fleets (high confidence). {23.4.6} Observed higher water temperatures have adversely
affected both wild and farmed freshwater salmon production in the southern part of their distribution (high confidence). {23.4.6} High
temperatures may increase the frequency of harmful algal blooms (low confidence). {23.4.6}
Climate change will affect bioenergy cultivation patterns in Europe by shifting northward their potential area of production
(medium confidence). {23.4.5}
Elevated atmospheric carbon dioxide (CO
2
) can improve drought tolerance of bioenergy crop species due to
improved plant water use, maintaining high yields in future climate scenarios in temperate regions (low confidence). {23.4.5}
Climate change is likely to affect human health in Europe. Heat-related deaths and injuries are likely to increase, particularly in Southern
Europe (medium confidence). {23.5.1} Climate change may change the distribution and seasonal pattern of some human infections, including
those transmitted by arthropods (medium confidence), and increase the risk of introduction of new infectious diseases (low confidence).
{23.5.1}
Climate change and sea level rise may damage European cultural heritage, including buildings, local industries, landscapes, archaeological
sites, and iconic places (medium confidence), and some cultural landscapes may be lost forever (low confidence). {23.5.4; Table 23-3}
Climate change may adversely affect background levels of tropospheric ozone (low confidence; limited evidence, low agreement),
assuming no change in emissions, but the implications for future particulate pollution (which is more health-damaging) are very
uncertain. {23.6.1}
Higher temperatures may have affected trends in ground level tropospheric ozone (low confidence). {23.6.1} Climate
change is likely to decrease surface water quality due to higher temperatures and changes in precipitation patterns (medium confidence),
{23.6.3} and is likely to increase soil salinity in coastal regions (low confidence). {23.6.2} Climate change may also increase soil erosion (from
increased extreme events) and reduce soil fertility (low confidence, limited evidence). {23.6.2}
1273
Europe Chapter 23
23
Observed climate change is affecting a wide range of flora and fauna, including plant pests and diseases (high confidence)
{23.4.1, 23.4.4, 23.6.4} and the disease vectors and hosts (medium confidence). {23.4.2}
Climate change is very likely to cause
changes in habitats and species, with local extinctions (high confidence) and continental-scale shifts in species distributions (medium confidence).
{23.6.4} The habitat of alpine plants is very likely to be significantly reduced (high confidence). {23.6.4} Phenological mismatch will constrain
both terrestrial and marine ecosystem functioning under climate change (high confidence), {23.6.4-5} with a reduction in some ecosystem services
(low confidence). {23.6.4; Box 23-1} The introduction and expansion of invasive species, especially those with high migration rates, from outside
Europe is likely to increase with climate change (medium confidence). {23.6.4} Climate change is likely to entail the loss or displacement of
coastal wetlands (high confidence). {23.6.5} Climate change threatens the effectiveness of European conservation areas (low confidence),
{23.6.4} and stresses the need for habitat connectivity through specific conservation policies. {23.6.4}
Adaptation
The capacity to adapt in Europe is high compared to other world regions, but there are important differences in impacts and in
the capacity to respond between and within the European sub-regions.
In Europe, adaptation policy has been developed at international
(European Union), national, and local government levels, {23.7} including the prioritization of adaptation options. There is limited systematic
information on current implementation or effectiveness of adaptation measures or policies. {Box 23-3} Some adaptation planning has been
integrated into coastal and water management, as well as disaster risk management. {23.7.1-3} There is limited evidence of adaptation planning
in rural development or land use planning. {23.7.4-5}
Adaptation will incur a cost, estimated from detailed bottom-up sector-specific studies for coastal defenses, energy production, energy use,
and agriculture. {23.7.6} The costs of adapting buildings (houses, schools, hospitals) and upgrading flood defenses increase under all scenarios
relative to no climate change (high confidence). {23.3.2} Some impacts will be unavoidable owing to limits (physical, technological, social,
economic, or political). {23.7.7; Table 23-3}
There is also emerging evidence regarding opportunities and unintended consequences of policies, strategies, and measures that
address adaptation and/or mitigation goals. {23.8} Some agricultural practices can reduce greenhouse gas (GHG) emissions and also
increase resilience of crops to temperature and rainfall variability. {23.8.2} There is evidence for unintended consequences of mitigation policies
in the built environment (especially dwellings) and energy sector (medium confidence). {23.8.1} Low-carbon policies in the transport and energy
sectors to reduce emissions are associated with large benefits to human health (high confidence). {23.8.3}
1274
Chapter 23 Europe
23
23.1. Introduction
This chapter reviews the scientific evidence published since the IPCC
Fourth Assessment Report (AR4) on observed and projected impacts of
a
nthropogenic climate change in Europe and adaptation responses. The
geographical scope of this chapter is the same as in AR4 with the
inclusion of Turkey. Thus, the European region includes all countries from
Iceland in the west to the Russian Federation (west of the Urals) and
the Caspian Sea in the east, and from the northern shores of the
Mediterranean and Black Seas and the Caucasus in the south to the
Arctic Ocean in the north. Impacts above the Arctic Circle are addressed
in Chapter 28 and impacts in the Baltic and Mediterranean Seas in
Chapter 30. Impacts in Malta, Cyprus, and other island states in Europe
are discussed in Chapter 29. The European region has been divided into
five sub-regions (see Figure 23-1): Atlantic, Alpine, Southern, Northern,
and Continental. The sub-regions are derived by aggregating the climate
zones developed by Metzger et al. (2005) and therefore represent
geographical and ecological zones rather than political boundaries. The
scientific evidence has been evaluated to compare impacts across
(rather than within) sub-regions, although this was not always possible
depending on the scientific information available.
23.1.1. Scope and Route Map of Chapter
The chapter is structured around key policy areas. Sections 23.3 to 23.6
summarize the latest scientific evidence on sensitivity climate, observed
i
mpacts and attribution, projected impacts, and adaptation options, with
respect to four main categories of impacts:
Production systems and physical infrastructure
Agriculture, fisheries, forestry, and bioenergy production
Health protection and social welfare
Protection of environmental quality and biological conservation.
The benefit of assessing evidence in a regional chapter is that impacts
across sectors can be described, and interactions between impacts can
be identified. Further, the cross-sectoral decision making required to
address climate change can be reviewed. The chapter also includes
sections that were not in AR4. As adaptation and mitigation policy
develops, the evidence for potential co-benefits and unintended
consequences of such strategies is reviewed (Section 23.8). The final
section synthesizes the key findings with respect to: observed impacts
of climate change, key vulnerabilities, and research and knowledge gaps.
The chapter evaluates the scientific evidence in relation to the five sub-
regions highlighted above. The majority of the research in the Europe
region is for impacts in countries in the European Union due to targeted
research funding through the European Commission and national
governments, which means that countries in Eastern Europe and the
Russian Federation are less well represented in this chapter. Further,
regional assessments may be reported for the EU15, EU27, or EEA (32)
group of countries (Table SM23-1).
23.1.2. Policy Frameworks
Since AR4, there have been significant changes in Europe in responses
to climate change. More countries now have adaptation and mitigation
policies in place. An important force for climate policy development in the
region is the European Union (EU). EU member states have mitigation
targets, as well as the overall EU target, with both sectoral and regional
aspects to the commitments.
Adaptation policies and practices have been developed at international,
national, and local levels although research on implementation of such
policies is limited. Owing to the vast range of policies, strategies, and
measures it is not possible to describe them extensively here. However,
adaptation in relation to cross-sectoral decision making is discussed in
Section 23.7 (see also Box 23-3 on national adaptation policies). The
European Climate Adaptation Platform (Climate-ADAPT) catalogs
adaptation actions reported by EU Member States (EC, 2013a). The EU
Adaptation Strategy was adopted in 2013 (EC, 2013b). See Chapter 15
for a more extensive discussion of institutions and governance in
relation to adaptation planning and implementation.
23.1.3. Conclusions from Previous Assessments
AR4 documented a wide range of impacts of observed climate change
in Europe (WGII AR4 Chapter 12). The IPCC Special Report on Managing
the Risks of Extreme Events and Disasters to Advance Climate Change
Adaptation (SREX) confirmed increases in warm days and warm nights
and decreases in cold days and cold nights since 1950 (high confidence;
SREX Section 3.3.1). Extreme precipitation increased in part of the
Alpine Atlantic
Continental Northern
Southern
Figure 23-1 | Sub-regional classification of the IPCC Europe region. Based on
Metzger et al., 2005.
1275
Europe Chapter 23
23
c
ontinent, mainly in winter over Western-Central Europe and European
Russia (medium confidence; SREX Section 3.3.2). Dryness has increased
mainly in Southern Europe (medium confidence; SREX Section 3.3.2).
Climate change is expected to magnify regional differences within
Europe for agriculture and forestry because water stress was projected
to increase over Central and Southern Europe (WGII AR4 Section 12.4.1;
SREX Sections 3.3.2, 3.5.1). Many climate-related hazards were
projected to increase in frequency and intensity, but with significant
variations within the region (WGII AR4 Section12.4).
The AR4 identified that climate changes would pose challenges to many
economic sectors and was expected to alter the distribution of economic
activity within Europe (high confidence). Adaptation measures were
evolving from reactive disaster response to more proactive risk
management. A prominent example was the implementation of heat
health warning systems following the 2003 heat wave event (WGII AR4
Section 12.6.1; SREX Section 9.2.1). National adaptation plans were
developed and specific plans were incorporated in European and
national policies (WGII AR4 Sections 12.2.3, 12.5), but these were not
yet evaluated (WGII AR4 Section 12.8).
23.2. Current and Future Trends
23.2.1. Non-Climate Trends
European countries are diverse in both demographic and economic
trends. Population health and social welfare have improved everywhere
in Europe, with reductions in adult and child mortality rates, but social
inequalities both within and between countries persist (Marmot et al.,
2012). Population has increased in most EU27 countries, primarily as a
result of net immigration (Eurostat, 2011a), although population growth
is slow (total and working age population; Rees et al., 2012). Aging of
the population is a significant trend in Europe. This will have both
economic and social implications, with many regions experiencing a
decline in the labor force (Rees et al., 2012). Since AR4, economic
growth has slowed or become negative in many countries, leading to a
reduction in social protection measures and increased unemployment
(Eurostat, 2011b). The longer term implications of the financial crisis in
Europe are unclear, although it may lead to a modification of the economic
outlook and affect future social protection policies with implications for
adaptation.
Europe is one of the world’s largest and most productive suppliers of
food and fiber (Easterling et al., 2007). Agriculture is an important land
use across the European region; for example, it covers about 35% of
the total land area of western Europe (Rounsevell et al., 2006). After
1945, an unprecedented increase in agricultural productivity occurred,
but also declines in agricultural land use areas. This intensification had
several negative impacts on the ecological properties of agricultural
systems, such as carbon sequestration, nutrient cycling, soil structure and
functioning, water purification, and pollination. Pollution from agriculture
has led to eutrophication and declines in water quality in some areas
(Langmead et al., 2007). Most scenario studies suggest that agricultural
land areas will continue to decrease in the future (see also Busch,
2006, for a discussion). Agriculture accounts for 24% of total national
freshwater abstraction in Europe and more than 80% in some Southern
E
uropean countries (EEA, 2009). Economic restructuring in some
Eastern European countries has led to a decrease in water abstraction
for irrigation, suggesting the potential for future increases in irrigated
agriculture and water use efficiency (EEA, 2009).
Forest in Europe covers approximately 34% of the land area (Eurostat,
2009). The majority of forests now grow faster than in the early 20th
century as a result of advances in forest management practices, genetic
improvement, and, in Central Europe, the cessation of site-degrading
practices such as litter collection for fuel. Increasing temperatures and
carbon dioxide (CO
2
) concentrations, nitrogen deposition, and the
reduction of air pollution (sulfur dioxide (SO
2
)) have also had a positive
effect on forest growth. Scenario studies suggest that forested areas will
increase in Europe in the future on land formerly used for agriculture
(Rounsevell et al., 2006). Soil degradation is already intense in parts of
the Mediterranean and Central-Eastern Europe and, together with
prolonged drought periods and fires, is already contributing to an
increased risk of desertification. Projected risks for future desertification
are the highest in these areas (EEA, 2012).
Urban development is projected to increase all over Europe (Reginster
and Rounsevell, 2006), but especially rapidly in Eastern Europe, with the
magnitude of these increases depending on population growth, economic
growth, and land use planning policy. Although changes in urban land
use will be relatively small in area terms, urban development has major
impacts locally on environmental quality. Outdoor air quality has,
however, been improving (Langmead et al., 2007). Peri-urbanization is
an increasing trend in which residents move out of cities to locations
with a rural character, but retain a functional link to cities by commuting
to work (Reginster and Rounsevell, 2006; Rounsevell and Reay, 2009).
Several European scenario studies have been undertaken to describe
European future trends with respect to socioeconomic development (de
Mooij and Tang, 2003), land use change (Verburg et al., 2010; Haines-
Young et al., 2012; Letourneau et al., 2012), land use and biodiversity
(Spangenberg et al., 2011), crop production (Hermans et al., 2010),
demographic change (Davoudi et al., 2010), economic development
(Dammers, 2010), and European policy (Lennert and Robert, 2010;
Helming et al., 2011). Many of these scenarios also account for the
effects of future climate change (see Rounsevell and Metzger, 2010,
for a review). Long-term projections (to the end of the century) are
described under the new Shared Socioeconomic Pathway scenarios
(SSPs) (Kriegler et al., 2010). Detailed country and regional scale
socioeconomic scenarios have also been produced for the Netherlands
(WLO, 2006), the UK (UK National Ecosystem Assessment, 2011), and
Scotland (Harrison et al., 2013). The probabilistic representation of
socioeconomic futures has also been developed for agricultural land
use change (Hardacre et al., 2013). There is little evidence to suggest,
however, that probabilistic futures or scenarios more generally are being
used in policy making (Bryson et al., 2010).
23.2.2. Observed and Projected Climate Change
23.2.2.1. Observed Climate Change
Theaverage temperature in Europe has continued to increase, with
regionally and seasonally different rates of warming being greatest
1276
Chapter 23 Europe
23
i
n high latitudes in Northern Europe (Chapter 28). Since the 1980s,
warming has been strongest over Scandinavia, especially in winter,
whereas the Iberian Peninsula warmed mostly in summer (EEA, 2012).
The decadal average temperature over land area for 2002–2011 is 1.3°
± 0.11°C above the 1850–1899 average, based on Hadley Centre/Climatic
Research Unit gridded surface temperature data set 3 (HadCRUT3; Brohan
et al., 2006), Merged Land-Ocean Surface Temperature (MLOST; Smith
et al., 2008), and Goddard Institute of Space Studies (GISS) Temp (Hansen
et al., 2010). See WGI AR5 Section 2.4 for a discussion of data and
uncertainties and Chapter 21 for observed regional climate change.
Since 1950, high-temperature extremes (hot days, tropical nights, and
heat waves) have become more frequent, while low-temperature
extremes (cold spells, frost days) have become less frequent (WGI AR5
Section 2.6; SREX Chapter 3; EEA, 2012). The recent cold winters in
Northern and Atlantic Europe reflect the high natural variability in the
region (Peterson et al., 2012; see also WGI AR5 Section 2.7), and do not
contradict the general warming trend. In Eastern Europe, including the
European part of Russia, summer 2010 was exceptionally hot, with an
amplitude and spatial extent that exceeded the previous 2003 heat
wave (Barriopedro et al., 2011). Table 23-1 describes the impacts of
major extreme events in Europe in the last decade.
Since 1950, annual precipitation has increased in Northern Europe (up
to +70 mm per decade), and decreased in parts of Southern Europe (EEA,
2012, based on Haylock et al., 2008). Winter snow cover extent has a
high interannual variability and a nonsignificant negative trend over the
period 1967–2007 (Henderson and Leathers, 2010). Regional observed
changes in temperature and precipitation extremes are also described
in Table 3-2 of SREX and in Berg et al. (2013). Mean wind speeds have
declined over Europe over recent decades (Vautard et al., 2010) with
low confidence because of problematic anemometer data and climate
variability (SREX Section 3.3). Bett et al. (2013) did not find any trend
in windspeed using the Twentieth Century Reanalysis.
Europe is marked by increasing mean sea level with regional variations,
except in the northern Baltic Sea, where the relative sea level decreased
due to vertical crustal motion (Haigh et al., 2010; Menendez and Wood-
Worth, 2010; Albrecht et al., 2011; EEA, 2012). Extreme sea levels have
increased due to mean sea level rise (medium confidence; SREX Section
3.5; Haigh et al., 2010; Menendez and WoodWorth, 2010). Variability in
waves is related to internal climate variability rather than climate trends
(SREX Section 3.5; Charles et al., 2012).
23.2.2.2. Projected Climate Changes
Sub-regional information from global (see Chapter 21 supplementary
material; see also WGI AR5 Section 14.8.6, Annex I) and regional high-
resolution climate model output (Chapters 21, 23; see also WGI AR5
Section 14.8.6) provide more knowledge about the range of possible
future climates under the Special Report on Emissions Scenarios (SRES)
and Representative Concentration Pathway (RCP) emission scenarios.
Within the recognized limitations of climate projections (Chapter 21;
WGI AR5 Chapter 9), new research on inter-model comparisons has
provided a more robust range of future climates to assess future impacts.
Since AR4, climate impact assessments are more likely to use a range
f
or the projected changes in temperature and rainfall. Access to
comprehensive and detailed sets of climate projections for decision
making exist in Europe (SREX Section 3.2.1; Mitchell et al., 2004; Fronzek
et al., 2012; Jacob et al., 2013).
Climate models show significant agreement for all emission scenarios in
warming (magnitude and rate) all over Europe, with strongest warming
projected in Southern Europe in summer, and in Northern Europe in winter
(Goodess et al., 2009; Kjellström et al., 2011). Even under an average
global temperature increase limited to 2°C compared to preindustrial
times, the climate of Europe is simulated to depart significantly in the
next decades from today’s climate (Van der Linden and Mitchell, 2009;
Jacob and Podzun, 2010).
Precipitation signals vary regionally and seasonally. Trends are less clear
in Continental Europe, with agreement in increase in Northern Europe
and decrease in Southern Europe (medium confidence; Kjellström et al.,
2011). Precipitation is projected to decrease in the summer months up
to southern Sweden and increase in winter (Schmidli et al., 2007), with
more rain than snow in mountainous regions (Steger et al., 2013). In
Northern Europe, a decrease of long-term mean snowpack (although
snow-rich winters will remain) toward the end of the 21st century
(Räisänen and Eklund, 2012) is projected. There is lack of information
about past and future changes in hail occurrence in Europe. Changes
in future circulation patterns (Ulbrich et al., 2009; Kreienkamp et al.,
2010) and mean wind speed trends are uncertain in sign (Kjellström et
al., 2011; McInnes et al., 2011).
Regional coupled simulations over the Mediterranean region provide a
more realistic characterization of impact parameters (e.g., snow cover,
aridity index, river discharge), which were not revealed by Coupled
Model Intercomparison Project Phase 3 (CMIP3) global simulations
(Dell’Aquila et al., 2012).
For 2081–2100 compared to 1986–2005, projected global mean sea
level rises (meters) are in the range 0.29 to 0.55 for RCP2.6, 0.36 to
0.63 for RCP4.5, 0.37 to 0.64 for RCP6.0, and 0.48 to 0.82 for RCP8.5
(medium confidence; WGIII AR5 Chapter 5). There is a low confidence
on projected regional changes (Slangen et al., 2012; WGI AR5 Section
13.6). Low-probability/high-impact estimates of extreme mean sea
level rise projections derived from the SRES A1FI scenario for the
Netherlands (Katsman et al., 2011) indicate that the mean sea level
could rise globally between 0.55 and 1.15 m, and locally (Netherlands)
by 0.40 to 1.05 m, by 2100. Extreme (very unlikely) scenarios for the
UK vary from 0.9 to 1.9 m by 2100 (Lowe et al., 2009).
23.2.2.3. Projected Changes in Climate Extremes
There will be a marked increase in extremes in Europe, in particular, in
heat waves, droughts, and heavy precipitation events (Beniston et al.,
2007; Lenderink and Van Meijgaard, 2008; see also Chapter 21
supplementary material). There is a general high confidence concerning
changes in temperature extremes (toward increased number of warm
days, warm nights, and heat waves; SREX Table 3-3). Figure 23-2c
shows projected changes in the mean number of heat waves in May to
September for 2071–2100 compared to 1971–2000 for RCP4.5 and
1277
Europe Chapter 23
23
Significant change
Robust change
Seasonal changes in heavy
precipitation in percent
(a) DJF seasonal changes in heavy precipitation (%), 2071–2100 compared to 1971–2000
(b) JJA seasonal changes in heavy precipitation (%), 2071–2100 compared to 1971–2000
RCP4.5 RCP8.5
RCP4.5
RCP8.5
Continue
d
next page
Continued
next
page
Figure 23-2 | (a) and (b): Projected seasonal changes in heavy precipitation defined as the 95th percentile of daily precipitation (only days with precipitation >1 mm day
–1
are
considered) for the period 2071–2100 compared to 1971–2000 (in %) in the months December to February (DJF) and June to August (JJA). (c) Projected changes in the mean
number of heat waves occurring in the months May to September for the period 2071–2100 compared to 1971–2000 (number per 30 years). Heat waves are defined as periods
of more than 5 consecutive days with daily maximum temperature exceeding the mean maximum temperature of the May to September season of the control period
(1971–2000) by at least 5°C. (d) Projected changes in the 95th percentile of the length of dry spells for the period 2071–2100 compared to 1971–2000 (in days). Dry spells are
defined as periods of at least 5 consecutive days with daily precipitation below 1 mm. Hatched areas indicate regions with robust (at least 66% of models agree in the sign of
change) and/or statistically significant change (significant on a 95% confidence level using Mann–Whitney U test). For the eastern parts of Black Sea, eastern Anatolia, and
southeast Anatolia (Turkey), no regional climate model projections are available. Changes represent the mean over 8 (RCP4.5, left side) and 9 (RCP8.5, right side) regional model
simulations compiled within the Coordinated Downscaling Experiment – European Domain (EURO-CORDEX) initiative. Adapted from Jacob et al., 2013.
–25
–15 –5
5
15
25 35 45
1278
Chapter 23 Europe
23
RCP4.5 RCP8.5
–1 1 2 3 4 5 6 7 8 9
Changes in mean number
of heat waves
–4 –2 –1 1 2 4 8 16 24 32
Changes in the 95th percentile of
the length of dry spells in days
(c) Changes in mean number of heat waves for MJJAS, 2071–2100 compared to 1971–2000
(d) Changes in the 95th percentile of the length of dry spells (days) 2071–2100 compared to 1971–2000
RCP4.5 RCP8.5
Significant change
Robust change
Significant change
Robust change
F
igure 23-2 (continued)
1279
Europe Chapter 23
23
R
CP8.5 with large differences depending on the emission scenario. The
increase in likelihood of some individual events due to anthropogenic
change has been quantified for the 2003 heat wave (Schär and
Jendritzky, 2004), the warm winter of 2006/2007, and warm spring of
2007 (Beniston, 2007).
Changes in extreme precipitation depend on the region, with a high
confidence of increased extreme precipitation in Northern Europe (all
seasons) and Continental Europe (except summer). Future projections
are regionally and seasonally different in Southern Europe (SREX Table
3-3). Figure 23-2a,b shows projected seasonal changes of heavy
precipitation events for 2071–2100 compared to 1971–2000 for RCP4.5
and RCP8.5.
Projected changes of spatially averaged indices over the European sub-
regions are described in the supplemental information (Tables SM23-2
and SM23-3 for sub-regions, and Table SM23-4 for three Alpine areas).
In winter, small increases in extreme wind speed are projected for
Central and Northern Europe (medium confidence; Section 21.3.3.1.6;
SREX Figure 3-8; Beniston et al., 2007; Rockel and Woth, 2007; Haugen
and Iversen, 2008; Rauthe et al., 2010; Schwierz et al., 2010), connected
to changes in storm tracks (medium confidence; Pinto et al., 2007a,b,
2010; Donat et al., 2010). Other parts of Europe and seasons are less
clear in sign with a small decreasing trend in Southern Europe (low
confidence; Donat et al., 2011; McInnes et al., 2011).
Extreme sea level events will increase (high confidence; WGI AR5 Section
13.7; SREX Section 3.5.3), mainly dominated by the global mean sea
level increase. Storm surges are expected to vary along the European
coasts. Significant increases are projected in the eastern North Sea
(increase of 6 to 8% of the 99th percentile of the storm surge residual,
2071–2100 compared to 1961–1990, based on the B2, A1B, and A2
SRES scenarios; Debernard and Rÿed, 2008) and west of UK and Ireland
(Debernard and Rÿed, 2008; Wang et al., 2008), except south of Ireland
(Wang et al., 2008). There is a medium agreement for the south of North
Sea and Dutch coast where trends vary from increasing (Debernard and
Rÿed, 2008) to stable (Sterl et al., 2009). There is a low agreement on
the trends in storm surge in the Adriatic Sea (Planton et al., 2006; Jordà
et al., 2012; Lionello et al., 2012; Troccoli et al., 2012b).
23.2.3. Observed and Projected Trends
in Riverflow and Drought
Streamflows have decreased in the south and east of Europe and
increased in Northern Europe (Stahl et al., 2010; Wilson et al., 2010; see
also Section 3.2.3). In general, few changes in flood trends can be
attributed to climate change, partly owing to the lack of sufficiently long
records (Kundzewicz et al., 2013). European mean and peak discharges
are highly variable (Bouwer et al., 2008); for instance, in France, upward
trends in low flows were observed over 1948–1988 and downward
trends over 1968–2008 (Giuntoli et al., 2013). Alpine glacier retreat
during the last 2 decades caused a 13% increase in glacier contribution
to August runoff of the four main rivers originating in the Alps, compared
to the long-term average (Huss, 2011). Increases in extreme river
discharge (peak flows) over the past 30 to 50 years have been observed
i
n parts of Germany (Petrow et al., 2007, 2009), the Meuse River basin
(Tu et al., 2005), parts of Central Europe (Villarini et al., 2011), Russia
(Semenov, 2011), and northeastern France (Renard et al., 2008).
Decreases in extreme river discharge have been observed in the Czech
Republic (Yiou et al., 2006), and no change observed in Switzerland
(Schmocker-Fackel and Naef, 2010), Germany (Bormann et al., 2011),
and the Nordic countries (Wilson et al., 2010). River regulation possibly
partly masks increasing peak flows in the Rhine (Vorogushyn et al.,
2012). One study (Pall et al., 2011) suggested that the UK 2000 flood
was partly due to anthropogenic forcing, although another showed a
weaker effect (Kay et al., 2011).
Climate change is projected to affect the hydrology of river basins
(Chapter 4; SREX Chapter 3). The occurrence of current 100-year return
period discharges is projected to increase in Continental Europe, but
decrease in some parts of Northern and Southern Europe by 2100 (Dankers
and Feyen, 2008; Rojas et al., 2012). In contrast, studies for individual
catchments indicate increases in extreme discharges, to varying degrees,
in Finland (Veijalainen et al., 2010), Denmark (Thodsen, 2007), Ireland
(Wang et al., 2006; Steele-Dunne et al., 2008; Bastola et al., 2011), the
Rhine basin (Görgen et al., 2010; te Linde et al., 2010a), Meuse basin
(Leander et al., 2008; Ward et al., 2011), the Danube basin (Dankers et
al., 2007), and France (Quintana-Segui et al., 2011; Chauveau et al.,
2013). Although snowmelt floods may decrease, increased autumn and
winter rainfall could lead to higher peak discharges in Northern Europe
(Lawrence and Hisdal, 2011). Declines in low flows are projected for the
UK (Christierson et al., 2012), Turkey (Fujihara et al., 2008), France
(Chauveau et al., 2013), and rivers fed by Alpine glaciers (Huss, 2011).
The analysis of trends in droughts is made complex by the different
categories or definitions of drought (meteorological, agricultural, and
hydrological) and the lack of long-term observational data (SREX Box
3-3). Southern Europe shows trends toward more intense and longer
meteorological droughts, but they are still inconsistent (Sousa et al., 2011).
Drought trends in all other sub-regions are not statistically significant
(SREX Section 3.5.1). Regional and global climate simulations project
(medium confidence) an increase in duration and intensity of droughts in
Central and Southern Europe and the Mediterranean up until the UK for
different definitions of drought (Gao and Giorgi, 2008; Feyen and Dankers,
2009; Vidal and Wade, 2009; Koutroulis et al., 2010; Tsanis et al., 2011;
Chapter 21). Even in regions where summer precipitation is expected
to increase, soil moisture and hydrological droughts may become more
severe as a result of increasing evapotranspiration (Wong et al., 2011).
Projected changes in the length of meteorological dry spells show that
the increase is large in Southern Europe (Figure 23-2d).
23.3. Implications of Climate Change
for Production Systems and
Physical Infrastructure
23.3.1. Settlements
23.3.1.1. Coastal Flooding
As the risk of extreme sea level events increases with climate change
(Section 23.2.3; Chapter 5), coastal flood risk will remain a key challenge
1280
Chapter 23 Europe
23
f
or several European cities, port facilities, and other infrastructure
(Hallegatte et al., 2008, 2011; Nicholls et al., 2008). With no adaptation,
coastal flooding in the 2080s is projected to affect an additional 775,000
and 5.5 million people per year in the EU27 (B2 and A2 scenarios,
respectively; Ciscar et al., 2011). The Atlantic, Northern, and Southern
European regions are projected to be most affected. Direct costs from
sea level rise in the EU27 without adaptation could reach €17 billion per
year by 2100 (Hinkel et al., 2010), with indirect costs also estimated for
land-locked countries (Bosello et al., 2012). Countries with high absolute
damage costs include Netherlands, Germany, France, Belgium, Denmark,
Spain, and Italy (Hinkel et al., 2010). Upgrading coastal defenses would
substantially reduce impacts and damage costs (Hinkel et al., 2010).
However, the amount of assets and populations that need to be protected
by coastal defenses is increasing; thus, the magnitude of losses when
floods do occur will also increase in the future (Hallegatte et al., 2013).
An increase in future flood losses due to climate change have been
estimated for Copenhagen (Hallegatte et al., 2011), UK coast (Mokrech
et al., 2008; Purvis et al., 2008; Dawson et al., 2011), the North Sea
coast (Gaslikova et al., 2011), cities including Amsterdam and Rotterdam
(Hanson et al., 2011), and the Netherlands (Aerts et al., 2008). A 1 m sea
level rise in Turkey could affect 3 million additional people and put
US$12 billion capital value at risk, with around US$20 billion adaptation
costs (10% of GNP; Karaca and Nicholls, 2008). In Poland, up to 240,000
p
eople would be affected by increasing flood risk on the Baltic coast
(Pruszak and Zawadzka, 2008). The increasing cost of insurance and
unwillingness of investors to place assets in affected areas is a potential
growth impediment to coastal and island economies (Day et al., 2008).
23.3.1.2. River and Pluvial Flooding
Recent major flood events in Europe include the 2007 floods in the UK
(Table 23-1; Chatterton et al., 2010) and the 2013 floods in Germany.
The observed increase in river flood events and damages in Europe is
well documented (see Section 18.4.2.1); however, the main cause is
increased exposure of persons and property in flood risk areas (Barredo,
2009). Since AR4, new studies provide a wider range of estimates of
future economic losses from river flooding attributable to climate
change, depending on the modeling approach and climate scenario
(Bubeck et al., 2011). Studies now also quantify risk under changes in
population and economic growth, generally indicating this contribution
to be about equal or larger than climate change per se (Feyen et al.,
2009; Maaskant et al., 2009; Bouwer et al., 2010; Rojas et al., 2013; te
Linde et al., 2011). Some regions may see increasing risks, but others
may see decreases or little to no change (ABI, 2009; Feyen et al., 2009,
2012; Lugeri et al., 2010; Mechler et al., 2010; Bubeck et al., 2011; Lung
et al., 2012). In the EU15, river flooding could affect 250,000 to 400,000
1281
Europe Chapter 23
23
a
dditional people by the 2080s (SRES A2 and B2 scenarios, respectively)
more than doubling annual average damages, with Central and Northern
Europe and the UK most affected (Ciscar, 2009; Ciscar et al., 2011). When
economic growth is included, economic flood losses in Europe could
increase 17-fold under the A1B climate scenario (Rojas et al., 2013).
Few studies have estimated future damages from inundation in response
to an increase in intense rainfall (Hoes, 2006; Willems et al., 2012).
Processes that influence flash flood risk include increasing exposure from
urban expansion, and forest fires that lead to erosion and increased
surface runoff (Lasda et al., 2010). Some studies have costed adaptation
measures but these may only partly offset anticipated impacts (Zhou et
al., 2012).
23.3.1.3. Windstorms
Several studies project an overall increase in storm hazard in northwest
Europe (Section 23.2.2.3) and in economic and insured losses (Section
17.7), but natural variations in frequencies are large. There is no
evidence that the observed increase in European storm losses is due to
anthropogenic climate change (Barredo, 2010). There is a lack of
information for other storm types, such as tornadoes and thunderstorms.
23.3.1.4. Mass Movements and Avalanches
In the European Alps, the frequency of rock avalanches and large rock
slides has apparently increased over the period 1900–2007 (Fischer et
al., 2012). The frequency of landslides may also have increased in some
locations (Lopez Saez et al., 2013). Mass movements are projected to
become more frequent with climate change (Huggel et al., 2010; Stoffel
and Huggel, 2012), although several studies indicate a more complex or
stabilizing response of mass movements to climate change (Dixon and
Brook, 2007; Jomelli et al., 2007, 2009; Huggel et al., 2012; Melchiorre
and Frattini, 2012). Some land use practices have led to conditions
favorable to increased landslide risk, despite climate trends that would
result in a decrease of landslide frequency, as reported in Calabria
(Polemio and Petrucci, 2010) and in the Apennines (Wasowski et al.,
2010). Snow avalanche frequency changes in Europe are dominated by
climate variability; studies based on avalanche observations (Eckert et
al., 2010) or favorable meteorological conditions (Castebrunet et al.,
2012; Teich et al., 2012) show contrasting variations, depending on the
region, elevation, season, and orientation.
23.3.2. Built Environment
Built infrastructure in Europe is vulnerable to extreme weather events,
including overheating of buildings (houses, hospitals, schools) during hot
weather (Crump et al., 2009; DCLG, 2012). Buildings that were originally
designed for certain thermal conditions will need to function in warmer
climates in the future (WHO, 2008). Climate change in Europe is
expected to increase cooling energy demand (Dolinar et al., 2010; see
also Section 23.3.4), with implications for mitigation and adaptation
policies (Section 23.8.1). A range of adaptive strategies for buildings
are available, including effective thermal mass and solar shading
(
Three Regions Climate Change Group, 2008). Climate change may also
increase the frequency and intensity of drought-induced soil subsidence
and associated damage to dwellings (Corti et al., 2009).
With respect to the outdoor built environment, there is limited evidence
regarding the potential for differential rates of radiatively forced climate
change in urban compared to rural areas (McCarthy et al., 2010). Climate
change may exacerbate London’s nocturnal urban heat island (UHI)
(Wilby, 2008); however, the response of different cities may vary. For
example, a study of Paris (Lemonsu et al., 2013) indicated a future
reduction in strong urban heat island events when increased soil
dryness was taken into effect. Modification of the built environment,
via enhanced urban greening, for example, can reduce temperatures in
urban areas, with co-benefits for health and well-being (Sections 23.7.4,
23.8.1).
23.3.3. Transport
Systematic and detailed knowledge on climate change impacts on
transport in Europe remains limited (Koetse and Rietveld, 2009).
On road transport, in line with AR4, more frequent but less severe
collisions due to reduced speed are expected in case of increased
precipitation (Kilpeläinen and Summala, 2007; Brijs et al., 2008).
However, lower traffic speed may cause welfare losses due to additional
time spent driving (Sabir et al., 2010). Severe snow and ice-related
accidents will also decrease, but the effect of fewer frost days on total
accidents is unclear (Andersson and Chapman, 2011a,b). Severe accidents
caused by extreme weather are projected to decrease by 63 to 70% in
2040–2070 compared to 2007 as a result of modified climate and
expected developments in vehicle technology and emergency systems
(Nokkala et al., 2012).
For rail, consistent with AR4, increased buckling in summer, as occurred
in 2003 in the UK, is expected to increase the average annual cost of
heat-related delays in some regions, while the opposite is expected for
ice and snow-related delays (Lindgren et al., 2009; Dobney et al., 2010;
Palin et al., 2013). Effects from extreme precipitation, as well as the net
overall regional impact of climate change remain unclear. Efficient
adaptation comprises proper maintenance of track and track bed.
Regarding inland waterways, the case of Rhine shows that, for 1°C to
2°C increases by 2050, more frequent high water levels are expected
in winter, while after 2050 days with low water levels in summer will
also increase (te Linde, 2007; Hurkmans et al., 2010; Jonkeren et al.,
2011; te Linde et al., 2011). Low water levels will reduce the load factor
of inland ships and consequently increase transport prices, as in the
Rhine and Moselle in 2003 (Jonkeren et al., 2007; Jonkeren, 2009).
Adaptation includes modal shifts, increased navigational hours per
day under low water levels, and infrastructure modifications (e.g.,
canalization of river parts) (Jonkeren et al., 2011; Krekt et al., 2011).
For long range ocean routes, the economic attractiveness of the
Northwest Passage and the Northern Sea Route depends also on passage
fees, bunker prices, and cost of alternative sea routes (Verny and
Grigentin, 2009; Liu and Kronbak, 2010; Lasserre and Pelletier, 2011).
1282
Chapter 23 Europe
23
R
egarding air transport, for Heathrow airport (UK), future temperature
and wind changes were estimated to cause a small net annual increase
but much larger seasonal changes on the occurrence of delays (Pejovic
et al., 2009).
23.3.4. Energy Production, Transmission, and Use
On wind energy, no significant changes are expected before 2050, at
least in Northern Europe (Pryor and Barthelmie, 2010; Pryor and Schoof,
2010; Seljom et al., 2011; Barstad et al., 2012; Hueging et al., 2013).
After 2050, in line with AR4, the wind energy potential in Northern,
Continental, and most of Atlantic Europe may increase during winter and
decrease in summer (Rockel and Woth, 2007; Harrison et al., 2008; Nolan
et al., 2012; Hueging et al., 2013). For Southern Europe, a decrease in
both seasons is expected, except for the Aegean Sea and Adriatic coast,
where a significant increase during summer is possible (Bloom et al.,
2008; Najac et al., 2011; Pašičko et al., 2012; Hueging et al., 2013).
For hydropower, electricity production in Scandinavia is expected to
increase by 5 to 14% during 2071–2100 compared to historic or present
levels (Haddeland et al., 2011; Golombek et al., 2012); for 2021–2050,
increases by 1 to 20% were estimated (Haddeland et al., 2011; Seljom
et al., 2011; Hamududu and Killingtveit, 2012). In Continental and part
of Alpine Europe, reductions in electricity production by 6 to 36% were
estimated (Schaefli et al., 2007; Stanzel and Nachtnebel, 2010; Paiva
et al., 2011; Pašičko et al., 2012; Hendrickx and Sauquet, 2013). For
Southern Europe, production is expected to decrease by 5 to 15% in
2050 compared to 2005 (Hamududu and Killingtveit, 2012; Bangash
et al., 2013). Adaptation consists of improved water management,
including pump storage if appropriate (Schaefli et al., 2007; García-Ruiz
et al., 2011).
B
iofuel production is discussed in Section 23.4.5. There are few studies
of impacts on solar energy production. Crook et al. (2011) estimated an
increase of the energy output from photovoltaic panels and especially
from concentrated solar power plants in most of Europe under the A1B
scenario.
On thermal power, in line with AR4, van Vliet et al. (2012) estimated a
6 to 19% decrease of the summer average usable capacity of power
plants by 2031–2060 compared to 1971–2000, while smaller decreases
have been also estimated (Förster and Lilliestam, 2010; Linnerud et al.,
2011). Closed-cooling circuits are efficient adaptation choices for new
plants (Koch and Vögele, 2009). In power transmission, increasing
lightning and decreasing snow-sleet and blizzard faults for 2050–2080
were estimated for the UK (McColl et al., 2012).
By considering both heating and cooling, under a +3.7°C scenario by
2100 a decrease of total annual energy demand in Europe as a whole
during 2000–2100 was estimated (Isaac and van Vuuren, 2009).
Seasonal changes will be prominent, especially for electricity (see Figure
23-3), with summer peaks arising also in countries with moderate
summer temperatures (Hekkenberg et al., 2009). Heating degree days
are expected to decrease by 11 to 20% between 2000 and 2050 due
solely to climate change (Isaac and van Vuuren, 2009). For cooling, very
large percentage increases up to 2050 are estimated by the same
authors for most of Europe as the current penetration of cooling
devices is low; then, increases by 74 to 118% in 2100 (depending on
the region) from 2050 are expected under the combined effect of
climatic and non-climatic drivers. In Southern Europe, cooling degree
days by 2060 will increase, while heating degree days will decrease but
with substantial spatial variations (Giannakopoulos et al., 2009).
Consequently, net annual electricity generation cost will increase in most
of the Mediterranean and decrease in the rest of Europe (Mirasgedis
–10
–5
0
5
10
15
20
25
DecNovOctSeptAugJulyJuneMayAprMarFebJan
% Change in electricity demand due to climate change
Climate B2 / Economy B2,
convergence with OECD-2100
Climate B2 / Economy B2,
convergence with OECD-2070
Climate B2 / Economy B2,
OECD general trend
Climate A2 / Economy A2,
convergence with OECD-2100
Climate A2 / Economy A2,
convergence with OECD-2070
Climate A2 / Economy A2,
OECD general trend
Figure 23-3 | Percentage change in electricity demand in Greece attributable to climate change, under a range of climate scenarios and economic assumptions. Source:
Mirasgedis et al., 2007.
1283
Europe Chapter 23
23
e
t al., 2007; Eskeland and Mideksa, 2010; Pilli-Sihlova et al., 2010;
Zachariadis, 2010).
Future building stock changes and retrofit rates are critical for impact
assessment and adaptation (Olonscheck et al., 2011). Energy-efficient
buildings and cooling systems, and demand-side management, are
effective adaptation options (Artmann et al., 2008; Jenkins et al., 2008;
Day et al., 2009; Breesch and Janssens, 2010; Chow and Levermore,
2010).
23.3.5. Industry and Manufacturing
Research on the potential effects of climate change in industry is limited.
Modifications in future consumption of food and beverage products
have been estimated on the basis of current sensitivity to seasonal
temperature (Mirasgedis et al., 2013). Higher temperatures may favor
the growth of food-borne pathogens or contaminants (Jacxsens et al.,
2010; Popov Janevska et al., 2010; see also Section 23.5.1). The quality
of some products, such as wine (Section 23.4.1; Box 23-2), is also likely
to be affected. In other sectors, the cumulative cost of direct climate
change impacts in the Greek mining sector for 2021–2050 has been
estimated at €0.245 billion, in 2010 prices (Damigos, 2012). Adaptation
to buildings or work practices are likely to be needed to maintain labor
productivity during hot weather (Kjellstrom et al., 2009; see also Section
11.6.2.2).
23.3.6. Tourism
In line with AR4, the climate for general tourist activities especially after
2070 is expected to improve significantly during summer and less during
autumn and spring in northern Continental Europe, Finland, southern
Scandinavia, and southern England (Amelung et al., 2007; Nicholls and
Amelung, 2008; Amelung and Moreno, 2012). For the Mediterranean,
climatic conditions for light outdoor tourist activities are expected to
deteriorate in summer mainly after 2050, but improve during spring and
autumn (Amelung et al., 2007; Amelung and Moreno, 2009; Hein et al.,
2009; Perch-Nielsen et al., 2010; Giannakopoulos et al., 2011). Others
concluded that before 2030 (or even 2060) this region as a whole will
not become too hot for beach or urban tourism (Moreno and Amelung,
2009; Rutty and Scott, 2010), while surveys showed that beach tourists
are deterred mostly by rain (De Freitas et al., 2008; Moreno, 2010).
Thus, from 2050, domestic tourism and tourist arrivals at locations in
Northern and parts of Continental Europe may be enhanced at the
expense of southern locations (Hamilton and Tol, 2007; Hein et al., 2009;
Amelung and Moreno, 2012; Bujosa and Roselló, 2012). The age of
tourists, the climate in their home country, and local economic and
environmental conditions (e.g., water stress, tourist development) are
also critical (Hamilton and Tol, 2007; Lyons et al., 2009; Moreno and
Amelung, 2009; Rico-Amoros et al., 2009; Eugenio-Martin and Campos-
Soria, 2010; Perch-Nielsen et al., 2010).
Tourism in mountainous areas may benefit from improved climatic
conditions in summer (Endler et al., 2010; Perch-Nielsen et al., 2010;
Endler and Matzarakis, 2011; Serquet and Rebetez, 2011). However, in
a
greement with AR4, natural snow reliability and thus ski season length
will be adversely affected, especially where artificial snowmaking is
limited (Moen and Fredman, 2007; OECD, 2007; Steiger, 2011). Low-
lying areas will be the most vulnerable (Uhlmann et al., 2009; Endler et
al., 2010; Endler and Matzarakis, 2011; Serquet and Rebetez, 2011;
Steiger, 2011). Tourist response to marginal snow conditions remains
largely unknown, while changes in weather extremes may also be critical
(Tervo, 2008). Up to 2050, demographic changes (e.g., population
declines in source countries, aging populations) may have a higher
impact than climate change (Steiger, 2012). Artificial snowmaking has
physical and economic limitations, especially in small sized and low-
altitude ski stations (Steiger and Mayer, 2008; Sauter et al., 2010; Steiger,
2010, 2011), and increases water and energy consumption. Shifts to
higher altitudes, operational/ technical measures, and year-round tourist
activities may not fully compensate for adverse impacts.
23.3.7. Insurance and Banking
Insurance and banking face problems related to accurate pricing of risks,
shortage of capital after large loss events, and by an increasing burden
of losses that can affect markets and insurability, within but also outside
the European region (CEA, 2007; Botzen et al., 2010a,b; see also Section
10.7). However, risk transfer, including insurance, also holds potential
for adaptation by providing incentives to reduce losses (Botzen and van
den Bergh, 2008; CEA, 2009; Herweijer et al., 2009).
Banking is potentially affected through physical impacts on assets and
investments, as well as through regulation and/or mitigation actions by
changing demands regarding sustainability of investments and lending
portfolios. Few banks have adopted climate strategies that also address
adaptation (Cogan, 2008; Furrer et al., 2009).
Windstorm losses are well covered in Europe by building and motor
policies, and thus create a large exposure to the insurance sector. Flood
losses in the UK in 2000, 2007, and 2009 have put the insurance market
under further pressure, with increasing need for the government to reduce
risk (Ward et al., 2008; Lamond et al., 2009). Other risks of concern to
the European insurance industry is building subsidence related to
drought (Corti et al., 2009), and hail damage to buildings and agriculture
(Kunz et al., 2009; Botzen et al., 2010b; GDV, 2011).
The financial sector can adapt by adjusting premiums, restricting or
reducing coverage, spreading risk further, and importantly incentivizing
risk reduction (Crichton, 2006, 2007; Clemo, 2008; Botzen et al., 2010a;
Surminski and Philp, 2010; Wamsler and Lawson, 2011). Public attitudes
in Scotland and the Netherlands would support insurance of private
property and public infrastructure damages in the case of increasing
flood risk (Botzen et al., 2009; Glenk and Fisher, 2010). Government
intervention is, however, often needed to provide compensation and
back-stopping in the event of major losses (Aakre and Rübbelke, 2010;
Aakre et al., 2010). Hochrainer et al. (2010) analyzed the performance of
the European Union Solidarity Fund that supports European governments
in large events, and argue there is a need to increase its focus on risk
reduction. Current insurance approaches present in Europe are likely to
remain, as they are tailored to local situations and preferences (Schwarze
et al., 2011).
1284
Chapter 23 Europe
23
23.4. Implications of Climate Change for
Agriculture, Fisheries, Forestry,
and Bioenergy Production
23.4.1. Plant (Food) Production
In AR4, Alcamo et al. (2007) reported that crop suitability is likely to
change throughout Europe. During the 2003 and 2010 summer heat
waves, grain-harvest losses reached 20 and 25-30% in affected regions
of Europe and Russia, respectively (Ciais et al., 2005; Barriopedro et al.,
2011; see also Table 23-1). Cereals production fell on average by 40%
in the Iberian Peninsula during the intense 2004/2005 drought (EEA,
2010a). Climate-induced variability in wheat production has increased
in recent decades in Southern and Central Europe (Ladanyi, 2008;
Brisson et al., 2010; Hawkins et al., 2013), but no consistent reduction
has been recorded in the northernmost areas of Europe (Peltonen-Sainio
et al., 2010). Country-scale rainfed cereals yields are below agro-climatic
potentials (Supit et al., 2010), and wheat yield increases have leveled
off in several countries over 1961–2009 (Olesen et al., 2011). High
temperatures and droughts during grain filling have contributed to the
lack of yield increase of winter wheat in France despite improvements
in crop breeding (Brisson et al., 2010; Kristensen et al., 2011). In contrast,
in eastern Scotland, warming has favored an increase in potato yields
since 1960 (Gregory and Marshall, 2012). In northeast Spain, grape yield
was reduced by an increased water deficit in the reproductive stage
since the 1960s (Camps and Ramos, 2012).
Insight into the potential effect of climate change on crops requires the
combination of a wide range of emission scenarios, Global Climate
Models (GCMs), and impact studies (Trnka et al., 2007; Soussana et al.,
2010). In the EU27, a 2.5°C regional temperature increase in the 2080s
under the B2 scenario could lead to small changes (on average +3%)
in crop yields, whereas a 5.4°C regional warming under the A2 scenario
could reduce mean yields by 10% according to a study based on
regional climate models (Ciscar et al., 2011). An initial benefit from the
increasing CO
2
concentration for rainfed crop yields would contrast by
the end of the century with yield declines in most European sub-regions,
although wheat yield could increase under the A2 scenario (three GCMs,
B1, A2 scenarios; Supit et al., 2012). Disease-limited yields of rainfed
wheat and maize in the 2030s does not show consistent trends across
two GCMs (Donatelli et al., 2012). For a global temperature increase of
5°C, agroclimatic indices show an increasing frequency of extremely
unfavorable years in European cropping areas (Trnka et al., 2011). Under
the A2 and B2 scenarios, crop production shortfalls, defined as years
with production below 50% of its average climate normal production
would double by 2020 and triple by 2070 as compared to a current
frequency of 1 to 3 years per decade in the currently most productive
southern European regions of Russia (Alcamo et al., 2007).
The regional distribution of climate change impacts on agricultural
production is likely to vary widely (Donatelli et al., 2012; Iglesias et al.,
2012; see also Figure 23-4). Southern Europe would experience the
largest yield losses (–25% by 2080 under a 5.4°C warming; Ciscar et
al., 2011), with increased risks of rainfed summer crop failure (Ferrara
et al., 2010; Bindi and Olesen, 2011; Ruiz-Ramos et al., 2011). Warmer
and drier conditions by 2050 (Trnka et al., 2010, 2011) would cause
moderate declines in crop yields in Central Europe regions (Ciscar et al.,
2
011). In Western Europe, increased heat stress around flowering could
cause considerable yield losses in wheat (Semenov, 2009). For Northern
Europe, there is diverging evidence concerning future impacts. Positive
yield changes combined with the expansion of climatically suitable
areas could lead to crop production increases (between 2.5°C and 5.4°C
regional warming) (Bindi and Olesen, 2011). However, increased climatic
variability would limit winter crops expansion (Peltonen-Sainio et al.,
2010) and cause at high latitudes high risk of marked cereal yield loss
(Rötter et al., 2011). Spring crops from tropical origin like maize for
silage could become cultivated in Finland by the end of the century
(Peltonen-Sainio et al., 2009). Cereal yield reduction from ozone (Fuhrer,
2009) could reach 6 and 10 % in 2030 for the European Union with the
B1 and A2 scenarios, respectively (Avnery et al., 2011a,b). Because of
limited land availability and soil fertility outside of Chernozem (black
earth) areas, the shift of agriculture to the boreal forest zone would not
compensate for crop losses owing to increasing aridity in South European
regions of Russia with the best soils (Dronin and Kirilenko, 2011).
With generally warmer and drier conditions, deep rooted weeds (Gilgen
et al., 2010) and weeds with contrasting physiology, such as C
4
species,
could pose a more serious threat (Bradley et al., 2010) to crops than
shallow rooted C
3
weeds (Stratonovitch, 2012). Arthropod-borne diseases
(viruses and phytoplasmas), winter infection root and stem diseases
(phoma stem canker of oilseed rape and eyespot of wheat; Butterworth
et al., 2010; West et al., 2012), Fusarium blight (Madgwick et al., 2011),
grapevine moth (Caffarra et al., 2012), and a black rot fungus in fruit
trees (Weber, 2009) could create increasing damages in Europe under
climate change. However, other pathogens such as cereal stem rots (e.g.,
Puccinia striiformis; Luck et al., 2011) and grapevine powdery mildew
(Caffarra et al., 2012) could be limited by increasing temperatures.
Increased damages from plant pathogens and insect pests are projected
by 2050 in Nordic countries, which have hitherto been protected by cold
winters and geographic isolation (Hakala et al., 2011; Roos et al., 2011).
Some pests, such as the European corn borer (Trnka et al., 2007), could
also extend their climate niche in Central Europe. Pests and disease
management will be affected with regard to timing, preference, and
efficacy of chemical and biological measures of control (Kersebaum et
al., 2008).
Autonomous adaptation by farmers, through the advancement of
sowing and harvesting dates and the use of longer cycle varieties
(Howden et al., 2007; Moriondo et al., 2010a, 2011; Olesen et al., 2011)
could result in a general improvement of European wheat yields in the
2030s compared to the 2000s (Donatelli et al., 2012; see also Figure
23-4). However, farmer sowing dates seem to advance slower than crop
phenology (Menzel et al., 2006; Siebert and Ewert, 2012), possibly
because earlier sowing is often prevented by lack of soil workability and
frost-induced soil crumbling (Oort et al., 2012). Simulation studies that
anticipate on earlier sowing in Europe may thus be overly optimistic.
Further adaptation options include changes in crop species, fertilization,
irrigation, drainage, land allocation, and farming system (Bindi and
Olesen, 2011). At the high range of the projected temperature changes,
only plant breeding aimed at increasing yield potential jointly with
drought resistance and adjusted agronomic practices may reduce risks
of yield shortfall (Olesen et al., 2011; Rötter et al., 2011; Ventrella et al.,
2012). Crop breeding is, however, challenged by temperature and
rainfall variability, since (1) breeding has not yet succeeded in altering
1285
Europe Chapter 23
23
c
rop plant development responses to short-term changes in temperature
(Parent and Tardieu, 2012), and (2) distinct crop drought tolerance traits
are required for mild and severe water deficit scenarios (Tardieu, 2012).
Adaptation to increased climatic variability may require an increased
use of between and within species genetic diversity in farming systems
(
Smith and Olesen, 2010) and the development of insurance products
against weather-related yield variations (Musshoff et al., 2011). Adaptive
capacity and long-term economic viability of farming systems may vary
given farm structural change induced by climate change (Moriondo et
al., 2010b; Mandryk et al., 2012). In Southern Europe, the regional welfare
Percent difference of water-limited yield for wheat with adaptation
A1B scenario, ECHAM5, 2030–2000 (baseline)
Percent difference of water-limited yield for wheat with adaptation
A1B scenario, HadCM3, 2030–2000 (baseline)
Percent difference of water-limited yield for wheat
A1B scenario, ECHAM5, 2030–2000 (baseline)
Percent difference of water-limited yield for wheat
A1B scenario, HadCM3, 2030–2000 (baseline)
Percent difference
2
0
3
0
−10
−5
5 10
20 30
Figure 23-4 | Percentage change in simulated water-limited yield for winter wheat in 2030 with respect to the 2000 baseline for the A1B scenario using European Centre for
Medium Range Weather Forecasts and Hamburg 5 (ECHAM5; left column) and Hadley Centre Coupled Model version 3 (HadCM3; right) General Circulation Models (GCMs).
Upper maps do not take adaptation into account. Bottom maps include adaptation. Analysis developed at the Joint Research Centre of the European Commission. Source:
Donatelli et al., 2012.
1286
Chapter 23 Europe
23
l
oss caused by changes in the agriculture sector under a high warming
scenario (+5.4°C) was estimated at 1% of gross domestic product
(GDP). Northern Europe was the single sub-region with welfare gains
(+0.7%) from agriculture in this scenario (Ciscar et al., 2011).
23.4.2. Livestock Production
Livestock production is adversely affected by heat (Tubiello et al., 2007;
see also Section 7.2.1.3). With intensive systems, heat stress reduced
dairy production and growth performance of large finishing pigs at daily
mean air temperatures above 18°C and 21°C, respectively (André et al.,
2011; Renaudeau et al., 2011). High temperature and air humidity
during breeding increased cattle mortality risk by 60% in Italy (Crescio
et al., 2010). Adaptation requires changes in diets and in farm buildings
(Renaudeau et al., 2012) as well as targeted genetic improvement
programs (Hoffmann, 2010).
With grass-based livestock systems, model simulations (A1B scenario,
ensemble of downscaled GCMs) show by the end of the 21st century
increases in potential dairy production in Ireland and France, with,
however, higher risks of summer-autumn production failures in Central
Europe and at French sites (Trnka et al., 2009; Graux et al., 2012).
Climate conditions projected for the 2070s in central France (A2 scenario)
reduced significantly grassland production in a 4-year experiment under
elevated CO
2
(Cantarel et al., 2013). At the same site, a single experimental
summer drought altered production during the next 2 years (Zwicke et
al., 2013).
Resilience of grassland vegetation structure was observed to prolonged
experimental heating and water manipulation (Grime et al., 2008).
However, weed pressure from tap-rooted forbs was increased after severe
experimental summer droughts (Gilgen et al., 2010). Mediterranean
populations could be used to breed more resilient and better adapted
forage plant material for livestock production (Poirier et al., 2012).
Climate change has affected animal health in Europe (high confidence).
The spread of bluetongue virus in sheep across Europe has been partly
attributed to climate change (Arzt et al., 2010; Guis et al., 2012) through
increased seasonal activity of the Culicoides vector (Wilson and Mellor,
2009). The distribution of this vector is unlikely to expand but its
abundance could increase in Southern Europe (Acevedo et al., 2010).
Ticks, the primary arthropod vectors of zoonotic diseases in Europe (e.g.,
Lyme disease and tick-borne encephalitis), have changed distributions
towards higher altitudes and latitudes with climate change (Randolph
and Rogers, 2010; van Dijk et al., 2010; Petney et al., 2012; see also
Section 23.5). Exposure to fly strike could increase in a warmer climate
but adaptation in husbandry practices would limit impacts on livestock
(Wall and Ellse, 2011). The overall risk of incursion of Crimean-Congo
hemorrhagic fever virus in livestock through infected ticks introduced
by migratory bird species would not be increased by climate change
(Gale et al., 2012). The probability of introduction and large-scale spread
of Rift Valley fever in Europe is also very low (Chevalier et al., 2010).
Epidemiological surveillance and increased coordinated regional
monitoring and control programs have the potential to reduce the
incidence of vector-borne animal diseases (Wilson and Mellor, 2009;
Chevalier et al., 2010).
23.4.3. Water Resources and Agriculture
Future projected trends confirm the widening of water resource
differences between Northern and Southern Europe reported in AR4
(Alcamo et al., 2007). In Southern Europe, soil water content will
decline, saturation conditions and drainage will be increasingly rare and
restricted to periods in winter and spring, and snow accumulation and
melting will change, especially in the mid-mountain areas (García-
Ruiz et al., 2011). Across most of Northern and Continental Europe, an
increase in flood hazards (Falloon and Betts, 2010; see also Section
23.3.1) could increase damages to crops and plant growth, complicate
soil workability, and increase yield variability (Olesen et al., 2011).
Groundwater recharge and/or water table level would be significantly
reduced by the end of the 21st century under A2 scenario for river basins
located in southern Italy, Spain, northern France, and Belgium (Ducharne
et al., 2010; Goderniaux et al., 2011; Guardiola-Albert and Jackson,
2011; Senatore et al., 2011). However, nonsignificant impacts were
found for aquifers in Switzerland and in England (Jackson et al., 2011;
Stoll et al., 2011). Less precipitation in summer and higher rainfall
during winter could increase nitrate leaching (Kersebaum et al., 2008)
with negative impacts on water quality (Bindi and Olesen, 2011).
Even with reduced nitrogen fertilizer application, groundwater nitrate
concentrations would increase by the end of the century in the Seine
river basin (Ducharne et al., 2007). More robust water management,
pricing, and recycling policies to secure adequate future water supply
and prevent tensions among users could be required in Southern Europe
(García-Ruiz et al., 2011).
Reduced suitability for rainfed agricultural production (Henriques et al.,
2008; Daccache and Lamaddalena, 2010; Trnka et al., 2011; Daccache
et al., 2012) will increase water demand for crop irrigation (Savé et al.,
2012). However, increased irrigation may not be a viable option,
especially in the Mediterranean area, because of projected declines in
total runoff and groundwater resources (Olesen et al., 2011). In a number
of catchments water resources are already over-licensed and/or over-
abstracted (Daccache et al., 2012) and their reliability is threatened by
climate change-induced decline in groundwater recharge and to a lesser
extent by the increase in potential demand for irrigation (Ducharne et
al., 2010; Majone et al., 2012). To match this demand, irrigation system
costs could increase by 20 to 27% in southern Italy (Daccache and
Lamaddalena, 2010) and new irrigation infrastructures would be
required in some regions (van der Velde et al., 2010). However, since the
economic benefits are expected to be small, the adoption of irrigation
would require changes in institutional and market conditions (Finger et
al., 2011). Moreover, since aquatic and terrestrial ecosystems are
affected by agricultural water use (Kløve et al., 2011), irrigation demand
restrictions are projected in environmentally focussed future regional
scenarios (Henriques et al., 2008). Earlier sowing dates, increased soil
organic matter content, low-energy systems, deficit irrigation, and
improved water use efficiency of irrigation systems and crops can be
used as adaptation pathways (Gonzalez-Camacho et al., 2008; Lee et
al., 2008; Daccache and Lamaddalena, 2010; Schutze and Schmitz,
2010), especially in Southern and southeastern regions of Europe (Trnka
et al., 2009; Falloon and Betts, 2010). Improved water management in
upstream agricultural areas could mitigate adverse impacts downstream
(Kløve et al., 2011), and groundwater recharge could be targeted in
areas with poor water-holding soils (Wessolek and Asseng, 2006).
1287
Europe Chapter 23
23
23.4.4. Forestry
Observed and future responses of forests to climate change include
changes in growth rates, phenology, composition of animal and plant
communities, increased fire and storm damage, and increased insect
and pathogen damage. Tree mortality and forest decline due to severe
drought events were observed in forest populations in Southern Europe
(Bigler et al., 2006; Raftoyannis et al., 2008; Affolter et al., 2010), including
Italy (Giuggiola et al., 2010; Bertini et al., 2011), Cyprus (ECHOES Country
Report: Cyprus, 2009), and Greece (Raftoyannis et al., 2008), as well as
in Belgium (Kint et al., 2012), Switzerland (Rigling et al., 2013), and the
pre-Alps in France (Rouault et al., 2006; Allen et al., 2010; Charru et al.,
2010). Declines have also been observed in wet forests not normally
considered at risk of drought (Choat et al., 2012). An increase in forest
productivity has been observed in the Russian Federation (Sirotenko
and Abashina, 2008).
Future projections show that, in Northern and Atlantic Europe, increasing
atmospheric CO
2
and higher temperatures are expected to increase
forest growth and wood production, at least in the short to medium
term (Lindner et al., 2010). On the other hand, in Southern and Eastern
Europe, increasing drought and disturbance risks will cause adverse
effects and productivity is expected to decline (Sirotenko and Abashina,
2008; Lavalle et al., 2009; Lindner et al., 2010; Hlásny et al., 2011; Keenan
e
t al., 2011; Silva et al., 2012). By 2100, climate change is expected to
reduce the economic value of European forest land depending on
interest rate and climate scenario, which equates to potential damages
of several hundred billion euros (Hanewinkel et al., 2013).
In Southern Europe, fire frequency and wildfire extent significantly
increased after the 1970s compared with previous decades (Pausas and
Fernández-Muñoz, 2012) as a result of fuel accumulation (Koutsias et al.,
2012), climate change (Lavalle et al., 2009), and extreme weather events
(Camia and Amatulli, 2009; Hoinka et al., 2009; Carvalho et al., 2011;
Koutsias et al., 2012; Salis et al., 2013), especially in the Mediterranean
basin (Fernandes et al., 2010; Marques et al., 2011; Koutsias et al., 2012;
Pausas and Fernández-Muñoz, 2012). The most severe events in France,
Greece, Italy, Portugal, Spain, and Turkey in 2010 were associated with
strong winds during a hot dry period (EEA, 2010c). However, for the
Mediterranean region as a whole, the total burned area has decreased
since 1985 and the number of wildfires has decreased from 2000 to 2009,
with large interannual variability (Marques et al., 2011; San-Miguel-Ayanz
et al., 2012; Turco et al., 2013). Megafires, triggered by extreme climate
events, had caused record maxima of burnt areas in some Mediterranean
countries during the last decades (San-Miguel-Ayanz et al., 2013).
Future wildfire risk is projected to increase in Southern Europe (Lindner
et al., 2010; Carvalho et al., 2011; Dury et al., 2011; Vilén and Fernandes,
1288
Chapter 23 Europe
23
2
011), with an increase in the occurrence of high fire danger days (Arca
et al., 2012; Lung et al., 2012) and in fire season length (Pellizzaro et al.,
2010). The annual burned area is projected to increase by a factor of 3 to
5 in Southern Europe compared to the present under the A2 scenario by
2100 (Dury et al., 2011). In NorthernEurope, fires are projected to become
less frequent due to increased humidity (Rosan and Hammarlund, 2007).
Overall, the projected increase in wildfires is likely to lead to a significant
increase in greenhouse gas (GHG) emissions due to biomass burning
(Pausas et al., 2008; Vilén and Fernandes, 2011; Chiriacò et al., 2013),
even if often difficult to quantify (Chiriacò et al., 2013).
Wind storm damage to forests in Europe has recently increased (Usbeck
et al., 2010). Boreal forests will become more vulnerable to autumn/
early spring storm damage due to expected decrease in period of frozen
soil (Gardiner et al., 2010). Increased storm losses by 8 to 19% under
A1B and B2 scenarios, respectively, is projected in western Germany for
2060–2100 compared to 1960–2000, with the highest impacts in the
mountainous regions (Pinto et al., 2010; Klaus et al., 2011).
An increase in the incidence of diseases has been observed in many
European forests (Marcais and Desprez-Loustau, 2007; FAO, 2008b). In
Continental Europe, some species of fungi benefit from milder winters
and others spread during drought periods from south to north (Drenkhan
et al., 2006; Hanso and Drenkhan, 2007). Projected increased late summer
warming events will favor diffusion of bark beetle in Scandinavia, in
l
owland parts of Central Europe, and Austria (Jönsson et al., 2009, 2011;
Seidl et al., 2009).
Possible response approaches to the impacts of climate change on
forestry include short- and long-term strategies that focus on enhancing
ecosystem resistance and resilience and responding to potential limits
to carbon accumulation (Millar et al., 2007; Nabuurs et al., 2013).
Fragmented small-scale forest ownership can constrain adaptive capacity
(Lindner et al., 2010). Landscape planning and fuel load management
may reduce the risk of wildfires but may be constrained by the higher
flammability owing to warmer and drier conditions (Moreira et al.,
2011). Strategies to reduce forest mortality include preference of species
better adapted to relatively warm environmental conditions (Resco de
Dios et al., 2007). The selection of tolerant or resistant families and
clones may also reduce the risk of damage by pests and diseases in
pure stands (Jactel et al., 2009).
23.4.5. Bioenergy Production
The potential distribution of temperate oilseeds (e.g., oilseed rape,
sunflower), starch crops (e.g., potatoes), cereals (e.g., barley), and solid
biofuel crops (e.g., sorghum, Miscanthus) is projected to increase in
Northern Europe by the 2080s, as a result of increasing temperatures,
and to decrease in Southern Europe due to increased drought frequency
Box 23-1 | Assessment of Climate Change Impacts on Ecosystem Services by Sub-region
Ecosystems provide a number of vital provisioning, regulating, and cultural services for people and society that flow from the stock of
natural capital (Stoate et al., 2009; Harrison et al., 2010). Provisioning services such as food from agro-ecosystems or timber from
forests derive from intensively managed ecosystems; regulating services underpin the functioning of the climate and hydrological
systems; and cultural services such as tourism, recreation, and aesthetic value are vital for societal well-being (see Section 23.5.4).
The table summarizes the potential impacts of climate change on ecosystem services in Europe by sub-region based on an assessment
of the published literature (2004–2013). The direction of change (increasing, decreasing, or neutral) is provided, as well as the number
of studies/papers on which the assessment was based (in parentheses). Empty cells indicate the absence of appropriate literature.
Unless otherwise stated, impacts assume no adaptation and are assessed for the mid-century (2050s). A decrease in natural hazard
regulation (e.g., for wildfires) implies an increased risk of the hazard occurring. Biodiversity is included here as a service (for
completeness), although it is debated whether biodiversity should be considered as a service or as part of the natural capital from
which services flow. What is agreed, however, is that biodiversity losses within an ecosystem will have deleterious effects on service
provision (Mouillot et al., 2013).
The provision of ecosystem services in Southern Europe is projected to decline across all service categories in response to climate
change (high confidence). Other European sub-regions are projected to have both losses and gains in the provision of ecosystem
services (high confidence). The Northern sub-region will have increases in provisioning services arising from climate change (high
confidence). Except for the Southern sub-region, the effects of climate change on regulating services are balanced with respect to
gains and losses (high confidence). There are fewer studies for cultural services, although these indicate a balance in service provision
for the Alpine and Atlantic regions, with decreases in service provision for the Continental, Northern, and Southern sub-regions (low
confidence).
Continued next page
1289
Europe Chapter 23
23
Box 23-1 (continued)
S
outhern Atlantic Continental Alpine Northern
P
rovisioning
services
Food production
i
(1)
i
(1)
i
(1)
No
i
(1)
(
4)
h
i
(1)
(
1)
Livestock production No
i
(1)
(1)
Fiber production
i
(1)
Bioenergy production
i
(1)
h
(1)
h
(1)
Fish production No
i
(1)
(2)
No
i
(1)
(1)
i
(
1)
No
i
(1)
(1)
Timber production
i
(2)
h
No
(
2)
(
3)
h
No
i
(1)
(2)
(
1)
h
No
i
(5)
(2)
(
5)
h
No
(
6)
(
1)
Non-wood forest products
i
(1)
h
No
(1)
(1)
Sum of effects on provisioning services
N
o
i
(
1)
(
7)
h
No
i
(
2)
(
4)
(2)
h
No
i
(
1)
(
2)
(3)
h
No
i
(
6)
(
4)
(11)
h
No
i
(
9)
(
3)
(2)
Regulating
services
Climate regulation
(carbon sequestration)
General / forests
h
i
(3)
(1)
h
No
(4)
(1)
h
No
(3)
(1)
h
No
i
(4)
(
1)
(3)
h
No
i
(4)
(
1)
(1)
Wetland No
i
(1)
(1)
No
i
(1)
(1)
i
(
1)
No
i
(1)
(1)
Soil carbon stocks No
i
(1)
(1)
No
i
(1)
(2)
No
i
(1)
(1)
No
i
(1)
(2)
i
(3)
Pest control
i
(1)
h
(1)
h
(1)
h
(1)
Natural hazard
regulation
a
Forest fi res /wildfi res
i
(1)
i
(1)
i
(2)
Erosion, avalanche,
landslide
h
i
(2)
(1)
Flooding
i
(1)
Drought No
i
(1)
(1)
i
(1)
Water quality regulation
i
(1)
i
(1)
Biodiversity
h
i
(1)
(8)
h
No
i
(2)
(1)
(4)
h
i
(2)
(4)
h
i
(2)
(4)
h
i
(3)
(2)
Sum of effects on regulating services
h
No
i
(4)
(3)
(14)
h
No
i
(6)
(4)
(9)
h
No
i
(6)
(2)
(9)
h
No
i
(9)
(2)
(11)
h
No
i
(8)
(2)
(8)
Cultural
services
Recreation (fi shing, nature enjoyment)
h
(1)
i
(1)
h
i
(1)
(2)
Tourism (skiing)
h
(1)
h
(1)
Aesthetic / heritage (landscape character,
cultural landscapes)
i
(1)
i
(1)
No
i
(1)
(1)
h
(1)
Sum of effects on cultural services
i
(2)
h
i
(1)
(1)
No
i
(1)
(1)
h
i
(1)
(1)
h
i
(1)
(3)
a
A decline in ecosystem services implies an increased risk of the specifi ed natural hazard.
Entries for biodiversity are those that were found during the literature search for climate change impacts on ecosystem services. A wider discussion of the impacts of
climate change on biodiversity can be found in Sections 4.3.4 and 23.6.
References: Wessel et al. (2004); Schroter et al. (2005); Fuhrer et al. (2006); Koca et al. (2006); Gret-Regamy et al. (2008); Hemery (2008); Metzger et al. (2008); Palahi
et al. (2008); Bolte et al. (2009); Garcia-Fayos and Bochet (2009); Johnson et al. (2009); Albertson et al. (2010); Canu et al. (2010); Clark et al. (2010a); Lindner et al.
(2010); Lorz et al. (2010); Milad et al. (2011); Okruszko et al. (2011); Seidl et al. (2011); Briner et al. (2012); Civantos et al. (2012); Rusch (2012); Bastian (2013); Forsius
et al. (2013); Gret-Regamy et al. (2013); Seidl and Lexer (2013).
Numbers in brackets refer to the number of studies supporting
the change (increasing, decreasing, neutral) in ecosystem service.
= Climate change impacts are
= decreasing ecosystem service
i
= Climate change impacts are
= increasing ecosystem service
i
No = Neutral effect
(1) =
1290
Chapter 23 Europe
23
(
Tuck et al., 2006). Mediterranean oil and solid biofuel crops, currently
restricted to Southern Europe, are likely to extend further north (Tuck et
al., 2006). The physiological responses of bioenergy crops, in particular C
3
Salicaceae trees, to rising atmospheric CO
2
concentration may increase
drought tolerance because of improved plant water use; consequently
yields in temperate environments may remain high in future climate
scenarios (Oliver et al., 2009).
A future increase in the northward extension of the area for short rotation
coppice (SRC) cultivation leading to GHG neutrality is expected (Liberloo
et al., 2010). However, the northward expansion of SRC would erode
the European terrestrial carbon sink due to intensive management and
high turnover of SRC compared to conventional forest where usually
harvesting is less than annual growth (Liberloo et al., 2010).
23.4.6. Fisheries and Aquaculture
In AR4, Easterling et al. (2007) reported that the recruitment and
production of marine fisheries in the North Atlantic are likely to increase.
In European seas, warming causes a displacement to the north and/or
in depth of fish populations (Daufresne et al., 2009; see also Chapter 6;
Section 23.6.4), which has a direct impact on fisheries (Tasker, 2008;
Cheung et al., 2010, 2013). For instance, in British waters, the lesser
sandeel (Ammodytes marinus), which is a key link in the food web,
shows declining recruitments since 2002 and is projected to further
decline in the future with a warming climate (Heath et al., 2012). In the
Baltic Sea, although some new species would be expected to immigrate
because of an expected increase in sea temperature, only a few of these
would be able to successfully colonize the Baltic because of its low
salinity (Mackenzie et al., 2007). In response to climate change and
intensive fishing, widespread reductions in fish body size (Daufresne et
al., 2009) and in the mean size of zooplankton (Beaugrand and Reid,
2012) have been observed over time and these trends further affect the
sustainability of fisheries (Pitois and Fox, 2006; Beaugrand and Kirby,
2010; see also Chapter 6). Aquaculture can be affected as the areal
extent of some habitats that are suitable for aquaculture can be reduced
by sea level rise. Observed higher water temperatures have adversely
affected both wild and farmed freshwater salmon production in the
southern part of the distribution areas (Jonsson and Jonsson, 2009). In
addition, ocean acidification may disrupt the early developmental
stages of shellfish (Callaway et al., 2012).
Numerous studies confirm the amplification through fishing of the
effects of climate change on population dynamics and consequently on
fisheries (Planque et al., 2010). The decline of the North Sea cod during
the 1980–2000 period resulted from the combined effects of overfishing
and of an ecosystem regime shift due to climate change (Beaugrand
and Kirby, 2010). Over the next decade, this stock was not restored from
its previous collapse (Mieszkowska et al., 2009; ICES, 2010). In the North
and Celtic Seas, the steep decline in boreal species (Henderson, 2007)
was compensated for by the arrival of southern (Lusitanian) species (ter
Hofstede et al., 2010; Engelhard et al., 2011; Lenoir et al., 2011).
Climate change may reinforce parasitic diseases and impose severe risks
for aquatic animal health (see Chapter 6). As water temperatures increase,
a number of endemic diseases of both wild and farmed salmonid
p
opulations are likely to become more prevalent and threats associated
with exotic pathogens may rise (Marcos-Lopez et al., 2010). In the Iberian
Atlantic, the permitted harvesting period for the mussel aquaculture
industry was reduced because of harmful algal blooms resulting from
changes in phytoplankton communities linked to a weakening of the
Iberian upwelling (Perez et al., 2010). With freshwater systems, summer
heat waves boost the development of harmful cyanobacterial blooms
(Johnk et al., 2008). For oysters in France, toxic algae may be linked to
both climate warming and direct anthropogenic stressors (Buestel et
al., 2009).
Fishery management thresholds will have to be reassessed as the
ecological basis on which existing thresholds have been established
changes, and new thresholds will have to be developed for immigrant
species (Mackenzie et al., 2007; Beaugrand and Reid, 2012). These
changes may lead to loss of productivity, but also the opening of new
fishing opportunities, depending on the interactions between climate
impacts, fishing grounds, and fleet types. They will also affect fishing
regulations, the price of fish products, and operating costs, which in
turn will affect the economic performance of the fleets (Cheung et al.,
2012). Climate change impacts on fisheries profits range from negative
for sardine fishery in the Iberian Atlantic fishing grounds (Garza-Gil et
al., 2010; Perez et al., 2010) to nonsignificant for the Bay of Biscay (Le
Floc’h et al., 2008) and positive on the Portuguese coast, since most of
the immigrant fish species are marketable (Vinagre et al., 2011). Human
social fishing systems dealing with high variability upwelling systems
with rapidly reproducing fish species may have greater capacities to
adjust to the additional stress of climate change than human social fishing
systems focused on longer-lived and generally less variable species
(Perry et al., 2010, 2011). Climate change adaptation is being considered
for integration in European maritime and fisheries operational programs
(EC, 2013c).
23.5. Implications of Climate Change
for Health and Social Welfare
23.5.1. Human Population Health
Climate change is likely to have a range of health effects in Europe.
Studies since AR4 have confirmed the effects of heat on mortality and
morbidity in European populations and particularly in older people and
those with chronic disease (Kovats and Hajat, 2008; Åström et al., 2011;
Corobov et al., 2012, 2013). With respect to sub-regional vulnerability,
populations in Southern Europe appear to be most sensitive to hot weather
(Michelozzi et al., 2009; D’Ippoliti et al., 2010; Baccini et al., 2011), and
also will experience the highest heat wave exposures (Figure 23-2).
However, populations in Continental (Hertel et al., 2009) and Northern
Europe (Rocklöv and Forsberg, 2010; Armstrong et al., 2011; Varakina et
al., 2011) are also vulnerable to heat wave events. Adaptation measures
to reduce heat health effects include heat wave plans (Bittner et al.,
2013) which have been shown to reduce heat-related mortality in Italy
(Schifano et al., 2012), but evidence of effectiveness is still very limited
(Hajat et al., 2010; Lowe et al., 2011). There is little information about
how future changes in housing and infrastructure (Section 23.3.2) would
reduce the regional or local future burden of heat-related mortality or
morbidity. Climate change is likely to increase future heat-related
1291
Europe Chapter 23
23
m
ortality (Baccini et al., 2011; Ballester et al., 2011; Huang et al., 2011)
and morbidity (Åström et al., 2013), although most published risk
assessments do not include consideration of adaptation (Huang et al.,
2011). For most countries in Europe, the current burden of cold-related
mortality (Analitis et al., 2008) is greater than the burden of heat
mortality. Climate change is likely to reduce future cold-related mortality
(Ballester et al., 2011; HPA, 2012; see also Section 11.4.1).
Mortality and morbidity associated with flooding is becoming better
understood, although the surveillance of health effects of disasters
remains inadequate (WHO, 2013). Additional flood mortality due to sea
level rise has been estimated in the Netherlands (Maaskant et al., 2009)
and in the UK for river flooding (Hames and Vardoulakis, 2012), but
estimates of future mortality due to flooding are highly uncertain.
There remains limited evidence regarding the long-term mental health
impacts of flood events (Paranjothy et al., 2011; WHO, 2013).
Evidence about future risks from climate change with respect to
infectious diseases is still limited (Semenza and Menne, 2009; Randolph
and Rogers, 2010; Semenza et al., 2012). There have been developments
in mapping the current and potential future distribution of important
disease vector species in Europe. The Asian tiger mosquito Aedes
albopictus (a vector of dengue and Chikungunya; Queyriaux et al., 2008)
is currently present in Southern Europe (ECDC, 2009) and may extend
eastward and northward under climate change (Fisher et al., 2011; Roiz
et al., 2011; Caminade et al., 2012). The risk of introduction of dengue
remains very low because it would depend on the introduction and
expansion of the Aedes aegypti together with the absence of effective
vector control measures (ECDC, 2012).
Climate change is unlikely to affect the distribution of visceral and
cutaneous leishmaniasis (currently present in the Mediterranean region)
in the near term (Ready, 2010). However, in the long term (15 to 20
years), there is potential for climate change to facilitate the expansion
of either vectors or current parasites northwards (Ready, 2010). The risk
of introduction of exotic Leishmania species was considered very low due
to the low competence of current vectors (Fischer, D. et al., 2010). The
effect of climate change on the risk of imported or locally transmitted
(autochthonous) malaria in Europe has been assessed in Spain (Sainz-
Elipe et al., 2010), France (Linard et al., 2009), and the UK (Lindsay et
al., 2010). Disease re-emergence would depend on many factors,
including the introduction of a large population of infectious people
or mosquitoes, high levels of people-vector contact, resulting from
significant changes in land use, as well as climate change (see Chapter
11).
Since AR4, there has been more evidence on implications of climate
change on food safety at all stages from production to consumption
(FAO, 2008a; Jacxsens et al., 2010; Popov Janevska et al., 2010). The
sensitivity of salmonellosis to ambient temperature has declined in
recent years (Lake et al., 2009) and the overall incidence of salmonellosis
is declining in most European countries (Semenza et al., 2012). Climate
change may also have effects on food consumption patterns. Weather
affects pre- and post-harvest mycotoxin production but the implications
of climate change are unclear. Cold regions may become liable to
temperate-zone problems concerning contamination with ochratoxin
A, patulin, and Fusarium toxins (Paterson and Lima, 2010). Control of
t
he environment of storage facilities may avoid post-harvest problems
but at additional cost (Paterson and Lima, 2010).
Other potential consequences concern marine biotoxins in seafood
following production of phycotoxins by harmful algal blooms and the
presence of pathogenic bacteria in foods following more frequent extreme
weather conditions (Miraglia et al., 2009). There is little evidence that
climate change will affect human exposures to contaminants in the soil
or water (e.g., persistent organic pollutants). Risk modeling is often
developed for single-exposure agents (e.g., a pesticide) with known
routes of exposure. These are difficult to scale up to the population level.
The multiple mechanisms by which climate may affect transmission or
contamination routes also make this very complex (Boxall et al., 2009).
Adaptation in the health sector has so far been largely limited to the
development of heat health warning systems, with many research gaps
regarding effective adaptation options (HPA, 2012). A survey of national
infectious disease experts in Europe identified several institutional
changes that needed to be addressed to improve future responses to
climate change risks: ongoing surveillance programs, collaboration with
veterinary sector and manage ment of animal disease outbreaks, national
monitoring and control of climate-sensitive infec tious diseases, health
services during an infectious disease outbreak, and diagnostic support
during an epidemic (Semenza et al., 2012).
23.5.2. Critical Infrastructure
Critical national infrastructure is defined as assets (physical or electronic)
that are vital to the continued delivery and integrity of essential services
on which a country relies, the loss or compromise of which would lead
to severe economic or social consequences or to loss of life. Extreme
weather events, such as floods, heat waves, and wild fires are known
to damage critical infrastructure. The UK floods in 2007 led to significant
damage to power and water utilities, and to communications and transport
infrastructure (Chatterton et al., 2010; see also Table 23-1). Forest fires
can affect transport infrastructure, as well as the destruction of buildings.
Major storms in Sweden and Finland have led to loss of trees, with
damage to the power distribution network, leading to electricity blackouts
lasting weeks, as well as the paralysis of services such as rail transport
and other public services that depend on grid electricity.
Health system infrastructure (hospitals, clinics) is vulnerable to extreme
events, particularly flooding (Radovic et al., 2012). The heat waves of
2003 and 2006 had adverse effects on patients and staff in hospitals
from overheating of buildings. Evidence from France and Italy indicate
that death rates among in-patients increased significantly during heat
wave events (Ferron et al., 2006; Stafoggia et al., 2008). Further, higher
temperatures have had serious implications for the delivery of health
care, as well drug storage and transport (Carmichael et al., 2013).
23.5.3. Social Impacts
There is little evidence regarding the implications of climate change for
employment and/or livelihoods in Europe. However, the evidence so far
(as reviewed in this chapter) indicates that there are likely to be changes
1292
Chapter 23 Europe
23
t
o some industries (e.g., tourism, agriculture) that may lead to changes
in employment opportunities by sub-region and by sector.
Current damages from weather-related disasters (floods and storms)
are significant (Section 23.3.1). Disasters have long lasting effects on
the affected populations (Schnitzler et al., 2007). Households are often
displaced while their homes are repaired (Whittle et al., 2010). Little
research has been carried out on the impact of extreme weather events
such as heat waves and flooding on temporary or permanent displacement
in Europe. Coastal erosion associated with sea level rise, storm surges,
and coastal flooding will require coastal retreat in some of Europes
low-lying areas (Philippart et al., 2011). Managed retreat is also an
adaptation option in some coastal areas. Concerns have been raised
about equality of access to adaptation within coastal populations at
risk from climate change. For example, a study in the UK found that
vulnerability to climate change in coastal communities is likely to be
increased by social deprivation (Zsamboky et al., 2011).
In the European region, the indigenous populations present in the Arctic
are considered vulnerable to climate change impacts on livelihoods and
food sources (ACIA, 2005; see also Sections 12.3, 28.2.4). Research has
focused on indigenous knowledge, impacts on traditional food sources,
and community responses/adaptation (Mustonen and Mustonen,
2011a,b). However, these communities are also experiencing rapid
social, economic, and other non-climate-related environmental changes
(such as oil and gas exploration; see Section 28.2.4). There is evidence
that climate change has altered the seasonal behavior of pastoralist
populations, such as the Nenets reindeer herders in northern Russia
(
Amstislavski et al., 2013). However, socioeconomic factors may be more
important than climate change for the future sustainability of reindeer
husbandry (Rees et al., 2008; see also Section 28.2.3.5).
23.5.4. Cultural Heritage and Landscapes
Climate change will affect culturally valued buildings (Storm et al., 2008)
through extreme events and chronic damage to materials (Brimblecombe
et al., 2006; Brimblecombe and Grossi, 2010; Brimblecombe, 2010a,
2010b; Grossi et al., 2011; Sabbioni et al., 2012). Cultural heritage is a
non-renewable resource and impacts from environmental changes are
assessed over long time scales (Brimblecombe and Grossi, 2008, 2009,
2010; Grossi et al., 2008; Bonazza et al., 2009a,b). Climate change may
also affect indoor environments where cultural heritage is preserved
(Lankester and Brimblecombe, 2010) as well as visitor behavior at
heritage sites (Grossi et al., 2010). There is also evidence to suggest that
climate change and sea level rise will affect maritime heritage in the
form of shipwrecks and other submerged archaeology (Björdal, 2012).
Surface recession on marble and compact limestone will be affected by
climate change (Bonazza et al., 2009a). Marble monuments in Southern
Europe will continue to experience high levels of thermal stress (Bonazza
et al., 2009b) but warming is likely to reduce frost damage across Europe,
except in Northern and Alpine Europe and permafrost areas (Iceland)
(Grossi et al., 2007; Sabbioni et al., 2008). Damage to porous materials
due to salt crystallization may increase all over Europe (Benavente et
al., 2008; Grossi et al., 2011). In Northern and Eastern Europe, wood
Box 23-2 | Implications of Climate Change for European Wine and Vineyards
Wine production in Europe accounts for more than 60% of the global total (Goode, 2012) and makes an important contribution to
cultural identity. Apart from impacts on grapevine yield, higher temperatures are also expected to affect wine quality in some regions
and grape varieties by changing the ratio between sugar and acids (Duchêne et al., 2010; Bock et al., 2011; Santos et al., 2011). In
Western and Central Europe, projected future changes could benefit wine quality, but might also demarcate new potential areas for
viticulture (Malheiro et al., 2010). Adaptation measures are already occurring in some vineyards (e.g., vine management, technological
measures, production control, and to a smaller extent relocation; Battaglini et al., 2009; Holland and Smit, 2010; Malheiro et al.,
2010; Duarte Alonso and O’Neill, 2011; Moriondo et al., 2011; Santos et al., 2011).Vineyards may be displaced geographically beyond
their traditional boundaries (“terroir” linked to soil, climate, and traditions; Metzger and Rounsevell, 2011) and, in principle, wine
producers could adapt to this problem by growing grape varieties that are more suited to warmer climates. Such technical solutions,
however, do not account for the unique characteristics of wine production cultures and consumer perceptions of wine quality that
strongly affect the prices paid for the best wines (White et al., 2009; Metzger and Rounsevell, 2011). It would become very difficult,
for example, to produce fine wines from the cool-climate Pinot Noir grape within its traditional “terroir” of Burgundy under many
future climate scenarios, but consumers may not be willing to pay current day prices for red wines produced from other grape varieties
(Metzger and Rounsevell, 2011). An additional barrier to adaptation is that wine is usually produced within rigid, regionally specific,
regulatory frameworks that often prescribe, among other things, what grapes can be grown where, for example, the French AOC
(Appellation d'Origine Controlee) or the Italian DOC (Denominazione di Origine Controllata) and DOCG (Denominazione di Origine
Controllata e Garantita) designations. Suggestions have been made to replace these rigid concepts of regional identity with a
geographically flexible “terroir” that ties a historical or constructed sense of culture to the wine maker and not to the region (White
et al., 2009).
1293
Europe Chapter 23
23
s
tructures will need additional protection against rainwater and high
winds (Sabbioni et al., 2012). AR4 concluded that current flood defenses
would not protect Venice from climate change. Venice now has a flood
forecasting system, and is introducing the MOSE (MOdulo Sperimentale
Elettromeccanico) system of flood barriers (Keskitalo, 2010). Recent
evidence suggests, however, that climate change may lead to a decrease
in the frequency of extreme storm surges in this area (Troccoli et al.,
2012a).
Europe has many unique rural landscapes, which reflect the cultural
heritage that has evolved from centuries of human intervention, for
example, the cork oak based Montado in Portugal, the Garrigue of southern
France, Alpine meadows, grouse moors in the UK, machair in Scotland,
peatlands in Ireland, the polders of Belgium and the Netherlands, and
vineyards. Many, if not all, of these cultural landscapes are sensitive
to climate change and even small changes in the climate could have
significant impacts (Gifford et al., 2011). Alpine meadows, for example,
are culturally important within Europe, but although there is analysis
of the economics (tourism, farming) and functionality (water runoff,
flooding, and carbon sequestration) of these landscapes there is very
little understanding of how climate change will affect the cultural
aspects on which local communities depend. Because of their societal
value, cultural landscapes are often protected and managed through
rural development and environmental policies. The peat-rich uplands of
Northern Europe, for example, have begun to consider landscape
management as a means of adapting to the effects of climate change
(e.g., the moors for the future partnership in the Peak District National
Park, UK). For a discussion of the cultural implications of climate change
for vineyards, see Box 23-2.
23.6. Implications of Climate Change for the
Protection of Environmental Quality and
Biological Conservation
Terrestrial and freshwater ecosystems provide a number of vital services
for people and society, such as biodiversity, food, fiber, water resources,
carbon sequestration, and recreation (Box 23-1).
23.6.1. Air Quality
Climate change will have complex and local effects on pollution chemistry,
transport, emissions, and deposition. Outdoor air pollutants have adverse
effects on human health, biodiversity, crop yields, and cultural heritage.
The main outcomes of concern are both the average (background) levels
and peak events for tropospheric ozone, particulates, sulfur oxides (SO
x
),
and nitrogen oxides (NO
x
). Future pollutant concentrations in Europe
have been assessed using atmospheric chemistry models, principally for
ozone (Forkel and Knoche, 2006, 2007). Reviews have concluded that
GCM/Chemical Transport Model (CTM) studies find that climate change
per se (assuming no change in future emissions or other factors) is likely
to increase summer tropospheric ozone levels (range 1 to 10 ppb) by
2050s in polluted areas (i.e., where concentrations of precursor nitrogen
oxides are higher) (AQEG, 2007; Jacob and Winner, 2009; see also
Section 21.3.3.6). The effect of future climate change alone on future
concentrations of particulates, nitrogen oxides, and volatile organic
c
ompounds (VOCs) is much more uncertain. Higher temperatures also
affect natural VOC emissions, which are ozone precursors (Hartikainen
et al., 2012). One study has projected an increase in fire-related air
pollution (ozone and particulate matter with aerodynamic diameter
<10 μm (PM
10
)) in Southern Europe (Carvalho et al., 2011).
Overall, the model studies are inconsistent regarding future projections
of background level and exceedances. Recent evidence has shown
adverse impacts on agriculture from even low concentrations of ozone;
however, there is more consistent evidence now regarding the threshold
for health (mortality) impacts of ozone. Therefore, it is unclear whether
increases in background levels below health-related thresholds would
be associated with an increased burden of ill health.
Some studies have attributed an observed increase in European ozone
levels to observed warming (Meleux et al., 2007), which appears to be
driven by the increase in extreme heat events (Solberg et al., 2008).
High ozone levels were observed during the major heat waves in Europe
in multiple countries (Table 23-1). Wildfire events have had an impact
on local and regional air quality (Hodzic et al., 2007; Liu et al., 2009;
Miranda et al., 2009), with implications for human health (Analitis et al.,
2012; Table 23-1).
23.6.2. Soil Quality and Land Degradation
The current cost of soil erosion, organic matter decline, salinization,
landslides, and contamination is estimated to be €38 billion annually
for the EU (JRC and EEA, 2010), in the form of damage to infrastructures,
treatment of water contaminated through the soil, disposal of sediments,
depreciation of land, and costs related to the ecosystem functions of
soil (JRC and EEA, 2010). Projections show significant reductions in
summer soil moisture in the Mediterranean region, and increases in the
northeastern part of Europe (Calanca et al., 2006). Climate change
impacts on erosion shows diverging evidence under the A2 scenario. In
Tuscany, even with a decline in precipitation volume until 2070, in some
months higher erosion rates would occur because of higher rainfall
erosivity (Marker et al., 2008). For two Danish river catchments, assuming
a steady-state land use, suspended sediment transport would increase
by 17 to 27% by 2071–2100 (Thodsen, 2007; Thodsen et al., 2008). In
Upper Austria, with the regional climate model HadRM3H, a small
reduction in average soil losses is projected for croplands in all tillage
systems, however, with high uncertainty (Scholz et al., 2008). In Northern
Ireland, erosion decreases are generally projected with downscaled
GCMs for a case study hillslope (Mullan et al., 2012).
Adaptive land use management can reduce the impact of climate change
through soil conservation methods such as zero tillage and conversion
of arable land to grasslands (Klik and Eitzinger, 2010). In central Europe,
compared to conventional tillage, conservation tillage systems reduced
modeled soil erosion rates under future climate scenarios by between
49 and 87% (Scholz et al., 2008). Preserving upland vegetation reduced
both erosion and loss of soil carbon and favored the delivery of a high-
quality water resource (McHugh, 2007; House et al., 2011). Maintaining
soil water retention capacity, for example, through adaptation measures
(Post et al., 2008), contributes to reduce risks of flooding as soil organic
matter absorbs up to 20 times its weight in water.
1294
Chapter 23 Europe
23
23.6.3. Water Quality
Climate change may affect water quality in several ways, with implications
for food production and forestry (Section 23.4), ecosystem functioning
(Box 23-1), human and animal health, and compliance with environmental
quality standards, including those of the Water Framework Directive.
Shallower waters will witness a more rapid temperature increase than
deeper waters, since heat is absorbed mainly in the upper water layers
and turbulent mixing is truncated by shallow depth. In parallel, a decrease
in saturating oxygen concentrations occurs. Since AR4, there is further
evidence of adverse effects caused by extreme weather events:
reductions in dissolved oxygen, algal blooms (Mooij et al., 2007; Ulén
and Weyhenmeyer, 2007) during hot weather, and contamination of
surface and coastal waters with sewage and/or chemicals (pesticides)
after rainfall (Boxall et al., 2009). A reduction in rainfall may lead to
low flows that increase concentrations of biological and chemical
contaminants. Reduced drainage can also enhance sedimentation in
drainage systems and hence enhance particle-bound phosphorous
retention and reduce phosphorous load to downstream higher order
streams (Hellmann and Vermaat, 2012).
Variability in changes in rainfall and runoff, as well as water temperature
increases, will lead to differences in water quality impacts by sub-region.
Climate change is projected to increase nutrient loadings: In Northern
Europe this is caused by increased surface runoff, and in Southern Europe
by increased evapotranspiration and increased concentrations due to
reduced volumes of receiving lakes (Jeppesen et al., 2011). Local studies
generally confirm this pattern. Increased nutrient loads are foreseen in
Danish watersheds (Andersen et al., 2006), and in France (Delpla et al.,
2011) and the UK (Whitehead et al., 2009; Howden et al., 2010; Macleod
et al., 2012; see also Section 4.3.3.3). In larger rivers, such as the Meuse,
increased summer temperature and drought can lead to more favorable
conditions for algal blooms and reduced dilution capacity of effluent
from industry and sewage works (van Vliet and Zwolsman, 2008).
23.6.4. Terrestrial and Freshwater Ecosystems
Currentand projected future climatechanges, including CO
2
increase,
are determining negativeeffectsofhabitatlosson species density
and diversity (Rickebusch et al., 2008; Mantyka-pringle et al., 2012).
Projected habitat loss is greater for species at higher elevations (Castellari,
2009; Engler et al., 2011; Dullinger et al., 2012) and suitable habitats
for Europes breeding birds are projected to shift nearly 550 km northeast
by the end of the 21st century (Huntley et al., 2007). Aquatic habitats
and habitat connectivity in river networks may become increasingly
fragmented (Fronzek et al., 2006, 2010, 2011; Elzinga et al., 2007; Della
Bella et al., 2008; Harrison et al., 2008; Blaustein et al., 2010; Gallego-
Sala et al., 2010; mez-Rodguez et al., 2010; Hartel et al., 2011; Mon-
López et al., 2012). Despite some local successes and increasing
responses, the rate ofbiodiversityloss does not appear to be slowing
(Butchart et al., 2010). The effectiveness of Natura 2000 areas to respond
to climate change has been questioned (Araújo et al., 2011). However,
when considering connectivity related to the spatial properties of the
network, the Natura 2000 network appears rather robust (Mazaris et
al., 2013). Several studies now highlight the importance of taking into
account climate change projections in the selection of conservation
a
reas (Araújo et al., 2011; Ellwanger et al., 2011; Filz et al., 2013;
Virkkala et al., 2013).
Observed changes in plant communities in European mountainous
regions show a shift of species ranges to higher altitudes resulting in
species richness increase in boreal-temperate mountain regions and
decrease in Mediterranean mountain regions (Gottfried et al., 2012;
Pauli et al., 2012). In Southern Europe, a great reduction in phylogenetic
diversity of plant, bird, and mammal assemblages will occur, and gains
are expected in regions of high latitude or altitude for 2020, 2050, and
2080. However, losses will not be offset by gains and a trend toward
homogenization across the continent will be observed (Alkemade et al.,
2011; Thuiller et al., 2011). Large range contractions due to climate
change are projected for several populations of Pinus cembra and Pinus
Sylvestris (Casalegno et al., 2010; Giuggiola et al., 2010) while for the
dominant Mediterranean tree species, holm oak, a substantial range
expansion is projected under the A1B emissions scenario (Cheaib et al.,
2012). The human impacts on distribution of tree species landscape may
make them more vulnerable to climate change (del Barrio et al., 2006;
Hemery et al., 2010).
Observed climate changes are altering breeding seasons, timing of spring
migration, breeding habitats, latitudinal distribution, and migratory
behavior of birds (Jonzén et al., 2006; Lemoine et al., 2007a,b; Rubolini
et al., 2007a,b; Feehan et al., 2009). A northward shift in bird community
composition has been observed (Devictor et al., 2008). Common species
of European birds with the lowest thermal maxima have showed the
sharpest declines between 1980 and 2005 (Jiguet et al., 2010).
Projections for 120 native terrestrial non-volant European mammals
suggest that 5 to 9% are at risk of extinction, assuming no migration,
during the 21st century due to climate change, while 70 to 78% may
be severely threatened under A1 and B2 climatic scenarios (Levinsky et
al., 2007). Those populations not showing a phenological response to
climate change may decline (Moller et al., 2008), such as amphibian
and reptile species (Araújo et al., 2006), or experience ecological
mismatches (Saino et al., 2011). Climate change can affect trophic
interactions, as co-occurring species may not react in a similar manner.
Novel emergent ecosystems composed of new species assemblages
arising from differential rates of range shifts of species can occur (Keith
et al., 2009; Montoya and Raffaelli, 2010; Schweiger et al., 2012).
Since invasive alien species rarely change their original climatic niches
(Petitpierre et al., 2012), climate change can exacerbate the threat
posed by invasive species to biodiversity in Europe (West et al., 2012),
amplifying the effects of introduction of the exotic material such as
alien bioenergy crops (EEA, 2012), pest and diseases (Aragòn and Lobo,
2012), tropical planktonic species (Cellamare et al., 2010), and tropical
vascular plants (Skeffington and Hall, 2011; Taylor et al., 2012).
23.6.5. Coastal and Marine Ecosystems
Climate change will affect Europe’s coastal and marine ecosystems by
altering the biodiversity, functional dynamics, and ecosystem services
of coastal wetlands, dunes, inter-tidal and subtidal habitats, offshore
shelves, seamounts, and currents (Halpern et al., 2008) through changes
1295
Europe Chapter 23
23
i
n eutrophication, invasive species, species range shifts, changes in fish
stocks, and habitat loss (EEA, 2010d; Doney et al., 2011). The relative
magnitude of these changes will vary temporally and spatially, requiring
a range of adaptation strategies that target different policy measures,
audiences, and instruments (Airoldi and Bec, 2007; Philippart et al.,
2011).
Europe’s northern seas are experiencing greater increases in sea surface
temperatures (SSTs) than the southern seas, with the Baltic, North, and
Black Seas warming at two to four times the mean global rate (Belkin,
2009; Philippart et al., 2011). In the Baltic, decreased sea ice will expose
coastal areas to more storms, changing the coastal geomorphology
(HELCOM, 2007; BACC Author Team, 2008). Warming SSTs will influence
biodiversity and drive changes in depth and latitudinal range for inter-
tidal and subtidal marine communities, particularly in the North and
Celtic Seas (Sorte et al., 2010; Hawkins et al., 2011; Wethey et al., 2011).
Warming is affecting food chains and changing phenological rates
(Durant et al., 2007). For example, changes in the timing and location
of phytoplankton and zooplankton are affecting North Sea cod larvae
(Beaugrand et al., 2010; Beaugrand and Kirby, 2010). Temperature
changes have affected the distribution of fisheries in all seas over the
past 30 years (Beaugrand and Kirby, 2010; Hermant et al., 2010). Warmer
waters also increase the rate of the establishment and spread of
invasive species, further altering trophic dynamics and the productivity
of coastal marine ecosystems (Molnar et al., 2008; Rahel and Olden,
2008). Changes in the semi-enclosed seas could be indicative of future
c
onditions in other coastal-marine ecosystems (Lejeusne et al., 2009).
In the Mediterranean, invasive species have arrived in recent years at
the rate of one introduction every 4 weeks (Streftaris et al., 2005). While
in this case the distribution of endemic species remained stable, most
non-native species have spread northward by an average of 300 km
since the 1980s, resulting in an area of spatial overlap with invasive
species replacing natives by nearly 25% in 20 years.
Dune systems will be lost in some places due to coastal erosion from
combined storm surge and sea level rise, requiring restoration (Day et
al., 2008; Magnan et al., 2009; Ciscar et al., 2011). In the North Sea,
the Iberian coast, and Bay of Biscay, a combination of coastal erosion,
infrastructure development, and sea defenses may lead to narrower
coastal zones (“coastal squeeze”) (EEA, 2010d; OSPAR, 2010; Jackson
and McIlvenny, 2011).
23.7. Cross-Sectoral Adaptation
Decision Making and Risk Management
Studies on impacts and adaptation in Europe generally consider single
sectors or outcomes, as described in the previous sections of this chapter.
For adaptation decision making, more comprehensive approaches are
required. Considerable progress has been made to advance planning
and development of adaptation measures, including economic analyses
(Section 23.7.6; see Box 23-3), and the development of climate services
(WMO, 2011; Medri et al., 2012). At the international level, the European
Box 23-3 | National and Local Adaptation Strategies
The increasing number of national (EEA, 2013) and local (Heidrich et al., 2013) adaptation strategies in Europe has led to research on
their evaluation and implementation (Biesbroek et al., 2010). Many adaptation strategies were found to be agendas for further
research, awareness raising, and/or coordination and communication for implementation (e.g., Pfenniger et al., 2010; Dumollard and
Leseur, 2011). Actual implementation often was limited to disaster risk reduction, environmental protection, spatial planning (Section
23.7.4), and coastal zone and water resources management. The implementation of planned adaptation at the national level was
attributed to political will and good financial and information capacity (Westerhoff et al., 2011). Analysis of seven national adaptation
strategies (Denmark, Finland, France, Germany, Netherlands, Spain, UK) found that although there is a high political commitment to
adaptation planning and implementation, evaluation of the strategies and actual implementation is yet to be defined (Swart et al.,
2009b; Biesbroek et al., 2010; Westerhoff et al., 2011). One of the earliest national adaptation strategies (Finland) has been evaluated,
in order to compare identified adaptation measures with those launched in different sectors. It has found that although good
progress has been made on research and identification of options, few measures have been implemented except in the water
resources sector (Ministry of Agriculture and Forestry, 2009).
At the local government level, adaptation plans are being developed in several cities (EEA, 2013), including London (GLA, 2010),
Madrid, Manchester, Copenhagen, Helsinki, and Rotterdam. Adaptation in general is a low priority for many European cities, and
many plans do not have adaptation priority as the main focus (Carter, 2011). Many studies are covering sectors sensitive to climate
variability, as well as sectors that are currently under pressure from socioeconomic development. A recent assessment found a lack of
cross-sector impact and adaptation linkages as an important weakness in the city plans (Hunt and Watkiss, 2011). Flexibility in
adaptation decision making needs to be maintained (Hallegatte et al., 2008; Biesbroek et al., 2010).
1296
Chapter 23 Europe
23
U
nion has started adaptation planning, through information sharing
(Climate-ADAPT platform) and legislation (EC, 2013b). National and
local governments are also beginning to monitor progress on adaptation,
including the development of a range of indicators (UK-ASC, 2011).
23.7.1. Coastal Zone Management
Coastal zone management and coastal protection plans that integrate
adaptation concerns are now being implemented. Underlying scientific
studies increasingly assess effectiveness and costs of specific options
(Hilpert et al., 2007; Kabat et al., 2009; Dawson et al., 2011; see also
Section 23.7.6). Early response measures are needed for floods and
coastal erosion, to ensure that climate change considerations are
incorporated into marine strategies, with mechanisms for regular update
(OSPAR, 2010; UNEP, 2010).
In the Dutch plan for flood protection, adaptation to increasing river
runoff and sea level rise plays a prominent role (Delta Committee, 2008).
It also includes synergies with nature conservation and freshwater
storage (Kabat et al., 2009), and links to urban renovation (cost
estimates are included in Section 23.7.6). Though that plan mostly
relies on large-scale measures, new approaches such as small-scale
containment of flood risks through compartmentalization are also
studied (Klijn et al., 2009). The UK government has developed extensive
adaptation plans (TE2100) to adjust and improve flood defenses for the
protection of London from future storm surges and flooding (EA, 2009).
An elaborate analysis has provided insight in the pathways for different
adaptation options and decision-points that will depend on the eventual
sea level rise (Box 5-1).
23.7.2. Integrated Water Resource Management
Water resources management in Europe has experienced a general shift
from “hard” to “soft” measures that allow more flexible responses to
environmental change (Pahl-Wostl, 2007). Integrated water resource
management explicitly includes the consideration of environmental
and social impacts (Wiering and Arts, 2006). Climate change has been
incorporated into water resources planning in England and Wales
(Arnell, 2011; Charlton and Arnell, 2011; Wade et al., 2013) and in the
Netherlands (de Graaff et al., 2009). The robustness of adaptation
strategies for water management in Europe has been tested in England
(Dessai and Hulme, 2007) and Denmark (Haasnoot et al., 2012;
Refsgaard et al., 2013). Other studies have emphasized the search for
robust pathways, for instance, in the Netherlands (Kwadijk et al., 2010;
Haasnoot et al., 2012).
Public participation has also increased in decision making, for example,
river basin management planning (Huntjens et al., 2010), flood defense
plans (e.g., TE2100), and drought contingency plans (Iglesias et al.,
2007). Guidance has been developed on the inclusion of adaptation in
water management (UNECE, 2009) and river basin management plans
(EC, 2009b). Adaptation in the water sector could also be achieved
through the EU Water Framework and Flood Directives (Quevauviller,
2011), but a study of decision makers, including local basin managers,
identified several important barriers to this (Brouwer et al., 2013). Water
a
llocation between upstream and downstream countries is challenging
in regions exposed to prolonged droughts such as the Euphrates-Tigris
river basin, where Turkey plans to more than double water abstraction
by 2023 (EEA, 2010a).
23.7.3. Disaster Risk Reduction and Risk Management
A series of approaches to disaster risk management are employed in
Europe, in response to national and European policy developments to
assess and reduce natural hazard risks. New developments since the
AR4 include assessment and protection efforts in accordance with the
EU Floods Directive (European Parliament and EU Council, 2007), the
mapping of flood risks, and improvement of civil protection response
and early warning systems (Ciavola et al., 2011). Most national policies
address hazard assessment and do not include analyses of possible
impacts (de Moel et al., 2009). The effectiveness of flood protection
(Bouwer et al., 2010) and also non-structural or household level measures
to reduce losses from river flooding has been assessed (Botzen et al.,
2010a; Dawson et al., 2011). Some studies show that current plans may
be insufficient to cope with increasing risks from climate change, as
shown, for instance, for the Rhine River basin (te Linde et al., 2010a,b).
Other options that are being explored are the reduction of consequences,
response measures, and increasing social capital (Kuhlicke et al., 2011),
as well as options for insuring and transferring losses (Section 23.3.7).
The Netherlands carried out a large-scale analysis and simulation
exercise to study the possible emergency and evacuation response for
a worst-case flood event (ten Brinke et al., 2010). Increasing attention
is also being paid in Europe to non-government actions that can reduce
possible impacts from extreme events. Terpstra and Gutteling (2008)
found through a survey that individual citizens are willing to assume
some responsibility for managing flood risk, and they are willing to
contribute to preparations in order to reduce impacts. Survey evidence
is available for Germany and the Netherlands that, under certain
conditions, individuals can be encouraged to adopt loss prevention
measures (Thieken et al., 2006; Botzen et al., 2009). Small businesses
can reduce risks when informed about possibilities immediately after
an event (Wedawatta and Ingirige, 2012).
23.7.4. Land Use Planning
Spatial planning policies can build resilience to the impacts of climate
change (Bulkeley, 2010). However, the integration of adaptation into
spatial planning is often limited to a general level of policy formulation
that can sometimes lack concrete instruments and measures for
implementation in practice (Mickwitz et al., 2009; Swart et al., 2009a).
There is evidence to suggest the widespread failure of planning
policy to account for future climate change (Branquart et al., 2008).
Furthermore, a lack of institutional frameworks to support adaptation
is, potentially, a major barrier to the governance of adaptation through
spatial planning (ESPACE, 2007; Chapter 16). Climate change adaptation
is often treated as a water management or flooding issue, which omits
other important aspects of the contribution of land use planning to
adaptation (Wilson, 2006; Mickwitz et al., 2009; Van Nieuwaal et al.,
2009). For example, in the UK, houses were still being built in flood risk
1297
Europe Chapter 23
23
a
reas (2001–2011) because of competing needs to increase the housing
stock (ARUP, 2011).
City governance is also dominated by the issues of climate mitigation and
energy consumption rather than adapting to climate change (Bulkeley,
2010; Heidrich et al., 2013). Some cities, for example, Rotterdam, have
started to create climate adaptation plans and this process tends to be
driven by the strong political leadership of mayors (Sanchez-Rodriguez,
2009). The Helsinki Metropolitan Area’s Climate Change Adaptation
Strategy (HSY, 2010) is a regional approach focusing on the built
environment in the cities of Helsinki, Espoo, Vantaa, and Kauniainen,
and their surroundings. It includes approaches for dealing with
increasing heat waves, more droughts, milder winters, increasing
(winter) precipitation, heavy rainfall events, river floods, storm surges,
drainage water floods, and sea level rise.
Green infrastructure provides both climate adaptation and mitigation
benefits as well as offering a range of other benefits to urban areas,
including health improvements, amenity value, inward investment, and
the reduction of noise and outdoor air pollution. Green infrastructure
is an attractive climate adaptation option since it also contributes to the
sustainable development of urban areas (Gill et al., 2007; James et al.,
2009). Urban green space and green roofs can moderate temperature
and decrease surface rainwater runoff (Gill et al., 2007). Despite the
benefits of urban green space, conflict can occur between the use of land
for green space and building developments (Hamin and Gurran, 2009).
European policies for biodiversity (e.g., the European Biodiversity Strategy
(EC, 2011)) look to spatial planning to help protect and safeguard
internationally and nationally designated sites, networks, and species,
as well as locally valued sites in urban and non-urban areas, and to create
new opportunities for biodiversity through the development process
(Wilson, 2008). Conservation planning in response to climate change
impacts on species aims to involve several strategies to better manage
isolated habitats, increase colonization capacity of new climate zones,
and optimize conservation networks to establish climate refugia (Vos
et al., 2008).
23.7.5. Rural Development
Rural development is one of the key policy areas for Europe, yet there
is little or no discussion about the role of climate change in affecting
future rural development. The EU White Paper on adapting to climate
change (EC, 2009a) encourages member states to embed climate
change adaptation into the three strands of rural development aimed
at improving competitiveness, the environment, and the quality of life
in rural areas. It appears however that little progress has been made in
achieving these objectives.
For example, the EUs Leader program was designed to help rural actors
improve the long-term potential of their local areas by encouraging the
implementation of sustainable development strategies. Many Leader
projects address climate change adaptation, but only as a secondary
or in many cases a non-intentional by-product of the primary rural
development goals. The World Bank’s community adaptation project has
seen a preponderance of proposals from rural areas in Eastern Europe
and Central Asia (Heltberg et al., 2012), suggesting that adaptation-
based development needs in Eastern Europe are currently not being
met by policy.
23.7.6. Economic Assessments of Adaptation
Compared to studies assessed in AR4 (WGII AR4 Section 17.2.3), cost
estimates for Europe are increasingly derived from bottom-up and
sector-specific studies, aimed at costing response measures (Watkiss and
Hunt, 2010), in addition to the economy-wide assessments (Aaheim et
al., 2012). The evidence base, however, is still fragmented and incomplete.
The coverage of adaptation costs and benefit estimates is dominated
by structural (physical) protection measures, where effectiveness and cost
components can be more easily identified. For energy, agriculture, and
infrastructure, there is medium coverage of cost and benefit categories.
There is a lack of information regarding adaptation costs in the health
and social care sector. Table 23-2 summarizes some of the more
comprehensive cost estimates for Europe for sectors at regional and
1298
Chapter 23 Europe
23
n
ational levels. It is stressed that the costing studies use a range of
methods and metrics and relate to different time periods and sectors,
which renders robust comparison difficult. As an example, there are
large differences between the cost estimates for coastal and river
protection in the Netherlands and other parts of Europe (Table 23-2),
which is due to the objectives for adaptation and the large differences
in the level of acceptable risk. For example, Rojas et al. (2013) assess a
1-in-100 year level of protection for Europe, while the Netherlands has
set standards up to 1-in-4000 and 10,000-year level return periods.
More detailed treatment of the economics of adaptation is provided in
Chapter 17.
23.7.7. Barriers and Limits to Adaptation
Implementation of adaptation options presents a range of opportunities,
constraints, and limits. Constrains (barriers) to implementation are financial,
technical, and political (see discussion in Chapter 16). Some impacts will
be unavoidable due to physical, technological, social, economic, or political
limits. Examples of limits in the European context are described by sector
in Table 23-3. For example, the contraints on building or extending flood
defenses would include pressure for land, conservation needs, and
amenity value of coastal areas (Section 5.5.6).
Toward the end of the century, it is likely that adaptation limits will be
reached earlier under higher rates of warming. Opportunities and co-
benefits of adaptation are also discussed in Section 23.8.
23.8. Co-Benefits and Unintended Consequences
of Adaptation and Mitigation
Scientific evidence for decision making is more useful if impacts are
considered in the context of impacts on other sectors and in relation to
adaptation, mitigation, and other important policy goals. The benefits
o
f adaptation and mitigation policies can be felt in the near term and
in the local population, although benefits relating to GHG emissions
reduction may not be apparent until the longer term. The benefits of
adaptation measures are often assessed using conventional economic
analyses, some of which include non-market costs and benefits
(externalities) (Watkiss and Hunt, 2010). This section describes policies,
strategies, and measures where there is good evidence regarding
mitigation/adaptation costs and benefits. Few studies have quantified
directly the trade-offs/synergies for a given policy.
23.8.1. Production and Infrastructure
Mitigation policies (decarbonization strategies) are likely to have
important implications for dwellings across Europe. The unintended
consequences of mitigation in the housing sector include changes to
household energy prices and adverse effects from decreased ventilation
in dwellings (Jenkins et al., 2008; Jenkins, 2009; Davies and Oreszczyn,
2012; Mavrogianni et al., 2012). The location, type, and dominant
energy use of the building will determine its overall energy gain or loss
to maintain comfort levels. Adaptation measures such as the use of
cooling devices will probably increase a building’s energy consumption
if no other mitigation measures are applied. The potential for cooling
dwellings without increased energy consumption, and with health
benefits is large (Wilkinson et al., 2009).
When looking at the broader context of urban infrastructures, despite
existing efforts to include both adaptation and mitigation into sustainable
development strategies at the city level (e.g., Hague, Rotterdam,
Hamburg, Madrid, London, Manchester), priority on adaptation still
remains low (Carter, 2011). There is potential to develop strategies that
can address both mitigation and adaptation solutions, as well as have
health and environmental benefits (Milner et al., 2012). In energy supply,
the adverse effect of climate change on water resources in some coastal
regions in Southern Europe may further enhance the development of
1299
Europe Chapter 23
23
d
esalination plants as an adaptation measure, possibly increasing
energy consumption and thus GHG emissions. Coastal flood defense
measures may alter vector habits and have implications for local vector-
borne disease transmission (Medlock and Vaux, 2013).
In tourism, adaptation and mitigation may be antagonistic, as in the
case of artificial snowmaking in European ski resorts, which requires
significant amounts of energy and water (OECD, 2007; Rixen et al.,
2011), and the case of desalination for potable water production, which
also requires energy. However, depending on the location and size of
the resort, implications are expected to differ and thus need to be
investigated on a case-by-case basis. A similar relationship between
adaptation and mitigation may hold for tourist settlements in Southern
Europe, where expected temperature increases during the summer may
require increased cooling to maintain tourist comfort and thus increase
GHG emissions and operating costs. Furthermore, a change of tourist
flows as a result of tourists adapting to climate change may affect
transport emissions, while mitigation in transport could also lead to a
change in transport prices and thus possibly affect tourist flows.
23.8.2. Agriculture, Forestry, and Bioenergy
Agriculture and forestry face two challenges under climate change,
both to reduce emissions and to adapt to a changing and more variable
climate (Lavalle et al., 2009; Smith and Olesen, 2010). The agriculture
sector contributes about 10% of the total anthropogenic GHG emissions
in the EU27 (EEA, 2010b). Estimates of European CO
2
, methane, and NO
x
fluxes between 2000and 2005 suggest that methane emissions from
livestock and NO
x
emissions from agriculture are fully compensated for
by the CO
2
sink provided by forests and by grassland soils (Schulze et
al., 2010). However, projections following a baseline scenario suggest
a significant decline (–25 to –40%) of the forest carbon sink of the EU
until 2030 compared to 2010. Using wood for bioenergy results initially
in a carbon debt due to reduced storage in forests, which affects the
net GHG balance depending on the energy type that is replaced and
the time span considered (McKechnie et al., 2011). Including additional
bioenergy targets of EU member states has an effect on the development
of the European forest carbon sink (and on the carbon stock), which is
not accounted for in the EU emission reduction target (Bottcher et al.,
2012).
In arable production systems, adapting to climate change by increasing
the resilience of crop yields to heat and to rainfall variability would
have positive impacts on mitigation by reducing soil erosion, as well as
soil organic carbon and nitrogen losses. Improving soil water holding
capacity through the addition of crop residues and manure to arable
soils, or by adding diversity to the crop rotations, may contribute both
to adaptation and to mitigation (Smith and Olesen, 2010). There are
also synergies and trade-offs between mitigation and adaptation options
for soil tillage, irrigation, and livestock breeding (Smith and Olesen,
2010). Reduced tillage (and no-till) may contribute to both adaptation
and mitigation as it tends to reduce soil erosion and runoff (Soane et
al., 2012) and fossil-fuel use (Khaledian et al., 2010), while increasing
in some situations soil organic carbon stock (Powlson et al., 2011).
However, increased N
2
O emission may negate the mitigation effect of
reduced tillage (Powlson et al., 2011). Irrigation may enhance soil carbon
s
equestration in arable systems (Rosenzweig and Tubiello, 2007;
Rosenzweig et al., 2008), but increased irrigation under climate change
would increase energy use and may reduce water availability for hydro-
power (reduced mitigation potential) (Wreford et al., 2010). In intensive
livestock systems, warmer conditions in the coming decades might
trigger the implementation of enhanced cooling and ventilation in farm
buildings (Rosenzweig and Tubiello, 2007), thereby increasing energy
use and associated GHG emissions. In grass-based livestock systems,
adaptation by adjusting the mean annual animal stocking density to
the herbage growth potential (Graux et al., 2012) is likely to create a
positive feedback on GHG emissions per unit area (Soussana and
Luscher, 2007; Soussana et al., 2010).
Land management options may also create synergies and trade-offs
between mitigation and adaptation. Careful adaptation of forestry and
soil management practices will be required to preserve a continental
ecosystem carbon sink in Europe (Schulze et al., 2010) despite the
vulnerability of this sink to climatic extremes (Ciais et al., 2005) and first
signs of carbon sink saturation in European forest biomass (Nabuurs et
al., 2013). In areas that are vulnerable to extreme events (e.g., fires,
storms, droughts) or with high water demand, the development of
bioenergy production from energy crops and from agricultural residues
(Fischer, G. et al., 2010; De Wit et al., 2011) could further increase demands
on adaptation (Wreford et al., 2010). Conversely, increased demands
on mitigation could be induced by the potential expansion of agriculture
at high latitudes, which may release large amounts of carbon and
nitrogen from organic soils (Rosenzweig and Tubiello, 2007).
23.8.3. Social and Health Impacts
Significant research has been undertaken since AR4 on the health co-
benefits of mitigation policies (see Chapter 11 and WGIII AR5 Chapters
7, 8, 9). Several assessments have quantified benefits in terms of lives
saved by reducing particulate air pollution. Policies that improve health
from changes in transport and energy can be said to have a general
benefit to population health and resilience (Haines et al., 2009a,b).
Changes to housing and energy policies also have indirect implications
for human health. Research on the benefits of various housing options
(including retrofitting) has been intensively addressed in the context of
low-energy, healthy, and sustainable housing (see WGIII AR5 Chapters
9, 12).
23.8.4. Environmental Quality and Biological Conservation
There are several conservation management approaches that can address
mitigation, adaptation, and biodiversity objectives (Lal et al., 2011).
Some infrastructure adaptation strategies—such as desalinization, sea
defenses, and flood control infrastructure—may have negative effects
on both mitigation and biodiversity. However, approaches, such as
forest conservation and urban green space (Section 23.7.4) have
multiple benefits and potentially significant effects. There has been
relatively little research about the impacts of future land use demand
for bioenergy production, food production, and urbanization on nature
conservation.
1300
Chapter 23 Europe
23
Figure 23-6 (Paterson et al., 2008) summarizes the evidence regarding
mitigation and adaptation options on biodiversity assessed from the
literature. The figure shows that the options that come closest to
being win-win-win are green rooftops, urban tree planting, forest
conservation, and low-till cultivation. Other options with clear benefits
are afforestation, forest pest control, increased farmland irrigation, and
species translocation.
23.9. Synthesis of Key Findings
23.9.1. Key Vulnerabilities
Climate change will have adverse impacts in nearly all sectors and across
all sub-regions. Table 23-4 describes the range of impacts projected in
2050 on infrastructure, settlements, environmental quality, and the
health and welfare of the European population. The projected impacts
of climate change on ecosystem services (including food production)
are described in Box 23-1. A key finding is that all sub-regions are
vulnerable to some impacts from climate change but these impacts differ
significantly in type between the sub-regions. Impacts in neighboring
regions (inter-regional) may also redistribute economic activities across
the European landscape. The sectors most likely to be affected by climate
change, and therefore with implications for economic activity and
population movement (changes in employment opportunities), include
tourism (Section 23.3.6), agriculture (Section 23.4.1), and forestry
(Section 23.4.4).
The majority of published assessments are based on climate projections
in the range of 1°C to 4°C global mean temperature per century. Under
these scenarios, regions in Europe may experience higher rates of
warming (in the range 1°C to 4°C per century), due to climate variability
(Jacob et al., 2013). Limited evidence exists on the potential impacts in
Europe under very high rates of warming (>4°C above preindustrial levels)
but these would lead to a large increase in coastal flood risk as well as
impacts on global cereal yields and other effects on the global economy
(Section 19.5.1).
Many key vulnerabilities are already well known since the AR4, but
some new vulnerabilities are emerging based on the evidence reviewed
in this report. The policy/governance context in Europe is extremely
important in determining (reducing or exacerbating) key vulnerabilities
since Europe is a highly regulated region. Further, vulnerability will be
strongly affected by changes in the non-climate drivers of change
(e.g., economic, social protection measures, governance, technological
drivers).
1301
Europe Chapter 23
23
Extreme events affect multiple sectors and have the potential to cause
systemic impacts from secondary effects (Chapter 19). Past events indicate
the vulnerability of transport, energy, agriculture, water resources, and
health systems. Resilience to very extreme events varies by sector, and
by country (Pitt, 2008; Ludwig et al., 2011; Ulbrich et al., 2012). Extreme
events (heat waves and droughts) have had significant impacts on
populations as well as multiple economic sectors (Table 23-1), and
resilience to future heat waves has been addressed only within some
sectors. However, there is surprisingly little evidence regarding the
impacts of major extreme events (e.g., Russian heat wave of 2010) and
Increasing
No change
Decreasing
A range from no change to increasing
A range from no change to decreasing
A range from increasing to decreasing
Green arrows mean a “beneficial change”
Red arrows means a “harmful change”
? means no relevant literature found
1302
Chapter 23 Europe
23
o
n responses implemented post-event to increase resilience. Future
vulnerability will also be strongly affected by cross-sectoral (indirect)
interactions, for example, flooding-ecosystems, agriculture-species,
agriculture-cultural landscapes, and so on.
Climate change is likely to have significant impacts on future water
availability, and the increased risks of water restrictions in Southern,
Central, and Atlantic sub-regions. Studies indicate a significant reduction
in water availability from river abstraction and from groundwater
resources, combined to increased demands from a range of sectors
(irrigation, energy and industry, domestic use) and to reduced water
drainage and runoff (as a result of increased evaporative demand)
(Ludwig et al., 2011).
Climate change will affect rural landscapes by modifying relative land
values, and hence competition, between different land uses (Smith et
al., 2010). This will occur directly, for example, through changes in the
productivity of crops and trees (Section 23.4), and indirectly through
climate change impacts on the global supply of land-based commodities
and their movement through international trade (Section 23.9.2).
Climate change will have a range of impacts in different European sub-
regions. The adaptive capacity of populations is likely to vary significantly
within Europe. Adaptive capacity indicators have been developed based
on future changes in socioeconomic indicators and projections (Metzger
et al., 2008; Lung et al., 2012; Acosta et al., 2013). These studies concluded
t
hat the Nordic countries have higher adaptive capacity than most of the
Southern European countries, with countries around the Mediterranean
having a lower capacity than the countries around the Baltic Sea region.
Some regions or areas are particularly vulnerable to climate change:
Populations and infrastructure in coastal regions are likely to be
adversely affected by sea level rise, particularly after mid-century
(Sections 23.3.1, 23.5.3).
Urban areas are also vulnerable to weather extremes owing to high
density of people and built infrastructure (Sections 23.3, 23.5.1).
Owing to high impact of climate change on natural hazard, and
water and snow resources, and the lack of migration possibilities
for plant species, mountain regions concentrate vulnerabilities in
infrastructure for transport and energy sectors, as well as for
tourism, agriculture, and biodiversity.
The Mediterranean region will suffer multiple stresses and systemic
failures due to climate changes. Changes in species composition,
increase of alien species, habitat losses, and degradation both in
land and sea together with agricultural and forests production
losses due to increasing heat waves and droughts exacerbated also
by the competition for water will increase vulnerability (Ulbrich et
al., 2012).
The following risks have emerged from observations of climate sensitivity
and observed adaptation:
There is new evidence to suggest that arable crop yields and
production may be more vulnerable as a result of increasing climate
1303
Europe Chapter 23
23
v
ariability. This will limit the potential poleward expansion of
agricultural production. Limits to genetic progress to adapt are
increasingly reported.
New evidence has emerged regarding implications during summer
on inland waterways (decreased access) and long-range ocean
transport (increased access).
Terrestrial and freshwater species are vulnerable from climate
change shifts in habitats. There is new evidence that species cannot
populate new habitat due to habitat fragmentation (urbanization).
Observed migration rates are less than that assumed in modeling
studies. There are legal barriers to introducing new species (e.g.,
forest species in France). New evidence reveals that phenological
mismatch will cause additional adverse effects on some species.
A positive (and emerging) effect that may reduce vulnerability is that
many European governments (and individual cities) have become
aware of the need to adapt to climate change and so are developing
and/or implementing adaptation strategies and measures.
Additional risks have emerged from the assessed literature:
Increased summer energy demand, especially in Southern Europe,
requires additional power generation capacity (underutilized during
the rest of the year), entailing higher supply costs.
Housing will be affected, with increased overheating under no
adaptation and damage from subsidence and flooding. Passive
cooling measures alone are unlikely to be sufficient to address
adaptation in all regions and types of buildings. Retrofitting current
housing stock will be expensive.
The vulnerability of cultural heritage, including monuments/buildings
and cultural landscapes, is an emerging concern. Some cultural
landscapes will disappear. Grape production is highly sensitive to
climate, but production (of grape varieties) is strongly culturally
dependent and adaptation is potentially limited by the regulatory
context.
There is strong evidence that climate change will increase the
distribution and seasonal activity of pests and diseases, and limited
evidence that such effects are already occurring. Increased threats
to plant and animal health are noted. Public policies are in place
to reduce pesticide use in agriculture use and antibiotics in livestock,
and this will increase vulnerability to the impact of climate change
on agriculture and livestock production.
Lack of institutional frameworks is a major barrier to adaptation
governance, in particular, the systematic failure in land use planning
policy to account for climate change.
23.9.2. Climate Change Impacts Outside Europe
and Inter-regional Implications
With increasing globalization, the impacts of climate change outside
the European region are likely to have implications for countries within
the region. For example, the Mediterranean region (Southern Europe and
non-European Mediterranean countries) has been considered highly
vulnerable to climate change (Navarra, 2013).
Eastern European countries have, in general, lower adaptive capacity
than Western or Northern European countries. The high volume of
international travel increases Europe’s vulnerability to invasive species,
i
ncluding the vectors of human and animal infectious diseases. The
transport of animals and animal products has facilitated the spread of
animal diseases (Conraths and Mettenleiter, 2011). Important “exotic”
vectors that have become established in Europe include the vector
Aedes albopictus (Becker, 2009; see also Section 23.5.1).
Another inter-regional implication concerns the changes in the location
of commercial fish stocks shared between countries. Such changes may
render existing international agreements regarding the sharing of yield
from these stocks obsolete, giving rise to international disputes (Arnason,
2012). For instance, the North Sea mackerel stock has recently been
extending westwards beyond the EU jurisdiction into the Exclusive
Economic Zones of Iceland and the Faroe Islands, which unilaterally
claimed quota for mackerel. Territorial disagreements of this type could
increase in the future with climate change.
Although several studies have proposed a role for climate change in
increasing migration pressures in low- and middle-income countries in
the future, there is little robust information regarding the respective
roles of climate change, environmental resource depletion, and weather
disasters in future inter-continental population movements. The effect
of climate change on external migration flows into Europe is highly
uncertain (see Section 12.4.1 for a more complete discussion). Modeling
future migration patterns is complex, and so far no robust approaches
have been developed.
23.9.3. Effects of Observed Climate Change in Europe
Table 23-6 summarizes the evidence with respect to key indicators in
Europe for the detection of a trend and the attribution of that trend to
observed changes in climate factors. The attribution of local warming
to anthropogenic climate change is less certain (see Chapter 18 for a
full discussion).
Further and better quality evidence since 2007 supports the conclusion
of AR4 (Alcamo et al., 2007) that climate change is affecting land,
freshwater, and marine ecosystems in Europe. Observed warming has
caused advancement in the life cycles of many animal groups, including
frogs spawning, birds nesting, and the arrival of migrant birds and
butterflies (see Chapter 4 and review by Feehan et al., 2009). There is
further evidence that observed climate change is already affecting
agricultural, forest, and fisheries productivity (see Section 23.4).
The frequency of river flood events, and annual flood and windstorm
damages, in Europe have increased over recent decades, but this increase
is attributable mainly to increased exposure and the contribution of
observed climate change is unclear (high confidence, based on robust
evidence, high agreement; SREX Section 4.5.3; Barredo, 2010).
The observed increase in the frequency of hot days and hot nights (high
confidence) is likely to have increased heat-related health effects in
Europe (medium confidence), as well as a decrease in cold-related health
effects (medium confidence; Christidis et al., 2010). Multiple impacts
on health, welfare, and economic sectors were observed due to the
major heat wave events of 2003 and 2010 in Europe (Table 23-1; see
Chapter 18 for discussion on attribution of events).
1304
Chapter 23 Europe
23
23.9.4. Key Knowledge Gaps and Research Needs
There is a clear mismatch between the volume of scientific work on
climate change since the AR4 and the insights and understanding
required for policy needs, as many categories of impacts are still under-
studied. Some specific research needs have been identified:
Little information is available on integrated and cross-sectoral
climate change impacts in Europe, as the impact studies typically
describe a single sector (see Sections 23.3-6). This also includes a
lack of information on cross-sector vulnerabilities, and the indirect
effects of climate change impacts and adaptation responses. This is
a major barrier in developing successful evidence-based adaptation
strategies that are cost-effective.
Climate change impact models are difficult to validate (Sections
23.3-6); proper testing of the characteristics of baseline impact
estimates against baseline information and data would improve
their reliability, or the development of alternative methods where
baseline data are not available.
There is little knowledge on co-benefits and unintended consequences
of adaptation options across a range of sectors (Sections 23.3-6).
There is a need to better monitor and evaluate local and national
adaptation and mitigation responses to climate change, in both
public and private sectors (Section 23.7; Box 23-3). This includes
policies and strategies—as well as the effectiveness of individual
adaptation measures. Evaluation of adaptation strategies, over a range
of time scales, would better support decision making. Although
some means for reporting of national actions exist in Europe (e.g.,
EU Climate-ADAPT), there is no consistent method of monitoring
or a mechanism for information exchange (Section 23.7).
There are now more economic methods and tools available for the
costing and valuation of specific adaptation options, in particular
for flood defenses, water, energy, and agriculture sectors (Section
23.7.6). However, for other sectors—such as biodiversity, business
and industry, and population health costs—cost estimates are still
lacking or incomplete. The usefulness of this costing information
in decision making needs to be evaluated and research can be
undertaken to make economic evaluation more relevant to decision
making.
The need for local climate information to inform decision making
also needs to be evaluated.
1305
Europe Chapter 23
23
Further research is needed on the effects of climate change on
critical infrastructure, including transport, water and energy supplies,
and health services (Section 23.5.2).
Further research is needed on the role of governance in adaptation
(local and national institutions) with respect to implementation of
measures in the urban environment, including flood defenses, over-
heating, and urban planning.
•The impacts from high end scenarios of climate change (>4°C global
average warming, with higher temperature change in Europe) are
n
ot yet known. Such scenarios have only recently become available,
and related impact studies still need to be undertaken for Europe.
More study of the implications for rural development would inform
policy in this area (Section 23.7.5). There is also a lack of information
on the resilience of cultural landscapes and communities, and how
to manage adaptation, particularly in low-technology (productively
marginal) landscapes.
More research is needed for the medium- and long-term monitoring
of forest responses and adaptation to climate change and on the
Fr
e
que
nt
l
y
As
k
e
d
Q
ue
s
t
i
ons
FAQ 23.1 | Will I still be able to live on the coast in Eur
ope?
Coa
s
t
a
l
a
r
e
a
s
a
ff
e
ct
e
d
by
s
t
or
m
s
ur
ge
s
w
i
l
l
f
a
ce
i
ncr
e
a
s
e
d
r
i
s
k
bot
h
be
ca
us
e
of
t
he
i
ncr
e
a
s
i
ng
f
r
e
que
ncy
of
s
t
or
m
s
a
nd
be
ca
us
e
of
hi
ghe
r
s
e
a
l
e
v
e
l
.
M
os
t
of
t
hi
s
i
ncr
e
a
s
e
i
n
r
i
s
k
w
i
l
l
occur
a
f
t
e
r
t
he
m
i
ddl
e
of
t
hi
s
ce
nt
ur
y
.
M
ode
l
s
of
t
h
e
co
a
s
t
lin
e
s
u
g
g
e
s
t
t
h
a
t
p
o
p
u
la
t
io
n
s
in
t
h
e
n
o
r
t
h
w
e
s
t
e
r
n
r
e
g
io
n
o
f
E
u
r
o
p
e
a
r
e
mo
s
t
a
f
f
e
ct
e
d
a
n
d
ma
n
y
co
u
n
t
r
ie
s
,
i
n
c
l
udi
ng
the
Netherl
ands,
Germ
any,
Franc
e,
Bel
gi
um
,
Den
m
ark,
Spai
n,
and
I
tal
y,
w
i
l
l
need
to
strengthen
thei
r
coastal
defenses.
S
om
e countri
es hav
e al
ready
rai
sed thei
r coastal
defense standards.
The com
bi
nati
on of rai
sed
sea defenses and coastal
erosi
on m
ay
l
ead to narrower coastal
z
ones i
n the N
orth S
ea,
the I
beri
an coast,
and the
Bay
of
Bi
scay
.
A
dapti
ng
dwel
l
i
ngs
and
com
m
erci
al
bui
l
di
ngs
to
occasi
onal
oodi
ng
i
s
another
response
to
cl
i
m
ate
change.
But though adapti
ng bui
l
di
ngs i
n coastal
com
m
uni
ti
es and upgradi
ng coastal
defenses can si
gni
cantl
y
reduce
adv
erse
i
m
pacts
of
sea
l
ev
el
ri
se
and
storm
surges,
they
cannot
el
i
m
i
nate
these
ri
sks,
especi
al
l
y
as
sea
l
ev
el
s
wi
l
l
conti
nue
to
ri
se
ov
er
ti
m
e.
I
n
som
e
l
ocati
ons,
“m
anaged
retreat”
i
s
l
i
kel
y
to
becom
e
a
necessary
response.
Fr
equent
l
y
As
ked
Ques
t
i
ons
FAQ 23.2 | Will climate change introduce new infectious diseases into Europe?
M
any
factors pl
ay
a rol
e i
n the i
ntroducti
on of i
nfecti
ous di
seases i
nto new areas.
F
actors that determ
i
ne whether
a di
sease changes di
stri
buti
on i
ncl
ude:
i
m
portati
on from
i
nternati
onal
trav
el
of peopl
e,
v
ectors or hosts (
i
nsects,
agri
cul
tural
products)
,
changes i
n v
ector or host suscepti
bi
l
i
ty
,
drug resi
stance,
and env
i
ronm
ental
changes,
such
as l
and use change or cl
i
m
ate change.
One area of concern that has gai
ned attenti
on i
s the potenti
al
for cl
i
m
ate
c
h
an
g
e
to
fac
ilitate
th
e
sp
read
o
f
tro
p
ic
al
d
iseases,
su
c
h
as
malaria,
in
to
Eu
ro
p
e.
Malaria
w
as
o
n
c
e
en
d
emic
in
Europe.
Ev
en though i
ts m
osqui
to v
ectors are sti
l
l
present and i
nternati
onal
trav
el
i
ntroduces fresh cases,
m
al
ari
a
ha
s not be
come
e
sta
blishe
d in E
urope
be
ca
use
infe
cte
d pe
ople
a
re
quick
ly de
te
cte
d a
nd tre
a
te
d. M
a
inta
ining good
heal
th
surv
ei
l
l
ance
and
good
heal
th
sy
stem
s
are
therefore
essenti
al
to
prev
ent
di
seases
from
spreadi
ng.
When
an
outbre
a
k
ha
s oc
c
urre
d (i.e
.,
the
introduc
tion of a new disease) determining the causes is often difficult. It is likely
tha
t a
c
ombina
tion of fa
c
tors will be
important. A suitable climate is a necessary but not a sufficient factor for the
introduc
tion of ne
w infe
c
tious dise
a
ses.
Freque
ntly Aske
d Q
uestions
FAQ 23.3 | Will Europe need to import more food because of climate change?
E
urope
is one
of the
world’s la
rge
st a
nd most productive suppliers of food, but also imports large amounts of some
agricultural commodities. A reduction in crop yields, particularly wheat in Southern Europe, is expected under fu
ture
climate scenarios. A shift in cultivation areas of high-value crops, such as grapes for wine, may also occur. Loss of
food production may be compensated by increases in other European sub-regions. However, if the capacity of the
European food production system to sustain climate shock events is exceeded, the region would require exceptional
food importation.
1306
Chapter 23 Europe
23
p
redictive modeling of wildfire distribution to better address
adaptation and planning policies. There is also a lack of information
on the impact of climate changes and climate extremes on carbon
sequestration potential of agricultural and forestry systems (Section
23.4.4).
More research is needed on impacts of climate change on transport,
especially on the vulnerability of road and rail infrastructure, and
on the contribution of climatic and non-climatic parameters in the
vulnerability of air transport (e.g., changes in air traffic volumes,
airport capacities, air traffic demand, weather at the airports of
origin, intermediate and final destination; Section 23.3.3).
Improved monitoring of droughts is needed to support the
management of crop production (Section 23.4). Remote sensing could
be complemented by field experiments that assess the combined
effects of elevated CO
2
and extreme heat and drought on crops and
pastures.
Research is needed on resilience of human populations to extreme
events (factors that increase resilience), including responses to flood
and heat wave risks. Research is also needed on how adaptation
policies may increase or reduce social inequalities (Section 23.5).
Improved risk models need to be developed for vector-borne disease
(human and animal diseases) to support health planning and
surveillance (Sections 23.4.2, 23.5.1).
A major barrier to research is lack of access to data, which is variable
across regions and countries (especially with respect to socioeconomic
data, climate data, forestry, and routine health data). There is a need
for long-term monitoring of environmental and social indicators and to
ensure open access to data for long-term and sustainable research
programs. Cross-regional cooperation could also ensure compatibility
and consistency of parameters across the European region.
References
Aaheim, A., H. Amundsen, T. Dokken, and T. Wei, 2012: Impacts and adaptation to
climate change in European economies. Global Environmental Change, 22(4),
959-968.
Aakre, S. and D.T.G. Rübbelke, 2010: Adaptation to climate change in the European
Union: efficiency versus equity considerations. Environmental Policy and
Governance, 20(3), 159-179.
Aakre, S., I. Banaszak, R. Mechler, D. Rübbelke, A. Wreford, and H. Kalirai, 2010:
Financial adaptation to disaster risk in the European Union: identifying roles
for the public sector. Mitigation and Adaptation Strategies for Global Change,
15(7), 721-736.
ABI, 2009: The Financial Risk of Climate Change: Examining the Financial Implications
of Climate Change Using Climate Models and Insurance Catastrophe Risk
Models [Dailey, P., M. Huddleston, S. Brown, and D. Fasking (eds.)]. ABI Research
Paper No. 19, Report by AIR Worldwide Corporation and the Met Office Hadley
Centre for the Association of British Insurers (ABI), ABI, London, UK, 107 pp.
Acevedo, P., F. Ruiz-Fons, R. Estrada, A.L. Márquez, M.A. Miranda, C. Gortázar, and
J. Lucientes, 2010: A broad assessment of factors in determining Culicoides
imicola abundance: modelling the present and forecasting its future in climate
change scenarios. PloS One, 5(12), e14236, doi:10.1371/journal.pone.0014236.
ACIA, 2005: Arctic Climate Impact Assessment. ACIA Scientific Report, Cambridge
University Press, New York, NY, USA, 1042 pp., www.acia.uaf.edu/pages/
scientific.html.
Acosta, L., R.J.T. Klein, P. Reitsma, M.J. Metzger, M.D.A. Rounsevell, R. Leemans, and
D. Schroter, 2013: A spatially explicit scenario-driven model of adaptive capacity
to global change in Europe. Global Environmental Change, 23(5), 1211-1224,
doi.org/10.1016/j.gloenvcha.2013.03.008.
Aerts, J., T. Sprong, B. Bannink, J. Bessembinder, E. Koomen, C. Jacobs, N. van der
Hoeven, D. Huitema, S. van ’t Klooster, J. Veraart, A. Walraven, S.N. Jonkman, B.
Maaskant, L.M. Bouwer, K. de Bruijn, E. Oosterveld, H. Schuurman, K. Peters,
W. Ottevanger, W. Immerzeel, P. Droogers, J. Kwadijk, J. Kind, L. Voogt, H. van
der Klis, R. Dellink, F. Affolter, P. Bubeck, M. van der Meulen, G. de Lange, B.
Bregman, H. van den Brink, H. Buiteveld, S. Drijfhout, A. Feijt, W. Hazeleger, B.
van den Hurk, C. Katsman, A. Kattenberg, G. Lenderink, E. Meijgaard, P. Siegmund,
M. de Wit, M. Naples, E. van Velzen, and J. van Zetten, 2008: Aandacht voor
Veiligheid [Aerts, J., T. Sprong, and B. Bannink (eds.)]. Leven met Water, Klimaat
voor Ruimte, DG Water, Netherlands, 198 pp. (in Dutch).
Affolter, P., U. Büntgen, J. Esper, A. Rigling, P. Weber, J. Luterbacher, and D. Frank,
2010: Inner Alpine conifer response to 20
th
century drought swings. European
Journal of Forest Research, 129, 289-298.
AGRESTE, 2011: Agreste Infos Rapides-Grandes Cultures et Fourrages, N°6/7: Les
Prairies Vues par ISOP en Septembre 2011 [Cassagne, J.P. (ed.)]. Ministère de
l’Agriculture, de l’Alimentation, de la Pêche, de la Ruraliet de l’Aménagement
du Territoire, Secrétariat Général, Montreuil, France, 4 pp. (in French).
Airoldi, L. and M.W. Bec, 2007: Loss, status and trends for coastal marine habitats
of Europe. Oceanography and Marine Biology Annual Review, 45, 345-405.
Albertson, K., J. Aylen, G. Cavan, and J. McMorrow, 2010: Climate change and the
future occurrence of moorland wildfires in the Peak District of the UK. Climate
Research, 45, 105-118.
Albrecht, F., T. Wahl, J. Jensen, and R. Weisse, 2011: Determining sea level change in
the German Bight. Ocean Dynamics, 61(12), 2037-2050.
Alcamo, J., J.M. Moreno, B. Novaky, M. Bindi, R. Corobov, R.J.N. Devoy, C.
Giannakopoulos, E. Martin, J.E. Olesen, and A. Shvidenko, 2007: Europe. 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 Yoork, NY, USA, pp. 541-580.
Alkemade, R., M. Bakkenes, and B. Eickhout, 2011: Towards a general relationship
between climate change and biodiversity: an example for plant species in
Europe. Regional Environmental Change, 11(1 Suppl.), S143-S150.
Allen, C.D., A.K. Macalady, H. Chenchouni, D. Bachelet, N. McDowell, M. Vennetier, T.
Kitzberger, A. Rigling, D.D. Breshears, E.H. Hogg, P. Gonzalez, R. Fensham, Z.
Zhang, J. Castro, N. Demidova, J. Lim, G. Allard, S.W. Running, A. Semerci, and
N. Cobb, 2010: A global overview of drought and heat-induced tree mortality
reveals emerging climate change risks for forests. Forest Ecology and
Management, 259(4), 660-684.
Amelung, B. and A. Moreno, 2009: Impacts of Climate Change in Tourism in Europe.
PESETA-Tourism Study. EUR 24114 EN 2009, JRC Scientific and Technical
Reports, Projection of Economic Impacts of Climate Change in Sectors of the
European Union (PESETA) Study, Joint Research Centre (JRC) Institute for
Prospective Technological Studies of the European Communities (EC), Publications
Office of the European Union, Luxembourg, Luxembourg, 44 pp.
Amelung, B. and A. Moreno, 2012: Costing the impact of climate change on tourism
in Europe: results of the PESETA project. Climatic Change, 112(1), 83-100.
Amelung, B., S. Nicholls, and D. Viner, 2007: Implications of global climate change
for tourism flows and seasonality. Journal of Travel Research, 45, 285-296.
Amstislavski, P., L. Zubov, H. Chen, P. Ceccato, J.F. Pekel, and I.J. Weedon, 2013: Effects
of increase in temperature and open water on transmigration and access to
health care by the Nenets reindeer herders in northern Russia. International
Journal of Circumpolar Health, 72(Suppl. 1), 21183, doi:10.3402/ijch.v72i0.21183.
Amthor, J.S., 2001: Effects of atmospheric CO
2
concentration on wheat yield:
review of results from experiments using various approaches to control CO
2
concentration. Field Crops Research, 73(1), 1-34.
Analitis, A., K. Katsouyanni, A. Biggeri, M. Baccini, B. Forsberg, L. Bisanti, U. Kirchmayer,
F. Ballester, E. Cadum, P.G. Goodman, A. Hojs, J. Sunyer, P. Tiittanen, and P.
Michelozzi, 2008: Effects of cold weather on mortality: results from 15 European
cities within the PHEWE project. American Journal of Epidemiology, 168(12),
1397-1408.
Analitis, A., I. Georgiadis, and K. Katsouyanni, 2012: Forestfiresare associated with
elevated mortality in a dense urban setting. Occupational and Environmental
Medicine, 69(3), 158-162.
Andersen, H.E., B. Kronvang, S. Larsen, C.C. Hoffmann, T.S. Jensen, and E.K. Rasmussen,
2006: Climate-change impacts on hydrology and nutrients in a Danish lowland
river basin. Science of the Total Environment, 365(1-3), 223-237.
1307
Europe Chapter 23
23
Andersson, A.K. and L. Chapman, 2011a: The impact of climate change in winter
road maintenance and traffic accidents in West Midlands, UK. Accident Analysis
and Prevention, 43(1), 284-289.
Andersson, A.K. and L. Chapman, 2011b: The use of a temporal analogue to predict
future traffic accidents and winter road conditions in Sweden. Meteorological
Applications, 18, 125-136.
André, G., B. Engel, P.B.M. Berentsen, T. Vellinga, and A.G.J.M. Oude Lansink, 2011:
Quantifying the effect of heat stress on daily milk yield and monitoring dynamic
changes using an adaptive dynamic model. Journal of Dairy Science, 94(9),
4502-4513.
AQEG, 2007: Air Quality and Climate Change: A UK Perspective.Third Report of the Air
Quality Expert Group (AQEG), prepared for the UK Department for Environment,
Food and Rural Affairs (DEFRA), Scottish Executive, Welsh Assembly Government,
and Department of the Environment in Northern Ireland, DEFRA, London, UK,
272 pp.
Aragòn, P. and J.M. Lobo, 2012: Predicted effect of climate change on the invasibility
and distribution of the western corn root-worm. Agricultural and Forest
Entomology, 14, 13-18.
Araújo, M.B., W. Thuiller, and R.G. Pearson, 2006: Climate warming and the decline
of amphibians and reptiles in Europe. Journal of Biogeography, 33, 1712-1728.
Araújo, M.B., D. Alagador, M. Cabeza, D. Nogués-Bravo, and W. Thuiller, 2011: Climate
change threatens European conservation areas. Ecology Letters, 14, 484-492.
Arca, B., G. Pellizzaro, P. Duce, M. Salis, V. Bacciu, D. Spano, A. Ager, and E. Scoccimarro,
2012: Potential changes in fire probability and severity under climate change
scenarios in Mediterranean areas. In: Modelling Fire Behaviour and Risk [Spano,
D., V. Bacciu, M. Salis, and C. Sirca (eds.)]. Nuova Stampa Color, Muros, Italy,
pp. 92-98.
Armstrong, B.G., Z. Chalab, B. Fenn, S. Hajat, S. Kovats, A. Milojevic, and P. Wilkinson,
2011: Association of mortality with high temperatures in a temperate climate:
England and Wales. Journal of Epidemology and Community Health, 65, 340-
345.
Arnason, R., 2012: Global warming: new challenges for the common fisheries policy?
Ocean & Coastal Management, 70, 4-9.
Arnell, N., 2011: Incorporating climate change into water resources planning in
England and Wales. Journal of the American Water Resources Association,
47(3), 541-549.
Artmann, N., D. Gyalistras, H. Manz, and P. Heiselberg, 2008: Impact of climate
warming on passive night cooling potential. Building Research & Information,
36(2), 111-128.
ARUP, 2011: Analysis of How Land Use Planning Decisions Affect Vulnerability to
Climate Risks. Final Report. Commissioned by the Adaptation Sub-Committee
of the Committee on Climate Change, Ove Arup and Partners Ltd., London, UK,
132 pp.
Arzt, J., W.R. White, B.V. Thomsen, and C.C. Brown, 2010: Agricultural diseases on
the move early in the third millennium. Veterinary Pathology, 47(1), 15-27.
Åström, D., B. Forsberg, and J. Rocklöv, 2011: Heat wave impact on morbidity and
mortality in the elderly population: a review of recent studies. Maturitas, 69(2),
99-105.
Åström, D., H. Orru, J. Rocklöv, G. Strandberg, K.L. Ebi, and B. Forsberg, 2013: Heat-
related respiratory hospital admissions in Europe in a changing climate: a
health impact assesssment. BMJ Open, 3, e001842, doi:10.1136/bmjopen-
2012-001842.
Avnery, S., D.L. Mauzerall, J. Liu, and L.W. Horowitz, 2011a: Global crop yield
reductions due to surface ozone exposure: 1. Year 2000 crop production losses
and economic damage. Atmospheric Environment, 45, 2284-2296.
Avnery, S., D.L. Mauzerall, J. Liu, and L.W. Horowitz, 2011b: Global crop yield
reductions due to surface ozone exposure: 2. Year 2030 potential crop production
losses and economic damage under two scenarios of O
3
pollution. Atmospheric
Environment, 45, 2297-2309.
BACC Author Team, 2008: Assessment of Climate Change for the Baltic Sea Basin.
Springer-Verlag, Berlin Heidelberg, Germany, 474 pp.
Baccini, M., T. Kosatsky, A. Analitis, H.R. Anderson, M. D’Ovidio, B. Menne, P.
Michelozzi, and A. Biggeri, 2011: Impact of heat on mortality in 15 European
cities: attributable deaths under different weather scenarios. Journal of
Epidemiology and Community Health, 65(1), 64-70.
Ballester, J., J. Robine, F.R. Herrmann, and X. Rodo, 2011: Long-term projections and
acclimatization scenarios of temperature-related mortality in Europe. Nature
Communications, 2(1), 358, doi:10.1038/ncomms1360.
Bangash, R.F., A. Passuello, M. Sanchez-Canales, M. Terrado, A. Lopez, F.J. Elorza, G.
Ziv, V. Acuna, and M. Schuhmacher, 2013: Ecosystem services in Mediterranean
river basin: climate change impact on water provisioning and erosion control.
The Science of the Total Environment, 458-460, 246-255.
Bank of Greece, 2011: The Environmental, Economic and Social Impacts of Climate
Change in Greece. Climate Change Impacts Study Committee, Bank of Greece,
Economic Research Department – Secretariat, Athens, Greece, 470 pp.
Barredo, J.I., 2009: Normalised flood losses in Europe: 1970-2006. Natural Hazards
and Earth System Sciences, 9(1), 97-104.
Barredo, J.I., 2010: No upward trend in normalised windstorm losses in Europe:
1970-2008. Natural Hazards and Earth System Sciences, 10(1), 97-104.
Barredo, J.I., D. Saurí, and M.C. Llasat, 2012: Assessing trends in insured losses from
floods in Spain 1971-2008. Natural Hazards and Earth System Science, 12(5),
1723-1729.
Barriopedro, D., E.M. Fischer, J. Luterbacher, R.M. Trigo, and R. García-Herrera, 2011:
The hot summer of 2010: redrawing the temperature record map of Europe.
Science, 332(6026), 220-224.
Barstad, I., A. Sorteberg, and M. Dos-Santos, 2012: Present and future offshore wind
power potential in Northern Europe based on downscaled global climate runs
with adjusted SST and sea ice cover. Renewable Energy, 44, 398-405.
Bastian, O., 2013: The role of biodiversity in supporting ecosystem services in Natura
2000 sites. Ecological Indicators, 24, 12-22.
Bastola, S., C. Murphy, and J. Sweeney, 2011: The sensitivity of fluvial flood risk in
Irish catchments to the range of IPCC AR4 climate change scenarios. Science
of the Total Environment, 409(24), 5403-5415.
Battaglini, A., G. Barbeau, M. Bindi, and F.W. Badeck, 2009: European winegrowers’
perceptions of climate change impact and options for adaptation. Regional
Environmental Change, 9(2), 61-73.
Beaugrand, G. and R.R. Kirby, 2010: Climate, plankton and cod. Global Change
Biology, 16(4), 1268-1280.
Beaugrand, G. and P.C. Reid, 2012: Relationships between North Atlantic salmon,
plankton, and hydroclimatic change in the Northeast Atlantic. ICES Journal of
Marine Science, 69(9), 1549-1562.
Beaugrand, G., P. Reid, F. Ibañez, J. Lindley, and M. Edwards, 2002: Reorganization
of North Atlantic marine copepod biodiversity and climate. Science, 296(5573),
1692-1694.
Beaugrand, G., M. Edwards, and L. Legendre, 2010: Marine biodiversity, ecosystem
functioning, and carbon cycles. Proceedings of the National Academy of
Sciences of the United States of America, 107(22), 10120-10124.
Becker, N., 2009: The impact of globalization and climate change on the development
of mosquitoes and mosquito-borne diseases in Central Europe [Die Rolle der
Globalisierung und Klimaveränderung auf die Entwicklung von Stechmücken und
von ihnen übertragenen Krankheiten in Zentral-Europa]. Umweltwissenschaften
und Schadstoff-Forschung, 21(2), 212-222.
Belkin, I.M., 2009: Rapid warming of large marine ecosystems. Progress in
Oceanography, 81, 207-213.
Benavente, D., P. Brimblecombe, and C.M. Grossi, 2008: Salt weathering and climate
change. In: New Trends in Analytical, Environmental and Cultural Heritage
Chemistry [Colombini, M.P. and L. Tassi (eds.)]. Transworld Research Network,
Trivandrum, Kerala, India, pp. 277-286.
Beniston, M., 2007: Entering into the ‘greenhouse century’: recent record temperatures
in Switzerland are comparable to the upper temperature quantiles in a
greenhouse climate. Geophysical Research Letters, 34(16), L16710, doi:10.1029/
2007GL030144.
Beniston, M., D.B. Stephenson, O.B. Christensen, C.A.T. Ferro, C. Frei, S. Goyette, K.
Halsnaes, T. Holt, K. Jylhä, B. Koffi, J. Palutikof, R. Schöll, T. Semmler, and K. Woth,
2007: Future extreme events in European climate: an exploration of regional
climate model projections. Climatic Change, 81(Suppl. 1), 71-95.
Berg, P., C. Moseley, and J.O. Haerter, 2013: Strong increase in convective precipitation
in response to higher temperatures. Nature Geoscience, 6(3), 181-185.
Bertini, G., T. Amoriello, G. Fabbio, and M. Piovosi, 2011: Forest growth and climate
change: evidences from the ICP-Forests intensive monitoring in Italy. Journal
of Biogeosciences and Forestry, 4, 262-267.
Bett, P.E., H.E. Thornton, and R.T. Clark, 2013: European wind variability over 140 yr.
Advances in Science and Research, 10, 51-58.
Biesbroek, G.R., R.J. Swart, T.R. Carter, C. Cowan, T. Henrichs, H. Mela, M.D. Morecroft,
and D. Rey, 2010: Europe adapts to climate change: comparing National
Adaptation Strategies. Global Environmental Change, 20(3), 440-450.
1308
Chapter 23 Europe
23
Bigler, C., O. Bräker, H. Bugmann, M. Dobbertin, and A. Rigling, 2006: Drought as an
inciting mortality factor in Scots pine stands of the Valais, Switzerland.
Ecosystems, 9(3), 330-343.
Bindi, M. and J.E. Olesen, 2011: The responses of agriculture in Europe to climate
change. Regional Environmental Change, 11(Suppl. 1), 151-158.
Bittner, M., E.F. Matthies, D. Dalbokova, and B. Menne, 2013: Are European countries
prepared for the next big heat-wave? European Journal of Public Health,
doi:10.1093/eurpub/ckt121.
Björdal, C.G., 2012: Evaluation of microbial degradation of shipwrecks in the Baltic
Sea. International Biodeterioration & Biodegradation, 70, 126-140.
Blaustein, A.R., S.C. Walls, B.A. Bancroft, J.J. Lawler, C.L. Searle, and S.S. Gervasi, 2010:
Direct and indirect effects of climate change on amphibian populations.
Diversity, 2(2), 281-313.
Bloom, A., V. Kotroni, and K. Lagouvardos, 2008: Climate change impact of wind
energy availability in the Eastern Mediterranean using the regional climate
model PRECIS. Natural Hazards and Earth System Sciences, 8(6), 1249-1257.
Bock, A., T. Sparks, N. Estrella, and A. Menzel, 2011: Changes in the phenology and
composition of wine from Franconia, Germany. Climate Research, 50, 69-81.
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.
Bonazza, A., P. Messina, C. Sabbioni, C.M. Grossi, and P. Brimblecombe, 2009a:
Mapping the impact of climate change on surface recession of carbonate
buildings in Europe. Science of the Total Environment, 407(6), 2039-2050.
Bonazza, A., C. Sabbioni, P. Messina, C. Guaraldi, and P. De Nuntiis, 2009b: Climate
change impact: mapping thermal stress on Carrara marble in Europe. Science
of the Total Environment, 407(15), 4506-4512.
Bondur, V.G., 2011: Satellite monitoring of wildfires during the anomalous heat wave
of 2010 in Russia. Izvestiya, Atmospheric and Oceanic Physics, 47(9), 1039-
1048.
Bormann, H., N. Pinter, and S. Elfert, 2011: Hydrological signatures of flood trends
on German rivers: flood frequencies, flood heights and specific stages. Journal
of Hydrology, 404, 50-66.
Bosello, F., R.J. Nicholls, J. Richards, R. Roson, and R.S.J. Tol, 2012: Economic impacts
of climate change in Europe: sea-level rise. Climatic Change, 112(1), 63-81.
Bottcher, H., P.J. Verkerk, M. Gusti, P. Havllk, and G. Grassi, 2012: Projection of the
future EU forest CO
2
sink as affected by recent bioenergy policies using two
advanced forest management models. GCB Bioenergy, 4(6), 773-783.
Botzen, W.J.W. and J.C.J.M. van den Bergh, 2008: Insurance against climate change
and flooding in the Netherlands: present, future, and comparison with other
countries. Risk Analysis, 28, 413-426.
Botzen, W.J.W., J.C.J.H. Aerts, and J.C.J.M. van den Bergh, 2009: Willingness of
homeowners to mitigate climate risk through insurance. Ecological Economics,
68, 2265-2277.
Botzen, W.J.W., J.C.J.M. van den Bergh, and L.M. Bouwer, 2010a: Climate change
and increased risk for the insurance sector: a global perspective and an
assessment for the Netherlands. Natural Hazards, 52, 577-598.
Botzen, W.J.W., L.M. Bouwer, and J.C.J.M. van den Bergh, 2010b: Climate change
and hailstorm damage: empirical evidence and implications for agriculture and
insurance. Resource and Energy Economics, 32(3), 341-362.
Bouwer, L.M., J.E. Vermaat, and J.C.J.H. Aerts, 2008: Regional sensitivities of mean
and peak river discharge to climate variability in Europe. Journal of Geophysical
Research, 113(D19), D19103, doi:10.1029/2008JD010301.
Bouwer, L.M., P. Bubeck, and J.C.J.H. Aerts, 2010: Changes in future flood risk due
to climate and development in a Dutch polder area. Global Environmental
Change, 20(3), 463-471.
Boxall, A., A. Hardy, S. Beulke, T. Boucard, L. Burgin, P.D. Falloon, P.M. Haygarth, T.
Hutchinson, S. Kovats, G. Leonardi, L.S. Levy, G. Nichols, S.A. Parsons, L. Potts,
D. Stone, E. Topp, D.B. Turley, K. Walsh, E.M.H. Wellington, and R.J. Williams,
2009: Impacts of climate change on indirect human exposure to pathogens
and chemicals from agriculture: Environmental Health Perspectives, 117(4),
508-514.
Bradley, B.A., D.M. Blumenthal, D.S. Wilcove, and L.H. Ziska, 2010: Predicting plant
invasions in an era of global change. Trends in Ecology and Evolution, 25(5),
310-318.
Branquart, E., K. Verheyen, and J. Latham, 2008: Selection criteria of protected forest
areas in Europe: the theory and the real world. Biological Conservation,
11(141), 2795-2806.
Breesch, H. and A. Janssens, 2010: Performance evaluation of passive cooling in
office buildings based on uncertainty and sensitivity analysis. Solar Energy,
84(8), 1453-1467.
Brijs, T., D. Karlis, and G. Wets, 2008: Studying the effect of weather conditions on
daily crash counts using a discrete time-series model. Accident Analysis and
Prevention, 40(3), 1180-1190.
Brimblecombe, P., 2010a: Heritage climatology. In: Climate Change and Cultural
Heritage [Lefèvre, R.-A. and C. Sabbioni (eds.)]. Edipuglia, Bari, Italy, pp. 49-56.
Brimblecombe, P., 2010b: Mapping heritage climatologies. In: Effect of Climate
Change on Built Heritage [Bunnik, T., H. de Clercq, R. van Hees, H. Schellen, and
L. Schueremans (eds.)]. Wissenschaftlich-Technische Arbeitsgemeinschaft für
Bauwerkserhaltung und Denkmalpflege (WTA) Publications, Pfaffenhofen,
Germany, pp. 18-30.
Brimblecombe, P. and C.M. Grossi, 2008: Millennium-long recession of limestone
facades in London. Environmental Geology, 56(3-4), 463-471.
Brimblecombe, P. and C.M. Grossi, 2009: Millennium-long damage to building
materials in London. Science of the Total Environment, 407(4), 1354-1361.
Brimblecombe, P. and C.M. Grossi, 2010: Potential damage to modern building
materials from 21
st
century air pollution. Scientific World Journal, 10, 116-125.
Brimblecombe, P., M.C. Grossi, and I. Harris, 2006: Climate change critical to cultural
heritage. In: Heritage Weathering and Conservation: Proceedings of the
International Heritage, Weathering and Conservation Conference (HWC-2006),
21-24 June 2006, Madrid, Spain – Two Volumes [Fort, R., M. Alvarez de Buergo,
M. Gomez-Heras, and C. Vazquez-Calvo (eds.)]. Taylor and Francis, London, UK,
pp. 387-393.
Briner, S., C. Elkin, R. Huber, and A. Gret-Regamy, 2012: Assessing the impacts of
economic and climate changes on land-use in mountain regions: spatial dynamic
modelling approach. Agriculture, Ecosystems and Environment, 149, 50-63.
Brisson, N., P. Gate, D. Gouache, G. Charmet, F. Oury, and F. Huard, 2010: Why are
wheat yields stagnating in Europe? A comprehensive data analysis for France.
Field Crops Research, 119(1), 201-212.
Brohan, P., J.J. Kennedy, I. Harris, S.F.B. Tett, and P.D. Jones, 2006: Uncertainty estimates
in regionaland global observed temperature changes: a new data set from
1850. Journal of Geophysical Research, 111(D12), D12106, doi:10.1029/
2005JD006548.
Brouwer, S., T. Rayner, and D. Huitema, 2013: Mainstreaming climate policy: the case
of climate adaptation and the implementation of EU water policy. Environment
and Planning C: Government and Policy, 31(1), 134-153.
Bryson, J., J. Piper, and M. Rounsevell, 2010: Envisioning futures for climate change
policy development: scenarios use in European environmental policy institutions.
Environmental Policy and Governance, 20(5), 283-294.
Bubeck, P., H. De Moel, L.M. Bouwer, and J.C.J. H. Aerts, 2011: How reliable are
projections of future flood damage? Natural Hazards and Earth System Science,
11(12), 3293-3306.
Buestel, D., M. Ropert, J. Prou, and Goulletquer, 2009: History, status and future of
oyster culture in France. Journal of Shellfish Research, 28(4), 813-820.
Bujosa, A. and J. Roselló, 2012: Climate change and summer mass tourism: the case
of Spanish domestic tourism. Climatic Change, 117(1-2), 363-375.
Bulkeley, H., 2010: Cities and the governing of climate change. Annual Review of
Environment and Resources, 35, 229-253.
Busch, G., 2006: Future European agricultural landscapes – what can we learn from
existing quantitative land use scenario studies? Agriculture, Ecosystems and
Environment, 114(1), 121-140.
Butchart, S.H.M., M. Walpole, B. Collen, A. Van Strien, J.P.W. Scharlemann, R.E.A.
Almond, J.E.M. Baillie, B. Bomhard, C. Brown, J. Bruno, K.E. Carpenter, G.M. Carr,
J. Chanson, A.M. Chenery, J. Csirke, N.C. Davidson, F. Dentener, M. Foster, A.
Galli, J.N. Galloway, P. Genovesi, R.D. Gregory, M. Hockings, V. Kapos, J.-F.
Lamarque, F. Leverington, J. Loh, M.A. McGeoch, L. McRae, A. Minasyan, M.
Hernández Morcillo, T.E.E. Oldfield, D. Pauly, S. Quader, C. Revenga, J.R. Sauer,
B. Skolnik, D. Spear, D. Stanwell-Smith, S.N. Stuart, A. Symes, M. Tierney, T.D.
Tyrrell, J.-C. Vié, and R. Watson, 2010: Global biodiversity: indicators of recent
declines. Science, 328(5982), 1164-1168.
Butterworth, M.H., M.A. Semenov, A. Barnes, D. Moran, J.S. West, and B.D.L. Fitt,
2010: North-South divide: contrasting impacts of climate change on crop yields
in Scotland and England. Journal of the Royal Society Interface, 7(42), 123-130.
Caffarra, A., M. Rinaldi, E. Eccela, V. Rossi, and I. Pertota, 2012: Modelling the impact
of climate change on the interaction between grapevine and its pests and
pathogens: European grapevine moth and powdery mildew. Agriculture,
Ecosystems & Environment, 148, 89-101.
1309
Europe Chapter 23
23
Calanca, P., A. Roesch, J. Karsten, and M. Wild, 2006: Global warming and the
summertime evapotranspiration regime of the Alpine region. Climatic Change,
79(1-2), 65-78.
Callaway, R., A.P. Shinn, S.E. Grenfell, J.E. Bron, G. Burnell, E.J. Cook, M. Crumlish, S.
Culloty, K. Davidson, R.P. Ellis, K.J. Flynn, C. Fox, D.M. Green, G.C. Hays, A.D.
Hughes, E. Johnston, C.D. Lowe, I. Lupatsch, S. Malham, A.F. Mendzil, T. Nickell,
T. Pickerell, A.F. Rowley, M.S. Stanley, D.R. Tocher, J.F. Turnbull, G. Webb, E.
Wootton, and R.J. Shields, 2012: Review of climate change impacts on marine
aquaculture in the UK and Ireland. Aquatic Conservation: Marine and Freshwater
Ecosystems, 22(3), 389-421.
Camia, A. and G. Amatulli, 2009: Weather factors and fire danger in the Mediterranean.
In: Earth Observation of Wildland Fires in Mediterranean Ecosystems [Chuvieco,
E. (ed.)]. Springer-Verlag, Berlin Heidelberg, Germany, pp. 71-82.
Caminade, C., J.M. Medlock, E. Ducheyne, K.M. McIntyre, S. Leach, M. Baylis, and
A.P. Morse, 2012: Suitability of European climate for the Asian tiger mosquito
Aedes Albopictus: recent trends and future scenarios. Journal of the Royal
Society Interface, 9(75), 2707-2717.
Camps, J.O. and M.C. Ramos, 2012: Grape harvest and yield responses to inter-annual
changes in temperature and precipitation in an area of north-east Spain with a
Mediterranean climate. International Journal of Biometeorology, 56(5), 853-864.
Cantarel, A.M., J.M.G. Bloor, and J. Soussana, 2013: Four years of simulated climate
change reduces above-ground productivity and alters functional diversity in a
grassland ecosystem. Journal of Vegetation Science, 24(1), 113-126.
Canu, D., C. Solidoro, G. Cossarini, and F. Giorgi, 2010: Effect of global change on
bivalve rearing activity and the need for adaptive management. Climate
Research, 42, 13-26.
Carmichael, C., G. Bickler, R.S. Kovats, D. Pencheon, V. Murray, C. West, and Y. Doyle,
2013: Overheating and hospitals: what do we know? Journal of Hospital
Administration, 2(1), doi:10.5430/jha.v2n1p1.
Carter, J.G., 2011: Climate change adaptation in European cities. Current Opinion in
Environmental Sustainability, 3(3), 193-198.
Carvalho, A., A. Monteiro, M. Flannigan, S. Solman, A.I. Miranda, and C. Borrego,
2011: Forest fires in a changing climate and their impacts on air quality.
Atmospheric Environment, 45(31), 5545-5553.
Casalegno, S., G. Amatulli, A. Bastrup-Birk, and T. Houston, 2007: Modelling current
and future distribution of European forest categories. In: Proceedings of the
6th European Conference on Ecological Modelling (ECEM ’07): “Challenges for
Ecological Modelling in a Changing World: Global Changes, Sustainability and
Ecosystem Based Management, 27-30, November 2007, Trieste, Italy. ECEM
‘07 Secretariat, Istituto Nazionale di Oceanografia di Geofisica Sperimentale
(OGS), Trieste Italy, pp. 1-2.
Castebrunet, H., N. Eckert, and G. Giraud, 2012: Snow and weather climatic control
on snow avalanche occurrence fluctuations over 50 yr in the French Alps.
Climate of the Past, 8(2), 855-875.
Castellari, S., 2009: Climate change, impacts and adaptation strategies in the Alpine
space: some results from the INTERREG III B Project ClimChAlp. In: Proceedings
of the International Conference “Mountains as Early Indicators of Climate
Change, 17-18 Aprile 2008, Padova, Italy. United Nations Environment
Programme (UNEP) Regional Office for Europe – Vienna, Vienna, Austria, pp.
81-91.
CEA, 2007: Reducing the Social and Economic Impact of Climate Change and Natural
Catastrophes: Insurance Solutions and Public-Private Partnerships. Report
prepared by CEA, Insurers of Europe, the European Insurance and Reinsurance
Federation (now Insurance Europe), Brussels, Belgium, 47 pp.
CEA, 2009: Tackling Climate Change: The Vital Contribution of Insurers. Report
prepared by CEA, Insurers of Europe, the European Insurance and Reinsurance
Federation (now Insurance Europe), Brussels, Belgium, 63 pp.
Cellamare, M., M. Leitao, M. Coste, A. Dutartre, and J. Haury, 2010: Tropical
phytoplankton taxa in Aquitaine lakes (France). Hydrobiologia, 639(1), 129-145.
Charles, E., D. Idier, J. Thiebot, G. Le Cozannet, R. Pedreros, F. Ardhuin, and S. Planton,
2012: Present wave climate in the Bay of Biscay: spatiotemporal variability and
trends from 1958 to 2001. Journal of Climate, 25, 2020-2035.
Charlton, M.B. and N.W. Arnell, 2011: Adapting to climate change impacts on water
resources in England – an assessment of draft Water Resources Management
Plans. Global Environmental Change, 21(1), 238-248.
Charru, M., I. Seynave, F. Morneau, and J. Bontemps, 2010: Recent changes in forest
productivity: an analysis of national forest inventory data for common beech
(Fagus sylvatica L.) in north-eastern France. Forest Ecology and Management,
260, 864-874.
Chatterton, J., C. Viavattene, J. Morris, E. Penning-Rowsell, and S. Tapsell, 2010:
The Costs of the Summer 2007 Floods in England.Project: SC070039/R1,
Environment Agency (EA), Bristol, UK, 41 pp.
Chauveau, M., S. Chazot, C. Perrin, P.-Y. Bourgin, E. Sauquet, J.-P. Vidal, N. Rouchy, E.
Martin, J. David, T. Norotte, P. Maugis, and X. de Lacaze, 2013: What impacts of
climate change on surface hydrology in France by 2070?La Houille Blanche,
(4), 5-15.
Cheaib, A., V. Badeau, J. Boe, I. Chuine, C. Delire, E. Dufrêne, C. François, E.S. Gritti,
M. Legay, C. Pagé, W. Thuiller, N. Viovy, and P. Leadley, 2012: Climate change
impacts on tree ranges: model intercomparison facilitates understanding and
quantification of uncertainty. Ecology Letters, 15(6), 533-544.
Cheung, W.W.L., V.W.Y. Lam, J.L. Sarmiento, K. Kearney, R. Watson, D. Zeller, and D.
Pauly, 2010: Large-scale redistribution of maximum fisheries catch potential in
the global ocean under climate change. Global Change Biology, 16(1), 24-35.
Cheung, W.W.L., J. Pinnegar, G. Merino, M.C. Jones, and M. Barange, 2012: Review
of climate change impacts on marine fisheries in the UK and Ireland. Aquatic
Conservation: Marine and Freshwater Ecosystems, 22(3), 368-388.
Cheung, W.W.L., J.L. Sarmiento, J. Dunne, T.L. Frölicher, V.W.Y. Lam, M.L.D. Palomares,
R. Watson, and D. Pauly, 2013: Shrinking of fishes exacerbates impacts of global
ocean changes on marine ecosystems. Nature Climate Change, 3(3), 254-258.
Chevalier, V., M. Pepin, L. Plee, and R. Lancelot, 2010: Rift Valley fever – a threat for
Europe? Eurosurveillance, 15(10), 18-28.
Chiriacò, M.V., L. Perugini, D. Cimini, E. D’Amato, R. Valentini, G. Bovio, P. Corona,
and A. Barbati, 2013: Comparison of approaches for reporting forest fire-related
biomass loss and greenhouse gas emissions in southern Europe. International
Journal of Wildland Fire, 22(6), 730-738.
Choat, B., S. Jansen, T.J. Brodribb, H. Cochard, S. Delzon, R. Bhaskar, S.J. Bucci, T.S.
Feild, S.M. Gleason, U.G. Hacke, A.L. Jacobsen, F. Lens, H. Maherali, J. Martínez-
Vilalta, S. Mayr, M. Mencuccini, P.J. Mitchell, A. Nardini, J. Pittermann, R.B. Pratt,
J.S. Sperry, M. Westoby, I.J. Wright, and A.E. Zanne, 2012: Global convergence
in the vulnerability of forests to drought. Nature, 491(7426), 752-755.
Chow, D.H. and G.J. Levermore, 2010: The effects of future climate change on heating
and cooling demands in office buildings in the UK. Building Services Engineering
Research and Technology, 31(4), 307-323.
Christidis, N., G.C. Donaldson, and P.A. Stott, 2010: Causes for the recent changes
in cold- and heat-related mortality in England and Wales. Climatic Change,
102(3-4), 539-553.
Christierson, B.V., J. Vidal, and S.D. Wade, 2012: Using UKCP09 probabilistic climate
information for UK water resource planning. Journal of Hydrology, 424-425,
48-67.
Ciais, P., M. Reichstein, N. Viovy, A. Granier, J. Ogee, V. Allard, M. Aubinet, N.
Buchmann, C. Bernhofer, A. Carrara, F. Chevallier, N. De Noblet, A.D. Friend, P.
Friedlingstein, T. Grunwald, B. Heinesch, P. Keronen, A. Knohl, G. Krinner, D.
Loustau, G. Manca, G. Matteucci, F. Miglietta, J.M. Ourcival, D. Papale, K.
Pilegaard, S. Rambal, G. Seufert, J.F. Soussana, M.J. Sanz, E.D. Schulze, T. Vesala,
and R. Valentini, 2005: Europe-wide reduction in primary productivity caused
by the heat and drought in 2003. Nature, 437(7058), 529-533.
Ciavola, P., O. Ferreira, P. Haerens, M. Van Koningsveld, and C. Armaroli, 2011: Storm
impacts along European coastlines. Part 2: lessons learned from the MICORE
project. Environmental Science & Policy, 14(7), 924-933.
Ciscar, J.-C., A. Iglesias, L. Feyen, C.M. Goodess, L. Sza, O.B. Christensen, R. Nicholls,
B. Amelung, P. Watkiss, F. Bosello, R. Dankers, L. Garrote, A. Hunt, L. Horrocks,
M. Moneo, A. Moreno, S. Pye, S. Quiroga, D. van Regemorter, J. Richards, R.
Roson, and A. Soria, 2009: Climate Change Impacts in Europe. Final Report of
the PESETA Research Project. EUR 24093 EN, EUR Scientific and Technical
Research Series, Projection of Economic Impacts of Climate Change in Sectors
of the European Union (PESETA) Project, Joint Research Centre (JRC) – Institute
for Prospective Technological Studies (IPTS) of the European Communities (EC),
Publications Office of the European Union, Luxembourg, Luxembourg, 114 pp.
Ciscar, J.-C., A. Iglesias, L. Feyen, L. Szabó, D. Van Regemorter, B. Amelung, R. Nicholls,
P. Watkiss, O.B. Christensen, R. Dankers, L. Garrote, C.M. Goodess, A. Hunt, A.
Moreno, J. Richards, and A. Soria, 2011: Physical and economic consequences
of climate change in Europe. Proceedings of the National Academy of Sciences
of the United States of America, 108(7), 2678-2683.
Civantos, E., W. Thuiller, L. Maiorano, A. Guisan, and M. Araujo, 2012: Potential
impacts of climate change on ecosystem services in Europe: the case of pest
control by vertebrates. BioScience, 62, 658-666.
Clemo, K., 2008: Preparing for climate change: insurance and small business. Geneva
Papers on Risk and Insurance: Issues and Practice, 33(1), 110-116.
1310
Chapter 23 Europe
23
Cogan, D.G., 2008: Corporate Governance and Climate Change: The Banking Sector.
A Ceres Report, Prepared for Ceres by the Climate Change Research Team,
RiskMetrics Group, Ceres, Boston, MA, USA, 58 pp.
Conraths, F.J. and T.C. Mettenleiter, 2011: Globalisation and change of climate:
growing risk for livestock epidemics in Germany [Globalisierung und
klimawandel: steigendes risiko für tierseuchen in Deutschland]. Zuchtungskunde,
83(1), 21-26.
Corobov, R., S. Sheridan, N. Opopol, and K. Ebi, 2012: Heat-related mortality in
Moldova: the summer of 2007. International Journal of Climatology, 33(11),
2551-2560.
Corobov, R., S. Sheridan, K. Ebi, and N. Opopol, 2013: Warm season temperature-
mortality relationships in Chisinau (Moldova). International Journal of
Atmospheric Sciences, 2013, 346024, doi:10.1155/2013/346024.
Corti, T., V. Muccione, P. Kollner-Heck, D. Bresch, and S.I. Seneviratne, 2009: Simulating
past droughts and associated building damages in France. Hydrology and Earth
System Sciences, 13(9), 1739-1747.
Coumou, D. and S. Rahmstorf, 2012: A decade of weather extremes. Nature Climate
Change, 2(7), 491-496.
Crescio, M.I., F. Forastiere, C. Maurella, F. Ingravalle, and G. Ru, 2010: Heat-related
mortality in dairy cattle: a case crossover study. Preventative Veterinary
Medicine, 97, 191-197.
Crichton, D., 2006: Climate Change and its Effects on Small Business in the UK.
Commissioned and published by AXA Insurance UK, London, UK, 41 pp.
Crichton, D., 2007: The Hull Floods of June 2007. Some Insurance Industry
Implications. Benfield UCL Hazard Research Centre, Univeristy College London
(UCL), London, UK.
Crook, J.A., L.A. Jones, P.M. Forster, and R. Crook, 2011: Climate change impacts on
future photovoltaic and concentrated solar power energy output. Energy &
Environmental Science, 4(9), 3101-3109.
Crump, D., A. Dengel, and M. Swainson, 2009: Indoor Air Quality in Highly Energy
Efficient Homes – A Review. NF18, National House-Building Council (NHBC)
Foundation, IHS BRE Press, Bracknell, UK, 82 pp.
Daccache, A. and N. Lamaddalena, 2010: Climate change impacts on pressurised
irrigation systems. Proceedings of the Institution of Civil Engineers-Engineering
Sustainability, 163(2), 97-105.
Daccache, A., C. Keay, R.J.A. Jones, E.K. Weatherhead, M.A. Stalham, and J.W. Knox,
2012: Climate change and land suitability for potato production in England
and Wales: impacts and adaptation. Journal of Agricultural Science, 150(2),
161-177.
Damigos, D., 2012: Monetizing the impacts of climate change on the Greek mining
sector. Mitigation and Adaptation Strategies for Global Change, 17(8), 865-
878.
Dammers, E., 2010: Making territorial scenarios for Europe. Futures, 42, 785-793.
Dankers, R. and L. Feyen, 2008: Climate change impact on flood hazard in Europe:
an assessment based on high-resolution climate simulations. Journal of
Geophysical Research, 113(D19), D19105, doi:10.1029/2007JD009719.
Dankers, R., O.B. Christensen, L. Feyen, M. Kalas, and A. de Roo, 2007: Evaluation of
very high-resolution climate model data for simulating flood hazards in the
Upper Danube Basin. Journal of Hydrology, 347(3-4), 319-331.
Daufresne, M., P. Bady, and J.F. Fruget, 2007: Impacts of global changes and extreme
hydroclimatic events on macroinvertebrate community structures in the French
Rhône River. Oecologia, 151(3), 544-559.
Daufresne, M., K. Lengfellner, and U. Sommer, 2009: Global warming benefits the
small in aquatic ecosystems. Proceedings of the National Academy of Sciences
of the United States of America, 106(31), 12788-93.
Davies, M. and T. Oreszczyn, 2012: The unintended consequences of decarbonising
the built environment: a UK case study. Energy and Buildings, 46, 80-85.
Davoudi, S., M. Wishardt, and I. Strange, 2010: The ageing of Europe: demographic
scenarios of Europe’s futures. Futures, 42, 794-803.
Dawson, R.J., T. Ball, J. Werritty, A. Werritty, J.W. Hall, and N. Roche, 2011: Assessing
the effectiveness of non-structural flood management measures in the Thames
Estuary under conditions of socio-economic and environmental change. Global
Environmental Change, 21, 628-646.
Day, A.R., P.G. Jones, and G.G. Maidment, 2009: Forecasting future cooling demand
in London. Energy and Buildings, 41(9), 942-948.
Day, J.W., R.R. Christian, D.M. Boesch, A. Yáñez-Arancibia, J. Morris, R.R. Twilley, L.
Naylor, L. Schaffner, and C. Stevenson, 2008: Consequences of climate change
on the ecogeomorphology of coastal wetlands. Estuaries and Coasts, 31(3),
477-491.
DCLG, 2012: Investigation into Overheating in Homes: Literature Review. UK
Department of Communities and Local Government, London, UK, 124 pp.
De Freitas, C.R., D. Scott, and G. McBoyle, 2008: A second generation Climate Index
for Tourism (CIT): specification and verification. International Journal of
Biometeorology, 52(5), 399-407.
de Graaff, M., C. Van Kessel, and J. Six, 2009: Rhizodeposition-induced decomposition
increases N availability to wild and cultivated wheat genotypes under elevated
CO
2
. Soil Biology & Biochemistry, 41(6), 1094-1103.
de Moel, H., J. van Alphen, and J.C.J.H. Aerts, 2009: Flood maps in Europe methods,
availability and use. Natural Hazards and Earth System Sciences, 9(2), 289-
301.
de Mooij, R. and P. Tang, 2003: Four Futures of Europe. Centraal Planbureau (CPB),
The Hague, Netherlands, 220 pp.
De Wit, M., M. Londo, and A. Faaij, 2011: Productivity developments in European
agriculture: relations to and opportunities for biomass production. Renewable
and Sustainable Energy Reviews, 15(5), 2397-2412.
Debernard, J.B. and L.P. Rÿed, 2008: Future wind, wave and storm surge climate in
the Northern Seas: a revisit. Tellus A, 60(3), 427-438.
del Barrio, G., P.A. Harrison, P.M. Berry, N. Butt, M.E. Sanjuan, R.G. Pearson, and T.
Dawson, 2006: Integrating multiple modelling approaches to predict the
potential impacts of climate change on species’ distributions in contrasting
regions: comparison and implications for policy. Environmental Science and
Policy, 9(2), 129-147.
Della Bella, V., M. Bazzanti, M.G. Dowgiallo, and M. Iberite, 2008: Macrophyte
diversity and physico-chemical characteristics of Tyrrhenian coast ponds in
central Italy: implications for conservation. Hydrobiologia, 597(1), 85-95.
Dell’Aquila, A., S. Calmanti, P.M. Ruti, M.V. Struglia, G. Pisacane, A. Carillo, and G.
Sannino, 2012: Effects of seasonal cycle fluctuations in an A1B scenario over
the Euro-Mediterranean region. Climate Research, 52, 135-157.
Delpla, I., E. Baurès, A.-V. Jung, and O. Thomas, 2011: Impacts of rainfall events on
runoff water quality in an agricultural environment in temperate areas. Science
of the Total Environment, 409, 1683-1688.
Delta Committee, 2008: Working Together With Water. A Living Land Builds for its
Future [ten Brinke, W.B.M. (ed.)]. Findings of the Deltacommissie 2008,
commissioned by the Dutch Secretary of Public Works and Water Management,
Secretariat Delta Committee, The Hague, Netherlands, 134 pp.
Dessai, S. and M. Hulme, 2007: Assessing the robustness of adaptation decisions to
climate change uncertainties: a case study on water resources management in
the East of England. Global Environmental Change, 17(1), 59-72.
Devictor, V., R. Julliard, D. Couvet, and F. Jiguet, 2008: Birds are tracking climate
warming, but not fast enough. Proceedings of the Royal Society B, 275, 2743-
2748.
D’Ippoliti, D., P. Michelozzi, C. Marino, F. de’Donato, B. Menne, K. Katsouyanni, U.
Kirchmayer, A. Analitis, M. Medina-Ramon, A. Paldy, R. Atkinson, S. Kovats, L.
Bisanti, A. Schneider, A. Lefranc, C. Iniguez, and C.A. Perucci, 2010: The impact of
heat waves on mortality in 9 European cities: results from the EuroHEAT project.
Environmental Health: A Global Access Science Source, 9, 37, doi:10.1186/
1476-069X-9-37.
Dixon, N. and E. Brook, 2007: Impact of predicted climate change on landslide
reactivation: case study of Mam Tor. UK Landslides, 4, 137-147.
Dobney, K., C.J. Baker, L. Chapman, and A.D. Quinn, 2010: The future cost to the
United Kingdom’s railway network of heat-related delays and buckles caused
by the predicted increase in high summer temperatures owing to climate
change. Proceedings of the Institution of Mechanical Engineers, Part F. Journal
of Rail and Rapid Transit, 224(1), 25-34.
Dolinar, M., B. Vidrih, L. Kajfež-Bogataj, and S. Medvec, 2010: Predicted changes in
energy demands for heating and cooling due to climate change. Physics and
Chemistry of the Earth, 35(1-2), 100-106.
Donat, M.G., G.C. Leckebusch, J.G. Pinto, and U. Ulbrich, 2010: European storminess
and associated circulation weather types: future changes deduced from a multi-
model ensemble of GCM simulations. Climate Research, 42(1), 27-43.
Donat, M.G., G.C. Leckebusch, S. Wild, and U. Ulbrich, 2011: Future changes in
European winter storm losses and extreme wind speeds inferred from GCM
and RCM multi-model simulations. Natural Hazards and Earth System Sciences,
11(5), 1351-1370.
Donatelli, M., A.K. Srivastava, G. Duveiller, and S. Niemeyer, 2012: Estimating impact
assessment and adaptation strategies under climate change scenarios for crops
at EU27 scale. In: International Environmental Modelling and Software Society
(iEMSs) 2012 International Congress on Environmental Modelling and Software,
1311
Europe Chapter 23
23
“Managing Resources of a Limited Planet: Pathways and Visions under
Uncertainty,Sixth Biennial Meeting, 1-5 July 2012, Leipzig, Germany [Seppelt,
R., A.A. Voinov, S. Lange, and D. Bankamp (eds.)]. International Environmental
Modelling and Software Society (iEMSs) Secretariat, Manno, Switzerland pp.
404-411, www.iemss.org/sites/iemss2012/proceedings.html.
Doney, S.C., M. Ruckelshaus, J. Emmett Duffy, J.P. Barry, F. Chan, C.A. English, H.M.
Galindo, J.M. Grebmeier, A.B. Hollowed, N. Knowlton, J. Polovina, N.N. Rabalais,
W.J. Sydeman, and L.D. Talley, 2011: Climate change impacts on marine
ecosystems. Annual Review of Marine Science, 4, 11-37.
Drenkhan, R., T. Kurkela, and M. Hanso, 2006: The relationship between the needle
age and the growth rate in Scots pine (Pinus sylvestris): a retrospective analysis
by needle trace method (NTM). European Journal of Forest Research, 125, 397-
405.
Dronin, N. and A. Kirilenko, 2011: Climate change, food stress, and security in Russia.
Regional Environmental Change, 11(Suppl. 1), 167-178.
Duarte Alonso, A. and M.A. O’Neill, 2011: Climate change from the perspective of
Spanish wine growers: a three-region study. British Food Journal, 113(2), 205-
221.
Ducharne, A., C. Baubion, N. Beaudoin, M. Benoit, G. Billen, N. Brisson, J. Garnier, H.
Kieken, S. Lebonvallet, E. Ledoux, B. Mary, C. Mignolet, X. Poux, E. Sauboua, C.
Schott, S. Thery, and P. Viennot, 2007: Long term prospective of the Seine River
system: confronting climatic and direct anthropogenic changes. Science of the
Total Environment, 375(1-3), 292-311.
Ducharne, A., F. Habets, C. Pagé, E. Sauquet, P. Viennot, M. qué, S. Gascoin, A.
Hachour, E. Martin, L. Oudin, L. Terray, and D. Thiéry, 2010: Climate change
impacts on water resources and hydrological extremes in northern France. In:
Proceedings of XVIII International Conference on Computational Methods in
Water Resources, CIMNE, 21-24, June 2010, Barcelona, Spain [Carrera, J. (ed.)].
The International Center for Numerical Methods in Engineering (CIMNE),
Technical University of Catalonia (UPC), Barcelona, Spain, 8 pp.,
congress.cimne.com/cmwr2010/Proceedings/Start.html.
Duchêne, E., F. Huard, V. Dumas, C. Schneider, and D. Merdinoglu, 2010: The challenge
of adapting grapevine varieties to climate change. Climate Research, 41(3),
193-204.
Dullinger, S., A. Gattringer, W. Thuiller, D. Moser, N.E. Zimmermann, A. Guisan, W.
Willner, C. Plutzar, M. Leitner, T. Mang, M. Caccianiga, T. Dirnböck, S. Ertl, A.
Fischer, J. Lenoir, J. Svenning, A. Psomas, D.R. Schmatz, U. Silc, P. Vittoz, and K.
Hülber, 2012: Extinction debt of high-mountain plants under twenty-first-
century climate change. Nature Climate Change, 2(8), 619-622.
Dumollard, G. and A. Leseur, 2011: Drawing Up A National Adaptation Policy:
Feedback on Five European Case Studies. Climate Report: Research on the
Economics of Climate Change, Climate Study No. 27, CDC Climat Research,
Paris, France, 32 pp.
Durant, J.M., D.O. Hjermann, G. Ottersen, and N.C. Stenseth, 2007: Climate and the
match or mismatch between predator requirements and resource availability.
Climate Research, 33(3), 271-283.
Dury, M., A. Hambuckers, P. Warnant, A. Henrot, E. Favre, M. Ouberdous, and L.
François, 2011: Responses of European forest ecosystems to 21
st
century
climate: assessing changes in interannual variability and fire intensity. iForest,
4, 82-99.
EA, 2009: Thames Estuary 2100: Managing Flood Risk through London and the
Thames Estuary. TE2100 Plan Consultation Document. Thames Estuary 2100,
Environmental Agency (EA), London, UK, 214 pp.
EA, 2011: TE2100 Strategic Outline Programme. Thames Estuary 2100, Environment
Agency (EA), London, UK.
Easterling, W.E., P.K. Aggarwal, P. Batima, K.M. Brander, L. Erda, S.M. Howden, A.
Kirilenko, J. Morton, J. Soussana, J. Schmidhuber, and F.N. Tubiello, 2007: Food,
fibre and forest products. In: Climate Change 2007: Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani,
J.P. Palutikof, P.J. van der Linden, and C.E. Hanson (eds.)]. Cambridge University
Press, Cambridge, UK and New York, NY, USA, pp. 273-313.
EC, 2009a: White Paper: Adapting to Climate Change: Towards a European Framework
for Action. Brussels, 1.4.2009, COM(2009) 147 Final, European Commission
(EC), Brussels, Belgium, 16 pp.
EC, 2009b: Guidance Document No. 24: River Basin Management in a Changing
Climate. Common Implementation Strategy for the Water Framework Directive
(2000/60/EC), European Commission (EC), Publications Office of the European
Union, Luxembourg, Luxembourg, 132 pp.
EC, 2011: Our Life Insurance, Our Natural Capital: An EU Biodiversity Strategy to
2020.Brussels, 3.5.2011 COM(2011) 244 Final, Communication from the
Commission to the European Parliament, The Council, The Economic and Social
Committee and The Committee of the Regions, European Commission (EC),
Brussels, Belgium, 16 pp.
EC, 2013a: Climate-Adapt. European Climate Adaptation Platform.The European
Climate Adaptation Platform (CLIMATE-ADAPT), a partnership between the
European Commission (DG CLIMA, DG Joint Research Centre and other DGs)
and the European Environment Agency (EEA). European Commission (EC),
Brussels, Belguim, climate-adapt.eea.europa.eu/.
EC, 2013b: An EU Strategy on Adaptation to Climate Change. Brussels, 16.4.2013,
COM(2013) 216 Final, Communication from the Commission to the European
Parliament, The Council, The European Economic and Social Committee and the
Committee of the Regions, European Commission (EC), Brussels, Belguim, 11 pp.
EC, 2013c: Principles and Recommendations for Integrating Climate Change
Adaptation Considerations under the 2014-2020 European Maritime and
Fisheries Fund Operational Programmes. Brussels, 30.7.2013. SWD(2013) 299
Final, Commission Staff Working Document, European Commission (EC),
Brussels, Belguim, 20 pp.
ECDC, 2009: Technical Report: Development of Aedes albopictus Risk Map. European
Centre for Disease Prevention and Control (ECDC), Stockholm, Sweden, 45 pp.
ECDC, 2012: Technical Report: The Climatic Suitability for Dengue Transmission in
Continental Europe. European Centre for Disease Prevention and Control
(ECDC), Stockholm, Sweden, 22 pp.
ECHOES Country Report: Cyprus, 2009: COST Action FP0703 – ECHOES: Expected
Climate Change and Options for European Silviculture – Draft Country Report:
Cyprus [Cacot, E. and J. Peyron (eds.)], Cyprus Department of Forests, Ministry
of Agriculture, Natural Resources and Environment, Nicosia, Cyprus, 18 pp.
Eckert, N., E. Parent, R. Kies, and H. Baya, 2010: A spatio-temporal modelling
framework for assessing the fluctuations of avalanche occurrence resulting
from climate change: application to 60 years of data in the Northern French
Alps. Climatic Change, 101(3), 515-553.
Edwards, M. and A.J. Richardson, 2004: Impact of climate change on marine pelagic
phenology and trophic mismatch. Nature, 430(7002), 881-884.
EEA, 2009: Water Resources Across Europe Confronting Water Scarcity and
Drought. European Environment Agency (EEA), Publications Office of the
European Union, Luxembourg, Luxembourg, 55 pp.
EEA, 2010a: The European Environment, State and Outlook 2010. Water Resources:
Quantity and Flows. EEA Report No. 1/2010, European Environment Agency
(EEA), Publications Office of the European Union, Luxembourg, Luxembourg,
32 pp.
EEA, 2010b: Tracking Progress Towards Kyoto and 2020 Targets in Europe. EEA Report
No. 7/2010, European Environment Agency (EEA), Publications Office of the
European Union, Luxembourg, Luxembourg, 107 pp.
EEA, 2010c: Mapping the Impacts of Natural Hazards and Technological Accidents
in Europe: An Overview of the Last Decade. Technical report No 13/2010,
European Environment Agency (EEA), Publications Office of the European
Union, Luxembourg, Luxembourg, 144 pp.
EEA, 2010d: 10 Messages for 2010 Marine Ecosystems. European Environment
Agency (EEA), Copenhagen, Denmark, 13 pp.
EEA, 2012: Climate Change, Impacts and Vulnerability in Europe 2012, an Indicator-
Based Report. EEA Report No. 12/2012, European Environment Agency (EEA),
Copenhagen, Denmark, 304 pp.
EEA, 2013: Adaptation in Europe – Addressing Risks and Opportunities From Climate
Change in the Context of Socio-Economic Developments. EEA Report No
3/2013, European Environment Agency (EEA), Publications Office of the
European Union, Luxembourg, Luxembourg, 132 pp.
Ellwanger, G., A. Ssymank, A. Paulsch, and C. Paulsch, 2011: Natura 2000 and climate
change – state of knowledge: first results of an international workshop on the
Isle of Vilm. Natur und Landschaft, 86(1), 15-18.
Elzinga, J.A., S. van Nouhuys, D. van Leeuwen, and A. Biere, 2007: Distribution and
colonisation ability of three parasitoids and their herbivorous host in a
fragmented landscape. Basic and Applied Ecology, 8(1), 75-88.
Endler, C., K. Oehler, and A. Matzarakis, 2010: Vertical gradient of climate change
and climate tourism conditions in the Black Forest. International Journal of
Biometeorology, 54(1), 45-61.
Endler, C. and A. Matzarakis, 2011: Climatic potential for tourism in The Black Forest,
Germany winter season. International Journal of Biometeorology, 55(3),
339-351.
1312
Chapter 23 Europe
23
Engelhard, G.H., J.R. Ellis, M.R. Payne, R. Ter Hofstede, and J.K. Pinnegar, 2011:
Ecotypes as a concept for exploring responses to climate change in fish
assemblages. ICES Journal of Marine Science, 68(3), 580-591.
Engler, R., C. Randin, W. Thuiller, S. Dullinger, N.E. Zimmermann, M.B. Araújo, P.B.
Pearman, C.H. Albert, P. Choler, X. de Lamo, T. Dirnböck, D. Gómez-García, J.
Grytnes, E. Heegard, F. istad, G. Le Lay, D. Nogues-Bravo, S. Normand, C.
Piédalu, M. Puscas, M. Sebastià, A. Stanisci, J. Theurillat, M. Trivedi, P. Vittoz,
and A. Guisan, 2011: 21
st
century climate change threatens mountain flora
unequally across Europe. Global Change Biology, 17, 2330-2341.
Eskeland, G.S. and T.K. Mideksa, 2010: Electricity demand in a changing climate.
Mitigation and Adaptation Strategies for Global Change, 15(8), 877-897.
ESPACE, 2007: ESPACE Planning in a Changing Climate: The Strategy. European
Spatial Planning Adapting to Climate Events (ESPACE) project, The Environment
Department, Hampshire County Council, Winchester, UK, 8 pp.
Eugenio-Martin, J.L. and J.A. Campos-Soria, 2010: Climate in the region of origin
and destination choice in outbound tourism demand. Tourism Management,
31(6), 744-753.
European Parliament and EU Council, 2007: Directive of the European Parliament
and of the Council of 23 October 2007 on the Assessment and Management
of Flood Risks2007/60/EC. European Parliament, Strasbourg, France and EU
Council, Brussels, Begium, pp. L 288/27-L 288/34.
Eurostat, 2009: Forestry Statistics. Eurostat, Statistical Office of the European Union,
Publications Office of the European Union, Luxembourg, Luxembourg, 152 pp.
Eurostat, 2011a: Migrants in Europe. A Statistical Portrait of the First and Second
Generation. Eurostat, Statistical Office of the European Union, Luxembourg,
Luxembourg, 148 pp.
Eurostat, 2011b: Labour Market Statistics. Eurostat, Statistical Office of the European
Union, Publications Office of the European Union, Luxembourg, Luxembourg,
109 pp.
Falloon, P. and R. Betts, 2010: Climate impacts on European agriculture and water
management in the context of adaptation and mitigation the importance
of an integrated approach. Science of the Total Environment, 408(23), 5667-
5687.
FAO, 2008a: Climate Change: Implications for Food Safety. Food and Agriculture
Organization of the United Nations (FAO), Rome, Italy, 49 pp.
FAO, 2008b: Climate Change Impacts on Forest Health. Forest Health & Biosecurity
Working Papers FBS/34E, Food and Agriculture Organization of the United
Nations (FAO), Rome, Italy, 38 pp.
Feehan, J., M. Harley, and J. Van Minnen, 2009: Climate change in Europe. 1. Impact
on terrestrial ecosystems and biodiversity. A review. Agronomy for Sustainable
Development, 29(3), 409-421.
Fernandes, P.M., A. Luz, and C. Loureiro, 2010: Changes in wildfire severity from
maritime pine woodland to contiguous forest types in the mountains of
northwestern Portugal. Forest Ecology and Management, 260(5), 883-892.
Ferrara, R.M., P. Trevisiol, M. Acutis, G. Rana, G.M. Richter, and N. Baggaley, 2010:
Topographic impacts on wheat yields under climate change: two contrasted
case studies in Europe. Theoretical and Applied Climatology, 99(1-2), 53-65.
Ferron, C., D. Trewick, P. Le Conte, E.R. Batard, and L. Girard, 2006: Heat stroke in
hospital patients during the summer 2003 heat wave: a nosocomial disease.
Presse Medicale, 25(2), 196-199.
Feyen, L. and R. Dankers, 2009: Impact of global warming on streamflow drought in
Europe. Journal of Geophysical Research: Atmospheres, 114(D17), D17116,
doi:10.1029/2008JD011438.
Feyen, L., J.I. Barredo, and R. Dankers, 2009: Implications of global warming and
urban land use change on flooding in Europe. In:Water & Urban Development
Paradigms – Towards an Integration of Engineering, Design and Management
Approaches [Feyen, J., K. Shannon, and M. Neville (eds.)]. CRC Press/Balkema,
Leiden, Netherlands, pp. 217-225.
Feyen, L., R. Dankers, K. Bódis, P. Salamon, and J.I. Barredo, 2012: Fluvial flood risk
in Europe in present and future climates. Climatic Change, 112, 47-62.
Filz, K.J., J.O. Engler, J. Stoffels, M. Weitzel, and T. Schmitt, 2013: Missing the target?
A critical view on butterfly conservation efforts on calcareous grasslands in
south-western Germany. Biodiversity and Conservation, 22(10), 2223-2241.
Finger, R., W. Hediger, and S. Schmid, 2011: Irrigation as adaptation strategy to climate
change a biophysical and economic appraisal for Swiss maize production.
Climatic Change, 105(3-4), 509-528.
Fischer, D., S.M. Thomas, and C. Beierkuhnlein, 2010: Temperature-derived potential
for the establishment of phlebotomine sandflies and visceral leishmaniasis in
Germany. Geospatial Health, 5(1), 59-69.
Fischer, D., C. Thomas, F. Niemitz, and C. Reineking, 2011: Projection of climate
suitability for Aedes albopictus Skuse (Culicidae) in Europe under climate
change conditions. Global and Planetary Change, 78, 54-65.
Fischer, G., S. Prieler, H. van Velthuizen, G. Berndes, A. Faaij, M. Londo, and M. de
Wit, 2010: Biofuel production potentials in Europe: sustainable use of cultivated
land and pastures, Part II: land use scenarios. Biomass and Bioenergy, 34(2),
173-187.
Fischer, L., R. Purves, C. Huggel, J. Noetzli, and W. Haeberli, 2012: On the influence
of topographic, geologicaland cryospheric factors on rock avalanches and
rockfalls in high-mountain areas. Natural Hazards and Earth System Science,
12(1), 241-254.
Forkel, R. and R. Knoche, 2006: Regional climate change and its impact on photo-
oxidant concentrations in sourthern Germany: simulations with a coupled
regional climate-chemistry model. Journal of Geophysical Research Atmospheres,
111(D12), D12302, doi:10.1029/2005JD006748.
Forkel, R. and R. Knoche, 2007: Nested regional climate-chemistry simulations for
central Europe. Comptes Rendus Geoscience, 339(11-12), 734-746.
Forsius, M., S. Anttila, L. Arvola, I. Bergström, H. Hakola, H.I. Heikkinen, J. Helenius,
M. Hyvärinen, K. Jylhä, J. Karjalainen, T. Keskinen, K. Laine, E. Nikinmaa, P.
Peltonen-Sainio, K. Rankinen, M. Reinikainen, H. Setälä, and J. Vuorenmaa,
2013: Impacts and adaptation options of climate change on ecosystem services
in Finland: a model-based study. Current Opinion in Environmental Sustainability,
5(1), 26-40.
rster, H. and J. Lilliestam, 2010: Modeling thermoelectric power generation in view
of climate change. Regional Environmental Change, 10(4), 327-338.
Founda, D. and C. Giannakopoulos, 2009: The exceptionally hot summer of 2007 in
Athens, Greece a typical summer in the future climate? Global and Planetary
Change, 67(3-4), 227-236.
Fronzek, S., M. Luoto, and T.R. Carter, 2006: Potential effect of climate change on
the distribution of palsa mires in subarctic Fennoscandia. Climate Research,
32(1), 1-12.
Fronzek, S., T.R. Carter, J. Räisänen, L. Ruokolainen, and M. Luoto, 2010: Applying
probabilistic projections of climate change with impact models: a case study
for sub-arctic palsa mires in Fennoscandia. Climatic Change, 99(3), 515-534.
Fronzek, S., T.R. Carter, and M. Luoto, 2011: Evaluating sources of uncertainty in
modelling the impact of probabilistic climate change on sub-arctic palsa mires.
Natural Hazards and Earth System Science, 11(11), 2981-2995.
Fronzek, S., T.R. Carter, and K. Jylhä, 2012: Representing two centuries of past and
future climate for assessing risks to biodiversity in Europe. Global Ecology and
Biogeography, 21(1), 19-35.
Fuhrer, J., 2009: Ozone risk for crops and pastures in present and future climates.
Naturwissenschaften, 96(2), 173-194.
Fuhrer, J., M. Beniston, A. Fischlin, C. Frei, S. Goyette, K. Jasper, and C. Pfister, 2006:
Climate risks and their impact on agriculture and forests in Switzerland.
Climatic Change, 79(1-2), 79-102.
Fujihara, Y., K. Tanaka, T. Watanabe, T. Nagano, and T. Kojiri, 2008: Assessing the
impacts of climate change on the water resources of the Seyhan River Basin in
Turkey: use of dynamically downscaled data for hydrologic simulations. Journal
of Hydrology, 353(1-2), 33-48.
Furrer, B., V. Hoffmann, and M. Swoboda, 2009: Banking & Climate Change
Opportunities and Risks: An Analysis of Climate Strategies in more than 100
Banks Worldwide. Sustainable Asset Management (SAM), ETH Zurich (ETH),
and ZHAW Zurich University of Applied Sciences, supported by The World Wide
Fund for Nature (WWF), Zurich, Switzerland, 51 pp.
Gale, P., B. Stephenson, A. Brouwer, M. Martinez, A. de la Torre, J. Bosch, M. Foley-
Fisher, P. Bonilauri, A. Lindström, R.G. Ulrich, C.J. de Vos, M. Scremin, Z. Liu, L.
Kelly, and M.J. Muñoz, 2012: Impact of climate change on risk of incursion of
Crimean-Congo haemorrhagic fever virus in livestock in Europe through
migratory birds. Journal of Applied Microbiology, 112(2), 246-257.
Gallego-Sala, A.V., J.M. Clark, J.I. House, H.G. Orr, I.C. Prentice, P. Smith, T. Farewell,
and S.J. Chapman, 2010: Bioclimatic envelope model of climate change impacts
on blanket peatland distribution in Great Britain. Climate Research, 45, 151-
162.
Gao, X. and F. Giorgi, 2008: Increased aridity in the Mediterranean region under
greenhouse gas forcing estimated from high resolution simulations with a
regional climate model. Global and Planetary Change, 62(3-4), 195-209.
Garcia-Fayos, P. and E. Bochet, 2009: Indication of antagonistic interaction between
climate change and erosion on plant species richness and soil properties in
semiarid Mediterranean ecosystems. Global Change Biology, 15(2), 306-318.
1313
Europe Chapter 23
23
García-López, J.M. and C. Alluéa, 2011: Modelling phytoclimatic versatility as a large
scale indicator of adaptive capacity to climate change in forest ecosystems.
Ecological Modelling, 222(8), 1436-1447.
García-Ruiz, J.M., J.I. López-Moreno, S.M. Vicente-Serrano, T. Lasanta-Martínez, and
S. Baguería, 2011: Mediterranean water resources in a global change scenario.
Earth-Science Reviews, 105(3-4), 121-139.
Gardiner, B., K. Blennow, J. Carnus, P. Fleischer, F. Ingemarson, G. Landmann, M.
Lindner, M. Marzano, B. Nicoll, C. Orazio, J. Peyron, M. Reviron, M. Schelhaas,
A. Schuck, M. Spielmann, and T. Usbeck, 2010: Destructive Storms in European
Forests: Past and Forthcoming Impacts. Final Report to European Commission
DG Environment. European Forest Institute, Atlantic European Regional Office
(EFIATLANTIC), Bordeaux, France, 138 pp.
Garza-Gil, M., J. Torralba-Cano, and M. Varela-Lafuente, 2010: Evaluating the
economic effects of climate change on the European sardine fishery. Regional
Environmental Change, 11(1), 87-95.
Gaslikova, L., A. Schwerzmann, C.C. Raible, and T.F. Stocker, 2011: Future storm surge
impacts on insurable losses for the North Sea region. Natural Hazards and Earth
System Sciences, 11(4), 1205-1216.
GDV, 2011: The Climate Change Challenge: Answers and Demands of German Insurers.
Gesamtverband der Deutschen Versicherungswirtschaft e.V German Insurance
Association (GDV), Berlin, Germany, 18 pp.
Gehrig-Fasel, J., A. Guisan, and N.E. Zimmermann, 2007: Tree line shifts in the Swiss
Alps: climate change or land abandonment? Journal of Vegetation Science,
18(4), 571-582.
Giannakopoulos, C., P. Le Sager, M. Bindi, M. Moriondo, E. Kostopoulou, and C.M.
Goodess, 2009: Climatic changes and associated impacts in the Mediterranean
resulting from a 2
o
C global warming. Global and Planetary Change, 68(3),
209-224.
Giannakopoulos, C., E. Kostopoulou, K.V. Varotsos, K. Tziotziou, and A. Plitharas,
2011: An integrated assessment of climate change impacts for Greece in the
near future. Regional Environmental Change, 11(4), 829-843.
Gifford, R., L. Steg, and J.P. Reser, 2011: Environmental pyschology. In: The IAAP
Handbook of Applied Psychology [Martin, P.R., M.C. Cheung, L. Kyrios, M.
Littlefield, J.B. Knowles, M. Overmier, and J.M. Prieto (eds.)]. Wiley-Blackwell,
Chichester, UK, pp. 440-471.
Gilgen, A.K., C. Signarbieux, U. Feller, and N. Buchmann, 2010: Competitive advantage
of Rumex obtusifolius L. might increase in intensively managed temperate
grasslands under drier climate. Agriculture, Ecosystems & Environment, 135(1-2),
15-23.
Gill, S., J. Handley, R. Ennos, and S. Pauleit, 2007: Adapting cities for climate change:
the role of the green infrastructure. Built Environment, 33(1), 115-133.
Giuggiola, A., T.M. Kuster, and S. Saha, 2010: Drought-induced mortality of Scots pines
at the southern limits of its distribution in Europe: causes and consequences.
Journal of Biogeosciences and Forestry, 3, 95-97.
Giuntoli, I., B. Renard, J.-P. Vidal, and A. Bard, 2013: Low flows in France and their
relationship to large-scale climate indices. Journal of Hydrology, 482, 105-118.
GLA, 2010: The Draft Climate Change Adaptation Strategy for London, Public
Consultation Draft. Greater London Authority (GLA), London, UK, 136 pp.
Glenk, K. and A. Fisher, 2010: Insurance, prevention or just wait and see? Public
preferences for water management strategies in the context of climate change.
Ecological Economics, 69, 2279-2291.
Goderniaux, P., S. Brouyére, S. Blenkinsop, A. Burton, H.J. Fowler, P. Orban, and A.
Dassargues, 2011: Modeling climate change impacts on groundwater resources
using transient stochastic climatic scenarios. Water Resources Research,
47(12), W12516, doi:10.1029/2010WR010082.
Golombek, R., S. Kittlesen, and I. Haddeland, 2012: Climate change: impacts on
electricity markets in Western Europe. Climatic Change, 113, 357-370.
Gómez-Rodríguez, C., J. Bustamante, and C. Díaz-Paniagua, 2010: Evidence of
hydroperiod shortening in a preserved system of temporary ponds. Remote
Sensing, 2(6), 1439-1462.
Gonzalez-Camacho, J., J.C. Mailhol, and F. Ruget, 2008: Local impact of increasing
CO
2
in the atmosphere on maize crop water productivity in the Drome valley,
France. Irrigation and Drainage, 57(2), 229-243.
Goode, J., 2012: Viticulture: fruity with a hint of drought. Nature, 492(7429), 351-353.
Goodess, C., D. Jacob, M. Déqué, J. Guttiérrez, R. Huth, E. Kendon, G. Leckebusch, P.
Lorenz, and V. Pavan, 2009: Downscaling methods, data and tools for input to
impacts assessments. In: ENSEMBLES: Climate Change and its Impacts:
Summary of Research and Results from the ENSEMBLES Project [van der Linden,
P. and J.F.B. Mitchell (eds.)]. Met Office Hadley Centre, Exeter, UK, pp. 59-78.
Görgen, K., J. Beersma, H. Buiteveld, G. Brahmer, M. Carambia, O.d. Keizer, P. Krahe,
E. Nilson, R. Lammersen, C. Perrin, and D. Volken, 2010: Assessment of Climate
Change Impacts on Discharge in the River Rhine Basin. Results of the
RheinBlick2050 Project. International Commission for the Hydrology of the
Rhine Basin, Lelystad, Netherlands, 228 pp.
Gottfried, M., H. Pauli, A. Futschik, M. Akhalkatsi, P. Barancok, J.L. Benito Alonso, G.
Coldea, J. Dick, B. Erschbamer, M.R. Fernández Calzado, G. Kazakis, J. Krajci, P.
Larsson, M. Mallaun, O. Michelsen, D. Moiseev, P. Moiseev, U. Molau, A.
Merzouki, L. Nagy, G. Nakhutsrishvili, B. Pedersen, G. Pelino, M. Puscas, G. Rossi,
A. Stanisci, J.-P. Theurillat, M. Tomaselli, L. Villar, P. Vittoz, I. Vogiatzakis, and G.
Grabherr, 2012: Continent-wide response of mountain vegetation to climate
change. Nature Climate Change, 2(2), 111-115.
Graux, A., R. Lardy, G. Bellocchi, and J. Soussana, 2012: Global warming potential of
French grassland-based dairy livestock systems under climate change. Regional
Environmental Change, 12(4), 751-763.
Gregory, P.J. and B. Marshall, 2012: Attribution of climate change: a methodology
to estimate the potential contribution to increases in potato yield in Scotland
since 1960. Global Change Biology, 18(4), 1372-1388.
Gret-Regamy, A., P. Bebi, I. Bishop, and W. Schmid, 2008: Linking GIS-based models
to value ecosystem services in an alpine region. Journal of Environmental
Management, 89, 197-208.
Gret-Regamy, A., S. Brunner, J. Altwegg, and P. Bebi, 2013: Facing uncertainties in
ecosystem services-based resource management. Journal of Environmental
Management, 127(Suppl.), S145-S154.
Grime, J.P., J.D. Fridley, A.P. Askew, K. Thompson, J.G. Hodgson, and C.R. Bennett,
2008: Long-term resistance to simulated climate change in an infertile
grassland. Proceedings of the National Academy of Sciences of the United
States of America, 105(29), 10028-10032.
Grossi, C.M., P. Brimblecombe, and I. Harris, 2007: Predicting long term freeze-thaw
risks on Europe built heritage and archaeological sites in a changing climate.
Science of the Total Environment, 377(2-3), 273-281.
Grossi, C.M., A. Bonazza, P. Brimblecombe, I. Harris, and C. Sabbioni, 2008: Predicting
21
st
century recession of architectural limestone in European cities. Environmental
Geology, 56(3-4), 455-461.
Grossi, C.M., P. Brimblecombe, and H. Lloyd, 2010: The effects of weather on visits
to historic properties. Views, 47, 69-71.
Grossi, C.M., P. Brimblecombe, B. Mendez, D. Benavente, I. Harris, and M. Deque,
2011: Climatology of salt transitions and implications for stone weathering.
Science of the Total Environment, 409(13), 2577-2585.
Guardiola-Albert, C. and C.R. Jackson, 2011: Potential Impacts of climate change
on groundwater supplies to the Doñana wetland, Spain. Wetlands, 31(5), 907-
920.
Guis, H., C. Caminade, C. Calvete, A.P. Morse, A. Tran, and M. Baylis, 2012: Modelling
the effects of past and future climate on the risk of bluetongue emergence in
Europe. Journal of the Royal Society Interface, 9(67), 339-350.
Haasnoot, M., H. Middelkoop, A. Offermans, E. van Beek, and W.P.A. van Deursen,
2012: Exploring pathways for sustainable water management in river deltas in
a changing environment. Climatic Change, 115(3-4), 795-819.
Haddeland, I., P.C. Røhr, and H. Udnæs, 2011: Effects of Climate Changes on Water
Resources in the Glomma River Basin, Norway. WATCH Technical Report No.
27, Water and Global Change (WATCH) project, Norwegian Water Resources
and Energy Directorate (NVE), Oslo, Norway and Glommen and Laagen Water
Management Association (GLB), Lillehammer, Norway, 17 pp.
Haigh, I., R. Nicholls, and N. Wells, 2010: Assessing changes in extreme sea levels:
application to the English Channel, 1900-2006. Continental Shelf Research,
30(9), 1042-1055.
Haines, A., P. Wilkinson, C. Tonne, and I. Roberts, 2009a: Aligning climate change and
public health policies. The Lancet, 374(9707), 2035-2038.
Haines, A., A.J. McMichael, K.R. Smith, I. Roberts, J. Woodcock, A. Markandya, B.G.
Armstrong, D. Campbell-Lendrum, A.D. Dangour, M. Davies, N. Bruce, C. Tonne,
M. Barrett, and P. Wilkinson, 2009b: Public health benefits of strategies to
reduce greenhouse-gas emissions: overview and implications for policy makers.
The Lancet, 374(9707), 2104-2114.
Haines-Young, R., M. Potschin, and F. Kienast, 2012: Indicators of ecosystem service
potential at European scales: mapping marginal changes and trade-offs.
Ecological Indicators, 21, 39-53.
Hajat, S., M. O’Connor, and T. Kosatsky, 2010: Health effects of hot weather: from
awareness of risk factors to effective health protection. Lancet, 375(9717),
856-863.
1314
Chapter 23 Europe
23
Hakala, K., A.O. Hannukkala, E. Huusela-Veistola, M. Jalli, and P. Peltonen-Sainio,
2011: Pests and diseases in a changing climate: a major challenge for Finnish
crop production. Agricultural and Food Science, 20(1), 3-14.
Hallegatte, S., F. Henriet, and J. Corfee-Morlot, 2008: The Economics of Climate
Change Impacts and Policy Benefits at City Scale: A Conceptual
Framework.OECD Environment Working Paper No. 4, OECD Publishing, Paris,
France, 48 pp.
Hallegatte, S., N. Ranger, O. Mestre, P. Dumas, J. Corfee-Morlot, C. Herweijer, and R.
Wood, 2011: Assessing climate change impacts, sea level rise and storm surge
risk in port cities: a case study on Copenhagen. Climatic Change, 104, 113-
137.
Hallegatte, S., C. Green, R.J. Nicholls, J. Corfree-Morlot, 2013: Future flood losses
in major coastal cities. Nature Climate Change, 3, 802-806, doi:10.1038/
nclimate1979.
Halpern, B.S., S. Walbridge, K.A. Selkoe, C.V. Kappel, F. Micheli, C. D’Agrosa, J.F. Bruno,
K.S. Casey, C. Ebert, H.E. Fox, R. Fujita, D. Heinemann, H.S. Lenihan, E.M.P.
Madin, M.T. Perry, E.R. Selig, M. Spalding, R. Steneck, and R. Watson, 2008: A
global map of human impact on marine ecosystems. Science, 319, 948-952.
Hames, J. and S. Vardoulakis, 2012: Climate Change Risk Assessment for the Health
Sector. Sector report for CCRA – UK Climate Change Risk Assessment 2012 –
GA0204, Department for Environment, Food and Rural Affairs (DEFRA), London,
UK, 240 pp.
Hamilton, J.M. and R.S.J. Tol, 2007: The impact of climate change on tourism in
Germany, the UK and Ireland: a simulation study. Regional Environmental
Change, 7(3), 161-172.
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.
Hamududu, B. and A. Killingtveit, 2012: Assessing climate change impacts on global
hydropower. Energies, 5, 305-322.
Hanewinkel, M., D.A. Cullmann, M. Schelhaas, G. Nabuurs, and N.E. Zimmermann,
2013: Climate change may cause severe loss in the economic value of European
forest land. Nature Climate Change, 3, 203-207.
Hansen, J., R. Ruedy, M. Sato, and K. Lo, 2010: Globalsurface temperature change.
Review of Geophysics, 48(4), RG4004, doi:10.1029/2010RG000345.
Hanso, M. and R. Drenkhan, 2007: Retrospective analysis of Lophodermium
seditiosum epidemics in Estonia. Acta Silvatica & Lignaria Hungarica, 2007(SI),
31-45.
Hanson, S., R. Nicholls, N. Ranger, S. Hallegatte, J. Corfee-Morlot, C. Herweijer, and J.
Chateau, 2011: A global ranking of port cities with high exposure to climate
extremes. Climatic Change, 104(1), 89-111.
Hardacre, C.J., P.I. Palmer, K. Baumanns, M. Rounsevell, and D. Murray-Rust, 2013:
Probabilistic estimation of future emissions of isoprene and surface oxidant
chemistry associated with land-use change in response to growing food needs.
Atmospheric Chemistry and Physics, 13, 5451-5472. doi:10.5194/acp-13-5451-
2013.
Harrison, G.P., L.C. Cradden, and J.P. Chick, 2008: Preliminary assessment of climate
change impacts on the UK onshore wind energy resource. Energy Sources, Part
A: Recovery, Utilization, and Environmental Effects, 30(14), 1286-1299.
Harrison, P.A., P.M. Berry, C. Henriques, and I.P. Holman, 2008: Impacts of socio-
economic and climate change scenarios on wetlands: linking water resource
and biodiversity meta-models. Climatic Change, 90(1-2), 113-139.
Harrison, P.A., M. Vandewalle, M.T. Sykes, P.M. Berry, R. Bugter, F. de Bello, C.K. Feld,
U. Grandin, R. Harrington, J.R. Haslett, R.H.G. Jongman, G.W. Luck, P.M. da Silva,
M. Moora, J. Settele, J.P. Sousa, and M. Zobel, 2010: Identifying and prioritising
services in European terrestrial and freshwater ecosystems. Biodiversity and
Conservation, 19(10), 2791-2821.
Harrison, P.A., I.P. Holman, G. Cojocaru, K. Kok, A. Kontogianni, M. Metzger, and M.
Gramberger, 2013: Combining qualitative and quantitative understanding for
exploring cross-sectoral climate change impacts, adaptation and vulnerability
in Europe. Regional Environmental Change, 13(4), 761-780.
Hartel, T., R. Bncil, and D. Coglniceanu, 2011: Spatial and temporal variability of
aquatic habitat use by amphibians in a hydrologically modified landscape.
Freshwater Biology, 56(11), 2288-2298.
Hartikainen, K., J. Riikonen, A. Nerg, M. Kivimäenpää, V. Ahonen, A. Tervahauta, S.
Kärenlampi, M. Mäenpää, M. Rousi, S. Kontunen-Soppela, E. Oksanen, and T.
Holopainen, 2012: Impact of elevated temperature and ozone on the emission
of volatile organic compounds and gas exchange of silver birch (Betula pendula
Roth). Environmental and Experimental Botany, 84, 33-43.
Haugen, J.E. and T. Iversen, 2008: Response in extremes of daily precipitation and
wind from a downscaled multi-model ensemble of anthropogenic global
climate change scenarios. Tellus A, 60(3), 411-426.
Hawkins, E., J. Robson, R. Sutton, D. Smith, and N. Keenlyside, 2011: Evaluating the
potential for statistical decadal predictions of sea surface temperatures with a
perfect model approach. Climate Dynamics, 37(11-12), 2495-2509.
Hawkins, E., T.E. Fricker, A.J. Challinor, C.A.T. Ferro, C.K. Ho, and T.M. Osborne, 2013:
Increasing influence of heat stress on French maize yields from the 1960s to
the 2030s. Global Change Biology, 19(3), 937-947.
Haylock, M.R., N. Hofstra, A.M.G. Klein Tank, E.J. Klok, P.D. Jones, and M. New, 2008:
A European daily high-resolution gridded data set of surface temperature and
precipitation for 1950-2006. Journal of Geophysical Research, 113(D20),
D20119, doi:10.1029/2008JD010201.
Heath, M.R., F.C. Neat, J.K. Pinnegar, D.G. Reid, D.W. Sims, and P.J. Wright, 2012: Review
of climate change impacts on marine fish and shellfish around the UK and Ireland.
Aquatic Conservation: Marine and Freshwater Ecosystems, 22(3), 337-367.
Heidrich, O., R.J. Dawson, D. Reckian, and C.L. Walsh, 2013: Assessment of the climate
preparedness of 30 urban areas in the UK. Climatic Change, 120(4), 771-784.
Hein, L., M.J. Metzger, and A. Moreno, 2009: Potential impacts of climate change on
tourism; a case study for Spain. Current Opinion in Environmental Sustainability,
1(2), 170-178.
Hekkenberg, M., R. Benders, H. Moll, and A. Schoot Uiterkamp, 2009: Indications
for a changing electricity demand pattern: the temperature dependence of
electricity in the Netherlands. Energy Policy, 37, 1542-1551.
HELCOM, 2007: Climate Change in the Baltic Sea Area – HELCOM Thematic
Assessment in 2007. Baltic Sea Environment Proceedings No. 111, Prepared
by Marine Environmental Consultants (MEC) in collaboration with HELCOM
Secretariat, Baltic Marine Environment Protection Commission Helsinki
Commission, Helsinki, Finland, 49 pp.
Hellmann, F. and J.E. Vermaat, 2012: Impact of climate change on water management
in Dutch peat polders. Ecological Modelling, 240, 74-83.
Helming, K., K. Diehl, T. Kuhlman, T. Jansson, P.H. Verburg, M. Bakker, M. Perez-Soba,
L. Jones, P.J. Verkerk, P. Tabbus, J. Breton Morris, Z. Drillet, J. Farrington, P.
LeMouël, P. Zagame, T. Stuczynski, G. Siebielec, S. Sieber, and H. Wiggering, 2011:
Ex ante impact assessment of policies affecting land use, Part B: application of
the analytical framework. Ecology and Society, 16(1), 29, www.ecologyand
society.org/vol16/iss1/art29/.
Heltberg, R., H. Gitay, and R.G. Prabhu, 2012: Community-based adaptation: lessons
from a grant competition. Climate Policy, 12(2), 143-163.
Hemery, G.E., 2008: Forest management and silvicultural responses to projected
climate change impacts on European broadleaved trees and forests.
International Forestry Review, 10(4), 591-607.
Hemery, G.E., J.R. Clark, E. Aldinger, H. Claessens, M.E. Malvolti, E. O’Connor, Y.
Raftoyannis, P.S. Savill, and R. Brus, 2010: Growing scattered broadleaved tree
species in Europe in a changing climate: a review of risks and opportunities.
Forestry, 83(1), 65-81.
Henderson, G.R. and D.J. Leathers, 2010: European snow cover extent variability and
associations with atmospheric forcings. International Journal of Climatology,
30(10), 1440-1451.
Henderson, P.A., 2007: Discrete and continuous change in the fish community of
the Bristol Channel in response to climate change. Journal of the Marine
Biological Association of the UK, 87(02), 589-598.
Hendrickx, F. and E. Sauquet, 2013: Impact of warming climate on water management
for the Ariège River basin (France). Hydrological Sciences Journal, 58(5), 976-
993, doi:10.1080/02626667.2013.788790.
Henriques, C., I.P. Holman, E. Audsley, and K. Pearn, 2008: An interactive multi-scale
integrated assessment of future regional water availability for agricultural
irrigation in East Anglia and North West England. Climatic Change, 90(1-2),
89-111.
Hermans, C.M.L., I.R. Geijzendorffera, F. Ewertb, M.J. Metzgera, P.H. Vereijkene, G.B.
Woltjerf, and A. Verhagene, 2010: Exploring the future of European crop
production in a liberalised market, with specific consideration of climate change
and the regional competitiveness. Ecological Modelling, 221, 2177-2187.
Hermant, M., J. Lobry, S. Bonhommeau, J. Poulard, and O. Le Pape, 2010: Impact of
warming on abundance and occurrence of flatfish populations in the Bay of
Biscay (France). Journal of Sea Research, 64(1-2), 45-53.
Hertel, S., A. Le Tertre, K. Jöckel, and B. Hoffmann, 2009: Quantification of the heat
wave effect on cause-specific mortality in Essen, Germany. European Journal
of Epidemiology, 24(8), 407-414.
1315
Europe Chapter 23
23
Herweijer, C., N. Ranger, and R.E.T. Ward, 2009: Adaptation to climate change:
threats and opportunities for the insurance industry. Geneva Papers on Risk
and Insurance: Issues and Practice, 34, 360-380.
Hilpert, K., F. Mannke, and P. Schmidt-Thome, 2007: Towards Climate Change
Adapatation Strategies in the Baltic Sea Region. Guidance for decision-makers,
produced in the context of the INTERREG IIIB project, “Developing Policies and
Adaptation Strategies to Climate Change in the Baltic Sea Region (ASTRA)”,
Geological Survey of Finland, Espoo, Finland, 55 pp.
Hinkel, J., R. Nicholls, A. Vafeidis, R. Tol, and T. Avagianou, 2010: Assessing risk of and
adaptation to sea-level rise in the European Union: an application of DIVA.
Mitigation and Adaptation Strategies for Global Change, 15(7), 703-719.
Hlásny, T., Z. Barcza, M. Fabrika, B. Balázs, G. Churkina, J. Pajtík, R. Sedmák, and M.
Turčáni, 2011: Climate change impacts on growth and carbon balance of forests
in Central Europe. Climate Research, 47(3), 219-236.
Hochrainer, S., J. Linnerooth-Bayer, and R. Mechler, 2010: European Union Solidarity
Fund. Mitigation and Adaptation Strategies for Global Change, 15(7), 797-810.
Hodzic, A., S. Madronich, B. Bohn, S. Massie, L. Menut, and C. Wiedinmyer, 2007:
Wildfire particulate matter in Europe during summer 2003: meso-scale
modeling of smoke emissions, transport and radiative effects. Atmospheric
Chemistry and Physics, 7, 4043-4064.
Hoes, O., 2006: Aanpak Wateroverlast in Polders op Basis van Risicobeheer. Technische
Universiteit Delft, Delft, Netherlands, 188 pp. (in Dutch).
Hoffmann, B., S. Häfele, and U. Karl, 2013: Analysis of performance losses of thermal
power plants in Germany a System Dynamics model approach using data
from regional climate modeling. Energy, 49, 193-203.
Hoffmann, I., 2010: Climate change and the characterization, breeding and
conservation of animal genetic resources. Animal Genetics, 41(Suppl. 1), 32-46.
Hoinka, K.P., A. Carvalho, and A.I. Miranda, 2009: Regional-scale weather patterns
and wildland fires in central Portugal. International Journal of Wildland Fire,
18(1), 36-49.
Holland, T. and B. Smit, 2010: Climate change and the wine industry: current research
themes and new directions. Journal of Wine Research, 21(2-3), 125-136.
House, J.I., H.G. Orr, J.M. Clark, A. Gallego-Sala, C. Freeman, I.C. Prentice, and P. Smith,
2011: Climate change and the British Uplands: evidence for decision-making.
Climate Research, 45, 3-12.
Howden, N.J.K., T.P. Burt, F. Worrall, M.J. Whelan, and M.Z. Bieroza, 2010: Nitrate
concentrations and fluxes in the River Thames over 140 years (1868 - 2008):
are increases irreversible? Hydrological Processes, 24, 2657-2662.
Howden, S.M., J.F. Soussana, F.N. Tubiello, N. Chhetri, M. Dunlop, and H. Meinke, 2007:
Adapting agriculture to climate change. Proceedings of the National Academy
of Sciences of the United States of America, 104(50), 19691-19696.
HPA, 2012: Health Effects of Climate Change in the UK 2012 Current Evidence,
Recommendations and Research Gaps [Vardoulakis, S. and C. Heaviside, (eds.)].
Health Protection Agency (HPA), Didcot, UK, 242 pp.
HSY, 2010: Helsinki Metropolitan Area Adaptation to Climate Change Strategy.
Helsinki Region Environmental Services Authority [Helsingin seudun
ympäristöpalvelut -kuntayhtymä (HSY)], Helsinki, Finland, 88 pp.
Huang, C., A.G. Barnett, X. Wang, P. Vaneckova, G. FitzGerald, and S. Tong, 2011:
Projecting future heat related mortality under climate change scenarios: a
systematic review. Environmental Health Perspectives, 119(12), 1681-1690.
Hueging, H., R. Haas, K. Born, D. Jacob, and J.G. Pinto, 2013: Regional changes in
wind energy potential over Europe using regional climate model ensemble
projections. Journal of Applied Meteorology and Climatology, 52(4), 903-917
Huggel, C., N. Salzmann, S. Allen, J. Caplan-Auerbach, L. Fischer, W. Haeberli, C.
Larsen, D. Schneider, and R. Wessels, 2010: Recent and future warm extreme
events and high-mountain slope stability. Philosophical Transactions of the
Royal Society A, 368, 1919, 2435-2459, doi:10.1098/rsta.2010.0078.
Huggel, C., J.J. Clague, and O. Korup, 2012: Is climate change responsible for changing
landslide activity in high mountains? Earth Surface Processes and Landforms,
37(1), 77-91.
Hunt, A. and P. Watkiss, 2011: Climate change impacts and adaptation in cities: a
review of the literature. Climatic Change, 104(1), 13-49.
Huntjens, P., C. Pahl-Wostl, and J. Grin, 2010: Climate change adaptation in European
river basins. Regional Environmental Change, 10, 263-284.
Huntley, B., R.E. Green, Y.C. Collingham, and S.G. Willis, 2007: A Climatic Atlas of
European Breeding Birds. Published as a partnership between Durham University,
The Royal Society for the Protection of Birds (RSPB) and Lynx Edicions in
association with the University of Cambridge, BirdLife International, and the
European Bird Census Council (EBCC), Lynx Edicions, Barcelona, Spain, 521 pp.
Hurkmans, R., W. Terink, R. Uijlenhoet, P. Torfs, D. Jacob, and P.A. Troch, 2010: Changes
in streamflow dynamics in the Rhine basin under three high-resolution regional
climate scenarios. Journal of Climate, 23(3), 679-699.
Huss, M., 2011: Present and future contribution of glacier storage change to runoff
from macroscale drainage basins in Europe. Water Resources Research, 47(7),
W07511, doi:10.1029/2010WR010299.
ICES, 2010: 6.4.2. Cod in Subarea IV (North Sea) and Division IIIa West (Skagerrak).
In: Report of the ICES Advisory Committee 2010. ICES Advice, 2010. Book 6: North
Sea. International Council for the Exploration of the Sea (ICES), Copenhagen,
Denmark, pp. 35-54.
Iglesias, A., L. Garrote, F. Flores, and M. Moneo, 2007: Challenges to manage the risk
of water scarcity and climate change in the Mediterranean. Water Resources
Management, 21(5), 775-788.
Iglesias, A., L. Garrote, A. Diz, J. Schlickenrieder, and M. Moneo, 2012: Water and people:
assessing policy priorities for climate change adaptation in the Mediterranean.
In: Regional Assessment of Climate Change in the Mediterranean [Navarra, A.
and L. Tubiana (eds.)]. Springer, Dordrecht, Netherlands, pp. 201-233.
Isaac, M. and D.P. van Vuuren, 2009: Modelling global residential sector energy
demand for heating and air conditioning in the context of climate change.
Energy Policy, 37(2), 507-521.
Jackson, A.C. and J. McIlvenny, 2011: Coastal squeeze on rocky shores in northern
Scotland and some possible ecological impacts. Journal of Experimental Marine
Biology and Ecology, 400(1-2), 314-321.
Jackson, C.R., R. Meister, and C. Prudhomme, 2011: Modelling the effects of climate
change and its uncertainty on UK Chalk groundwater resources from an
ensemble of global climate model projections. Journal of Hydrology, 399(1-2),
12-28.
Jacob, D. and R. Podzun, 2010: Global warming below 2°C relative to pre-industrial
level: how might climate look like in Europe. Nova Acta Leopoldina, 384, 71-
76.
Jacob, D.J. and D.A. Winner, 2009: Effect of climate change on air quality. Atmospheric
Environment, 43(1), 51-63.
Jacob, D., J. Petersen, B. Eggert, A. Alias, O. Bøssing Christensen, L.M. Bouwer, A.
Braun, A. Colette, M. Déqué, G. Georgievski, E. Georgopoulou, A. Gobiet, L.
Menut, G. Nikulin, A. Haensler, N. Hempelmann, C. Jones, K. Keuler, S. Kovats,
N. Kröner, S. Kotlarski, A. Kriegsmann, E. Martin, E. van Meijgaard, C. Moseley,
S. Pfeifer, S. Preuschmann, C. Radermacher, K. Radtke, D. Rechid, M. Rounsevell,
P. Samuelsson, S. Somot, J.-F. Soussana, C. Teichmann, R. Valentini, R. Vautard,
B. Weber, and P. Yiou, 2013: EURO-CORDEX: New high-resolution climate
changeprojections for European impact research. Regional Environmental
Change (in press), doi:10.1007/s10113-013-0499-2.
Jactel, H., B.C. Nicoll, M. Branco, J. Gonzalez-Olabarria, W. Grodzki, B. Långström, F.
Moreira, S. Netherer, C. Orazio, D. Piou, H. Santos, M.J. Schelhaas, K. Tojic, and
F. Vodde, 2009: The influences of forest stand management on biotic and abiotic
risks of damage. Annals of Forest Science, 66(7), doi:10.1051/forest/2009054.
Jacxsens, L., P.A. Luning, J.G.A.J. van der Vorst, F. Devlieghere, R. Leemans, and M.
Uyttendaele, 2010: Simulation modelling and risk assessment as tools to
identify the impact of climate change on microbiological food safety the case
study of fresh produce supply chain. Food Research International, 43(7), 1925-
1935.
James, P., K. Tzoulas, M.D. Adams, A. Barber, J. Box, J. Breuste, T. Elmqvist, M. Frith, C.
Gordon, K.L. Greening, J. Handley, S. Haworth, A.E. Kazmierczak, M. Johnston,
K. Korpela, M. Moretti, J. Niemelä, S. Pauleit, M.H. Roe, J.P. Sadler, and C. Ward
Thompson, 2009: Towards an integrated understanding of green space in the
European built environment. Urban Forestry and Urban Greening, 8(2), 65-75.
Jenkins, D.P., 2009: The importance of office internal heat gains in reducing cooling
loads in a changing climate. International Journal of Low-Carbon Technologies,
4(3), 134-140.
Jenkins, D.P., Y. Liu, and A.D. Peacock, 2008: Climatic and internal factors affecting
future UK office heating and cooling energy consumptions. Energy and
Buildings, 40(5), 874-881.
Jeppesen, E., B. Kronvang, J.E. Olesen, J. Audet, M. Søndergaard, C.C. Hoffmann, H.E.
Andersen, T.L. Lauridsen, L. Liboriussen, S.E. Larsen, M. Beklioglu, M. Meerhoff,
A. Özen, and K. Özkan, 2011: Climate change effects on nitrogen loading from
cultivated catchments in Europe: implications for nitrogen retention, ecological
state of lakes and adaptation. Hydrobiologia, 663(1), 1-21.
Jiguet, F., R.D. Gregory, V. Devictor, R.E. Green, P. Vorisek, A. Van Strien, and D.
Couvet, 2010: Population trends of European common birds are predicted by
characteristics of their climatic niche. Global Change Biology, 16(2), 497-505.
1316
Chapter 23 Europe
23
Johnk, K.D., J. Huisman, J. Sharples, B. Sommeijer, P.M. Visser, and J.M. Stroom, 2008:
Summer heatwaves promote blooms of harmful cyanobacteria. Global Change
Biology, 14(3), 495-512.
Johnson, A., M. Acreman, M. Dunbar, S. Feist, A. Giacomello, R. Gozlan, S. Hinsley, A.
Ibbotson, H. Jarvie, J. Jones, M. Longshaw, S. Maberly, T. Marsh, C. Neal, J.
Newman, M. Nunn, R. Pickup, N. Reynard, C. Sullivan, J. Sumpter, and R. Williams,
2009: The British river of the future: how climate change and human activity
might affect two contrasting river ecosystems in England. Science of the Total
Environment, 407, 4787-4798.
Jomelli, V., D. Brunstein, D. Grancher, and P. Pech, 2007: Is the response of hill slope
debris flows to recent climate change univocal? A case study in the Massif des
Ecrins (French Alps). Climatic Change, 85(1-2), 119-137.
Jomelli, V., D. Brunstein, M. Déqué, M. Vrac, and D. Grancher, 2009: Impacts of future
climatic change (2070-2099) on the potential occurrence of debris flows: a case
study in the Massif des Ecrins (French Alps). Climatic Change, 97(1), 171-191.
Jonkeren, O.E., 2009: Adaptation to Climate Change in Inland Waterway Transport.
Diss. Ph.D., Vrije Universiteit, Amsterdam, Netherlands, 164 pp.
Jonkeren, O., P. Rietveld, and J. van Ommeren, 2007: Climate change and inland
waterway transport; welfare effects of low water levels on the river Rhine.
Journal of Transport Economics and Policy, 41(3), 387-411.
Jonkeren, O., B. Jourquin, and P. Rietveld, 2011: Modal-split effects of climate change:
the effect of low water levels on the competitive position of inland waterway
transport in the River Rhine area. Transportation Research Part A: Policy &
Practice, 45(10), 1007-1019.
nsson, A.M., G. Appelberg, S. Harding, and L. Bärring, 2009: Spatio-temporal impact
of climate change on the activity and voltinism of the spruce bark beetle,
Ips typographus. Global Change Biology, 15(2), 486-499.
Jönsson, A.M., S. Harding, P. Krokene, H. Lange, A. Lindelöw, B. Økland, H.P. Ravn,
and L.M. Schroeder, 2011: Modelling the potential impact of global warming
on Ips typographus voltinism and reproductive diapause. Climatic Change,
109(3-4), 695-718.
Jonsson, B. and N. Jonsson, 2009: A review of the likely effects of climate change on
anadromous Atlantic salmon Salmo salar and brown trout Salmo trutta, with
particular reference to water temperature and flow. Journal of Fish Biology,
75(10), 2381-2447.
Jonzén, N., A. Lindén, T. Ergon, E. Knudsen, J.O. Vik, D. Rubolini, D. Piacentini, C. Brinch,
F. Spina, L. Karlsson, M. Stervander, A. Andersson, J. Waldenström, A. Lehikoinen,
E. Edvardsen, R. Solvang, and N.C. Stenseth, 2006: Rapid advance of spring
arrival dates in long-distance migratory birds. Science, 312(5782), 1959-1961.
Jordà, G., D. Gomis, and M. Marcos, 2012: Comment on “Storm surge frequency
reduction in Venice under climate change” by Troccoli et al. Climatic Change,
113(3-4), 1081-1087.
JRC, 2008: Forest Fires in Europe 2007. JRC Scientific and Technical Reports, Report
No. 8, EUR 23492 EN 2008, European Commission, Joint Research Centre
(JRC), Institute for Environment and Sustainability, Publications Office of the
European Union, Luxembourg, Luxembourg, 77 pp.
JRC and EEA, 2010: The European Environment, State and Outlook 2010. Soil. European
Commission, Joint Research Centre (JRC) and European Environment Agency
(EEA), Publications Office of the European Union, Luxembourg, 44 pp.
Kabat, P., L.O. Fresco, M.J.F. Stive, C.P. Veerman, J.S.L.J. van Alphen, B.W.A.H. Parmet,
W. Hazeleger, and C.A. Katsman, 2009: Dutch coasts in transition. Nature
Geosciences, 2, 450-452.
Karaca, M. and R.J. Nicholls, 2008: Potential implications of accelerated sea-level
rise for Turkey. Journal of Coastal Research, 24(2), 288-298.
Katsman, C., A. Sterl, J. Beersma, H. van den Brink, J. Church, W. Hazeleger, R. Kopp,
D. Kroon, J. Kwadijk, R. Lammersen, J. Lowe, M. Oppenheimer, H. Plag, J. Ridley,
H. von Storch, D. Vaughan, P. Vellinga, L. Vermeersen, R. van de Wal, and R.
Weisse, 2011: Exploring high-end scenarios for local sea level rise to develop
flood protection strategies for a low-lying delta: the Netherlands as an example.
Climatic Change, 109, 617-645.
Kay, A.L., S.M. Crooks, P. Pall, and D.A. Stone, 2011: Attribution of Autumn/Winter
2000 flood risk in England to anthropogenic climate change: a catchment-
based study. Journal of Hydrology, 406(1-2), 97-112.
Keenan, T., J. Maria Serra, F. Lloret, M. Ninyerola, and S. Sabate, 2011: Predicting the
future of forests in the Mediterranean under climate change, with niche- and
process-based models: CO
2
matters! Global Change Biology, 17(1), 565-579.
Keith, S.A., A.C. Newton, R.J.H. Herbert, M.D. Morecroft, and C.E. Bealey, 2009: Non-
analogous community formation in response to climate change. Journal for
Nature Conservation, 17(4), 228-235.
Kersebaum, K.C., A.S. Nain, C. Nendel, M. Gandorfer, and M. Wegehenkel, 2008:
Simulated effect of climate change on wheat production and nitrogen
management at different sites in Germany. Journal of Agrometeorology, 10,
266-273.
Keskitalo, E.C.H., 2010: The Development of Adaptation Policy and Practice in
Europe: Multi-Level Governance of Climate Change. Springer, Dordrecht,
Netherlands, 376 pp.
Khaledian, M.R., J.C. Mailhol, P. Ruelle, I. Mubarak, and S. Perret, 2010: The impacts
of direct seeding into mulch on the energy balance of crop production system
in the SE of France. Soil and Tillage Research, 106(2), 218-226.
Kilpeläinen, M. and H. Summala, 2007: Effects of weather and weather forecasts
on driver behaviour. Transportation Research, 10(4), 288-299.
Kint, V., W. Aertsen, M. Campioli, D. Vansteenkiste, A. Delcloo, and B. Muys, 2012:
Radial growth change of temperate tree species in response to altered regional
climate and air quality in the period 1901-2008. Climatic Change, 115, 343-363.
Kjellström, E., G. Nikulin, U. Hansson, G. Strandberg, and A. Ullerstig, 2011: 21
st
century changes in the European climate: uncertainties derived from an
ensemble of regional climate model simulations. Tellus A, Series A, 63A(1), 24-
40, doi: 10.1111/j.1600-0870.2010.00475.x.
Kjellstrom, T., R.S. Kovats, S.J. Lloyd, T. Holt, and R.S. Tol, 2009: The direct impact of
climate change on regional labor productivity. Archives of Environmental and
Occupational Health, 64(4), 217-227.
Klaus, M., A. Holsten, P. Hostert, and J.P. Kropp, 2011: Integrated methodology to
assess windthrow impacts on forest stands under climate change. Forest
Ecology and Management, 261(11), 1799-1810.
Klijn, F., N. Asselman, and H. Van Der Most, 2009: Compartmentalisation: flood
consequence reduction by splitting up large polder areas. Journal of Flood Risk
Management, 3, 3-17.
Klik, A. and J. Eitzinger, 2010: Impact of climate change on soil erosion and the
efficiency of soil conservation practices in Austria. Journal of Agricultural
Science, 148, 529-541.
Kløve, B., P. Ala-aho, G. Bertrand, Z. Boukalova, A. Ertürk, N. Goldscheider, J. Ilmonen,
N. Karakaya, H. Kupfersberger, J. Kvœrner, A. Lundberg, M. Mileusnić, A.
Moszczynska, T. Muotka, E. Preda, P. Rossi, D. Siergieiev, J. Šimek, P. Wachniew,
V. Angheluta, and A. Widerlund, 2011: Groundwater dependent ecosystems.
Part I: hydroecological status and trends. Environmental Science and Policy,
14(7), 770-781.
Koca, D., B. Smith, and M. Sykes, 2006: Modelling regional climate change effects
on potential natural ecosystems in Sweden. Climatic Change, 78, 381-406.
Koch, H. and S. Vögele, 2009: Dynamic modeling of water demand, water availability
and adaptation strategies for power plants to global change. Ecological
Economics, 68(7), 2031-2039.
Koetse, M.J. and P. Rietveld, 2009: Impact of climate change and weather on transport:
an overview of empirical findings. Transportation Research, 14(3), 205-221.
Koutroulis, A.G., A. Vrochidou, and I.K. Tsanis, 2010: Spatiotemporal characteristics
of meteorological drought for the Island of Crete. Journal of Hydrometeorology,
12(2), 206-226, doi: 10.1175/2010JHM1252.1.
Koutsias, N., M. Arianoutsou, A.S. Kallimanis, G. Mallinis, J.M. Halley, and P.
Dimopoulos, 2012: Where did the fires burn in Peloponnisos, Greece the
summer of 2007? Evidence for a synergy of fuel and weather. Agricultural and
Forest Meteorology, 156, 41-53.
Kovats, R.S. and S. Hajat, 2008: Heat stress and public health: a critical review. Annual
Review of Public Health, 29, 41-55.
Kreienkamp, F., A. Spekat, and W. Enke, 2010: Stationarity of atmospheric waves
and blocking over Europe based on a reanalysis dataset and two climate
scenarios. Theoretical and Applied Climatology, 102(1-2), 205-212, doi:10.1007/
s00704-010-0261-3.
Krekt, A.H., T.J. van der Laan, R.A.E. van der Meer, B. Turpijn, O.E. Jonkeren, A. van
der Toorn, E. Mosselman, J. van Meijeren, and T. Groen, 2011: Climate Change
and Inland Waterway Transport: Impacts on the Sector, the Port of Rotterdam
and Potential Solutions. National Research Programme Knowledge for Climate/
Nationaal Onderzoekprogramma Kennis voor Klimaat (KvK), Knowledge for
Climate Programme Office Secretariat: c/o Utrecht University, Utrecht,
Netherlands, 74 pp.
Kriegler, E., B.C. O’Neill, S. Hallegatte, T. Kram, R.H. Moss, R. Lempert, and T.J.
Wilbanks, 2010: Socio-Economic Scenario Development for Climate Change
Analysis. CIRED Working Paper,DT/WPNo 2010-23, Centre International de
Recherche sur l’Environnement et le Développement (CIRED), CIRED, Nogent-
sur-Marne, France, 35 pp.
1317
Europe Chapter 23
23
Kristensen, K., K. Schelde, and J.E. Olesen, 2011: Winter wheat yield response to
climate variability in Denmark. Journal of Agricultural Science, 149(1), 33-47.
Kuhlicke, C., A. Steinführer, C. Begg, C. Bianchizza, M. Bründl, M. Buchecker, B. De
Marchi, M. Di Masso Tarditti, C. Höppner, B. Komac, L. Lemkow, J. Luther, S.
Mccarthy, L. Pellizzoni, O. Renn, A. Scolobig, M. Supramaniam, S. Tapsell, G.
Wachinger, G. Walker, R. Whittle, M. Zorn, and H. Faulkner, 2011: Perspectives
on social capacity building for natural hazards: outlining an emerging field of
research and practice in Europe. Environmental Science and Policy, 14(7), 804-
814.
Kundzewicz, Z.W., I. Pińskwar, and G.R. Brakenridge, 2013: Large floods in Europe,
1985-2009. Hydrological Sciences Journal, 58(1), 1-7.
Kunz, M., J. Sander, and C. Kottmeier, 2009: Recent trends of thunderstorm and
hailstorm frequency and their relation to atmospheric characteristics in
southwest Germany. International Journal of Climatology, 29, 2283-2297.
Kwadijk, J.C.J., M. Haasnoot, J.P.M. Mulder, M.M.C. Hoogvliet, A.B.M. Jeuken, R.A.A.
van der Krogt, N.G.C. van Oostrom, H.A. Schelfhout, E.H. van Velzen, H. van
Waveren, and M.J.M. de Wit, 2010: Using adaptation tipping points to prepare
for climate change and sea level rise, a case study in the Netherlands. Wiley
Interdisciplinary Reviews: Climate Change, 1(5), 729-740.
Ladanyi, M., 2008: Risk methods and their applications in agriculture. Applied
Ecology and Environmental Research, 6(1), 147-164.
Lake, I.R., I.A. Gillespie, G. Bentham, G.L. Nichols, C. Lane, G.K. Adak, and E.J. Threlfall,
2009: A re-evaluation of the impact of temperature and climate change on
foodborne illness. Epidemiology and Infection, 137(11), 1538-1547.
Lal, R., J.A. Delgado, P.M. Groffman, N. Millar, C. Dell, and A. Rotz, 2011: Management
to mitigate and adapt to climate change. Journal of Soil and Water Conservation,
66(4), 276-285.
Lamond, J.E., D.G. Proverbs, and F.N. Hammond, 2009: Accessibility of flood risk
insurance in the UK: confusion, competition and complacency. Journal of Risk
Research, 12(6), 825-841.
Langmead, O., A. McQuatters-Gollop, and L.D. Mee (eds.), 2007: European Lifestyles
and Marine Ecosystems: Exploring Challenges for Managing Europe’s Seas.
University of Plymouth Marine Institute, Plymouth, UK, 44 pp.
Lankester, P. and P. Brimblecombe, 2010: Predicting future indoor climate at Knole.
Views, 47, 71-73.
Lasda, O., A. Dikou, and E. Papapanagiotou, 2010: Flash flooding in Attika, Greece:
climatic change or urbanization? Ambio, 39, 608-611.
Lasserre, F. and S. Pelletier, 2011: Polar super seaways? Maritime transport in the
Arctic: an analysis of shipowners’ intentions. Journal of Transport Geography,
19, 1465-1473.
Lavalle, C., F. Micale, T.D. Houston, A. Camia, R. Hiederer, C. Lazar, C. Conte, G. Amatulli,
and G. Genovese, 2009: Climate change in Europe. 3. Impact on agriculture and
forestry. A review. Agronomy for Sustainable Development, 29(3), 433-446.
Lawrence, D. and H. Hisdal, 2011: Hydrological Projections for Floods in Norway
Under a Future Climate.NVE Report No. 2011-5, Norwegian Water Resources
and Energy Directorate (NVE), Oslo, Norway, 47 pp.
Le Floc’h, P.L., J. Poulard, O. Thébaud, F. Blanchard, J. Bihel, and F. Steinmetz, 2008:
Analyzing the market position of fish species subject to the impact of long-
term changes: a case study of French fisheries in the Bay of Biscay. Aquatic
Living Resources, 21(3), 307-316.
Leander, R., T.A. Buishand, B.J.J.M. van den Hurk, and M.J.M. de Wit, 2008: Estimated
changes in flood quantiles of the river Meuse from resampling of regional
climate model output. Journal of Hydrology, 351(3-4), 331-343.
Lee, H.C., R. Walker, S. Haneklaus, L. Philips, G. Rahmann, and E. Schnug, 2008:
Organic farming in Europe: a potential major contribution to food security in a
scenario of climate change and fossil fuel depletion. Landbauforschung
Volkenrode, 58(3), 145-151.
Lejeusne, C., P. Chevaldonne, C. Pergent-Martini, C.F. Boudouresque, and T. Perez,
2009: Climate change effects on a miniature ocean: the highly diverse, highly
impacted Mediterranean Sea. Trends in Ecology and Evolution, 25(4), 250-260.
Lemoine, N., H. Schaefer, and K.hning-Gaese, 2007a: Species richness of migratory
birds is influenced by global climate change. Global Ecology and Biogeography,
16(1), 55-65.
Lemoine, N., H. Bauer, M. Peintinger, and K. Böhning-Gaese, 2007b: Effects of climate
and land-use change on species abundance in a Central European bird
community. Conservation Biology, 21(2), 495-503.
Lemonsu, A., R. Kounkou-Arnaud, J. Desplat, J. Salagnac, and V. Masson, 2013:
Evolution of the Parisian urban climate under a global changing climate.
Climatic Change, 116(3-4), 679-692.
Lenderink, G. and E. Van Meijgaard, 2008: Increase in hourly precipitation extremes
beyond expectations from temperature changes. Nature Geoscience, 1(8), 511-
514.
Lennert, M. and J. Robert, 2010: The territorial futures of Europe: ‘Trends’, ‘Competition’
or ‘Cohesion’. Futures, 42(8), 833-845.
Lenoir, J., J.C. Gegout, P.A. Marquet, P. de Ruffray, and H. Brisse, 2008: A significant
upward shift in plant species optimum elevation during the 20
th
century.
Science, 320(5884), 1768-1771.
Lenoir, S., G. Beaugrand, and É. Lecuyer, 2011: Modelled spatial distribution of marine
fish and projected modifications in the North Atlantic Ocean. Global Change
Biology, 17(1), 115-129.
Letourneau, A., P.H. Verburg, and E. Stehfest, 2012: A land-use systems approach to
represent land-use dynamics at continental and global scales. Environmental
Modelling and Software, 33, 61-79.
Levinsky, I., F. Skov, J. Svenning, and C. Rahbek, 2007: Potential impacts of climate
change on the distributions and diversity patterns of European mammals.
Biodiversity and Conservation, 16(13), 3803-3816.
Liberloo, M., S. Luyssaert, V. Bellassen, S.N. Djomo, M. Lukac, C. Calfapietra, I.A.
Janssens, M.R. Hoosbeek, N. Viovy, G. Churkina, G. Scarascia-Mugnozza, and
R. Ceulemans, 2010: Bio-energy retains its mitigation potential under elevated
CO
2
. Public Library on Science, 5(7), e11648, doi:10.1371/journal.pone.
0011648.
Linard, C., N. Poncon, D. Fontenille, and E.F. Lambin, 2009: Risk of malaria reemergence
in Southern France: testing scenarios with a multiagent simulation model.
EcoHealth, 6(1), 135-147.
Lindgren, J., D.K. Johnsson, and A. Carlsson-Kanyama, 2009: Climate adaptation of
railways: lessons from Sweden. European Journal of Transport and Infrastructure
Research, 9(2), 164-181.
Lindner, M., M. Maroschek, S. Netherer, A. Kremer, A. Barbati, J. Garcia-Gonzalo, R.
Seidl, S. Delzon, P. Corona, M. Kolström, M.J. Lexer, and M. Marchetti, 2010:
Climate change impacts, adaptive capacity, and vulnerability of European forest
ecosystems. Forest Ecology and Management, 259(4), 698-709.
Lindsay, S.W., D.G. Hole, R.A. Hutchinson, S.A. Richards, and S.G. Willis, 2010:
Assessing the future threat from vivax malaria in the United Kingdom using
two markedly different modelling approaches. Malaria Journal, 9(1), 70-78.
Linnerud, K., T.H. Mideska, and G.S. Eskeland, 2011: The impact of climate change
on nuclear power supply. The Energy Journal, 32(1), 149-168.
Lionello, P., M.B. Galati, and E. Elvini, 2012: Extreme storm surge and wind wave
climate scenario simulations at the Venetian littoral. Physics and Chemistry of
the Earth, 40-41, 86-92.
Liu, M. and J. Kronbak, 2010: The potential economic viability of using the Northern
Sea Route (NSR) as an alternative route between Asia and Europe. Journal of
Transport Geography, 18, 434-444.
Liu, Y., R.A. Kahn, A. Chaloulakou, and P. Koutrakis, 2009: Analysis of the impact of
the forest res in August 2007 on air quality of Athens using multi-sensor
aerosol remote sensing data, meteorology and surface observations.
Atmospheric Environment, 43, 3310-3318.
Lobell, D.B., W. Schlenker, and J. Costa-Roberts, 2011: Climate trends and global crop
production since 1980. Science, 333(6042), 616-620.
Long, S.P., E.A. Ainsworth, A.D.B. Leakey, J. Nosberger, and D.R. Ort, 2006: Food
for thought: lower-than-expected crop yield stimulation with rising CO
2
concentrations. Science, 312(5782), 1918-1921.
Lopez Saez, J., C. Corona, M. Stoffel, and F. Berger, 2013: Climate change increases
frequency of shallow spring landslides in French Alps. Geology, 41(5), 619-622.
Lorz, C., C. Furst, Z. Galiz, D. Matijasic, V. Podrazky, N. Potocic, P. Simoncic, M. Strauch,
H. Vacik, and F. Makeschin, 2010: GIS-based probability assessment of natural
hazards in forested landscapes of Central and South-Eastern Europe.
Environmental Management, 46, 920-930.
Lowe, D., K.L. Ebi, and B. Forsberg, 2011: Heatwave early warning systems and
adaptation advice to reduce human health consequences of heatwaves.
International Journal of Environmental Research and Public Health, 8(12),
4623-4648.
Lowe, J.A., T.P. Howard, A. Pardaens, J. Tinker, J. Holt, S. Wakelin, G. Milne, J. Leake, J.
Wolf, K. Horsburgh, T. Reeder, G. Jenkins, J. Ridley, S. Dye, and S. Bradley, 2009:
UK Climate Projections Science Report: Marine and Coastal Projections. Met
Office Hadley Centre, Exeter, UK, 97 pp.
Luck, J., M. Spackman, A. Freeman, P. Trebicki, W. Griffiths, K. Finlay, and S. Chakraborty,
2011: Climate change and diseases of food crops. Plant Pathology, 60(1), 113-
121.
1318
Chapter 23 Europe
23
Ludwig, R., R. Roson, C. Zografos, and G. Kallis, 2011: Towards an inter-disciplinary
research agenda on climate change, water and security in Southern Europe and
neighboring countries. Environmental Science and Policy, 14(7), 794-803.
Lugeri, N., Z. Kundzewicz, E. Genovese, S. Hochrainer, and M. Radziejewski, 2010:
River flood risk and adaptation in Europe – assessment of the present status.
Mitigation and Adaptation Strategies for Global Change, 15(7), 621-639.
Lung, T., C. Lavalle, R. Hiederer, A. Dosio, and L.M. Bouwer, 2012: A multi-hazard
regional level impact assessment for Europe combining indicators of climatic
and non-climatic change. Global Environmental Change, 23(2), 522-536.
Luterbacher, J., D. Dietrich, E. Xoplaki, M. Grosjean, and H. Wanner, 2004: European
seasonal and annual temperature variability, trends, and extremes since 1500.
Science, 303(5663), 1499-1503.
Lyons, S., K. Mayor, and R. Tol, 2009: Holiday destinations: understanding the travel
choices of Irish tourists. Tourism Management, 30(5), 683-692.
Maaskant, B., S.N. Jonkman, and L.M. Bouwer, 2009: Future risk of flooding: an
analysis of changes in potential loss of life in South Holland (the Netherlands).
Environmental Science & Policy, 12(2), 157-169.
MacKenzie, B.R., H. Gislason, C. Mollmann, and F.W. Koster, 2007: Impact of 21
st
century climate change on the Baltic Sea fish community and fisheries. Global
Change Biology, 13(7), 1348-1367.
Macleod, C.J.A., P.D. Falloon, R. Evans, and P.M. Haygarth, 2012: Chapter 2: The effects
of climate change on the mobilization of diffuse substances from agricultural
systems. In: Advances in Agronomy, Vol. 115 [Sparks, D.L. (ed.)]. Elsevier Science
and Technology/Academic Press, Waltham, MA, USA, pp. 41-47.
Madgwick, J.W., J.S. West, R.P. White, M.A. Semenov, J.A. Townsend, J.A. Turner, and
B.D.L. Fitt, 2011: Impacts of climate change on wheat anthesis and fusarium
ear blight in the UK. European Journal of Plant Pathology, 130(1), 117-131.
Magnan, A., B. Garnaud, R. Billé, F. Gemenne, and S. Hallegatte, 2009: The Future of
the Mediterranean: From Impacts of Climate Change to Adaptation Issues.
Summary Report produced at the request of the Sustainable Development
General Commission (CGDD) of the French Ministry of Ecology, Energy,
Sustainable Development and Territorial Development (MEEDDAT) by the
Institut du Développement Durable et des Relations Internationales (IDDRI) in
collaboration with the Centre International de Recherche sur l’Environnement
et le Développement (CIRED), IDDRI, Paris, France, 42 pp.
Majone, B., C.I. Bovolo, A. Bellin, S. Blenkinsop, and H.J. Fowler, 2012: Modeling the
impacts of future climate change on water resources for the Gállego river basin
(Spain). Water Resources Research, 48(1), W01512, doi:10.1029/
2011WR010985.
Malheiro, A.C., J.A. Santos, H. Fraga, and J.G. Pinto, 2010: Climate change scenarios
applied to viticultural zoning in Europe. Climate Research, 43, 163-177.
Mandryk, M., P. Reidsma, and M. Ittersum, 2012: Scenarios of long-term farm
structural change for application in climate change impact assessment.
Landscape Ecology, 27(4), 509-527.
Mantyka-pringle, C.S., T.G. Martin, and J.R. Rhodes, 2012: Interactions between
climate and habitat loss effects on biodiversity: a systematic review and meta-
analysis. Global Change Biology, 18(4), 1239-1252, doi: 10.1111/j.1365-
2486.2011.02593.x.
Marcais, B. and M. Desprez-Loustau, 2007: Le réchauffement climatique a-t-il un
impact sur les maladies forestières? In: Forêts et Milieux Naturels Face aux
Changements Climatiques. Rendez-vous techniques de l’ONF, hors-série n° 3,
Décembre 2007, French Government Office National des Forêts (ONF), Paris,
France, pp. 47-52.
Marcos-Lopez, M., P. Gale, B.C. Oidtmann, and E.J. Peeler, 2010: Assessing the impact
of climate change on disease emergence in freshwater fish in the United
Kingdom. Transboundary and Emerging Diseases, 57(5), 293-304.
Marker, M., L. Angeli, L. Bottai, R. Costantini, R. Ferrari, L. Innocenti, and G. Siciliano,
2008: Assessment of land degradation susceptibility by scenario analysis: a
case study in Southern Tuscany, Italy. Geomorphology, 93(1-2), 120-129.
Marmot, M., J. Allen, R. Bell, E. Bloomer, P. Goldblatt, and Consortium for the
European Review of Social Determinants of Health and the Health Divide, 2012:
WHO European review of social determinants of health and the health divide.
Lancet, 380(9846), 1011-1029.
Marques, S., J.G. Borges, J. Garcia-Gonzalo, F. Moreira, J.M.B. Carreiras, M.M. Oliveira,
A. Cantarinha, B. Botequim, and J.M.C. Pereira, 2011: Characterization of
wildfires in Portugal. European Journal of Forest Research, 130(5), 775-784.
Mavrogianni, A., P. Wilkinson, M. Davies, P. Biddulph, and E. Oikonomou, 2012:
Building characteristics as determinants of propensity to high indoor summer
temperatures in London dwellings. Building and Environment, 55, 117-130.
Mazaris, A.D., A.D. Papanikolaou, M. Barbet-Massin, A.S. Kallimanis, F. Jiguet, D.S.
Schmeller, and J.D. Pantis, 2013: Evaluating the connectivity of a protected
areas’ network under the prism of global change: the efficiency of the European
Natura 2000 Network for four birds of prey. PLoS ONE, 8(3), e59640,
doi:10.1371/journal.pone.0059640.
McCarthy, M., M. Best, and R. Betts, 2010: Climate change in cities due to global
warming and urban effects. Geophysical Research Letters, 37(9), L09705,
doi:10.1029/2010GL042845.
McColl, L., E. Palin, H. Thornton, D. Sexton, R. Betts, and K. Mylne, 2012: Assessing
the potential impact of climate change on the UK’s electricity network. Climatic
Change, 115(3-4), 821-835.
McGrath, J. and D.B. Lobell, 2011: An independent method of deriving the carbon
dioxide fertilization effect in dry conditions using historical yield data from wet
and dry years. Global Change Biology, 17, 2689-2696.
McHugh, M., 2007: Short-term changes in upland soil erosion in England and Wales:
1999 to 2002. Geomorphology, 86(1-2), 204-213.
McInnes, K.L., T.A. Erwin, and J.M. Bathols, 2011: Global climate model projected
changes in 10 m wind speed and direction due to anthropogenic climate
change. Atmospheric Science Letters, 12(4), 325-333.
Mechler, R., S. Hochrainer, A. Aaheim, H. Salen, and A. Wreford, 2010: Modelling
economic impacts and adaptation to extreme events: insights from European
case studies. Mitigation and Adaptation Strategies for Global Change, 15(7),
737-762.
Medlock, J.M. and A.G.C. Vaux, 2013: Colonization of UK coastal realignment sites
by mosquitoes: implications for design, management, and public health. Journal
of Vector Ecology, 38(1), 53-62.
Medri, S., E. Banos de Guisasola, and S. Gualdi, 2012: Overview of the Main
International Climate Services. CMCC Research Paper, Issue RP0134, Euro-
Mediterranean Center on Climate Change, Lecce, Italy, 109 pp.
Medri, S., S. Venturini, and S. Castellari, 2013: Overview of Key Climate Change
Impacts, Vulnerabilities and Adaptation Action in Italy. CMCC Research Papers,
Issue RP0178, Euro-Mediterranean Center on Climate Change, Lecce, Italy, 75
pp.
Melchiorre, C. and P. Frattini, 2012: Modelling probability of rainfall-induced shallow
landslides in a changing climate, Otta, Central Norway. Climatic Change,
113(2), 413-436.
Meleux, F., F. Solmon, and F. Giorgi, 2007: Increase in summer European ozone
amounts due to climate change. Atmospheric Environment, 41(35), 7577-7587.
Menendez, M. and P.L. WoodWorth, 2010: Changes in extreme high water levels
based on quasi global tide-gauge data set. Journal of Geophysical Research,
115(C10), C10011, doi:10.1029/2009JC005997.
Menzel, A., T.H. Sparks, N. Estrella, E. Koch, A. Aasa, R. Ahas, K. Alm-Kübler, P. Bissolli,
O. Braslavská, A. Briede, F.M. Chmielewski, Z. Crepinsek, Y. Curnel, Å. Dahl, C.
Defila, A. Donnelly, Y. Filella, K. Jatczak, F. Måge, A. Mestre, Ø. Nordli, J. Peñuelas,
P. Pirinen, V. Remišová, H. Scheifinger, M. Striz, A. Susnik, A.J.H. van Vliet, F.-E.
Wielgolaski, S. Zach, and A. Zust, 2006: European phenological response to
climate change matches the warming pattern. Global Change Biology, 12(10),
1969-1976.
Metzger, M.J. and M.D.A. Rounsevell, 2011: A need for planned adaptation to climate
change in the wine industry. Environmental Research Letters, 6(3), 031001,
doi:10.1088/1748-9326/6/3/031001.
Metzger, M.J., R.G.H. Bunce, R.H.G. Jongman, C.A. Mücher, and J.W. Watkins, 2005:
A climatic stratification of the environment of Europe. Global Ecology and
Biogeography, 14(6), 549-563.
Metzger, M.J., D. Schroter, R. Leemans, and W. Cramer, 2008: A spatially explicit and
quantitative vulnerability assessment of ecosystem service change in Europe.
Regional Environmental Change, 8(3), 91-107.
Michelozzi, P., G. Accetta, M. De Sario, D. D’Ippoliti, C. Marino, M. Baccini, A. Biggeri,
H.R. Anderson, K. Katsouyanni, F. Ballester, L. Bisanti, E. Cadum, B. Forsberg, F.
Forastiere, P.G. Goodman, A. Hojs, U. Kirchmayer, S. Medina, A. Paldy, C.
Schindler, J. Sunyer, C.A. Perucci, and PHEWE Collaborative Group, 2009: High
temperature and hospitalizations for cardiovascular and respiratory causes in
12 European cities. American Journal of Respiratory and Critical Care Medicine,
179(5), 383-389.
Mickwitz, P., S. Beck, A. Jensenm, A.B. Pedersen, C. Görg, M. Melanen, N. Ferrand, C.
Kuhlicke, W. Kuindersma, M. Máñez, H. Reinert, and S. Bommel, 2009: Climate
policy integration as a necessity for an efficient climate policy. IOP Conference
Series: Earth and Environmental Science, 6, 582017, doi:10.1088/1755-
1307/6/58/582017.
1319
Europe Chapter 23
23
Mieszkowska, N., M.J. Genner, S.J. Hawkins, and D.W. Sims, 2009: Chapter 3. Effects
of climate change and commercial fishing on Atlantic cod Gadus morhua. In:
Advances in Marine Biology, Vol. 56 [Sims, D.W. (ed.)]. Elsevier Science and
Technology/Academic Press, Waltham, MA, USA, pp. 213-273.
Milad, M., H. Schaich, M. Bürgi, and W. Konold, 2011: Climate change and nature
conservation in Central European forests: a review of consequences, concepts
and challenges. Forest Ecology and Management, 261(4), 829-843.
Millar, C.I., N.L. Stephenson, and S.L. Stephens, 2007: Climate change and forests of
the future: managing in the face of uncertainty. Ecological Applications, 17(8),
2145-2151.
Milner, J., M. Davies, and P. Wilkinson, 2012: Urban energy, carbon management
(low carbon cities) and co-benefits for human health. Current Opinion in
Environmental Sustainability, 4(4), 338-404.
Ministry of Agriculture and Forestry, 2009: Evaluation of the Implementation of
Finland’s National Strategy for Adaptation to Climate Change 2009. Ministry
of Agriculture and Forestry, Helsinki, Finland, 44 pp.
Miraglia, M., H.J.P. Marvin, G.A. Kleter, P. Battilani, C. Brera, E. Coni, F. Cubadda, L.
Croci, B. De Santis, S. Dekkers, L. Filippi, R.W.A. Hutjes, M.Y. Noordam, M. Pisante,
G. Piva, A. Prandini, L. Toti, G.J. van den Born, and A. Vespermann, 2009: Climate
change and food safety: an emerging issue with special focus on Europe. Food
and Chemical Toxicology, 47(5), 1009-1021.
Miranda, A.I., E. Marchi, M. Ferretti, and M.M. Millan, 2009: Forest fires and air
quality issues in Southern Europe. In: Developments in Environmental Science
[Bytnerowicz, A., M. Arbaugh, A. Riebau, and C. Andersen (eds.)]. Elsevier,
Amsterdam, Netherlands, pp. 209-231.
Mirasgedis, S., Y. Sarafidis, E. Georgopoulou, V. Kotroni, K. Lagouvardos, and D.P.
Lalas, 2007: Modelling framework for estimating impacts of climate change
on electricity demand at regional level: case of Greece. Energy Conversion and
Management, 48(5), 1737-1750.
Mirasgedis, S., E. Georgopoulou, Y. Sarafidis, K. Papagiannaki, and D.P. Lalas, 2013:
The impact of climate change on the demand pattern of bottledwater and non-
alcoholic beverages. Business Strategy and the Environment, doi: 10.1002/
bse.1782.
Mitchell, T.D., T.R. Carter, P.D. Jones, M. Hulme, and M. New, 2004: A Comprehensive
Set of High-Resolution Grids of Monthly Climate for Europe and the Globe: The
Observed Record (1901-2000) and 16 Scenarios (2001-2100). Tyndall Centre
Research Working Paper 55, Tyndall Centre for Climate Change Research,
University of East Anglia, Norwich, UK, 25 pp.
Moen, J. and P. Fredman, 2007: Effects of climate change on alpine skiing in Sweden.
Journal of Sustainable Tourism, 15(4), 418-437.
Mokrech, M., R. Nicholls, J. Richards, C. Henriques, I. Holman, and S. Shackley, 2008:
Regional impact assessment of flooding under future climate and socio-
economic scenarios for East Anglia and North West England. Climatic Change,
90(1), 31-55.
Moller, A.P., D. Rubolini, and E. Lehikoinen, 2008: Populations of migratory bird
species that did not show a phenological response to climate change are
declining. Proceedings of the National Academy of Sciences of the United States
of America, 105(42), 16195-16200.
Molnar, J.L., R.L. Gamboa, C. Revenga, and M.D. Spalding, 2008: Assessing the global
threat of invasive species to marine biodiversity. Frontiers in Ecology and the
Environment, 6(9), 485-492.
Montoya, J.M. and D. Raffaelli, 2010: Climate change, biotic interactions and
ecosystem services. Philosophical Transactions of the Royal Society B,
365(1549), 2013-2018.
Mooij, W.M., J.H. Janse, L.N. De Senerpont Domis, S. Hülsmann, and B.W. Ibelings,
2007: Predicting the effect of climate change on temperate shallow lakes with
the ecosystem model PCLake. Hydrobiologica, 584(1), 443-454.
Morán-López, R., J.L. Pérez-Bote, E. da Silva, and A.B.P. Casildo, 2012: Hierarchical
large-scale to local-scale influence of abiotic factors in summer-fragmented
Mediterranean rivers: structuring effects on fish distributions, assemblage
composition and species richness. Hydrobiologia, 696(1), 137-158.
Moreira, F., O. Viedma, M. Arianoutsou, T. Curt, N. Koutsias, E. Rigolot, A. Barbati, P.
Corona, P. Vaz, G. Xanthopoulos, F. Mouillot, and E. Bilgili, 2011: Landscape –
wildfire interactions in southern Europe: implications for landscape management.
Journal of Environmental Management, 92(10), 2389-2402.
Moreno, A., 2010: Mediterranean tourism and climate (change): a survey-based
study. Tourism and Hospitality Planning & Development, 7(3), 253-265.
Moreno, A. and B. Amelung, 2009: Climate change and tourist comfort on Europe’s
beaches in summer: a reassessment. Coastal Management, 37(6), 550-568.
Moriondo, M., M. Bindi, Z.W. Kundzewicz, M. Szwed, A. Chorynski, P. Matczak, M.
Radziejewski, D. McEvoy, and A. Wreford, 2010a: Impact and adaptation
opportunities for European agriculture in response to climatic change and
variability. Mitigation and Adaptation Strategies for Global Change, 15(7), 657-
679.
Moriondo, M., C. Pacini, G. Trombi, C. Vazzana, and M. Bindi, 2010b: Sustainability
of dairy farming system in Tuscany in a changing climate. European Journal of
Agronomy, 32(1), 80-90.
Moriondo, M., M. Bindi, C. Fagarazzi, R. Ferrise, and G. Trombi, 2011: Framework for
high-resolution climate change impact assessment on grapevines at a regional
scale. Regional Environmental Change, 11(3), 553-567.
Mouillot, D., D.R. Bellwood, C. Baraloto, J. Chave, R. Galzin, M. Harmelin-Vivien, M.
Kulbicki, S. Lavergne, S. Lavorel, N. Mouquet, C.E. Paine, J. Renaud, and W. Thuiller,
2013: Rare species support vulnerable functions in high-diversity ecosystems.
PLoS Biology, 11(5), e1001569, doi:10.1371/journal.pbio.1001569.
Mullan, D., D. Favis-Mortlock, and R. Fealy, 2012: Addressing key limitations
associated with modelling soil erosion under the impacts of future climate
change. Agricultural and Forest Meteorology, 156, 18-30.
Musshoff, O., M. Odening, and W. Xu, 2011: Management of climate risks in
agriculture will weather derivatives permeate? Applied Economics, 43(9),
1067-1077.
Mustonen, T. and K. Mustonen, 2011a: Eastern Sámi Atlas. Snowchange Cooperative,
Helsinki, Finland, 331 pp.
Mustonen, T. and K. Mustonen, 2011b: Drowning Reindeers, Drowning Homes
Indigenous Saami and Hydroelectricity in Sompio, Finland. Snowchange
Cooperative, Helsinki, Finland, 115 pp.
Nabuurs, G.-J., M. Lindner, P.J. Verkerk, K. Gunia, P. Deda, R. Michalak, and G. Grassi,
2013: First signs of carbon sink saturation in European forest biomass. Nature
Climate Change, 3(9), 792-796.
Najac, J., C. Lac, and L. Terray, 2011: Impact of climate change on surface winds in
France using a statistical-dynamical downscaling method with mesoscale
modelling. International Journal of Climatology, 31(3), 415-430.
Navarra, A. and L. Tubiana (eds.), 2013: Regional Assessment of Climate Change in
the Mediterranean.Volume 3: Case Studies. Springer Science, Dordrecht,
Netherlands, 225 pp.
Nicholls, R., P. Wong, V. Burkett, C. Woodroffe, and J. Hay, 2008: Climate change and
coastal vulnerability assessment: scenarios for integrated assessment.
Sustainability Science, 3(1), 89-102.
Nicholls, S. and B. Amelung, 2008: Climate change and tourism in Northwestern
Europe: impacts and adaptation. Tourism Analysis, 13(1), 21-31.
Nokkala, M., P. Leviäkangas, and K. Oiva (eds.), 2012: The Costs of Extreme Weather
for the European Transport System. EWENT Project D4, Vtt Technology 36, VTT
Technical Research Centre of Finland, Espoo, Finland, 92 pp.
Nolan, P., P. Lynch, R. Mcgrath, T. Semmler, and S. Wang, 2012: Simulating climate
change and its effect on the the wind energy resource of Ireland. Wind Energy,
15, 593-608.
OECD, 2007: Climate Change in the European Alps: Adapting Winter Tourism and
Natural Hazards Management. Organisation for Economic Co-operation and
Development (OECD), Paris, France, 127 pp.
Okruszko, T., H. Duel, M. Acreman, M. Grygoruk, M. Flörke, and C. Schneider, 2011:
Broad-scale ecosystem services of European wetlands overview of the current
situation and future perspectivesunder different climate and water management
scenarios. Hydrological Sciences Journal, 56(8), 1501-1517.
Olesen, J.E., M. Trnka, K.C. Kersebaum, A.O. Skjelvåg, B. Seguin, P. Peltonen-Sainio, F.
Rossi, J. Kozyra, and F. Micale, 2011: Impacts and adaptation of European crop
production systems to climate change. European Journal of Agronomy, 34(2),
96-112.
Oliver, R.J., J.W. Finch, and G. Taylor, 2009: Second generation bioenergy crops and
climate change: a review of the effects of elevated atmospheric CO
2
and
drought on water use and the implications for yield. GCB Bioenergy, 1(2), 97-
114.
Olonscheck, M., A. Holsten, and J. Kropp, 2011: Heating and cooling demand and
related emissions of the German residential building stock under climate
change. Energy Policy, 39, 4795-4806.
Ommer, R., I.A. Perry, K. Cochrane, and P. Cury (eds.), 2011: World Fisheries: A
Social-Ecological Analysis. Wiley-Blackwell, Chichester, UK, 440 pp.
Oort, P.A.J., B.G.H. Timmermans, and A.C.P.M. van Swaaij, 2012: Why farmerssowing
dates hardly change when temperature rises. European Journal of Agronomy,
40, 102-111.
1320
Chapter 23 Europe
23
OSPAR, 2010: Chapter 12: Regional summaries. In: Quality Status Report 2010.
OSPAR Commission, London, UK, pp. 150-161.
Pahl-Wostl, C., 2007: Transitions towards adaptive management of water facing
climate and global change. Water Resources Management, 21(1), 49-62.
Paiva, R.C.D., W. Collischonn, E.B.C. Schettini, J. Vidal, F. Hendrickx, and A. Lopez,
2011: The case studies. In: Modelling The Impact of Climate Change on Water
Resources [Fung, F., A. Lopez, and M. New (eds.)]. John Wiley-Blackwell,
Chichester, UK, pp. 136-182.
Palahi, M., R. Mavsar, C. Gracia, and Y. Birot, 2008: Mediterranean forests under
focus. International Forestry Review, 10(4), 676-688.
Palin, E., H.E. Thornton, C.T. Mathison, R.E. McCarthy, R.T. Clark, and J. Dora, 2013:
Future projections of temperature-related climate change impacts on the
railway network of Great Britain. Climatic Change, 120(1-2), 71-93.
Pall, P., T. Aina, D.A. Stone, P.A. Stott, T. Nozawa, A.G.J. Hilberts, D. Lohmann, and M.R.
Allen, 2011: Anthropogenic greenhouse gas contribution to flood risk in
England and Wales in autumn 2000. Nature, 470 (7334), 382-385.
Paranjothy, S., J. Gallacher, R. Amlôt, G.J. Rubin, L. Page, T. Baxter, J. Wight, D. Kirrage,
R. McNaught, and S.R. Palmer, 2011: Psychosocial impact of the summer 2007
flood in England. BMC Public Health, 11, 145, doi:10.1186/1471-2458-11-145.
Parent, B. and F. Tardieu, 2012: Temperature responses of developmental processes
have not been affected by breeding in different ecological areas for 17 crop
species. New Phytologist, 194(3), 760-774.
Pašičko, R., Č. Branković, and Z. Simic, 2012: Assessment of climate change impacts
in energy generation from renewable sources in Croatia. Renewable Energy,
46, 224-231.
Paterson, J.S., M.B. Araújo, P.M. Berry, J.M. Piper, and M.D.A. Rounsevell, 2008:
Mitigation, adaptation, and the threat to biodiversity. Conservation Biology,
22(5), 1352-1355.
Paterson, R.R.M. and N. Lima, 2010: How will climate change affect mycotoxins in
food? Food Research International, 43(7), 1902-1914.
Pauli, H., M. Gottfried, S. Dullinger, O. Abdaladze, M. Akhalkatsi, J.L. Benito Alonso,
G. Coldea, J. Dick, B. Erschbamer, R. Fernández Calzado, D. Ghosn, J.I. Holten,
R. Kanka, G. Kazakis, J. Kollár, P. Larsson, P. Moiseev, D. Moiseev, U. Molau, J.
Molero Mesa, L. Nagy, G. Pelino, M. Puşc, G. Rossi, A. Stanisci, A.O. Syverhuset,
J.-P. Theurillat, M. Tomaselli, P. Unterluggauer, L. Villar, P. Vittoz, and G. Grabherr,
2012: Recent plant diversity changes on Europe’s mountain summits. Science,
336(6079), 353-355.
Pausas, J.G. and S. Fernández-Muñoz, 2012: Fire regime changes in the western
Mediterranean Basin: from fuel-limited to drought-driven fire regime. Climatic
Change, 110(1-2), 215-226.
Pausas, J.G., J. Llovet, A. Rodrigo, and R. Vallejo, 2008: Are wildfires a disaster in the
Mediterranean basin? A review. International Journal of Wildland Fire, 17(6),
713-723.
Pejovic, T., V.A. Williams, R.B. Noland, and R. Toumi, 2009: Factors affecting the
frequency and severity of airport weather delays and the implications of climate
change for future delays. Transportation Research Record, 2139, 97-106.
Pellizzaro, G., A. Ventura, B. Arca, A. Arca, P. Duce, V. Bacciu, and D. Spano, 2010:
Estimating effects of future climate on duration of fire danger season in
Sardinia. In: Proceedings of VI International Forest Fire Research Conference,
15-18, November 2010, University of Coimbra, Coimbra, Portugal, Session BP11
Climate and Fire Meteorology [Viegas, D.X. (ed.)]. Organized by the Associação
para o Desenvolvimento da Aerodinâmica Industrial (ADAI), Conference
Secretariat, ADAI, Coimbra, Portugal, 8 pp., www.cnr.it/istituti/ProdottoDella
Ricerca.html?cds=016&id=184241.
Peltonen-Sainio, P., L. Jauhiainen, and I.P. Laurila, 2009: Cereal yield trends in
northern European conditions: changes in yield potential and its realisation.
Field Crops Research, 110(1), 85-90.
Peltonen-Sainio, P., L. Jauhiainen, and K. Hakala, 2010: Crop responses to temperature
and precipitation according to long-term multi-location trials at high-latitude
conditions. The Journal of Agricultural Science, 149(01), 49-62.
Perch-Nielsen, S.L., B. Amelung, and R. Knutti, 2010: Future climate resources for
tourism in Europe based on the daily Tourism Climatic Index. Climatic Change,
103(3-4), 363-381.
Pereira, M., R. Trigo, C. da Camara, J. Pereira, and S. Leite, 2005: Synoptic patterns
associated with large summer forest fires in Portugal. Agricultural and Forest
Meteorology, 129(1-2), 11-25.
Perez, F.F., X.A. Padin, Y. Pazos, M. Gilcoto, M. Cabanas, P.C. Pardo, M.D. Doval, and
L. Farina-Busto, 2010: Plankton response to weakening of the Iberian coastal
upwelling. Global Change Biology, 16(4), 1258-1267.
Perry, R.I., R.E. Ommer, M. Barange, and F. Werner, 2010: The challenge of adapting
marine social-ecological systems to the additional stress of climate change.
Current Opinion in Environmental Sustainability, 2(5-6), 356-363.
Peterson, T., P. Scott, and S. Herring, 2012: Explaining extreme events of 2011 from
a climate perspective. Bulletin of the American Meteorological Society, 93,
1041-1067.
Petitpierre, B., C. Kueffer, O. Broennimann, C. Randin, C. Daehler, and A. Guisan,
2012: Climatic niche shifts are rare among terrestrial plant invaders. Science,
335(6074), 1344-1348.
Petney, T.N., J. Skuballa, S. Muders, M. Pfäffle, C. Zetlmeisl, and R. Oehme, 2012: The
changing distribution patterns of ticks (Ixodida) in Europe in relation to emerging
tick-borne diseases. Parasitology Research Monographs, 3, 151-166.
Petrow, T., B. Merz, K.E. Lindenschmidt, and A.H. Thieken, 2007: Aspects of seasonality
and flood generating circulation patterns in a mountainous catchment in south-
eastern Germany. Hydrology and Earth System Sciences, 11, 1455-1468.
Petrow, T., J. Zimmer, and B. Merz, 2009: Changes in the flood hazard in Germany
through changing frequency and persistence of circulation patterns. Natural
Hazards and Earth System Sciences, 9(4), 1409-1423.
Pfenniger, S., S. Hanger, M. Dreyfus, A. Dubel, N. Hernández-Mora, P. Esteve, C.
Varela-Ortega, P. Watkiss, and A. Patt, 2010: Report on Perceived Policy Needs
and Decision Contexts. MEDIATION Delivery Report, Deliverable 1.1 (Final
Draft) Subject to approval by the European Commission, Methodology for
Effective Decision-making on Impacts and AdaptaTION (MEDIATION) Research
Project, Seventh Framework Programme for Research and Technological
Development, European Commission (EC), Brussels, Belguim, 108 pp.
Philippart, C.J.M., R. Anadon, R. Danovaro, J.W. Dipper, K.F. Drinkwater, S.J. Hawkins,
T. Oguz, G. O’Sullivan, and P.C. Reid, 2011: Impacts of climate change on
European marine ecosystems: observations, expectations and indicators.
Journal of Experimental Marine Biology and Ecology, 400, 52-69.
Pilli-Sihlova, K., P. Aatola, M. Ollikainen, and H. Tuomenvirta, 2010: Climate change
and electricity consumption witnessing increasing or decreasing use and
costs? Energy Policy, 38(5), 2409-2419.
Pinto, J.G., U. Ulbrich, G.C. Leckebusch, T. Spangehl, M. Reyers, and S. Zacharias, 2007a:
Changes in storm track and cyclone activity in three SRES ensemble experiments
with the ECHAM5/MPI-OM1 GCM. Climate Dynamics, 29(2-3), 195-210.
Pinto, J.G., E.L. Fröhlich, G.C. Leckebusch, and U. Ulbrich, 2007b: Changing European
storm loss potentials under modified climate conditions according to ensemble
simulations of the ECHAM5/MPI-OM1 GCM. Natural Hazards and Earth System
Sciences, 7(1), 165-175.
Pinto, J.G., C.P. Neuhaus, G.C. Leckebusch, M. Reyers, and M. Kerschgens, 2010:
Estimation of wind storm impacts over Western Germany under future climate
conditions using a statistical-dynamical downscaling approach. Tellus A, 62(2),
188-201.
Pitois, S.G. and C.J. Fox, 2006: Long-term changes in zooplankton biomass concentration
and mean size over the Northwest European shelf inferred from Continuous
Plankton Recorder data. ICES Journal of Marine Science, 63(5), 785-798.
Pitt, M., 2008: The Pitt Review: Lessons Learned from the 2007 Floods. Final Report.
Cabinet Office, London, UK, 505 pp.
Planque, B., J. Fromentin, P. Cury, K.F. Drinkwater, S. Jennings, R.I. Perry, and S. Kifani,
2010: How does fishing alter marine populations and ecosystems sensitivity to
climate? Journal of Marine Systems, 79(3-4), 403-417.
Planton, S., P. Lionello, V. Artale, R. Aznar, A. Carillo, J. Colin, L. Congedi, C. Dubois, A.
Elizalde Arellano, S. Gualdi, E. Hertig, G. JorSanchez, L. Li, J. Jucundus, C.
Piani, P. Ruti, E. Sanchez-Gomez, G. Sannino, F. Sevault, and S. Somot, 2006: The
climate of the Mediterranean region under global warming. In: Mediterranean
Climate Variability [Lionello, P., P. Malanotte-Rizzoli, and R. Boscolo (eds.)].
Elsevier, Amsterdam, Netherlands and Oxford, UK, pp. 399-416.
Poirier, M., J.L. Durand, and F. Volaire, 2012: Persistence and production of perennial
grasses under water deficits and extreme temperatures: importance of intraspecific
vs. interspecific variability. Global Change Biology, 18(12), 3632-3646.
Polemio, M. and O. Petrucci, 2010: Occurrence of landslide events and the role of
climate in the twentieth century in Calabria, southern Italy. Quarterly Journal
of Engineering Geology and Hydrogeology, 43(4), 403-415.
Popov Janevska, D., R. Gospavic, E. Pacholewicz, and V. Popov, 2010: Application of
HACCP-QMRA approach for managing the impact of climate change on food
quality and safety. Food Research International, 43(7), 1915-1924.
Post, J., T. Conradt, F. Suckow, V. Krysanova, F. Wechsung, and F.F. Hattermann, 2008:
Integrated assessment of cropland soil carbon sensitivity to recent and future
climate in the Elbe River basin. Hydrological Sciences Journal, 53(5), 1043-1058.
1321
Europe Chapter 23
23
Powlson, D.S., A.P. Whitmore, and K.W.T. Goulding, 2011: Soil carbon sequestration
to mitigate climate change: a critical re-examination to identify the true and
the false. European Journal of Soil Science, 62(1), 42-55.
Pruszak, Z. and E. Zawadzka, 2008: Potential implications of sea-level rise for Poland.
Journal of Coastal Research, 24(2), 410-422.
Pryor, S.C. and R.J. Barthelmie, 2010: Climate change impacts on wind energy: a
review. Renewable and Sustainable Energy Reviews, 14(1), 430-437.
Pryor, S.C. and J.T. Schoof, 2010: Importance of the SRES in projections of climate
change impacts on near-surface wind regimes. Meteorologische Zeitschrift,
19(3), 267-274.
Purvis, M.J., P.D. Bates, and C.M. Hayes, 2008: A probabilistic methodology to
estimate future coastal flood risk due to sea level rise. Coastal Engineering,
55(12), 1062-1073.
Quevauviller, P., 2011: Adapting to climate change: reducing water-related risks in
Europe EU policy and research considerations. Environmental Science &
Policy, 14(7), 722-729.
Queyriaux, B., A. Armengaud, C. Jeannin, C. Coutourier, and F. Peloux-Petiot, 2008:
Chikungunya in Europe. The Lancet, 371(9614), 723-724.
Quintana-Segui, P., F. Habets, and E. Martin, 2011: Comparison of past and future
Mediterranean high and low extremes of precipitation and river flow projected
using different statistical downscaling methods. Natural Hazards and Earth
System Sciences, 11(5), 1411-1432.
Radovic, V., K. Vitale, and P.B. Tchounwou, 2012: Health facilities safety in natural
disasters: experiences and challenges from South East Europe. International
Journal of Environmental Research and Public Health, 9(5), 1677-1686.
Raftoyannis, Y., I. Spanos, and K. Radoglou, 2008: The decline of Greek fir (Abies
cephalonica Loudon): relationships with root condition. Plant Biosystems,
142(2), 386-390.
Rahel, F.J. and J.D. Olden, 2008: Assessing the effects of climate change on aquatic
invasive species. Conservation Biology, 22(13), 521-533.
Räisänen, J. and J. Eklund, 2012: 21
st
century changes in snow climate in Northern
Europe: a high-resolution view from ENSEMBLES regional climate models.
Climate Dynamics, 38(11-12), 2575-2591.
Randolph, S.E. and D.J. Rogers, 2010: The arrival, establishment and spread of exotic
diseases: patterns and predictions. Nature Reviews, 8(5), 361-371.
Rauthe, M., M. Kunz, and C. Kottmeier, 2010: Changes in wind gust extremes over
Central Europe derived from a small ensemble of high resolution regional
climate models. Meteorologische Zeitschrift, 19(3), 299-312.
Ready, P.D., 2010: Leishmaniasis emergence in Europe. Eurosurveillance, 15(10), 29-
39.
Rees, P., N. van der Gaag, J. de Beer, and F. Heins, 2012: European regional populations:
current trends, future pathways, and policy options. European Journal of
Population, 28(4), 385-416.
Rees, W.G., F.M. Stammler, F.S. Danks, and P. Vitebsky, 2008: Vulnerability of European
reindeer husbandry to global change. Climatic Change, 87(1-2), 199-217.
Refsgaard, J.C., K. Arnbjerg-Nielsen, M. Drews, K. Halsnæs, E. Jeppesen, H. Madsen,
A. Markandya, J.E. Olesen, J.R. Porter, and J.H. Christensen, 2013: The role of
uncertainty in climate change adaptation strategies. A Danish water management
example. Mitigation and Adaptation Strategies for Global Change, 18(3), 337-
359.
Reginster, I. and M. Rounsevell, 2006: Scenarios of future urban land use in Europe.
Environment and Planning B: Planning & Design, 33(4), 619-636.
Renard, B., M. Lang, P. Bois, A. Dupeyrat, O. Mestre, H. Niel, E. Sauquet, C.
Prudhomme, S. Parey, E. Paquet, L. Neppel, and J. Gailhard, 2008: Regional
methods for trend detection: assessing field significance and regional consistency.
Water Resources Research, 44(8), W08419, doi:10.1029/2007WR006268.
Renaudeau, D., J.L. Gourdine, and N.R. St-Pierre, 2011: Meta-analysis of the effect
of high ambient temperature on growing-finishing pigs. Journal of Animal
Science, 89(7), 2220-2230.
Renaudeau, D., A. Collin, S. Yahav, V. De Basilio, J.L. Gourdine, and R.J. Collier, 2012:
Adaptation to hot climate and strategies to alleviate heat stress in livestock
production. Animal, 6(5), 707-728.
Resco de Dios, V., C. Fischer, and C. Colinas, 2007: Climate change effects on
Mediterranean forests and preventive measures. New Forests, 33(1), 29-40.
Revich, B., 2011: Heat-wave, air quality and mortality in the Russian Federation’s
Europe, 2010: preliminary assessment. Human Ecology, 7, 3-9.
Revich, B. and D.A. Shaposhnikov, 2012: Climate change, heat waves, and cold spells
as risk factors for increased mortality in some regions of Russia. Studies on
Russian Economic Development, 23(2), 195-207.
Rickebusch, S., W. Thuiller, T. Hickler, M.B. Araújo, M.T. Sykes, O. Schweiger, and B.
Lafourcade, 2008: Incorporating the effects of changes in vegetation functioning
and CO
2
on water availability in plant habitat models. Biology Letters, 4(5),
556-559.
Rico-Amoros, A.M., J. Olcina-Cantosa, and D. Sauri, 2009: Tourist landuse patterns
and water demand: evidence from the Western Mediterranean. Land Use Policy,
26(2), 493-501.
Rigling, A., C. Bigler, B. Eilmann, E. Feldmeyer-Christe, U. Gimmi, C. Ginzler, U. Graf,
P. Mayer, G. Vacchiano, P. Weber, T. Wohlgemuth, R. Zweifel, and M. Dobbertin,
2013: Driving factors of a vegetation shift from Scots pine to pubescent oak in
dry Alpine forests. Global Change Biology, 19, 229-240.
Rixen, C., M. Teich, C. Lardelli, D. Gallati, M. Pohl, M. Pütz, and P. Bebi, 2011: Winter
tourism and climate change in the Alps: an assessment of resource consumption,
snow reliability, and future snowmaking potential. Mountain Research and
Development, 31(3), 229-236.
Robine, J.M., S.L.K. Cheung, S. Le Roy, H. Van Oyen, C. Griffiths, J.P. Michel, and F.R.
Herrmann, 2008: Death toll exceeded 70,000 in Europe during the summer of
2003. Comptes Rendus Biologies, 331(2), 171-178.
Rockel, B. and K. Woth, 2007: Extremes of near-surface wind speed over Europe and
their future changes as estimated from an ensemble of RCM simulations.
Climatic Change, 81(Suppl. 1), 267-280.
Rocklöv, J. and B. Forsberg, 2010: The effect of high ambient temperature on the
elderly population in three regions of Sweden. International Journal of
Environmental Research and Public Health, 7(6), 2607-2619.
Roiz, D., M. Neteler, C. Castellani, D. Arnoldi, and A. Rizzoli, 2011: Climatic factors
driving invasion of the tiger mosquito (Aedes albopictus) into new areas of
Trentino, northern Italy. PLoS One, 6(4), e14800, doi:10.1371/journal.pone.
0014800.
Rojas, R., L. Feyen, A. Bianchi, and A. Dosio, 2012: Assessment of future flood hazard
in Europe using a large ensemble of bias-corrected regional climate simulations.
Journal of Geophysical Research D: Atmospheres, 117(D17), D17109,
doi:10.1029/2012JD017461.
Rojas, R., L. Feyen, and P. Watkiss, 2013: Climate change and river floods in the
European Union: socio-economic consequences and the costs and benefits of
adaptation. Global Environmental Change, 23(6), 1737-1751.
Roos, J., R. Hopkins, A. Kvarnheden, and C. Dixelius, 2011: The impact of global
warming on plant diseases and insect vectors in Sweden. European Journal of
Plant Pathology, 129(1), 9-19.
Rosan, P. and D. Hammarlund, 2007: Effects of climate, fire and vegetation development
on Holocene changes in total organic carbon concentration in three boreal
forest lakes in northern Sweden. Biogeosciences, 4(6), 975-984.
Rosenzweig, C. and F.N. Tubiello, 2007: Adaptation and mitigation strategies in
agriculture: an analysis of potential synergies. Mitigation and Adaptation
Strategies for Global Change, 12(5), 855-873.
Rosenzweig, C., D. Karoly, M. Vicarelli, P. Neofotis, Q. Wu, G. Casassa, A. Menzel, T.L.
Root, N. Estrella, B. Seguin, P. Tryjanowski, C. Liu, S. Rawlins, and A. Imeson,
2008: Attributing physical and biological impacts to anthropogenic climate
change. Nature, 453, 353-357.
Rötter, R.P., T. Palosuo, N.K. Pirttioja, M. Dubrovski, T. Salo, S. Fronzek, R. Aikasalo,
M. Trnka, A. Ristolainen, and T. Carter, 2011: What would happen to barley
production in Finland if global warming exceeded 4 °C? A model-based
assessment. European Journal of Agronomy, 35(4), 205-214.
Rouault, G., J.N. Candau, F. Lieutier, L.M. Nageleisen, J.C. Martin, and N. Warzée,
2006: Effects of drought and heat on forest insect populations in relation to
the 2003 drought in Western Europe. Annals of Forest Science, 63(6), 613-
624.
Rounsevell, M.D.A. and M.J. Metzger, 2010: Developing qualitative scenario
storylines for environmental change assessment. Wiley Interdisciplinary
Reviews: Climate Change, 1(4), 606-619.
Rounsevell, M.D.A. and D.S. Reay, 2009: Land use and climate change in the UK.
Land Use Policy, 26(Suppl. 1), 160-169.
Rounsevell, M.D.A., I. Reginster, M.B. Araújo, T.R. Carter, N. Dendoncker, F. Ewert, J.I.
House, S. Kankaanpää, R. Leemans, M.J. Metzger, C. Schmit, P. Smith, and G.
Tuck, 2006: A coherent set of future land use change scenarios for Europe.
Agriculture, Ecosystems and Environment, 114(1), 57-68.
Rubolini, D., R. Ambrosini, M. Caffi, P. Brichetti, S. Armiraglio, and N. Saino, 2007a:
Long-term trends in first arrival and first egg laying dates of some migrant and
resident bird species in northern Italy. International Journal of Biometeorology,
51(6), 553-563.
1322
Chapter 23 Europe
23
Rubolini, D., A.P. Møller, K. Rainio, and E. Lehikoinen, 2007b: Intraspecific consistency
and geographic variability in temporal trends of spring migration phenology
among European bird species. Climate Research, 35(1-2), 135-146.
Ruiz-Ramos, D.V., E.A. Hernandez-Delgado, and N.V. Schizas, 2011: Population status
of the long-spined urchin Diadema antillarum in Puerto Rico 20 years after a
mass mortality event. Bulletin of Marine Science, 87(1), 113-127.
Rusch, G.M., 2012: Climate and Ecosystem Services.The Potential of Norwegian
Ecosystems for Climate Mitigation and Adaptation. NINA Report 791, Norwegian
Institute for Nature Research (NINA), Trondheim, Norway, 43 pp.
Rutty, M. and D. Scott, 2010: Will the Mediterranean become “too hot” for tourism?
A reassessment. Tourism Planning & Development, 7(3), 267-281.
Sabbioni, C., A. Bonazza, and P. Messina, 2008: Global climate change and
archaeological heritage: prevision, impact and mapping. In: ARCHAIA. Case
Studies on Research Planning, Characterisation, Conservation and Management
of Archaeological Sites [Marchetti, N. and I. Thuesen (eds.)]. Archaeopress,
Oxford, UK, pp. 295-300.
Sabbioni, C., P. Brimblecombe, and M. Cassar, 2012: Atlas of Climate Change Impact
on European Cultural Heritage: Scientific Analysis and Management Strategies.
Anthem Press, London, UK, 146 pp.
Sabir, M., J. Ommeren, M. Koetse, and P. Rietveld, 2010: Adverse weather and
commuting speed. Networks and Spatial Economics, 11(4), 701-712.
Saino, N., R. Ambrosini, D. Rubolini, J. Von Hardenberg, A. Provenzale, K. Hüppop, O.
Hüppop, A. Lehikoinen, E. Lehikoinen, K. Rainio, M. Romano, and L. Sokolov,
2011: Climate warming, ecological mismatch at arrival and population decline
in migratory birds. Proceedings of the Royal Society B, 278(1707), 835-842.
Sainz-Elipe, S., J.M. Latorre, R. Escosa, M. Masià, M.V. Fuentes, S. Mas-Coma, and
M.D. Bargues, 2010: Malaria resurgence risk in southern Europe: climate
assessment in an historically endemic area of rice fields at the Mediterranean
shore of Spain. Malaria Journal, 9(1), 221-237.
Salis, M., A. Ager, B. Arca, M. Finney, V. Bacciu, P. Duce, and D. Spano, 2013: Assessing
exposure of human and ecological values to wildfire in Sardinia, Italy.
International Journal of Wildland Fire, 22(4), 549-565.
Sanchez-Rodriguez, R., 2009: Learning to adapt to climate change in urban areas.
A review of recent contributions. Current Opinion in Environmental Sustainability,
1(2), 201-206.
San-Miguel-Ayanz, J., M. Rodrigues, S. Santos de Oliveira, C.K. Pacheco, F. Moreira,
B. Duguy, and A. Camia, 2012: Land cover change and fire regime in the
European Mediterranean region. In: Post-Fire Management and Restoration of
Southern European Forests [Moreira, F., M. Arianoutsou, P. Corona, and J. De
las Heras (eds.)]. Managing Forest Ecosystems Series, Book 24, Springer Science,
Dordrecht, Netherlands, pp. 21-43.
San-Miguel-Ayanz, J., J.M. Moreno, and A. Camia, 2013: Analysis of large fires in
European Mediterranean landscapes: lessons learned and perspectives. Forest
Ecology and Management, 294, 11-22.
Santos, J.A., A.C. Malheiro, M.K. Karremann, and J.G. Pinto, 2011: Statistical modelling
of grapevine yield in the Port Wine region under present and future climate
conditions. International Journal of Biometeorology, 55(2), 119-131.
Sauter, T., C. Weitzenkamp, and C. Schneider, 2010: Spatio-temporal prediction of snow
cover in the Black Forest mountain range using remote sensing and a recurrent
neural network. International Journal of Climatology, 30(15), 2330-2341.
Savé, R., F. de Herralde, X. Aranda, E. Pla, D. Pascual, I. Funes, and C. Biel, 2012:
Potential changes in irrigation requirements and phenology of maize, apple
trees and alfalfa under global change conditions in Fluvià watershed during
XXI
st
century: results from a modeling approximation to watershed-level water
balance. Agricultural Water Management, 114, 78-87.
Schaefli, B., B. Hingray, and A. Musy, 2007: Climate change and hydropower
production in the Swiss Alps: quantification of potential impacts and related
modeling uncertainties. Hydrology & Earth System Sciences, 11(3), 1191-1205.
Schär, C. and G. Jendritzky, 2004: Climate change: hot news from summer 2003.
Nature, 432(7017), 559-560.
Schifano, P., M. Leone, M. De Sario, F. de’Donato, A.M. Bargagli, D. D’Ippoliti, C.
Marino, and P. Michelozzi, 2012: Changes in the effects of heat on mortality
among the elderly from 1998-2010: results from a multicentre time series study
in Italy. Environmental Health: A Global Access Science Source, 11(1), 58,
doi:10.1186/1476-069X-11-58.
Schmidli, J., C.M. Goodess, C. Frei, M.R. Haylock, Y. Hundecha, J. Ribalaygua, and T.
Schmith, 2007: Statistical and dynamical downscaling of precipitation: an
evaluation and comparison of scenarios for the European Alps. Journal of
Geophysical Research, 112(D4), D04105, doi:10.1029/2005JD007026.
Schmocker-Fackel, P. and F. Naef, 2010: Changes in flood frequencies in Switzerland
since 1500. Hydrology and Earth System Sciences, 14(8), 1581-1594.
Schnitzler, J., J. Benzler, D. Altmann, I. Mucke, and G. Krause, 2007: Survey on the
population’s needs and the public health response during floods in Germany
2002. Journal of Public Health Management and Practice, 13(5), 461-464.
Scholz, G., J.N. Quinton, and P. Strauss, 2008: Soil erosion from sugar beet in Central
Europe in response to climate change induced seasonal precipitation variations.
Catena, 72(1), 91-105.
Schroter, D., W. Cramer, R. Leemans, I. Prentice, M. Araujo, N. Arnell, A. Bondeau, H.
Bugmann, T. Carter, A. Vega-Leinert, M. Erhard, F. Ewert, M. Glendining, J. House,
S. Kankaanpaa, R. Klein, S. Lavorel, M. Lindner, M. Metzger, J. Meyer, T. Mitchell,
I. Reginster, M. Rounsevell, S. Sabate, S. Sitch, B. Smith, J. Smith, P. Smith, M.
Sykes, K. Thonicke, W. Thuiller, G. Tuck, S. Saehle, and B. Zierl, 2005: Ecosystem
service supply and vulnerability to global change in Europe. Science, 310, 1333-
1337.
Schulze, E.D., S. Luyssaert, P. Ciais, A. Freibauer, I.A. Janssens, J.F. Soussana, P. Smith,
J. Grace, I. Levin, B. Thiruchittampalam, M. Heimann, A.J. Dolman, R. Valentini,
P. Bousquet, P. Peylin, W. Peters, C. Rödenbeck, G. Etiope, N. Vuichard, M.
Wattenbach, G.J. Nabuurs, Z. Poussi, J. Nieschulze, and J.H. Gash, 2010:
Importance of methane and nitrous oxide for Europe’s terrestrial greenhouse-
gas balance. Nature Geoscience, 2(12), 842-850.
Schutze, N. and G.H. Schmitz, 2010: OCCASION: new planning tool for optimal
climate change adaption strategies in irrigation. Journal of Irrigation and
Drainage Engineering, 136(12), 836-846.
Schwarze, R., M. Schwindt, H. Weck-Hannemann, P. Raschky, F. Zahn, and G. Wagner,
2011: Natural hazard insurance in Europe: tailored responses to climate change
are needed. Environmental Policy and Governance, 21, 14-30.
Schweiger, O., R.K. Heikkinen, A. Harpke, T. Hickler, S. Klotz, O. Kudrna, I. Kühn, J.
Pöyry, and J. Settele, 2012: Increasing range mismatching of interacting species
under global change is related to their ecological characteristics. Global Ecology
and Biogeography, 21(1), 88-99.
Schwierz, C., P. Köllner-Heck, E.Z. Mutter, D.N. Bresch, P.L. Vidale, M. Wild, and C.
Schär, 2010: Modelling European winter wind storm losses in current and future
climate. Climatic Change, 101(3), 485-514.
Seidl, R. and M. Lexer, 2013: Forest management under climatic and social
uncertainty: trade-offs between reducing climate change impacts and fostering
adaptive capacity. Journal of Environmental Management, 114, 461-469.
Seidl, R., M.-J. Schelhaas, M. Lindner, and M.J. Lexer, 2009: Modelling bark beetle
disturbances in a large scale forest scenario model to assess climate change
impacts and evaluate adaptive management strategies. Regional Environmental
Change, 9(2), 101-119.
Seidl, R., W. Rammer, and M. Lexer, 2011: Climate change vulnerability of sustainable
forest management in the Eastern Alps. Climatic Change, 106, 225-254.
Seljom, P., E. Rosenberg, A. Fidge, J. Haugen, M. Meir, J. Rekstad, and T. Jarlset, 2011:
Modelling the effects of climate change on the energy system – a case study
of Norway. Energy Policy, 39(11), 7310-7321.
Semenov, M.A., 2009: Impacts of climate change on wheat in England and Wales.
Journal of the Royal Society Interface, 6(33), 343-350.
Semenov, V.A., 2011: Climate-related changes in hazardous and adverse hydrological
events in the Russian rivers. Russian Meteorology and Hydrology, 36(2), 124-
129.
Semenza, J.C. and B. Menne, 2009: Climate change and infectious diseases in Europe.
Lancet Infectious Diseases, 9(6), 365-375.
Semenza, J., J. Suk, V. Estevez, K.L. Ebi, and E. Lindgren, 2012: Mapping climate
change vulnerabilities to infectious diseases in Europe. Environmental Health
Perspectives, 120(3), 385-392.
Senatore, A., G. Mendicino, G. Smiatek, and H. Kunstmann, 2011: Regional climate
change projections and hydrological impact analysis for a Mediterranean basin
in Southern Italy. Journal of Hydrology, 399(1-2), 70-92.
Serquet, G. and M. Rebetez, 2011: Relationship between tourism demand in the
Swiss Alps and hot summer air temperatures associated with climate change.
Climatic Change, 108(1), 291-300.
Shvidenko, A.Z., D.G. Shchepashchenko, E.A. Vaganov, A.I. Sukhinin, S.S. Maksyutov,
I. McCallum, and I.P. Lakyda, 2011: Impact of wildfire in Russia between 1998-
2010 on ecosystems and the global carbon budget. Doklady Earth Sciences,
441(2), 1678-1682.
Siebert, S. and F. Ewert, 2012: Spatio-temporal patterns of phenological development
in Germany in relation to temperature and day length. Agricultural and Forest
Meteorology, 152, 44-57.
1323
Europe Chapter 23
23
Silva, D.E., P. Rezende Mazzella, M. Legay, E. Corcket, and J.L. Dupouey, 2012: Does
natural regeneration determine the limit of European beech distribution under
climatic stress? Forest Ecology and Management, 266, 263-272.
Sirotenko, O.D. and E.V. Abashina, 2008: Modern climate changes of biosphere
productivity in Russia and adjacent countries. Russian Meteorology and
Hydrology, 33(4), 267-271.
Skeffington, M.S. and K. Hall, 2011: The ecology, distribution and invasiveness of
Gunnera L. species in Connemara, Western Ireland. Biology and Environment,
111(3), 157-175.
Slangen, A., C. Katsman, R. van de Wal, L. Vermeersen, and R. Riva, 2012: Towards
regional projections of twenty-first century sea-level change based on IPCC
SRES scenarios. Climate Dynamics, 38, 1191-1209.
Smith, P. and J.E. Olesen, 2010: Synergies between mitigation of, and adaptation to,
climate change in agriculture. Journal of Agricultural Science, 148(5), 543-
552.
Smith, P., P.J. Gregory, D. van Vuuren, M. Obersteiner, P. Havlik, M. Rounsevell, J.
Woods, E. Stehfest, and J. Bellarby, 2010: Competition for land. Philosophical
Transactions of the Royal Society B, 365(1554), 2941-2957.
Smith, T.M., R.W. Reynolds, T.C. Peterson, and J. Lawrimore, 2008: Improvements to
NOAAsHistorical Merged Land-Ocean Surface Temperature analysis (1880-
2006). Journal of Climate, 21(10), 2283-2293.
Soane, B.D., B.C. Ball, J. Arvidsson, G. Basch, F. Moreno, and J. Roger-Estrade, 2012:
No-till in northern, western and south-western Europe: a review of problems
and opportunities for crop production and the environment. Soil and Tillage
Research, 118, 66-87.
Solberg, S., Ø. Hov, A. Søvde, I.S.A. Isaksen, P. Coddeville, H. De Backer, C. Forster, Y.
Orsolini, and K. Uhse, 2008: European surface ozone in the extreme summer
2003. Journal of Geophysical Research, 113(D7), D07307, doi:10.1029/
2007JD009098.
Sorte, C.J.B., S.L. Williams, and R.A. Zerebecki, 2010: Ocean warming increases threat
of invasive species in a marine fouling community. Ecology, 91, 2198-2204.
Sousa, P.M., R.M. Trigo, P. Aizpurua, R. Nieto, L. Gimeno, and R. Garcia-Herrera, 2011:
Trends and extremes of drought indices throughout the 20
th
century in the
Mediterranean. Natural Hazards and Earth System Science, 11(1), 33-51.
Soussana, J.F. and A. Luscher, 2007: Temperate grasslands and global atmospheric
change: a review. Grass and Forage Science, 62(2), 127-134.
Soussana, J.F., A.I. Graux, and F.N. Tubiello, 2010: Improving the use of modelling for
projections of climate change impacts on crops and pastures. Journal of
Experimental Botany, 61(8), 2217-2228.
Spangenberg, L., F. Battke, M. Grana, K. Nieselt, and H. Naya, 2011: Identifying
associations between amino acid changes and meta information in alignments.
Bioinformatics, 27(20), 2782-2789.
Stafoggia, M., F. Forastiere, D. Agostini, N. Caranci, F. de’Donato, M. Demaria, P.
Michelozzi, R. Miglio, M. Rognoni, A. Russo, and C.A. Perucci, 2008: Factors
affecting in-hospital heat-related mortality: a multi-city case-crossover analysis.
Journal of Epidemiology and Community Health, 62(3), 209-215.
Stahl, K., H. Hisdal, J. Hannaford, L.M. Tallaksen, H.A.J. van Lanen, E. Sauquet, S.
Demuth, M. Fendekova, and J. dar, 2010: Streamflow trends in Europe:
evidence from a dataset of near-natural catchments. Hydrology & Earth System
Sciences, 14, 2367-2382.
Stanzel, P. and H.P. Nachtnebel, 2010: Mögliche auswirkungen des klimawandels auf
den wasserhaushalt und die wasserkraftnutzung in Österreich. Österreichische
Wasser- und Abfallwirtschaft, 62(9-10), 180-187.
Steele-Dunne, S., P. Lynch, R. McGrath, T. Semmler, S. Wang, J. Hanafin, and P. Nolan,
2008: The impacts of climate change on hydrology in Ireland. Journal of
Hydrology, 356(1-2), 28-45.
Steger, C., S. Kotlarski, T. Jonas, and C. Schär, 2013: Alpine snow cover in a changing
climate: a regional climate model perspective. Climate Dynamics, 41(3-4), 735-
754.
Steiger, R., 2010: The impact of climate change on ski season length and snowmaking
requirements in Tyrol, Austria. Climate Research, 43, 251-262.
Steiger, R., 2011: The impact of snow scarcity on ski tourism. An analysis of the
record warm season 2006/07 in Tyrol (Austria). Tourism Review, 66(3), 4-13.
Steiger, R., 2012: Scenarios for skiing tourism in Austria: integratingdemographics
with an analysis of climate change. Journal of Sustainable Tourism, 20(6), 867-
882.
Steiger, R. and M. Mayer, 2008: Snowmaking and climate change: future options for
snow production in Tyrolean ski resorts. Mountain Research and Development,
28(3-4), 292-298.
Sterl, A., H. van den Brink, H. de Vries, R. Haarsma, and E. van Meijgaard, 2009: An
ensemble study of extreme storm surge related water levels in the North Sea
in a changing climate. Ocean Science, 5, 369-378.
Stoate, C., A. Báldi, P. Beja, N.D. Boatman, I. Herzon, A. van Doorn, G.R. de Snoo, L.
Rakosy, and C. Ramwell, 2009: Ecological impacts of early 21
st
century agricultural
change in Europe a review. Journal of Environmental Management, 91(1),
22-46.
Stoffel, M. and C. Huggel, 2012: Effects of climate change on mass movements in
mountain environments. Progress in Physical Geography, 36(3), 421-439.
Stoll, S., H.J. Hendricks Franssen, M. Butts, and W. Kinzelbach, 2011: Analysis of the
impact of climate change on groundwater related hydrological fluxes: a multi-
model approach including different downscaling methods. Hydrology & Earth
System Sciences, 15, 21-38.
Storm, J., A.W. Cattaneo, and F. Trincardi, 2008: Coastal dynamics under conditions
of rapid sea-level rise: Late Pleistocene to Early Holocene evolution of barrier-
lagoon systems on the Northern Adriatic shelf (Italy). Quaternary Science
Reviews, 27(11-12), 1107-1123.
Stratonovitch, P., 2012: A process-based approach to modelling impacts of climate
change on the damage niche of an agricultural weed. Global Change Biology,
18(6), 2071-2080.
Streftaris, N., A. Zenetos, and E. Papathanassiou, 2005: Globalisation in marine
ecosystems: the story of non-indigenous marine species across European seas.
Oceanography and Marine Biology an Annual Review, 43, 419-453.
Supit, I., C.A. van Diepen, A.J.W. de Wit, P. Kabat, B. Baruth, and F. Ludwig, 2010:
Recent changes in the climatic yield potential of various crops in Europe.
Agricultural Systems, 103, 683-694.
Surminski, S. and A. Philp, 2010: Briefing: guidance on insurance issues for new
developments. Proceedings of the Institution of Civil Engineers: Engineering
Sustainability, 163(1), 3-6.
Swart, R., L. Bernstein, M. Ha-Duong, and A. Petersen, 2009a: Agreeing to disagree:
uncertainty management in assessing climate change, impacts and responses
by the IPCC. Climatic Change, 92(1), 1-29.
Swart, R., R. Biesbroek, S. Binnerup, T.R. Carter, C. Cowan, T. Henrichs, S. Loquen, H. Mela,
M. Morecroft, M. Reese, and D. Rey, 2009b: Europe Adapts to Climate Change:
Comparing National Adaptation Strategies. PEER Report No. 1, Partnership for
European Environmental Research (PEER), Helsinki, Finland, 280 pp.
Swedish Commission on Climate and Vulnerability, 2007: Sweden Facing Climate
Change Threats and Opportunities. Final report from the Swedish Commission
on Climate and Vulnerability, Swedish Government Official Reports SOU
2007:60, Stockholm, Sweden, 679 pp.
Tardieu, F., 2012: Any trait or trait-related allele can confer drought tolerance: just
design the right drought scenario. Journal of Experimental Botany, 63(1), 25-31.
Tasker, M.L. (ed.), 2008: The Effect of Climate Change on the Distribution and
Abundance of Marine Species in the OSPAR Maritime Area. ICES Cooperative
Research ReportNo. 293, International Council for the Exploration of the Sea
(ICES), Copenhagen, Denmark, 45 pp.
Taylor, S., L. Kumar, N. Reid, and D.J. Kriticos, 2012: Climate change and the potential
distribution of an invasive shrub, Lantana camara L. PLoS ONE, 7(4), e35565,
doi:10.1371/journal.pone.0035565.
te Linde, A.H., 2007: Effects of climate change on discharge behaviour of the river
Rhine. In: Proceedings of the Third International Conference on Climate and
Water, 3-6, September, 2007 [Heinonen, M. (ed.)]. Organized by the Finnish
Environment Institute (SYKE), Helsinki, Finland, pp. 296-301, protosh2o.act.be/
VIRTUELE_BIB/Water_in_de_Wereld/MIL-Milieu/W_MIL_E7_Proceedings_
conference.pdf.
te Linde, A.H., J.C.J.H. Aerts, A.M.R. Bakker, and J.C.J. Kwadijk, 2010a: Simulating
low-probability peak discharges for the Rhine basin using resampled climate
modeling data. Water Resources Research, 46(3), W03512, doi:10.1029/
2009WR007707.
te Linde, A.H., J.C.J.H. Aerts, and J.C.J. Kwadijk, 2010b: Effectiveness of flood
management measures on peak discharges in the Rhine basin under climate
change. Journal of Flood Risk Management, 3(4), 248-269.
te Linde, A.H., P. Bubeck, J.E.C. Dekkers, H. De Moel, and J.C.J.H. Aerts, 2011: Future
flood risk estimates along the river Rhine. Natural Hazards and Earth System
Sciences, 11(2), 459-473.
Teich, M., C. Marty, C. Gollut, A. Grêt-Regamey, and P. Bebi, 2012: Snow and weather
conditions associated with avalanche releases in forests: rare situations with
decreasing trends during the last 41years. Cold Regions Science and Technology,
83-84, 77-88.
1324
Chapter 23 Europe
23
ten Brinke, W.B.M., B. Kolen, A. Dollee, H. van Waveren, and K. Wouters, 2010:
Contingency planning for large-scale floods in the Netherlands. Journal of
Contingencies and Crisis Management, 18(1), 55-69.
ter Hofstede, R., J. Hiddink, and A. Rijnsdorp, 2010: Regional warming changes fish
species richness in the Eastern North Atlantic Ocean. Marine Ecology Progress
Series, 414, 1-9.
Terpstra, T. and J.M. Gutteling, 2008: Households’ perceived responsibilities in flood
risk management in the Netherlands. International Journal of Water Resources
Development, 24(4), 555-565.
Tervo, K., 2008: The operational and regional vulnerability of winter tourism to
climate variability and change: the case of the Finnish nature-based tourism
entrepreneurs. Scandinavian Journal of Hospitality and Tourism, 8(4), 317-332.
Thieken, A.H., T. Petrow, H. Kreibich, and B. Merz, 2006: Insurability and mitigation
of flood losses in private households in Germany. Risk Analysis, 26(2), 383-
395.
Thodsen, H., 2007: The influence of climate change on stream flow in Danish rivers.
Journal of Hydrology, 333(2-4), 226-238.
Thodsen, H., B. Hasholt, and J.H. Kjarsgaard, 2008: The influence of climate change
on suspended sediment transport in Danish rivers. Hydrological Processes,
22(6), 764-774.
Three Regions Climate Change Group, 2008: Your Home in a Changing Climate.
Retrofitting Existing Homes in a Changing Climate. Report for Policy Makers.
Prepared by ARUP for the Three Regions Climate Change Group, the Greater
London Authority, London, UK, 75 pp.
Thuiller, W., S. Lavergne, C. Roquet, I. Boulangeat, B. Lafourcade, and M.B. Araujo,
2011: Consequences of climate change on the tree of life in Europe. Nature,
470, 531-534.
Trnka, M., F. Muska, D. Semeradova, M. Dubrovsky, E. Kocmankova, and Z. Zalud, 2007:
European Corn Borer life stage model: regional estimates of pest development
and spatial distribution under present and future climate. Ecological Modelling,
207(2-4), 61-84.
Trnka, M., J. Eitzinger, P. Hlavinka, M. Dubrovska, D. Semerádová, P. Åtapanek, S.
Thaler, Z. Åsalud, M. Molna, and H. Formayer, 2009: Climate-driven changes of
production regions in Central Europe. Plant and Soil, 2009(521), 257-266.
Trnka, M., E. Kocmánková, J. Baleka, J. Eitzinger, F. Ruget, H. Formayer, P. Hlavinka,
M. Schaumberger, V. Horáková, M. Možný, and Z. Žaluda, 2010: Simple snow
cover model for agrometeorological applications. Agricultural and Forest
Meteorology, 150, 1115-1127.
Trnka, M., J.E. Olesen, K.C. Kersebaum, A.O. Skjelvåg, J. Eitzinger, B. Seguin, P.
Peltonen-Sainio, R. Rötter, A. Iglesias, S. Orlandini, M. Dubrovský, P. Hlavinka, J.
Balek, H. Eckersten, E. Cloppet, P. Calanca, A. Gobin, V. Vučetić, P. Nejedlik, S.
Kumar, B. Lalic, A. Mestre, F. Rossi, J. Kozyra, V. Alexandrov, D. Semerádová, and
Z. Žalud, 2011: Agroclimatic conditions in Europe under climate change. Global
Change Biology, 17(7), 2298-2318.
Troccoli, A., F. Zambon, K. Hodges, and M. Marani, 2012a: Storm surge frequency
reduction in Venice under climate change. Climatic Change, 113(3-4), 1065-
1079.
Troccoli, A., F. Zambon, K.I. Hodges, and M. Marani, 2012b: Reply to comment on
“Storm surge frequency reduction in Venice under climate change” by G. Jordà,
D. Gomis & M. Marcos. Climatic Change, 113(3-4), 1089-1095.
Tsanis, I.K., A.G. Koutroulis, I.N. Daliakopoulos, and D. Jacob, 2011: Severe climate-
induced water shortage and extremes in Crete. Climatic Change, 106(4), 667-
677.
Tu, M., M.J. Hall, P.J.M. de Laat, and M.J.M. de Wit, 2005: Extreme floods in the Meuse
river over the past century: aggravated by land-use changes? Physics and
Chemistry of the Earth, Parts A/B/C, 30(4-5), 267-276.
Tubiello, F.N., J.F. Soussana, and S.M. Howden, 2007: Crop and pasture response to
climate change. Proceedings of the National Academy of Sciences of the United
States of America, 104(50), 19686-19690.
Tuck, G., M.J. Glendining, P. Smith, J.I. House, and M. Wattenbach, 2006: The potential
distribution of bioenergy crops in Europe under present and future climate.
Biomass and Bioenergy, 30(3), 183-197.
Turco, M., M.C. Llasat, J. von Hardenberg, and A. Provenzale, 2013: Impact of climate
variability on summer fires in a Mediterranean environment (northeastern
Iberian Peninsula). Climatic Change, 116(3-4), 665-678.
Uhlmann, B., S. Goyette, and M. Beniston, 2009: Sensitivity analysis of snow patterns
in Swiss ski resorts to shifts in temperature, precipitation and humidity under
conditions of climate change. International Journal of Climatology, 29(8), 1048-
1055.
UK National Ecosystem Assessment, 2011: The UK National Ecosystem
Assessment: Technical Report. United Nations Environment Programme World
Conservation Monitoring Centre (UNEP-WCMC), UNEP-WCMC, Cambridge, UK,
1466 pp.
UK-ASC, 2011: Adapting to Climate Change in the UK: Measuring Progress. Adaptation
Sub-Committee (UK-ASC) Progress Report, Committee on Climate Change
(CCC), London, UK, 101 pp.
Ulbrich, U., G.C. Leckebusch, and J.G. Pinto, 2009: Extra-tropical cyclones in the
present and future climate: a review. Theoretical and Applied Climatology,
96(1-2), 117-131.
Ulbrich, U., P. Lionello, D. Belušic, J. Jacobeit, P. Knippertz, F.G. Kuglitsch, G.C.
Leckebusch, J. Luterbacher, M. Maugeri, P. Maheras, K.M. Nissen, V. Pavan, J.G.
Pinto, H. Saaroni, S. Seubert, A. Toreti, E. Xoplaki, and B. Ziv, 2012: Climate of
the Mediterranean: synoptic patterns, temperature, precipitation, winds, and
their extremes. In: The Climate of the Mediterranean Region: From the Past to
the Future [Lionello, P. (ed.)]. Elsevier, London, UK and Waltham, MA, USA, pp.
301-346.
Ulén, B.M. and G.A. Weyhenmeyer, 2007: Adapting regional eutrophication targets
for surface waters – influence of the EU Water Framework Directive, national
policy and climate change. Environmental Science & Policy, 10(7-8), 734-742.
Unbehaun, W., U. Pröbstl, and W. Haider, 2008: Trends in winter sport tourism:
challenges for the future. Tourism Review, 63(1), 36-47.
UNEP, 2010: Global Synthesis – A Report from the Regional Seas Conventions and
Action Plans for the Marine Biodiversity Assessment and Outlook Series. United
Nations Environment Programme (UNEP), Nairobi, Kenya, 55 pp.
Usbeck, T., T. Wohlgemuth, M. Dobbertin, C. Pfister, A. Burgi, and M. Rebetez, 2010:
Increasing storm damage to forests in Switzerland from 1858 to 2007.
Agricultural and Forest Meteorology, 150, 47-55.
Van der Linden, P. and J.F.B. Mitchell, 2009: ENSEMBLES: Climate Change and its
Impacts: Summary of Research and Results from The ENSEMBLES Project. Met
Office Hadley Centre, Exeter, UK, 160 pp.
van der Velde, M., G. Wriedt, and F. Bouraoui, 2010: Estimating irrigation use and
effects on maize yield during the 2003 heatwave in France. Agriculture
Ecosystems & Environment, 135(1-2), 90-97.
van Dijk, J., N.D. Sargison, F. Kenyon, and P.J. Skuce, 2010: Climate change and
infectious disease: helminthological challenges to farmed ruminants in temperate
regions. Animal, 4(3), 377-392.
Van Nieuwaal, K., P. Driessen, T. Spit, and C. Termeer, 2009: A State of the Art of
Governance Literature on Adaptation to Climate Change: Towards a Research
Agenda.Report No. 003/2009, Knowledge for Climate (KfC), KfC Secretariat,
Utrecht University, Utrecht, Netherlands, 43 pp.
van Vliet, M.T.H. and J.J.G. Zwolsman, 2008: Impact of summer droughts on the
water quality of the Meuse river. Journal of Hydrology, 353(1-2), 1-17.
van Vliet, M.T.H., J.R. Yearsley, F. Ludwig, S. Vögele, D.P. Lettenmaier, and P. Kabat,
2012: Vulnerability of US and European electricity supply to climate change.
Nature Climate Change, 2(9), 676-681.
Varakina, Z.L., D.A. Shaposhnikov, B.A. Revich, A.M. Vyazmin, E.D. Yurasova, J. Nurse,
and B. Menne, 2011: The projected impact of climate change on the daily
mortality: a case study in Archangelsk city in Northwest Russia. European
Journal of Public Health, 21(Suppl. 1), 159, eurpub.oxfordjournals.org/
content/21/suppl_1/10.full.pdf.
Vautard, R., J. Cattiaux, P. Yiou, J.N. Thepaut, and P. Ciais, 2010: Northern Hemisphere
atmospheric stilling partly attributed to an increase in surface roughness.
Nature Geoscience, 3(11), 756-761.
Veijalainen, N., E. Lotsari, P. Alho, B. Vehviläinen, and J. Käyhkö, 2010: National scale
assessment of climate change impacts on flooding in Finland. Journal of
Hydrology, 391(3-4), 333-350.
Ventrella, D., M. Charfeddine, and M. Bindi, 2012: Agronomic adaptation strategies
under climate change for winter durum wheat and tomato in southern Italy:
irrigation and nitrogen fertilization. Regional Environmental Change, 12(3),
407-419.
Verburg, P.H., D.B. van Berkel, A.M. van Doorn, M. van Eupen, and H.A.R.M. van den
Heiligenberg, 2010: Trajectories of land use change in Europe: a model-based
exploration of rural futures. Landscape Ecology, 25(2), 217-232.
Verny, J. and C. Grigentin, 2009: Container shipping on the Northern Sea Route.
International Journal of Production Economics, 122(1), 107-117.
Vidal, J.P. and S. Wade, 2009: A multimodel assessment of future climatological
droughts in the United Kingdom. International Journal of Climatology, 29(14),
2056-2071.
1325
Europe Chapter 23
23
Vilén, T. and P.M. Fernandes, 2011: Forest fires in Mediterranean countries: CO
2
emissions and mitigation possibilities through prescribed burning. Environmental
Management, 48, 558-567.
Villarini, G., J.A. Smith, F. Serinaldi, and A.A. Ntelekos, 2011: Analyses of seasonal
and annual maximum daily discharge records for Central Europe. Journal of
Hydrology, 399(3-4), 299-312.
Vinagre, C., F.D. Santos, H. Cabral, and M.J. Costa, 2011: Impact of climate warming
upon the fish assemblages of the Portuguese coast under different scenarios.
Regional Environmental Change, 11(4), 779-789.
Virkkala, R., R.K. Heikkinen, S. Fronzek, H. Kujala, and N. Leikola, 2013: Does the
protected area network preserve bird species of conservation concern in a
rapidly changing climate? Biodiversity and Conservation, 22(2), 459-482.
Vorogushyn, S., K.E. Lindenschmidt, H. Kreibich, H. Apel, and B. Merz, 2012: Analysis
of a detention basin impact on dike failure probabilities and flood risk for a
channel-dike-floodplain system along the river Elbe, Germany. Journal of
Hydrology, 436-437, 120-131.
Vos, C.C., P. Berry, P. Opdam, H. Baveco, B. Nijhof, J. O’Hanley, C. Bell, and H. Kuipers,
2008: Adapting landscapes to climate change: examples of climate-proof
ecosystem networks and priority adaptation zones. Journal of Applied Ecology,
45(6), 1722-1731.
Wade, S.D., J. Rance, and N. Reynard, 2013: The UK climate change risk assessment
2012: assessing the impacts on water resources to inform policy makers. Water
Resources Management, 27(4), 1085-1109.
Wall, R. and L.S. Ellse, 2011: Climate change and livestock parasites: integrated
management of sheep blowfly strike in a warmer environment. Global Change
Biology, 17(5), 1770-1777.
Walther, G.R., A. Roques, P.E. Hulme, M.T. Sykes, P. Pyšek, I. Kühn, M. Zobel, S. Bacher,
Z. Botta-Dukát, H. Bugmann, B. Czúcz, J. Dauber, T. Hickler, V. Jarošík, M. Kenis,
S. Klotz, D. Minchin, M. Moora, W. Nentwig, J. Ott, V.E. Panov, B. Reineking, C.
Robinet, V. Semenchenko, W. Solarz, W. Thuiller, M. Vilà, K. Vohland, and J.
Settele, 2009: Alien species in a warmer world: risks and opportunities. Trends
in Ecology & Evolution, 24(12), 686-693.
Wamsler, C. and N. Lawson, 2011: The role of formal and informal insurance
mechanisms for reducing urban disaster risk: a South-North comparison.
Housing Studies, 26(2), 197-223.
Wang, S., R. McGrath, T. Semmler, and P. Nolan, 2006: The impact of the climate
change on discharge of Suir River Catchment (Ireland) under different climate
scenarios. Natural Hazards and Earth System Science, 6(3), 387-395.
Wang, S., R. McGrath, J. Hanafin, P. Lynch, T. Semmler, and P. Nolan, 2008: The impact
of climate change on storm surges over Irish waters. Ocean Modelling, 25(1-2),
83-94.
Ward, P., H. Renssen, J. Aerts, and P. Verburg, 2011: Sensitivity of discharge and flood
frequency to twenty-first century and late Holocene changes in climate and
land use (River Meuse, northwest Europe). Climatic Change, 106(2), 179-
202.
Ward, R.E.T., C. Herweijer, N. Patmore, and R. Muir-Wood, 2008: The role of insurers
in promoting adaptation to the impacts of climate change. The Geneva Papers,
33, 133-139. doi:10.1057/palgrave.gpp.2510153.
Wasowski, J., C. Lamanna, and D. Casarano, 2010: Influence of land-use change and
precipitation patterns on landslide activity in the Daunia Apennines, Italy.
Quarterly Journal of Engineering Geology and Hydrogeology, 43(4), 387-401.
Watkiss, P. and A. Hunt, 2010: Review of Adaptation Costs and Benefit Estimates in
Europe for SOER 2010. Report prepared by Metroeconomica Economic and
Environmental Consultants and Paul Watkiss Associates for the European
Environment Agency (EEA) under the Framework contract for economic support
(EEA/IEA/09/002) Lot 2 ‘Climate change adaptation: costs and benefits’,
Metroeconomica, Bath, UK, 74 pp.
Weber, R.W.S., 2009: An evaluation of possible effects of climate change on pathogenic
fungi in apple production using fruit rots as examples. Erwerbs-Obstbau, 51(3),
115-120.
Wedawatta, G.S.D. and M.J.B. Ingirige, 2012: Resilience and adaptation of small and
medium-sized enterprises to flood risk. Disaster Prevention and Management,
21(4), 474-488.
Wessel, W., A. Tietema, C. Beier, B. Emmett, J. Penuelas, and T. Riis-Nielson, 2004: A
qualitative ecosystem assessment for different shrublands in Western Europe
under impact of climate change. Ecosystems, 7, 662-671.
Wessolek, G. and S. Asseng, 2006: Trade-off between wheat yield and drainage under
current and climate change conditions in northeast Germany. European Journal
of Agronomy, 24(4), 333-342.
West, J.S., J.A. Townsend, M. Stevens, and B.D.L. Fitt, 2012: Comparative biology of
different plant pathogens to estimate effects of climate change on crop diseases
in Europe. European Journal of Plant Pathology, 133, 315-331.
Westerhoff, L., E.H. Keskitalo, and S. Juhola, 2011: Capacities across scales: local to
national adaptation policy in four European countries. Climate Policy, 11(4),
1071-1085.
Wethey, D.S., S.A. Woodin, T.J. Hilbish, S.J. Jones, F.P. Lima, and P.M. Brannock, 2011:
Response of intertidal populations to climate: effects of extreme events verses
long term change. Journal of Experimental Marine Biology and Ecology,
400(1-2), 132-144.
White, M.A., P. Whalen, and G.V. Jones, 2009: Land and wine. Nature Geoscience, 2,
82-84.
Whitehead, P.G., R.L. Wilby, R.W. Battarbee, M. Kernan, and A.J. Wade, 2009: A
review of the potential impacts of climate change on surface water quality.
Hydrological Sciences Journal, 54(1), 101-123.
Whittle, R., W. Medd, H. Deeming, E. Kashefi, M. Mort, C. Twigger Ross, G. Walker,
and N. Watson, 2010: After the Rain Learning the Lessons from Flood
Recovery in Hull. Final Project Report for ‘Flood, Vulnerability and Urban
Resilience: a real-time study of local recovery following the floods of June 2007
in Hull’, Lancaster University, Lancaster, UK, 171 pp.
WHO, 2008: Heat-Health Action Plans: Guidance [Matthies, F., G. Bickler, N. Cardeñosa
Marín, and S. Hales (eds.)]. World Health Organisation (WHO) Regional Office
for Europe, Copenhagen, Denmark, 45 pp.
WHO, 2013: Floods in the WHO European Region: Health Effects and their Prevention
[Menne, B. and V. Murray (eds.)]. World Health Organization (WHO) Regional
Office for Europe, Copenhagen, Denmark, 134 pp.
Wiering, M.A. and B.J.M. Arts, 2006: Discursive shifts in Dutch river management:
‘deep’ institutional change or adaptation strategy? In: Living Rivers: Trends and
Challenges in Science and Management [Leuven, R.S.E.W., A.M.J. Ragas, A.J.M.
Smits, and G. Velde (eds.)]. Springer, Dordrecht, Netherlands, pp. 327-338.
Wilby, R.L., 2008: Constructing climate change scenarios of urban heat island intensity
and air quality. Environment and Planning B: Planning and Design, 35(5), 902-
919.
Wilkinson, P., K.R. Smith, M. Davies, H. Adair, B.G. Armstrong, M. Barrett, N. Bruce,
A. Haines, I. Hamilton, T. Oreszczyn, I. Ridley, C. Tonne, and Z. Chalabi, 2009:
Public health benefits of strategies to reduce greenhouse-gas emissions:
household energy. The Lancet, 374(9705), 1917-1929.
Willems, P., J. Olsson, K. Arnbjerg-Nielsen, S. Beecham, A. Pathirana, I. Bülow
Gregersen, H. Madsen, and V. Nguyen, 2012: Impacts of Climate Change
on Rainfall Extremes and Urban Drainage. IWA Publishing, London, UK,
226 pp.
Wilson, A.J. and P.S. Mellor, 2009: Bluetongue in Europe: past, present and future.
Philosophical Transactions of the Royal Society B, 364(1530), 2669-2681.
Wilson, D., H. Hisdal, and D. Lawrence, 2010: Has streamflow changed in the Nordic
countries? – Recent trends and comparisons to hydrological projections. Journal
of Hydrology, 394(3-4), 334-346.
Wilson, E., 2006: Adapting to climate change at the local level: the spatial planning
response. Local Environment, 11(6), 609-625.
Wilson, G., 2008: Our knowledge ourselves: Engineers (re)thinking technology in
development. Journal of International Development, 20(6), 739-750.
WLO, 2006: Welvaart en Leefomgeving: Een Scenariostudie voor Nederland in 2040
[Janssen, L.H.J.M., V.R. Okker, and J. Schuur (eds.)]. Central Planning Bureau,
Netherlands Environmental Assessment Agency and Spatial Planning Bureau,
The Hague, Netherlands, 239 pp. (in Dutch).
WMO, 2011: Climate Knowledge for Action: A Global Framework for Climate Services
– Empowering the Most Vulnerable.WMO-No. 1065, Report of the High-Level
Taskforce for the Global Framework for Climate Services, World Meteorological
Organization (WMO), Geneva, Switzerland, 240 pp.
Wong, W.K., B. Stein, E. Torill, H. Ingjerd, and H. Hege, 2011: Climate change effects
on spatiotemporal patterns of hydroclimatological summer droughts in Norway.
Journal of Hydrometeorology, 12(6), 1205-1220, doi: 10.1175/2011JHM1357.1.
Wreford, A., D. Moran, and N. Adger, 2010: Climate Change and Agriculture. Impacts,
Adaptation and Mitigation. OECD Publishing, Paris, France, 135 pp.
Yiou, P., P. Ribereau, P. Naveau, M. Nogaj, and R. Brazdil, 2006: Statistical analysis of
floods in Bohemia (Czech Republic) since 1825. Hydrological Sciences Journal,
51(5), 930-945.
Zachariadis, T., 2010: Forecast of electricity consumption in Cyprus up to the year
2030: the potential impact of climate change. Energy Policy, 38(2), 744-
750.
1326
Chapter 23 Europe
23
Zhou, Q., P.S. Mikkelsen, K. Halsnæs, and K. Arnbjerg-Nielsen, 2012: Framework for
economic pluvial flood risk assessment considering climate change effects and
adaptation benefits. Journal of Hydrology, 414-415, 539-549.
Zsamboky, M., A. Fernandez-Bilbao, D. Smith, J. Knight, and J. Allan, 2011: Impacts
of Climate Change on Disadvantaged UK Coastal Communities. Joseph
Rowntree Foundation, York, UK, 62 pp.
Zwicke, M., G.A. Alessio, L. Thiery, R. Falcimagne, R. Baumont, N. Rossignol, J.
Soussana, and C. Picon-Cochard, 2013: Lasting effects of climate disturbance
on perennial grassland above-ground biomass production under two cutting
frequencies. Global Change Biology, 19(11), 3435-3448, doi:10.1111/
gcb.12317.