Download 5.5.1 Introduction - Eionet Forum

Document related concepts

German Climate Action Plan 2050 wikipedia , lookup

2009 United Nations Climate Change Conference wikipedia , lookup

Mitigation of global warming in Australia wikipedia , lookup

Heaven and Earth (book) wikipedia , lookup

Climatic Research Unit email controversy wikipedia , lookup

Michael E. Mann wikipedia , lookup

Climate change in the Arctic wikipedia , lookup

Hotspot Ecosystem Research and Man's Impact On European Seas wikipedia , lookup

ExxonMobil climate change controversy wikipedia , lookup

Soon and Baliunas controversy wikipedia , lookup

Climate resilience wikipedia , lookup

Fred Singer wikipedia , lookup

Global warming controversy wikipedia , lookup

Climate change denial wikipedia , lookup

Climate engineering wikipedia , lookup

Global warming hiatus wikipedia , lookup

Climate governance wikipedia , lookup

Citizens' Climate Lobby wikipedia , lookup

Climatic Research Unit documents wikipedia , lookup

Politics of global warming wikipedia , lookup

Climate sensitivity wikipedia , lookup

General circulation model wikipedia , lookup

Carbon Pollution Reduction Scheme wikipedia , lookup

Economics of global warming wikipedia , lookup

Global warming wikipedia , lookup

Solar radiation management wikipedia , lookup

Climate change adaptation wikipedia , lookup

Media coverage of global warming wikipedia , lookup

Effects of global warming on human health wikipedia , lookup

Attribution of recent climate change wikipedia , lookup

Scientific opinion on climate change wikipedia , lookup

Instrumental temperature record wikipedia , lookup

Climate change and agriculture wikipedia , lookup

Global Energy and Water Cycle Experiment wikipedia , lookup

Climate change in Saskatchewan wikipedia , lookup

Climate change in Tuvalu wikipedia , lookup

Climate change feedback wikipedia , lookup

Public opinion on global warming wikipedia , lookup

Climate change in the United States wikipedia , lookup

Surveys of scientists' views on climate change wikipedia , lookup

Physical impacts of climate change wikipedia , lookup

Climate change and poverty wikipedia , lookup

Effects of global warming on humans wikipedia , lookup

Climate change, industry and society wikipedia , lookup

IPCC Fourth Assessment Report wikipedia , lookup

Transcript
EEA-JRC-WHO Report No X/2008
Impacts of Europe’s changing climate
2008 indicator based assessment
Contents
Acknowledgements
10
Summary
11
1 Introduction
18
1.1 Background and policy framework
18
1.2 Purpose and scope of this report
19
1.3 Outline
20
2 The Earth, its Climate and Man
22
3 Observed impacts: a cascade of effects with feedbacks
27
4 Climate change impacts: what the future has in store
31
4.1 Risks of climate change and the EU’s long-term goal
31
4.2 Climate change risks: probing the future
36
4.3 Can damage be avoided?
39
4.4 What is needed to meet the EU objective?
41
5 An Indicator-based assessment
43
5.1 Introduction
43
5.2 Atmosphere and climate
44
5.2.1 Introduction
5.2.2 Global and European temperature
5.2.3 European precipitation
5.2.4 Temperature extremes in Europe
5.2.5 Precipitation extremes in Europe
5.2.6 Storms and storm surges in Europe
5.2.7 Air pollution by ozone
44
46
49
51
54
58
61
5.3 Cryosphere
64
5.3.1 Introduction
5.3.2 Glaciers
5.3.3 Snow cover
5.3.4 Greenland ice sheet
5.3.5 Arctic sea ice
5.3.6 Mountain permafrost
64
66
69
72
75
79
81
5.4 Marine systems
5.4.1 Introduction
5.4.2 Sea level rise
5.4.3 Sea surface temperature
5.4.4 Marine Phenology
5.4.5 Northward movement of marine species
2
81
83
87
90
93
5.5 Terrestrial ecosystems, biodiversity
97
5.5.1 Introduction
5.5.2 Distribution of plant species
5.5.3 Plant phenology
5.5.4 Distribution of animal species
5.5.5 Animal phenology
5.5.6 Impacts on ecosystem functioning
97
98
103
106
110
112
5.6 Agriculture and forestry
115
5.6.1 Introduction
5.6.2 Crop yield variability
5.6.3 Timing of the cycle of agricultural crops (agrophenology)
5.6.4 Irrigation demand
5.6.5 Forest growth
5.6.6 Forest fire danger
5.6.7 Soil organic carbon
5.5.8 Growing season
5.7 Water quantity, droughts, floods
115
116
118
120
122
124
126
128
130
5.7.1 Introduction
5.7.2 River flow
5.7.3 River floods
5.7.4 River flow drought
130
131
134
137
5.8 Water quality and fresh water ecology
141
5.8.1 Introduction
5.8.2 Water temperature
5.8.3 Lake and river ice cover
5.8.4 Freshwater ecology
141
143
146
148
151
5.9 Human health
5.9.1 Introduction
5.9.2 Heat and health
5.9.3 Vector borne diseases
5.9.4 Water and food borne diseases
151
152
155
158
6 Economic consequences of climate change
161
6.1 Introduction
161
6.2 Direct loses from weather disasters
163
6.3 Normalised losses from river flood disasters
167
6.4 Energy
171
6.5 Coastal areas
177
6.6 Agriculture and forestry
179
6.7 Tourism and recreation
182
6.8 Health
185
6.9 Public water supply and drinking water management
187
6.10 Nature and Biodiversity
189
6.11 The Costs of climate change for society
190
3
7 Adaptation
192
7.1 Adaptation is required even if global greenhouse gas
concentrations are stabilized
193
7.2 Also Europe is vulnerable and will have to adapt
193
7.3 Different vulnerable systems at different geographic levels require
different approaches
195
7.4 From European and national plans to regional and local
implementation
196
8 Uncertainties, data availability, gaps and future needs 197
8.1 Sources of uncertainty
197
8.2 Uncertainties and data gaps in indicators
199
8.3 Gap filling and further needs
203
References
206
4
List of maps and graphs
Map S.1
Biogeographical regions in Europe
Figure 2.1
Scheme of the reflection of sunslight by aerosols and cluds and by
the surface of the Earth
Figure 2.2
Scheme of the greenhouse effect of the atmosphere.
Figure 2.3
Scheme of the radiation balance of the Earth and Atmosphere
system
Figure 2.4
Scheme of the carbon cycle
Figure 2.5
Antarctic temperature change and atmospheric CO2 concentration
over the last 420,000 years.
Figure 2.6
Several reconstructions of the Northern Hemispheric mean
temperature over the last 1000 years
Figure 2.7
Observed changes in global average surface temperature, global
average sea level and Northern Hemispheric snow cover from 1850
Map 3.1
Surface air temperature in Europe during the period 1970 - 2004
Figure 3.2
Selected relationships between some of the impacts discussed in
this report
Figure 4.1
Locations of significant changes in data series the Earth physical
and biological systems between 1970 and 2004
Figure 4.2
Examples of global impacts in various sectors associated with
different levels of climate change.
Figure 4.3
Examples of global impacts in various world regions associated with
different levels of climate change.
Figure 4.4
The probabilistic transient temperature implications for the
stabilization pathways at 450, 550, 650 ppm CO2-eq. concentration
and the pathways that peak at 510 ppm, 550 ppm, 650 ppm.
Figure 5.2.1.1
Mean winter (December–March) NAO index for period 1864 - 2002
Figure 5.2.2.1
Observed global and European annual average temperature
deviations, 1850-2007, relative to the 1850-1899 average
Map 5.2.2.2
Observed temperature change over Europe during last 50 years.
Map 5.2.2.3
Modelled change in mean daily temperature over Europe from
period 1961-1990 to 2071-2100 and seasonal changes for summer
and winter seasons
Map 5.2.3.1
Trend of annual precipitation amounts from 1946 to 2006
Map 5.2.3.2
Modelled precipitation amount change over Europe from period
1961-1990 to 2071-2100 and seasonal changes for summer and
winter seasons
Map 5.2.4.1
Changes in duration of warm spells in summer and frequency of
frost days in the period 1976-2006
Map 5.2.4.2
Modelled number of tropical nights over Europe from control period
(1961-1990) and scenario period (2071-2100) during summer and
change between periods.
Map 5.2.4.3
Summer 2003 temperature anomaly with respect to 1961–1990
period
5
Map 5.2.5.1
Trends in precipitation fraction due to very wet days between 1976
and 2006
Figure 5.2.5.2
Course (1860–2100) of European land average of maximum 5-day
precipitation sum in Europe
Figure 5.2.5.3
Course (1860–2100) of European land average of maximum number
of consecutive dry days
Figure 5.2.6.1
Storm index course in the period 1881 - 2004 for north-western,
north and central Europe
Map 5.2.6.2
Relative change in the annual maximum daily mean wind speed
between the +2°C scenario for 2050 and the reference climate
(1961-2000)
Map 5.2.6.3
Change in the height of an extreme water level event (with
probability of occurrence 2 times in 100 years)
Map 5.2.7.1
Linear trends of surface ozone concentrations at 90% significance
level
Figure 5.2.7.2
Change in number of ozone exceedance days per year from period
1990-1994 to the period 1999-2004
Figure 5.3.2.1
Cumulative ice loss of glaciers from all European regions with
glaciers
Figure 5.3.2.2
Shrinking of the Vernagt-ferner glacier, Austria
Figure 5.3.2.3
Modelled remains of the glacierisation in the European Alps
according to an increase in summer air temperature of +1 to +5 °C.
Map 5.3.3.1
Differences in the distribution of March-April average snow-cover in
Europe between periods1967-1987 and 1988-2004.
Figure 5.3.3.2
Northern Hemisphere snow-cover extent departures from monthly
means between 1966 and 2005
Map 5.3.3.3
Mean number of days with snow cover for the period 1961-1990 and
projected changes for the period 2071-2100.
Figure 5.3.4.1
Changes in glaciation of the ice mass changes in Greenland in the
period 1992 - 2006
Figure 5.3.4.2
Area of Greenland that undergoes melting in the period 1979 - 2007
Figure 5.3.5.1
Time series of arctic sea-ice extent (March, September) for the
period 1979 - 2006.
Map 5.3.5.2
Sea ice coverage on 16 Sept 2007
Figure 5.3.5.3
Arctic September sea ice extent between 1901 and 2100)
Figure 5.3.6.1
Temperature in a mountain range containing permafrost
Figure 5.3.6.2
Temperatures measured in different boreholes between 1987 and
2007
Figure 5.3.6.3
The Matterhorn
Figure 5.3.6.3
The Rock-Glacier Murtel-Corvatsch
Map 5.4.2.1
Sea level rise at different European tide gauge stations from 1896 to
2004
Figure 5.4.2.2
Trends in global sea level from 1870 to 2006
Map 5.4.2.3
Sea level variation trends in Europe from 1992 to 2007
Figure 5.4.2.4
Projected global average sea level rise from 1990 to 2100
6
Figure 5.4.3.1
Annual average SST difference from the 1982-2006 average in
different European seas.
Map 5.4.3.2
Spatial distribution of the SST linear trend of the last 25 years (19822006) for the European Seas.
Figure 5.4.4.1
Decapod abundance in the central North Sea highlighting the mean
seasonal peak in abundance for the period 1950-2005 and the
month of seasonal peak of decapod larvae for each year 1958-2005
Figure 5.4.4.2
Change in color index in southern North Sea from the 1950'ies until
2000s
Map 5.4.5.1
Northward movement of zooplankton between two time periods
(1958-1981) and (1982-1999).
Map 5.4.5.2
Recordings of tropical fish and warm-water copepods in different
years of the period 1963 - 1996.
Figure 5.4.5.3
Relative abundance of warm-water to cold-water flatfish species
depending on mean annual SST.
Figure 5.5.2.1
Change in species richness on Swiss Alpine mountain summits in
the last 100 years.
Map 5.5.2.2
Expansion of the climatically-limited species Holly (Ilex aquifolium)
between periods 1931 - 1960 and 1981 - 2000.
Map 5.5.2.3
Projected changes in plant species in Europe in 2050
Figure 5.5.2.4
Total number of species projected to have potentially suitable
climate space within each European region from present to 2080
Figure 5.5.3.1
Phenological sensitivity to temperature changes
Figure 5.5.3.2
Oak (Quercus sp) leafing date in Surey (UK) from 1950 to 2005
Figure 5.5.4.1
Latitudinal shifts in northern range margins in Britain for 16
taxonomic groups over the past 40 years
Map 5.5.4.2
Recorded occurrence of the comma butterfly (Polygonia c-album) in
the Netherlands from 1975 to 2000.
Map 5.5.4.3
Potential changes in climate space of reptiles and amphibians in
2050
Figure 5.5.5.1
Changes in egg laying dates of the flycatcher across Europe during
the period 1990-2002
Map 5.5.6.1
Current range of the butterfly Titiania fritillary (Boloria titania) and its
host plant American bistorpt (Polygonum bistorta) in Europe.
Map 5.5.6.2
Projected niche spaces of the butterfly Titiania fritillary (Boloria
titania) and its host plant American bistorpt (Polygonum bistorta) for
time horizon 2080
Map 5.6.2.1
Modelled suitability change for grain maize cultivation in the past
(1961–1990) and in the future (2071–2100)
Figure 5.6.2.2
Yield variation due to temperature increase for maize and wheat
Map 5.6.3.1
Rate of advance in the yearly date of flowering of winter wheat for
the period 1975-2007.
Figure 5.6.3.2
Evolution of potential alcohol levels at harvest for Riesling in Alsace
(F) from 1971 to 2003.
Map 5.6.4.1
Variation in the annual meteorological water balance between April
and October in the period 1975 - 2007.
7
Figure 5.6.4.2
Climatic water balance in various parts of Europe in the period 1975
- 2006
Map 5.6.5.1
Modelled change of habitat suitability of the 10 most dominant
European Forest Categories, for current (year 2000) and future (year
2100)
Map 5.6.6.1
Trends of fire danger level from 1958 to 2006 using the Seasonal
Severity Rating.
Map 5.6.6.2
Projected (2071-2100) and control (1961-1990) three-monthly fire
danger levels in Europe.
Map 5.6.7.1
Changes in soil organic carbon contents across England and Wales
between 1978 and 2003.
Map 5.6.7.2
Projected changes of organic carbon in the EU’s agricultural soil for
time horizon 2080.
Map 5.6.8.1
Variation of growing season length in the period 1975 - 2007
Figure 5.6.8.2
Duration of the frost-free period for various regions of Europe in the
period 1975 - 2006
Map 5.7.2.1
Change in annual river flow for the period 1971-98 relative to 19001970.
Map 5.7.2.2
Relative change in mean annual and seasonal stream low between
scenario (2071-2100) and control period (1961-1990).
Figure 5.7.2.3
Change in daily average river flow between 2071-2100 and 19611990.
Map 5.7.3.1
Recurrence of flood events in Europe for the period 1998-2005.
Map 5.7.3.2
Relative change in 100-year return level of river discharge between
scenario (2071-2100) and control period (1961-1990)
Map 5.7.4.1
Change in the severity of river flow drought (as expressed by annual
maximum deficit volume, AMV) for the period 1962-1990.
Map 5.7.4.2
Change in the severity of river flow drought in France for the period
1960-2000.
Map 5.7.4.3
Relative change in mean annual and summer minimum 7-day river
flow between 2071-2100 and 1961-1990.
Figure 5.8.2.1
Trend in annual water temperature in river Rhine (1909–2006),
Danube (1901–1998) and average water temperature in August in
Lake Saimaa, Finland (1924–2000).
Figure 5.8.2.2
Observed changes in annual average deepwater temperatures in
selected European lakes between 1951 and 2005.
Figure 5.8.3.1
Ice break-up dates from selected European lakes and rivers and the
North Atlantic Oscillation (NAO) index for winter (Dec. - Feb.) in the
period 1835 - 2005.
Figure 5.8.4.1
Northward shift of range margins of British Odonata between 1960–
1970 and 1985–1995 and observed occurrence of 7 types of
southern dragonflies in Belgium, 1980-2005
Map 5.8.4.2
The share of Trichoptera taxa sensitive to climate change in the
European ecoregions.
Map 5.9.2.1
Number of excess death in summer 2003 and for 2070
Figure 5.9.2.2
Factors affecting human thermoregulation and the risk of heat illness
Map 5.9.3.1
Recent (January 2007) and projected distribution of Aedes
albopictus in Europe
8
Map 5.9.4.1
some maps/graphs from Belgrade report?
Map 6.1.1
Summary of Economic Effects across Europe
Figure 6.2.1
Natural disasters in Europe 1980-2007 (number of events).
Figure 6.2.2
Natural disasters in Europe 1980-2007 (percentage distribution).
Figure 6.2.3
Overall and insured losses by weather disasters in Europe in the
period 1980-2007
Figure 6.3.1
EU flood losses per thousand of GDP between 1970 and 2005
Figure 6.3.2
Casualties caused by flood disasters in the EU between 1970 and
2005.
Map 6.3.3
Projected change in 100-year flood damage by 2100 time horizon
Figure 6.4.1
Heating Degree Days (HDD) in Europe in the period 1980 - 2005.
Map 6.4.2
Projections of energy demand for various time horizons in Europe
Figure 6.4.3
Projected changes in hydropower production in Scandinavia for
2070 - 2100
Map 6.4.4
Projected changes in hydropower production potential in Europe for
2070s
Figure 6.4.5
Number of days per year when the temperature of the water in the
Rhine was higher than 23°C during the period 1909–2003
Map 6.5.1
People actually flooded in coastal areas across Europe and for time
horizon 2080s
Map 6.6.1
Simulated crop yield changes by 2080s relative to the period 19611990
Map 6.7.1
Simulated conditions for summer tourism in Europe for 1961-1990
and 2071-2100
Figure 6.11.1
The coverage of marginal economic costs of climate change against
the risk matrix.
Figure 7.1
Emphasis on different types of adaptation polices in different
European regions
9
Acknowledgements
This report was prepared by the European Environment Agency the Commission’s
Joint Research centre (JRC-IES) and the Word Health Organization (European centre)
with close cooperation with EEA's European Topic Centres.
Note: List of authors, contributors and acknowledgments for the support of those who
contributed data, maps, graphs and comments will be further elaborated.
10
Summary
Introduction
Background and objective
This report is an update and extension of the 2004 EEA report “Impacts of Europe’s
changing climate”. Since 2004, much progress has been made in monitoring and
assessing the impacts of climate change in Europe. The objectives of this report are to
present this new information on past and projected climate change and its impacts
through indicators; identify sectors and regions most vulnerable to climate change
with a need for adaptation; and to highlight the need for enhanced monitoring and
reducing uncertainties in climate and impact modelling. To reflect the widening of the
coverage of the report and make use of the best available expertise, the report has
been developed jointly between EEA, JRC and WHO.
Global developments science and policy
The Intergovernmental Panel on Climate Change (IPCC) in its 4th Assessment report
reconfirmed and strengthened earlier scientific findings about key aspects of the
climate problem. Increased monitoring and research efforts have enhanced the
understanding of climate change impacts and vulnerability. At the 2007 Bali climate
summit, the urgency of responding effectively to climate change through both
adaptation and mitigation activities was recognized by a greater number of countries
than ever before. By end of 2009, a post-Kyoto regime is expected to be agreed that
would include both adaptation and mitigation. The implementation of the Nairobi
work programme on impacts, vulnerability and adaptation to climate change,
developed to help countries improve their understanding of climate change impacts,
has picked up speed.
European developments science and policy
European research on impacts and vulnerability in the context of national programmes
and the European 5th and 6th Framework Programmes has advanced considerably,
making a major contribution to international assessments such as those of the IPCC,
the Arctic Impact Assessment, and the UNEP Global Outlook for Ice and Snow. New
research programmes focusing on adaptation are currently being developed in many
member countries and in the context of the 7th Framework Programme. On the policy
side, in 2007 the European Commission published its Green Paper on adaptation, to
be followed by a White Paper by the end of 2008 with more concrete proposals for
action. National adaptation strategies have been or are being developed in many
member countries, usually on the basis of impact and vulnerability assessments.
This report
The main part of this report summarizes the relevance, past trends and future
projections for 36 indicators (from 22 in the 2004 report). The indicators address
atmosphere and climate; the cryosphere; marine systems; terrestrial systems and
biodiversity; agriculture and forestry; water quantity, foods and droughts; water
quantity and fresh water ecology; and human health. After a brief introduction of the
report, several chapters deal in a general way with the changes in the climate system,
the observed impacts and the projected impacts, respectively. The report concludes
11
with chapters on the economics of climate change impacts and adaptation; adaptation
strategies and policies; and data availability and uncertainty.
Key messages
Atmosphere and climate.
Recent observations confirm that the global mean temperature has increased (by
0.76oC as compared to pore-industrial times for land and oceans, by 0.98oC just for
land). Europe has warmed more than this (0.98 and 1.22oC, respectively), especially
in the south-west, the north-east and mountain areas. Projections suggest temperature
increases in Europe between 1-5.5oC by the end of this century. Whether the EU’s
2oC goal will be exceeded will depend on the effectiveness of international climate
policy. Annual precipitation changes already exacerbate differences between a wet
north (+10-40% since 1900) and a dry south (-20 %), a development that is projected
to continue. The past tendency of an increasing number of hot extremes and a
decreasing number of cold extremes is projected to continue. The same applies to the
tendency of increased intensity of precipitation and droughts. No clear trend in the
frequency and intensity of storms has yet been observed in Europe, but for the future
heavier storms are projected, albeit with slightly lower frequency. Uncertainties for
projected precipitation and extreme events continue to be larger than those for
temperature.
Cryosphere.
Much scientific and political attention has recently been given to the changes in the
cryosphere, the frozen world. European glaciers are rapidly melting: the glaciers in
the Alps lost 2/3 of their volume since 1850, accelerating since 1985, and projected to
continue their decline. Snow cover is decreasing with the greatest losses in spring (2
weeks earlier melt than 30 years ago) and autumn. Sea ice extent has fallen
considerably and may even disappear at the height of the melt season in the upcoming
decades, creating a feedback that will further increase climate change through the
albedo effect. Also mountain permafrost is reducing due to increasing temperatures.
New findings cast doubts on earlier conservative estimates of the stability of the
Greenland ice sheet with potentially very large consequences for sea level rise on the
longer term, but uncertainties remain large.
Marine systems.
According to satellite observations, the pace of global mean sea level rise has
increased to more than 3 mm/year in the last decade (as compared to a global average
in the 20st century of 1.7 mm/year). Because of ocean circulation and gravity effects
sea level is different across different seas. Also an acceleration of sea surface
temperature increases has been observed in recent decades. For the future, projections
suggest European sea level and sea surface temperature to rise more than the global
average. IPCC (2007) sea level rise estimates are possibly conservative because of the
abovementioned risks of more rapid changes in the Greenland ice sheet than assessed
so far. Changes in the phenology (periodic biological phenomena) and distribution of
marine species have been observed, such as earlier seasonal cycles and northward
12
movements. Because of the complex factors that influence behaviour, attribution of
such changes to climate change remains difficult.
Terrestrial ecosystems
Also for unmanaged terrestrial systems, changes in phenology and distribution of
plant and animal species have been observed in Europe, most of which are consistent
with observed climatic changes. This is particularly true for the vulnerable mountain
regions in Europe. Projections suggest northward shifts of some species by hundreds
of kilometers and in mountain areas hundreds of meters upward. Ecosystems are
linked in all kinds of complex ways: changing one component affects the behaviour of
other components and of the system as a whole. Landscape fragmentation hinders
migration and adaptation in unmanaged systems even more than in managed systems,
and extinctions of 5-35 % of plant species may occur in future.
Agriculture and forestry
In both agriculture and forestry, climate change affects average yields in parallel with
land-use and management changes, making it difficult to identify a clear climate
signal. The climate change signal is more evident for extreme conditions, including
prolonged droughts or heat waves, such as the one in 2003 that decreased yields
across Europe. For the future, more variable yields are projected. In the short-term in
some parts of Europe, benefits from carbon fertilization and growing season extension
may compensate for potentially negative climate change impacts, but as climate
continues to change, eventually adverse effects are projected to dominate. Climate
change is projected to increase forest fires risks in major parts of Europe. In
agriculture and forestry adaptation is possible, however with some limits. Some of the
adaptation options such as irrigation have trade-offs for mitigation, because of
increased energy consumption.
Water quantity: floods and droughts.
Global warming intensifies the hydrological cycle. This has implications for river
flows. In general, annual river flows have been observed to increase in the north and
decrease in the south, a difference projected to be exacerbated in the future. Strong
changes in seasonality are projected, with lower flows in the summer and larger flows
in the winter. As a consequence, water stress will be exacerbated particularly in the
south and in the summer. Particularly in winter and spring, more floods are projected.
Climate change is one of several factors affecting river flows (e.g., land use), so
attribution of changes in river flows remains difficult. In the past, the recorded
number of floods was influenced heavily by the improving monitoring and reporting
systems.
Water quality and freshwater ecology
Increased temperatures of lakes and rivers (by 1-3oC), have resulted in decrease of ice
cover by (on average 12 days in the last century in Europe). These changes can be at
least partly attributed to climate change, partly to other causes such as cooling and
sewage water discharges. Warming of surface water can have several effects on water
quality and hence on human use and aquatic ecosystems (e.g., reduced oxygen
13
content, changes in stratification, increased pollution load). Further monitoring is
needed to be able to analyze these changes.
Health
Increased temperatures can have various effects on human health. The large number
of additional deaths during the 2003 heat wave has highlighted the need for adaptation
actions, such as heat wave warning systems. Such heat waves are projected to become
much more common later in the century as the climate continue to change. There is
some evidence that the number of victims from extreme or prolonged cold has
decreased considerably in Europe and will continue to do so. Other health effects can
be vector-, water- and foodborne diseases. Changes in bird migration routes and insect
distribution increases the risks of related vector-borne diseases. A warmer climate can
also enhance diseases related to water and food. The risks are not only related to
climate change, but very much depend on human behaviour and the quality of health
care services and its ability to adapt to climate change.
Economic impacts
Economic costs of climate-related changes have increased in Europe and are projected
to increase further. The costs of weather-related disasters such as storms and floods
are known best, because of the insurance risks involved. Even though social change
and economic development are mainly responsible for increasing losses, there is
evidence that changing patterns of weather disasters are drivers too. However, it is
still not possible to determine the proportion of increase in damages that might be
attributed to climate change. Some studies project increased frequency and intensity
of extreme events, which would further increase cost. Economic costs and benefits of
other climate impacts, such as those on coastal safety, agriculture, forestry, fisheries,
energy supply, tourism and health have been quantified in some studies but factors
other than climate change usually have a dominant effect on costs, and the costs of
adaptation are poorly quantified.
Challenges
Improved monitoring and reporting
Over the last decades, the availability of data on climate change impacts has improved
across Europe, but for many indicators the data result from a limited amount of local
or regional projects and national or EU-wide research projects. No regular Europewide monitoring programmes exist. Some data are dependent on voluntary work of
non-governmental organizations. There is more robust information available for
observed and projected rate of climate change and impacts which are in many cases
very different for the European regions (Fig. S.1 and table S.1). To tackle these
changes and to develop adequate adaptation strategies requires more detailed
information. Through coordinated efforts from countries and Commission, monitoring
systems can be improved in a way that is consistent with SEIS. GMES projects could
fill key data and information gaps and also the INSPIRE directive may help in this
respect. It would be useful to have European agreement on the definition of key
14
climate change indicators, including on extreme weather events (for example “floods”
and “droughts”).
Improved attribution methods
Even if many of the observed changes in various natural and societal systems are
consistent with observed climatic changes, other factors also influence system
behaviour. Disentangling the climate signal from other signals for different indicators
remains a challenge. There is a need to improve in this area, in order to have better
projections of impacts and be able to develop more focused adaptation actions.
Improved understanding of socio-economic and institutional aspects of vulnerability
and adaptation
Much of the research and assessment activities to data have focused on the
climatological, physical and biological aspects of climate change impacts. A better
understanding of the socio-economic and institutional aspects of vulnerability and
adaptation, including costs and benefits, is urgently needed.
Improved and coordinated scenario analysis of impacts and vulnerability.
Scenarios for the climate change impacts and vulnerability indicators presented in this
report are as yet incomplete and different between indicators. Regular interaction is
needed between the climate modeling community and the user community analyzing
impacts, vulnerability and adaptation to develop high-resolution, tailor-made climate
change scenarios for the regional and local level. It would be useful if European
research projects would adopt the same contrasting set of climate scenarios for global
development, such as those used by IPCC. Both explorative research for the very long
term (centuries) would be needed as well as analysis of climate change impacts on the
medium term (decades) for which there is a high need to be able to better develop
adaptation actions.
Good practices in adaptation measures and their costs.
Only in the last few years, countries have started to prepare for climate change by
developing and implementing adaptation strategies. Information on the available
options at the national and especially sub-national level could be exchanged between
EU member states on a regular basis. The understanding how adaptation can be
integrated in other policy areas at the European and national level should be further
improved (e.g., environmental policies like water management and biodiversity
protection but also other like agriculture). Good practices can be developed from
“resilience bottom-up approaches’” in addition to “top down scenario approaches”.
Develop information exchange mechanisms.
Both at the national and European level planned research programmes will result in a
rapidly increasing amount of data and information on climate change impacts,
vulnerability and adaptation. A European Clearing house on climate change impacts,
vulnerability and adaptation can make this information widely available to potential
users across Europe. The information can include data on observed and projected
15
climatic changes, information on vulnerable systems, indicators, tools for impacts
assessments, and good practice adaptation measures.
Table S.1 : Observed (obs) and projected (scen) trends in climate and impacts for
Northern (arctic and boreal),temperate (Atlantic , Central, Eastern) and
Mediterranean (med.) region of Europe
Indicator
Northern
obs / scen
Atlantic
obs / scen
Central
obs / scen
East
obs / scen
Med.
obs / scen
5.2 Atmosphere and climate
5.2.2 Global and European temperature average
winter
summer
5.2.3 European precipitation
average
winter
summer
5.2.4 Heat waves in Europe
Number of days with frost?
5.2.5 Precipitation extremes in Europe
5.2.6 Storms and storm surges in Europe
5.2.7 Air pollution by ozone
↑/↑
↑/↑
↑/↑
↑/↑
↑/↑
?/↑
?/↑
↓/↓
↑/↑
o/↑
o/↑
↑/↑
↑/↑
↑/↑
↑/o
↑/↑
?/↓
?/↑
↓/↓
↑/↑
o / ↑?
o/↑
↑/↑
↑/↑
↑/↑
o/o
↓/↑
?/↓
?/↑
↓/↓
↑/↑
o/o
↑/↑
↑/↑
↑/↑
↑/↑
o/o
↓/↑
?/↓
?/↑
↓/↓
↑/↑
o/↑
o/↑
↑/↑
↑/↑
↑/↑
↓/↓
↓/o
?/↓
?/↑
↓/↓
↑/↑
o/↓
↑/o
5.3 Cryosphere
5.3.2 Mountain glaciers
5.3.3 Snow cover
5.3.4 Greenland ice sheet
5.3.5 Arctic sea ice
5.3.6 Mountain permafrost
↓/↓
↓/↓
↓/↓
↓/↓
↓/↓
n.a. / n.a.
↓/↓
n.a. / n.a.
n.a. / n.a.
↓/↓
↓/↓
o/↓
n.a. / n.a.
n.a. / n.a.
↓/↓
↓/↓
↓/↓
n.a. / n.a.
n.a. / n.a.
↓/↓
↓/↓
↑/↓
n.a. / n.a.
n.a. / n.a.
↓/↓
5.4 Marine systems
5.4.2 Sea level
5.4.3 Sea surface temperature
5.4.4 Phytoplankton biomass and growing season
5.4.5 Marine phenology
5.4.6 Marine northward movement
↑/↑
↑/↑
↑ / n.a.
↑ / ↑?
↑ / n.a.
↑/↑
↑/↑
↑ / n.a.
↑ / ↑?
↑ / n.a.
↑/↑
↑/↑
n.a. / n.a.
n.a. / n.a.
n.a. / n.a.
↑/↑
↑/↑
n.a. / n.a.
n.a. / n.a.
n.a. / n.a.
↑/↑
↑/↑
n.a. / n.a.
n.a. / n.a.
n.a. / n.a.
5.5 Terrestrial ecosystems, biodiversity
5.5.2 North- / upward shift of plant species
5.5.3 Plant phenology*
5.5.4 North- / upward shift of animal species
5.5.5 Animal phenology
5.5.6 Impacts on communities
n.a. / ↑
↑/?
↑/↑
↑ / n.a.
↓/↓
n.a. / ↑
↑/?
↑/↑
↑ / n.a.
↓/↓
↑ / ↑(m)
↑/?
↑/↑
↑ / n.a.
↓/↓
n.a. / ↑
↑/?
↑/↑
↑ / n.a.
↓/↓
n.a. / ↑
↑/?
↑/↑
↑ / n.a.
↓/↓
5.6 Agriculture and forestry
5.6.2 Crop yield
5.6.3 Agrophenology
5.6.4 Irrigation demand
5.6.5 Forest growth
5.6.6 Forest fire danger
5.6.7 Soil organic carbon
5.5.8 Growing season
↑/↑
↑/↑
↓/o
↑/↑
↑/↓
↓/↓
↑/↑
↑/↑
↑/↑
↓/↑
↑/↑
↓/↑
↓/↓
↑/↓
↑/↑
↑/↑
o/↑
↑/↑
↑/↑
↓/↓
↑/↓
↑/↑
↑/↑
o/↑
↑/↑
↑/↑
↓/↓
↑/↑
↑/↓
↑/↑
↑/↑
↑/↓
↑/↑
↓/↓
↓/↑
5.7 Water quantity, droughts, floods
5.7.2 River flow
5.7.4 River floods (number of events)
5.7.5 River flow drought
↑/↑
o/↓
o/o
o/↑
↑/↑
o/o
o/↑
↑/↑
o/o
o/↓
↑/↑
o/o
↓/↓
o/↑
o/o
5.8 Water quality and fresh water ecology
5.8.2 Water temperature
↑/↑
↑/↑
↑/↑
?↑ / ↑
?↑ / ↑
16
5.8.3 Lake and river ice coverage
Chemical quality (box)
5.8.5 Freshwater ecology (phenology and
northwards shift)
5.9 Human health
5.9.2 Heat and health
5.9.3 Vector borne diseases (case study)
5.9.4 Water and food borne diseases
↓/↓
↓/↓
↑/↑
↓/↓
↓/↓
↑/↑
↓/↓
↓/↓
↑/↑
n.a. / n.a.
↓/↓
↑/↑
n.a. / n.a.
↓/↓
↑/↑
↑/↑
o/o
n.a. / ↑
↑/↑
o/↑
↑/↑
↑/↑
o/o
↑/↑
↑/↑
o/o
↑/↑
↑/↑
↑/↑
↑/↑
↑/↑
↑ / ↑?
↑/↑
↑ / ↑?
↑/↑
↑ / ↑?
↑/↑
↑ / ↑?
6. Economic sectors
6.2 Direct losses from weather disasters
↑/↑
6.3 Normalised losses from river flood disasters
n.a. / ↑?
↑
= increasing
↓
= decreasing
o
= no significant changes
n.a. = not available
*
= only pan-European average available
m
= mountain regions
?
= to be clarified
Figure S.1: Biogeographical regions in Europe (Source: EEA)
17
Introduction
1.1 Background and policy framework
During the past decades, there have been notable changes in the global and European
climate. Sea level and temperatures are rising, precipitation is changing and weather
extremes show an increasing intensity and frequency in many regions.
The 2007 Fourth Assessment Report from the UN Intergovernmental Panel on
Climate Change (IPCC) concluded that “warming of the climate system is
unequivocal, as is now evident from observations of increases in global average air
and ocean temperatures, widespread melting of snow and ice and rising global
average sea level” (IPCC Synthesis Report, SPM, 2007).
The IPCC concluded further that “most of the observed increase in global average
temperatures since the mid-20th century is very likely due to the observed increase in
anthropogenic GHG concentrations” and “continued GHG emissions at or above
current rates would cause further warming and induce many changes in the global
climate system during the 21st century that would very likely be larger than those
observed during the 20th century (IPCC Synthesis Report, SPM, 2007).
Consequences of climate change include an increased risk of floods and droughts,
losses of biodiversity, threats to human health, and damage to economic sectors such
as energy, forestry, agriculture, and tourism. In some sectors, some new opportunities
might occur, at least for some time, although over a longer time period and with
increasing temperatures effects are likely to be adverse worldwide if no action is
taken to reduce emissions or to adapt to the consequences of climate change.
The United Nations Framework Convention on Climate Change (UNFCCC) came
into force in 1994. The ultimate objective of the UNFCCC is 'to achieve stabilisation
of greenhouse gas concentrations in the atmosphere at a level that would prevent
dangerous anthropogenic interference with the climate system”. To avoid “dangerous
climate change” the EU has proposed a target of a maximum global temperature
increase of 2ºC above the pre-industrial level. This will require global emissions to
stop rising within the next 10 to 15 years and then be reduced to levels below 50% of
1990 levels by 2050. Within UNFCCC an international post-2012 international
agreement is being negotiated, with the aim to reach an agreement at the climate
conference planned in Copenhagen end of 2009.
However, there is a growing awareness that, even if GHG emissions were stabilised
today, increases in temperature and associated impacts will continue for many
decades to come. Even if the EU target is achieved, the global warming already
incurred and embedded in unavoidable economic development will lead to climate
change impacts to which countries worldwide will need to adapt. Within the
UNFCCC and other UN organisations increasing attention is given to climate change
adaptation especially in developing countries, since these, often poor, countries will
suffer the earliest and most damaging effects, even though their greenhouse emissions
are low and thus they have contributed least to the problem (Human Development
18
Report 2007/2008, Fighting climate change: Human solidarity in a divided world,
Nov. 2007).
Across Europe the most vulnerable regions and sectors are different, but in all
countries the need to adapt to climate change has been recognised. The European
Commission’s Green Paper on Adaptation (2007) started the EU adaptation policy
process, while actions already take place at national level. A Commission White
Paper on adaptation will be published by the end of 2008. Integration of climate
change into other EU and national policy areas is already taking place, e.g. the Water
Framework Directive (aimed at improving water quality), the Floods Directive (aimed
at reducing damages from floods) and the European Commission’s Communication
on Water Scarcity and Droughts.
Policy makers and the public need reliable information and a key challenge is to
further develop the scientific understanding of climate change and impacts on a
regional scale so that the best adaptation options possible can be developed and
deployed. Some countries are developing or have finalised national vulnerability
assessments and/or national adaptation plans. However, more vulnerability and
adaptive capacity assessments across key economic sectors and environmental themes
are needed. There is very little quantified information on adaptation costs and further
work is needed to facilitate informed, cost effective and proportionate adaptation in
Europe. There are many EU and national results projects on climate change impacts,
vulnerability and adaptation. However results from such research programmes have
often not been fully shared with policy makers and other stakeholders in a form that
they can understand. There is a need for more projects that can help provide the right
policy guidance and tools and which will help to build effective trans-national and
sub-national networks.
1.2 Purpose and scope of this report
Taken into account the needs from policymakers, this report aims to present an
indicator-based assessment of recent and projected climate changes and their impacts
in Europe. The report is intended for a broad target audience consisting of
policymakers at EU and national and sub-national level and the interested public and
non-governmental organizations (e.g. environmental, businesses). The report is an
update of a previous EEA report on climate change impacts in Europe (2004). It
includes a number of additional indicators while some of the previous indicators have
not been retained.
The European Environment Agency, the Commission’s Joint Research Centre and the
World Health Organisation (European office) have joied forces to prepare this report.
The Agency in its preparation also cooperated closely with several of its European
Topic Centres (ETCs).
The report aims to provide short but comprehensive indicator information covering all
main impact categories, where feasible across Europe (EEA 32 member countries).
However for categories for which no Europe-wide data was available in some cases
indicators have been developed and presented for smaller scales, provided data was
available for at least several countries.
19
The main objectives of the report are, for Europe:
 Present past and projected climate change and its impacts through indicators
(easily understandable, scientifically sound and policy relevant)
 Identify sectors and regions most vulnerable to climate change with a high need
for adaptation
 Increase awareness of the need for global, EU and national action on both
mitigation (to achieve the EU global temperature target) as well as adaptation
 Highlight the need for enhanced monitoring, data collection and dissemination,
and reducing uncertainties in climate and impact modelling
The report presents results of key recent national and EU-wide research activities
(FP5-7 projects) and also builds on the fourth assessment of the IPCC (2007), and
other recent key international assessments, including the Arctic Climate Impact
Assessment (2004, and its 2007 follow-up) and UNEP’s Global Outlook for Ice and
Snow (2007). The report also uses information from national assessments from
various European countries. The main added value compared to these other reports is
the inclusion of the most recent scientific information and the specific focus on
Europe.
All indicators are also available on the web through the EEA web site indicator
management system. This will allow easy regular updating on the web of those
indicators for which regular (possibly annual) new data will become available and for
which trends are changing significantly in a relatively short period of a few years.
1.3 Outline
Chapter 2 of this report sets out the scientific background of climate change, its
causes and its impacts. It also provides an overview of the linkages between the
various indicator categories.
In chapter 3 an introduction and brief overview are presented of observed climate
change in Europe.
Chapter 4 gives an overview of projected climate change and also discusses possible
irreversible climate change with large potentially catastrophic risks. It also describes
some background on climate change scenarios and projected climate change
indicators.
The main part of the report is Chapter 5. The state of climate change and its impacts
in Europe are described by means of about 40 indicators, divided into eight different
categories:
 Atmosphere and climate
 Glaciers, snow and ice (cryosphere)
 Marine and coastal systems
 Terrestrial ecosystems and biodiversity
 Agriculture and forestry
 Water quantity
 Water quality and freshwater ecology
 Human health
20
The indicators present selected and measurable examples of climate change and its
impacts, which already show clear trends in response to climate change. Primarily
only indicators have been selected for which data is available for about 20 years,
although in some cases this period was shorter and explanations are provided why the
indicator was still included. The responses of the selected indicators can be
understood as being representative of the more complex responses of the whole
category. Furthermore, the results can give an indication of where, to what extent and
in which sectors Europe is vulnerable to climate change, now and in the future. Each
indicator is presented in a separate sub-chapter containing a summary of the key
messages, an explanation of the relevance of the indicator for the environment,
society and policy, a short description of main uncertainties and the analysis of past,
recent and future trends.
Chapter 6 addresses the effects of climate change on economic sectors, based on the
available limited knowledge. For almost all sectors no complete European wide
information is available, and therefore for many sectors information is provided from
either a few countries or over a relatively limited time period.
Chapter 7 discusses climate change adaptation strategies and actions and reviews the
current experiences.
Finally, Chapter 8 evaluates causes of uncertainties and discusses data availability and
quality. It also proposes potential indicators which could broaden future climate
impact assessments if monitoring would be performed and data would become
available.
21
2. The Earth, its Climate and Man
Climate change has hit the headlines of newspapers around the world. It got the
attention from Nobel Prize and Academy Award committees. We have come to
realize that we are responsible for climate change and that it is very likely to have
significant effects on the way we and people after us are going to live. But what is
climate? How does it work, how do we depend on it and how do we influence it? If
we don’t understand climate, it will be difficult to care for it.
The climate of the Earth is described in terms of the temperature at e.g. the Earths
surface, the strength of the winds and ocean currents, the presence of clouds and
precipitation, to name a few of its most important features. Weather as we experience
it day after day is obviously related to climate. However if we talk about climate we
do not talk about the weather on a given day in a certain place, but rather about the
weather averaged over a large geographical area and over a long time period, e.g. the
Mediterranean climate during the second half of the past century.
Climate exists primarily because the Earth is receiving energy from the sun in the
form of visible radiation: light.
The amount of light that reaches the Earths surface and that is absorbed by it depends
on a number of factors. Obviously it depends on the amount of light that the suns emit
and on the position of the Earth with respect to the Sun. It further depends on the
composition of the Earths atmosphere through which the light must travel before
reaching its surface. Certain atmospheric constituents, like aerosols (i.e. smoke, dust
and haze) and clouds, do prevent sun light to reach the surface by reflecting it back
into space (Fig. 2.1). Finally, very bright surfaces on the Earth, like snow and ice
fields, do also reflect light. The fraction of the incoming light that is eventually
absorbed by the surface will heat up the Earth; it will raise its temperature, will set in
motion the atmosphere, creating winds, clouds and precipitation, and it will also help
in maintaining the currents in the oceans.
It is a basic law of physics, and a fact and experience of every day life, that any object
that has a certain temperature will radiate heat in the form of invisible infra-red
radiation. We can feel the warmth of the object from a distance.
In case of the Earth, that infra-red radiation has again to pass through the atmosphere
before it is lost to space. Gases like water vapour, carbon dioxide, methane and others
absorb infrared radiation, and therefore keep the heat in the system (Fig. 2.2). These
gases are called greenhouse gases, because they act in a way somewhat similar to the
glass of a greenhouse.
The Earth, like the greenhouse, will have a constant “equilibrium” temperature when
the amount of radiation that comes in equals the amount of radiation that goes out
(Fig. 2.3).
The amount of greenhouse gases and aerosols, and the chemical composition of the
atmosphere in general, are controlled by natural processes that cycle specific
22
substances, like water, carbon, nitrogen, and sulfur, between the atmosphere, the
oceans and land. For understanding climate change and how it will develop in the
future it is of importance to understand these cycles, in particular that of carbon,
giving the primary role of carbon dioxide as a greenhouse gas (see Fig. 2.4).
in
Figure 2.1
Incoming sunlight is partly reflected
by aerosols and clouds in the
atmosphere and by the surface of the
Earth.
Figure 2.2
Heat radiating from the Earths surface
in the form of infra-red radiation will
be partly absorbed by greenhouse
gases in the atmosphere.
Figure 2.3
When the incoming radiation equals
the outgoing radiation for several
hundreds of years, the Earth is will
obtain a constant mean temperature
The complex interactions between the cycling of substances, radiation and other
processes lead to the existence of feedback loops, which might either amplify
(positive feedback) or dampen (negative feedback) an initial increase in greenhouse
gas emission, temperature or some other parameter. A simple example of a negative
feed-back is that a warmer climate will favour the growth of vegetation, leading to a
larger uptake by that vegetation of CO2 from the atmosphere, leading to a reduction of
the greenhouse effect and a cooler climate. A simple example of a positive feed-back
is that a warming of the ocean will enhance the transfer of CO2 from the ocean to the
atmosphere, leading to an additional greenhouse effect and further warming. A
warming will also lead to more evaporation of water from the ocean into the
atmosphere, and since water vapour is a greenhouse gas, it will also amplify the initial
warming. This is an important mechanism that will amplify, nearly double, any initial
global warming caused by man.
Looking at the 4.5 billion years history of the Earth, we see that a constant
temperature on Earth has been an exception rather than a rule. The global mean
temperature of the Earth has always been changing, because of changes in all of the
natural factors mentioned above. Climate change has been an important driver in the
evolution of the living species, including man.
23
out
Figure 2.4
The presence of CO2 in the atmosphere is a result of constant production and removal processes. These
processes are part of the carbon cycle, which describes the cycling of carbon through the various
compartments of the Earth System. During the past 10.000 years until about 150 years ago, the rate of
CO2 production in the atmosphere (through natural processes such as respiration by vegetation and
soils, natural fires, respiration from marine vegetation, volcanism …) has been roughly equal to the
rate of CO2 removal (through photosynthesis by terrestrial vegetation and uptake by the oceans), and
therefore the atmospheric CO2 concentration has been constant. Since 150 years this equilibrium is
disturbed by the burning of fossil fuels and man-made forest burning (red arrows). Production is now
larger than removal, and this is leading to a steady increase in the concentration of CO2, an
enhancement of the greenhouse effect and climate change.
The past 420,000 year and more. Bubbles of air trapped at various depths within the
ice sheets of Antarctica and Greenland, keep a memory of the atmospheric
composition and temperature back to almost one million year ago. Figure 2.5 shows a
record back to 420,000 years ago. It show us that that the climate on Earth has been
oscillating roughly every 100,000 years between so-called ice ages, during which the
global mean temperature has been about 5 degrees lower than present, and so-called
inter-glacials, during which the global mean temperature was about equal to that of
today. These transitions are triggered by predictable changes in the position of the
Earths axis with respect to the sun, followed by mechanisms within the Earth system
which are able to amplify the initial changes. The carbon cycle with its positive feedbacks play an important role in this amplification processes.
24
Figure 2.5
Antarctic temperature change and atmospheric carbon dioxide concentration (CO2) over the last
420,000 years., derived from and 3.6 km long ice core drilled in the Antarctic ice sheet. Nowadays,
such measurements go back to 650.000 years. They show that the temperature varied between ice ages
and inter-glacials. The last 10,000 years, i.e. the present inter-glacial (right end of the graph) is very
stable. The CO2 concentration measured today (370 ppm) and that projected in the next 100 years are
completely out of the regular pattern measured during the past 650.000 years. (based on Petit et al.,
Nature, 1999)
The oscillations in global mean temperature between ice ages and inter-glacials do
correspond with changes in the carbon dioxide, methane and nitrous oxide
concentration in the atmosphere, consistent with the greenhouse effect.
The past 10,000 years up till 150 years ago. The Earth is presently in an inter-glacial
period which started about 10,000 years ago. We know from a range of observations,
including ice-cores, tree rings, etc, that the concentrations of greenhouse gases and
aerosols and the global mean temperature have been relatively stable. A balance
between incoming and outgoing radiation was roughly established. Figure 2.6 shows
several reconstructions of the Northern Hemisphere mean temperature of the past
1,300 years showing that it stayed indeed within a range of only 0.5 degrees.
Variability within that range is explained by changes in the output of the Sun,
volcanic eruptions emitting large amounts of dust particles in the atmosphere, and
natural variations in the exchanges of carbon dioxide between atmosphere, oceans and
biosphere. A comparable stable climate period was recently discovered to exist about
600.000 years ago, and shows that periods with a relatively constant temperature have
been rather rare at least in the last million year. It is likely that the recent stable
climate has triggered the development of agriculture and consequently the building of
permanent settlements and civilization.
25
Figure 2.6
Several reconstructions of the Northern Hemispheric mean temperature, based on a range of
observations (e.g. ice-cores, lake sediments,
tree-rings, etc.). Taken together, these
reconstructions show the existence of a Medieval Warm Period and a little ice age in the 17th
centuries. Since the middle of the 18th century a measured temperature record exists (black
curve) which shows the exceptional warming during the past 150 years. (WikipediA, based on
10 peer reviewed reconstructions)
The past 150 years. Figure 6 also shows the exceptional increase in Northern
Hemispheric mean temperature during the past 150 years and in particular during the
past 50 years. Figure 2.7 shows this increase for the global mean temperature in
greater detail, together with the expected and observed sea level rise and change in
snow cover.
It is now widely accepted within the scientific community that these changes are
triggered mainly by man rather than by nature; there are insufficient natural changes
which can explain them. This has lead scientists to define a new geological epoch: the
Anthropocene (Crutzen et al. IGBP). Man has significantly changed the composition
of the atmosphere since the industrial and agricultural revolutions. The burning of
fossil fuels and deforestation, the raising of cattle and the use of synthetic fertilizers
have resulted in the emissions into the atmosphere of both (warming) greenhouse
gases and (cooling) aerosol particles, but with a clear net effect of warming.
The Intergovernmental Panel on Climate Change, in its 4th assessment report, (2007,
IPCC AR4) concludes that: “(there is) a very high confidence that the globally net
effect of human activities since 1750 has been one of warming”
26
Figure 2.7
Observed changes in (a) global average surface temperature, (b) global average sea level
and (c) Northern Hemispheric snow cover for March-April. All changes are relative to the
period 1961-1990. (IPCC 4AR)
27
3. Observed impacts: a cascade of effects with feedbacks
The change in the composition of the atmosphere and the resulting increase of the
global temperature are just two steps in a cascade of impacts caused by human
activities. Some selected aspects of the whole cascade is shown in Figure 3.2.
At the top of the cascade stands the use of fossil fuels for energy production and the
raising of cattle and the use of fertilizers for food production as the main sources of
greenhouse gases and aerosols. The availability of cheap fossil fuels and the mass
production of food has led to a higher quality of life, improved health, better
infrastructures etc. (red arrows to the right). However it has become apparent that the
way we produce energy and food has many collateral effects that now threaten that
very quality of life.
The existence of such a cascade could to some extent be foreseen based on our
knowledge of the underlying physical, chemical and biological processes. The
linkages between the various impacts, i.e. the consistency that Figure 3.2 implies, is
now increasingly confirmed by the observations such as the one presented in the
present report. 119 physical and 28115 biological impacts observed in Europe,
respectively 94% and 89% are consistent with a warming trend according to the IPCC
(see Figure 3.1).
Figure 3.1
Surface air temperature changes in Europe during the period 1970 – 2004, together with the
locations of significant changes in physical and biological systems. Most of these changes are
consistent with the observed warming (IPCC 2007).
28
Figure 3.2 also shows the existence of positive and negative feedback loops, as
already discussed in the previous chapter. Positive feedbacks are worrisome. For
instance, an increase in temperature will at one point in time start the melting of soils
which used to be permanently frozen. e.g. in Siberia or Northern Canada. When this
will happen, methane, which is trapped in these soils, will be released and cause even
more warming. Another example is that the increase in temperature will melt ice and
snow fields. This will reduce the reflectiveness of the Earths surface, increase the
absorption of incoming solar light, which will lead to even more warming. These
positive feedbacks would make control of climate extremely difficult. It is important
to note that some of these impacts, e.g. the melting of the Arctic sea ice, are occurring
earlier than was foreseen 10 years ago. Several related indicators are discussed in
Chapter 5 of this report.
We need to avoid the unmanageable - through the reduction of greenhouse gas
emissions - and manage the avoidable - through adaptation measures (Scientific
Expert Group on Climate Change, 2007). How much we have to do of each, depends
on how much warmer we will allow the Earth to become. A much warmer world will
result from little reductions in greenhouse gas emissions, but will require a lot of
adaptation. How much warmer we allow the world to become is eventually a decision
based on a risk assessment, which much necessarily be based on observations on the
impacts of global warming (see following chapter). More information is now
available about the nature and magnitude of climate change impacts than at the time
of the 3rd IPCC assessment in 2001 and the previous EEA report on climate change
impacts in Europe in 2004. As a consequence, the future risk can be assessed more
systematically for different levels of increased global annual average temperature.
The present report delves deeper in climate change impacts expected in Europe, and
goes further than what has been possible for the 4th IPCC assessment of 2007. E.g. it
clearly shows that climate change impacts are different for different European regions
(such as northern/Artic Europe, western/Atlantic Europe, central and eastern Europe,
southern and south-eastern Europe) and for different sectors (water, health, agriculture
and fisheries, nature and biodiversity, human settlements and infrastructure, etc.)
Human civilization can be seen as a process within the Earth System. Man can make
deliberate decisions about the course of its own future.It is still not clear whether,
overall, these decisions imply a positive or a negative feedback in the climate system.
Successful international negotiations that would avoid a dangerous interference with
the climate system would obviously be a negative, controlling feed-back. In other
words, it would turn the grey arrow to left in Fig. 3.2. blue.
29
Figure 3.2.
Selected relationships between some of the impacts discussed in this report. The availability of cheap
fossil fuels and the mass production of food has directly led to a higher quality of life; improved
health, better infrastructures etc. (red arrows to the right). However it has become apparent that the
very quality of life is now threatened by collateral effects of fossil fuel use (blue arrows arriving in the
lower box).
30
4. Climate change impacts: what the future has in
store
4.1 Risks of climate change and the EU’s long-term
goal
More and earlier impacts of climate change have been observed in Europe and
elsewhere than was foreseen 10 years ago. An arguable advantage of this is that this
knowledge also helps to understand future risks better. Most observed changes in the
world are from the extensive European research programmes, notably biological
studies (Figure 4.1). 89% of the many biological studies are consistent with warming,
whereas 94% of the observed physical changes in Europe are consistent with
warming. More information is now available about the nature and magnitude of
climate change risks than at the time of the 3rd IPCC assessment in 2001 and the
previous EEA report on climate change impacts in Europe in 2004. As a consequence,
the risk can be assessed more systematically for different levels of increased global
annual average temperatures (IPCC, 2007b).
Water, ecosystems, food, coastal areas and health are among the key vulnerable
sectors (see Figure 4.2 for an overview of potential impacts as a function of
temperature). The kind of dominant risks is different in different regions (see Figure
4.3). From a global perspective, the most vulnerable regions are in the developing
world, which has the lowest capacity to adapt. Impacts in those regions are likely to
have spill-over effects for Europe as well, through the interlinkages of economic
systems and through migration. These effects have not been quantified and are not
further discussed in this report.
The dominant risks of climate change are also different for different European regions
(such as northern/Artic Europe, western/Atlantic Europe, central and eastern Europe,
southern and south-eastern Europe) and for different vulnerable sectors (water, health,
agriculture and fisheries, nature and biodiversity, human settlements and
infrastructure, etc.). In the main body of this report, these impacts are presented in
detail, both in terms of observed impacts as well as in terms of future risks, to the
extent possible.
31
Figure 4.1: Locations of significant changes in data series of physical systems (snow, ice and frozen
ground; hydrology; and coastal processes) and biological systems (terrestrial, marine, and freshwater
biological systems), are shown together with surface air temperature changes over the period 19702004. A subset of about 29,000 data series was selected from about 80,000 data series from 577
studies. These met the following criteria: (1) ending in 1990or later; (2) spanning a period of at least 20
years; and (3) showing a significant change in either direction, as assessed in individual studies. These
data series are from about 75 studies (of which about 70 are new since the TAR) and contain about
29,000 data series, of which about 28,000 are from European studies. White areas do not contain
sufficient observational climate data to estimate a temperature trend. The 2 x 2 boxes show the total
number of dataseries with significant changes (top row) and the percentage of those consistent with
warming (bottom row) for (i) continental regions: North America (NAM), Latin America (LA), Europe
(EUR), Africa (AFR), Asia (AS), Australia and New Zealand (ANZ), and Polar Regions (PR) and (ii)
global-scale: Terrestrial (TER), Marine and Freshwater (MFW), and Global (GLO). The numbers of
studies from the seven regional boxes do not add up to the global (GLO) totals because numbers from
regions except Polar do not include the numbers related to Marine and Freshwater (MFW) systems.
Locations of large area marine changes are not shown on the map. Source: IPCC, 2007b
32
Figure 4.2: Examples of global impacts in various sectors associated with different levels of climate
change. Boxes indicate the range of temperature levels to which the impact relates. Arrows indicate
increasing impacts with increasing warming. Adaptation to climate change is not considered in this
overview. The black dashed line indicates the EU objective 2°C temperature change relative to preindustrial. Adopted from the Technical Summary of IPCC (2007b).
33
Figure 4.3: Examples of global impacts in various world regions associated with different levels of
climate change. Boxes indicate the range of temperature levels to which the impact relates. Arrows
indicate increasing impacts with increasing warming. Adaptation to climate change is not considered in
this overview. The black dashed line indicates the EU objective of 2°C temperature change relative to
pre-industrial. Adopted from the Technical Summary of IPCC WGII (2007b).
34
4.2 Climate change risks: probing the future
How are climate risks expected to develop over time? Both worldwide and for
Europe, since 2000 quantitative assessment of climate change impacts, both are
mostly based on the range of Special Report on Scenarios (SRES) scenarios A1FI,
A2, B1 and B2 (see Box 4.1). The SRES scenarios describe different ways in which
the world could develop without explicit climate policy.
Box 4.1 The Emissions Scenarios of the IPCC Special Report on Emissions
Scenarios (SRES)
A1. The A1 storyline and scenario family describes a future world of very rapid
economic growth, global population that peaks in mid-century and declines thereafter,
and the rapid introduction of new and more efficient technologies. Major underlying
themes are convergence among regions, capacity building and increased cultural and
social interactions, with a substantial reduction in regional differences in per capita
income. The A1 scenario family develops into three groups that describe alternative
directions of technological change in the energy system. The three A1 groups are
distinguished by their technological emphasis: fossil intensive (A1FI), non-fossil
energy sources (A1T), or a balance across all sources (A1B) (where balanced is
defined as not relying too heavily on one particular energy source, on the assumption
that similar improvement rates apply to all energy supply and end use technologies).
A2. The A2 storyline and scenario family describes a very heterogeneous world. The
underlying theme is self-reliance and preservation of local identities. Fertility patterns
across regions converge very slowly, which results in continuously increasing
population. Economic development is primarily regionally oriented and per capita
economic growth and technological change more fragmented and slower than other
storylines.
B1. The B1 storyline and scenario family describes a convergent world with the same
global population, that peaks in mid-century and declines thereafter, as in the A1
storyline, but with rapid change in economic structures toward a service and
information economy, with reductions in material intensity and the introduction of
clean and resource-efficient technologies. The emphasis is on global solutions to
economic, social and environmental sustainability, including improved equity, but
without additional climate initiatives.
B2. The B2 storyline and scenario family describes a world in which the emphasis is
on local solutions to economic, social and environmental sustainability. It is a world
with continuously increasing global population, at a rate lower than A2, intermediate
levels of economic development, and less rapid and more diverse technological
change than in the B1 and A1 storylines. While the scenario is also oriented towards
environmental protection and social equity, it focuses on local and regional levels.
An illustrative scenario was chosen for each of the six scenario groups A1B, A1FI,
A1T, A2, B1 and B2. All should be considered equally sound. The SRES scenarios do
not include additional climate initiatives, which means that no scenarios are included
that explicitly assume implementation of the United Nations Framework Convention
on Climate Change or the emissions targets of the Kyoto Protocol.
Source: quoted from IPCC, 2001
35
Many (European) research studies have assessed the impacts of these scenarios for
Europe (see Box 4.2). Results of many of these studies are included in the main body
of this report. Many of the projects are ongoing, but for a number of projects
intermediate results could be tapped for this report. While several impacts in Europe
can be considered potentially positive for lower levels of climate change (agricultural
production in temperate zones, increased water availability in some water basins,
navigation in polar regions), most impacts on the longer term, for higher levels of
climate change, are negative. In other words, there may be some winners on the short
term in some regions, but these are paralleled by losers in other regions (e.g. a shift in
tourism, agriculture). And on the longer term negative impacts dominate in most
regions.
Box 4.2 European climate change impacts and adaptation research
Various research projects of the European Commission focus on climate change
impacts in Europe or on providing the basis for assessing them. Fewer projects as yet
address adaptation issues. The projects aim at contributing to better understand Earth
system functioning, the origin and impacts of climate change and to predict its future
evolution, to guide and support to EU’s international commitments and EU policies,
and to provide a basis for effective mitigation and adaptation measures. Most FP6
projects have not yet been finalized, but some intermediate results could be used for
this report. The following are the most important ones:
*ATEAM (Advanced Terrestrial Ecosystem Analysis and Modelling). ATEAM’s
primary objective was to assess the vulnerability of human sectors relying on
ecosystem services with respect to global change. We consider vulnerability to be a
function of potential impacts and adaptive capacity to global change. Multiple,
internally consistent scenarios of potential impacts and vulnerabilities of the sectors
agriculture, forestry, carbon storage, water, nature conservation and mountain tourism
in the 21st century were mapped for Europe at a regional scale for four time slices
(1990, 2020, 2050, 2080).
* PRUDENCE ('Prediction of regional scenarios and uncertainties for defining
European climate change risks and effects'). Prudence used several regional models
to assess climate change at spatial scales between 30-50 kilometres. The project
developed scenarios for the variability of climate change with level of confidence for
the period 2071-2100, providing a quantitative basis for assessing the risks arising
from changes in regional weather and climate in different parts of Europe. Future
changes in extreme events such as drought, flooding and wind storms were estimated
and a robust estimation of the likelihood and magnitude of such changes provided.
FP5
* ACCELERATES (Assessing Climate Change Effects on Land use and
Ecosystems; from Regional Analysis to The European Scale). This project has
constructed regional and European-wide geo-referenced databases, developed models
that represent biophysical and socio-economic processes of agroecosystems at the
European scale and advanced methodologies at fine spatial and temporal resolutions.
In the project, adaptive responses to climate change of agroecosystems were analysed,
using the integrated models.FP5
36
* ENSEMBLES (Ensemble-based predictions of climate changes and their
impacts). This project aims to develop an ensemble prediction system for climate
change based on the principal state-of-the-art, high resolution, global and regional
Earth System models developed in Europe, validated against quality controlled, high
resolution gridded datasets for Europe. Eventually, the outputs of the ensemble
prediction system are intended to be used for a range of impacts analyses, including
agriculture, health, food security, energy, water resources, insurance and weather risk
management. FP6
* CIRCE (Climate Change and Impacts Research: the Mediterranean
Environment). The project will predict and quantify physical impacts of climate
change and evaluate the consequences of climate change for the society and the
economy of the populations located in the Mediterranean area. Adaptation and
mitigation strategies will be identified in collaboration with regional stakeholders.FP6
* ADAM (Adaptation and Mitigation Strategies: supporting European climate
policy).The project will lead to a better understanding of the synergies, trade-offs and
conflicts that exist between adaptation and mitigation policies at multiple scales.
ADAM will support EU policy development in the next stage of the development of
the Kyoto Protocol, in particular negotiations around a post-2012 global climate
policy regime, and will inform the emergence of new adaptation strategies for
Europe.FP6
* PESETA (Projection of Economic impacts of climate change in Sectors of the
European Union based on boTtom-up Analysis). PESETA aims to make a multisectoral assessment of the impacts of climate change in Europe for the 2011-2040 and
2071-2100 time horizons, focusing on the impacts of climate change on the following
sectors: Coastal systems, Energy demand, Human health, Agriculture, Tourism, and
Floods. The emphasis is on the economic costs of climate change in Europe based on
physical impact assessment and state-of-art high-resolution climate scenarios. FP6
A special kind of risk, which is particularly difficult to deal with from a policy point
of view, are impacts with a low likelihood of occurrence in the coming decades, but
with potentially very large consequences on the longer term (see Box 4.3 for
examples that may be of particular relevance for Europe.). Many impacts build up
slowly. However, there may be so-called “tipping points” beyond which large and fast
changes in the behaviour of natural or societal systems can occur. Some of these nonlinear changes are related to positive feedbacks in the climate system and can
accelerate change.
37
Box 4.3 What are the risks of non-linear climate change?
The risk of large-scale discontinuities has been identified by IPCC as one of five
“reasons for concern” and deserves some special attention, because of their potentially
very large consequences for the world, including Europe. What is non-linear, or
abrupt change? If a system has more than one equilibrium state, transitions to
structurally different states are possible. If and when a so-called “tipping point” would
be crossed, the development of the system is no longer controlled by the time scale of
the forcing, but rather determined by its internal dynamics, which can either be much
faster than the forcing, or significantly slower (IPCC, 2007a). A variety of different
tipping points has been identified. Below we discuss a few with potentially large
consequences for Europe.
The IPCC TAR suggested that a rapid warming of over 3ºC could trigger such largescale discontinuities in the climate system, such as changes in climate variability (e.g.,
ENSO changes), breakdown of the thermohaline circulation (THC, or equivalently,
meridional overturning circulation, MOC), deglaciation of the West-Antarctic
Icesheet (WAIS) and Greenland, climate biosphere-carbon cycle feedbacks, and
enhanced emissions of methane from melting permafrost or hydrates. However, there
was only low to medium confidence in this statement because of the complexity of the
interactions involved over long time scales. New research assessed in the AR4 has
enhanced the confidence in the projected consequences of partial or near complete
deglaciation of Greenland and West Antarctica, with medium confidence that already
as much as 1-2ºC sustained global warming versus present climate (thus 2-3 ºC versus
pre-industrial) is a threshold beyond which there will be a commitment to a large sealevel contribution due to at least partial deglaciation of both ice sheets (IPCC,
2007a,b). What would happen if these ice sheets would practically all disintegrate?
According to IPCC (IPCC, 2007b), sea level would rise on average by 7m and about
5m from Greenland and WAIS, respectively. This would basically alter the world’s
coast lines completely, in the course of the coming centuries. Because of ocean
circulation patterns, density and gravitational factors, the sea level rise will not be
evenly distributed over the globe.
There is less confidence in what may happen with the MOC. A slow down of the
North Atlantic ocean circulation may on the one hand counteract the global warming
trends in Europe, but on the other hand may have unexpected serious consequences
for the behaviour of the world’s climate system and exacerbated impacts elsewhere.
Other examples of possible non-linear effects are the progressive emission of methane
from permafrost melting and destabilization of hydrates, and rapid climate-driven
transitions from one ecosystem type to another (IPCC, 2007b). The understanding of
these processes is yet limited and the chance of major implications in the current
century is generally considered to be low.
38
4.3 Can damage be avoided?
Thus climate impacts have already been observed, and future risks for serious
damages are being mapped out by the scientific community. The question is whether
at least some of these damages can be forestalled. To limit impacts and guide policy
development, the EU has adopted a long-term climate goal of 2oC global mean
temperature increase as compared to pre-industrial levels. This goal will limit risks,
but not avoid all impacts (see Figures 4.1 and 4.2). However, at the same time, the
Figures also show that if temperatures are limited to the EU goal, many increasingly
serious impacts can be avoided. It should be noted that most of the studies underlying
the graphs do not take adaptation explicitly into account, and adaptation could further
reduce risks and economic costs. However, it should also be noted that there are limits
to adaptation, dependent on the type, magnitude and rate of change. Furthermore, the
Figures end by 2100, but climatic changes and their impacts do not stop by that time.
Because of delays in the climate system, emissions now and during the rest of this
century will have persisting effects in centuries to come. These considerations
illustrate, among others, the deeply ethical dimension of the climate problem in terms
of impacts over the current and future generations. In addition to ethical questions,
also economic considerations play a role. The costs of inaction as well as the costs of
action are very uncertain. Limits on quantification and valuation play a key role.
Economic effects of climate change for Europe are uncertain, but potentially very
significant (EEA, 2007). What is considered “dangerous interference with the climate
system” is therefore a political issue rather than a scientific one. The EU goal has
initially been established more than 10 years ago, when it was used to support the
Kyoto Protocol negotiations. Are the arguments of that time still valid today? The
answer is yes, as is further elaborated in Box 4.4.
Box 4.4 Limiting global temperature-rise to 2°C above pre-industrial level
The EU global temperature limit of 2C above the pre-industrial level was first established in
1996 before the Kyoto negotiations, and reaffirmed subsequently by the Environment Council
(2003) and the European Council (2005 and 2007). It was deduced from the evidence
available at the time and from the concern that adaptation rates of ecosystems are limited.
Since 1996, the understanding of the vulnerability to and impacts of climate change has
improved significantly. According to the AR4, some impacts are now projected to be stronger
and occur at lower temperatures than assessed in the TAR. For some cases the increase in
impacts will be relatively smooth. For other impacts, such as heat wave mortality, coral reef
losses and thawing of permafrost, a critical temperature limit, or threshold may be identified.
The expected climate changes and impacts vary regionally, making some thresholds regional
rather than global. Both kind of impacts should be taken into consideration when evaluating
the EU’s climate goal.
In some cases, exceedance of temperature thresholds could trigger climate feedbacks that
strongly accelerate climate change, initiate irreversible changes to the climate system, or
result in sudden and rapid exacerbation of certain impacts, requiring unachievable rates of
adaptation. The temperature changes at which these thresholds would be exceeded are not yet
clearly understood. At a temperature rise above 2C relative to pre-industrial levels, there is
an increase in the risk of a range of severe large scale events, such as shutdown of the
thermohaline circulation, but some thresholds may be passed at global average temperature
changes below 2C, such as the melting of the Greenland Ice sheet, which could be initiated
at a global temperature rise between 1-2ºC and could be irreversible if the temperature rise is
39
sustained for a sufficient length of time. This would mean a global sea-level rise of 7 metres
during the next 1,000 years or longer.
Also in Europe, the magnitude of impacts on human health, the water sector, ecosystems,
coastal zones, mountain regions and the economy is expected to increase as global
temperatures rise. A global mean temperature increase below 2°C above pre-industrial levels
is likely to allow adaptation to climate change for many human systems at moderate
economic, social and environmental costs. The ability of many natural ecosystems to adapt to
rapid climate change is limited and may be exceeded well before before 2°C is reached.
Beyond 2°C major increases in vulnerability, considerable impacts, very costly adaptation
needs, an unacceptably high and increasing risk of large scale irreversible effects, and
substantially increase the uncertainty of impacts are expected.
Further research is needed to better quantify risks in the context of Article 2. However, for the
present the 2C limit remains a reasonable level beyond which the risk of severe impacts
would increase markedly, recognising that it will not avoid all impacts. The original EU
stabilisation target related a 2C goal to stabilisation of atmospheric CO2 concentrations (i.e.
not all Kyoto gases) below 550ppm. More recent work indicates that a 2oC limit is more
likely to be consistent with stabilisation levels below 450ppm CO2eq (including all Kyoto
gases). There is a considerable uncertainty in the relationship between stabilisation levels and
temperature, due to both climate sensitivity and assumptions with respect to the non-CO2
gases as well as aerosols and their reduction potential. Preliminary estimates suggest that to
meet the EU temperature limit global GHG emissions will have to peak within the next 10 to
15 years, followed by substantial global emission reductions to at least 50% below 1990
levels by 2050 with emissions of non-CO2 GHG to be reduced by more than 80%.
To achieve those deep emissions reductions, a broad portfolio of technologies currently
available or expected to be commercialized in the next decades need to be deployed urgently
at high rates and at a large scale. Lock-in of carbon-intensive technologies needs to be
avoided, requiring a large shift in investment patterns. Mitigation costs are likely to rise more
quickly and delay economic growth more for lower stabilisation levels,. Cost estimates are
very uncertain, and many possibly over-estimated as factors like technological learning and
non-CO2 options are not taken into account. Estimates of the benefits of mitigation are still
very uncertain and tend to be underestimated. Allowing or not allowing “overshooting” the
goal temporarily has important implications for costs and feasibility.
Main sources: IPCC AR4 (2007), EU (2007, 2oC note for Portuguese chairmanship)
One way of looking at the rationale of different climate change response strategies is
by comparing mitigation efforts with adaptation actions. Mitigation is aiming
particularly at avoiding the serious impacts associated with continuing, longer-term
changes in the climate system as well as limiting the risks of large-scale
discontinuities in that system. Adaptation is aiming particularly at reducing
unavoidable negative impacts already at the shorter term, reducing the vulnerability to
present climate variability, and catching opportunities provided by climate change.
40
4.4 What is needed to meet the EU objective?
Current knowledge suggests that in order to have a 50 % chance of meeting the EU
climate objective, global GHG concentrations would have to be stabilized at 450 ppm
CO2 eq.(of which about 400 ppm CO2, see also den Elzen and Meinshausen, 2005,
van Vuuren et al., 2006, and Figure 4.4). The specific emissions profiles that are
consistent with the EU climate goal are dependent on assumptions that include the
climate sensitivity and the possible acceptance of a temporary peaking above the
objective. The achievement of a 450 ppm stabilization goal is generally considered to
be extremely ambitious, but feasible from a technical and macro-economic
perspective (IPCC, 2007c). Costs for some sectors and regions could be high, and
political, social and behavioural hurdles would have to be dealt with to allow for a
world-wide, effective mitigation response. The Bali Action Plan (UNFCCC, 2007)
and the “20 20 by 2020” climate package of the European Union (20 % reduction of
GHG emissions and a 20 % renewable energy share in total energy consumption by
2020, see EC, 2008) are encouraging signs that international action is mobilized to
stave off long-term climate change impacts. The need and options to adapt to the
remaining impacts is discussed in Chapter 5.
41
Figure 4.4: The probabilistic transient temperature implications for the stabilization pathways
at 450, 550 and 650 ppm CO2-eq. concentration (upper row) and the pathways that peak at
510 ppm, 550 ppm and 650 ppm (lower row). The FAIR-SiMCaP pathways shown are those
for the B2 baseline scenario based on a climate sensitivity that assumes the 1.5–4.5°C
uncertainty range for climate sensitivity (IPCC TAR), being a 90% confidence interval of a
lognormal distribution. Shown are the median (solid lines) and 90% confidence interval
boundaries (dashed lines), as well as the 1, 10, 33, 66, 90, and 99% percentiles (borders of
shaded areas). The historical temperature record and its uncertainty from 1900 to 2004 is
shown (blue shaded band). Source: van Vuuren et al., 2006.
42
5. An Indicator based assessment
5.1 Introduction
(to be further elaborated)
43
5.2 Atmosphere and climate
5.2.1 Introduction
The climate in Europe shows considerable regional variability. This variability is
related to the position of Europe in the Northern hemisphere and the influence of
neighbouring seas and continents, including the Arctic. Atmospheric circulation is an
important driver of the temporal and regional variances (see Box below).
This section describes the changing climatic and atmospheric conditions, while the
following sections deal with impacts of these changes on physical, ecological and
societal systems in Europe. The indicators included in this section are global and
European temperature, precipitation, temperature extremes and precipitation
extremes, storms and storm surges, and atmospheric ozone concentration. Whereas all
indicators focus on Europe, global temperature has been included because the policy
target of the European Council and confirmed by the European Parliament to limit
global average temperature increase to a maximum of 2 °C above pre-industrial
levels, in order to keep climate change at a manageable level and to reduce the
likelihood of irreversible disruptions. The indicators included in this section represent
different characteristics of the climate system that have different effects on physical
and biological systems as well as human society, which can be independent of each
other or, more often, have combined effects. High temperature and reduced
precipitation, for example, may lead to more intense droughts. Droughts are addressed
in section 5.2.2 (i.e. Precipitation extremes in Europe), including an enumeration of
different definitions of drought that have different consequences of sectors involved.
Temperature and precipitation extremes are both included here, because they have the
largest impacts on society and the environment. Storms are dangerous too. Coastal
areas are more vulnerable, damages caused by storms may be aggravated with storm
surges flooding. The surface ozone concentration can increase due to temperature
increase and chemical reactions of air pollutants emitted by human activities in the
lower atmosphere. Higher ozone concentrations have adverse effect on population and
environment particularly in large cities where global warming is amplified by urban
heat islands.
Data availability and accuracy has been one of the important criteria for the selection
of indicators. In general, the data availability of the climate indicators is good
compared to other indicators, although especially on the regional scale reliable data
can be more scarce. Data of the air temperature are the most reliable climatological
data, whereas precipitation and wind data are more variable. Their projections are
more uncertain than for temperature and dependent on future, still uncertain
circulation patterns. Climate models underestimate circulation changes, which is one
of the reasons of the low accuracy in regional climate projections (Gillet et al., 2003).
44
Box 5.2.1 Atmospheric circulation patterns in Europe
The atmospheric circulation moves air masses with their own typical characteristics
like temperature and humidity over long distances. The prevailing western circulation
at mid latitudes directing the oceanic air masses inland the continent is very important
for the European climate. Stronger western advection brings milder and wetter
weather and stronger winds to most of the Europe, especially in winter. Weaker
western circulation causes generally colder and dryer winters and hotter and dryer
summers. The changes in behaviour of this circulation pattern are one of the main
sources of the European climate variability. The intensity of western circulation in the
European region is expressed by the North Atlantic Oscillation (NAO) index. NAO is
the large-scale fluctuation in atmospheric pressure in the Atlantic ocean between the
high-pressure system near the Azores and the low pressure system near Iceland (Fig.
5.2.1.1).
Figure 5.2.1.1. Mean winter (December–March) NAO index . Positive indicates stronger
western flow. (Updated from Hurrell et al. (2003)).
The NAO oscillation has seasonal, inter-annual and interdecadal variations. The shortterm dynamics have driving mechanism connected with weather fluctuation, in longer
time scales especially with atmosphere-ocean-ice interactions.
The seasonal anomalies have direct human impacts, often being associated with
droughts, floods, heat- and cold-waves and other changes. The NAO behaviour in the
past appears to have been considerably more variable, with more extreme values in
the late 18th and early 19th centuries than in the 20th century. More recently, there has
been a large increase in the NAO index between 1970 and 1990, followed by a
decrease back to about normal. The relationship with anthropogenic climate change is
as yet unclear. Scenarios for future circulation patterns are very uncertain, because of
the limited predictive ability of current models.
In the southern hemisphere, the El Niño-Southern Oscillation (ENSO) in the Pacific
Ocean has global impacts on decadal and longer-term variability and can cause
precipitation and temperature changes over very large distances, including Europe.
Generally, for Europe the effects of ENSO on precipitation and temperature is much
weaker than those of variations in the NAO.
45
5.2.2 Global and European temperature
Key Messages
Global

The global (land and ocean) average temperature increase up to 2007 was
0.76 °C compared to pre-industrial levels. For land only, the observed
increase has been 0.98 oC.
 The rate of global average temperature change has increased from 0.08°C
per decade over last 100 years, to 0.13 °C per decade in last 50 years and
0.20°C in last decade.
 The best estimates for projected global warming this century are a further
rise in the global average temperature from 1.8 to 4.0°C or different
scenarios that assume no further/additional action to limit emissions.
Europe

Europe has warmed up more than the global average. The increase of the
annual average temperature for the European land area and combined land
and ocean areas has been 1.22°C and 0.98°C, respectively, comparing the
trend towards 2007 with pre-industrial times. Eight of the last 12 years from
the period 1996 – 2007 were in the 12 warmest years since 1850 in Europe.
The annual average temperature is projected in Europe to rise this century 15.5°C (best estimate) with the greatest warming over eastern and northern
Europe in winter, and over south-western and Mediterranean Europe in
summer.

2.00
2.00
Annual deviations (land and sea)
10-year moving average (land and sea)
1.50
Annual deviations (land only)
10-year moving average (land only)
10-year moving average (land and sea)
1.75
10-year moving average (land only)
Temperature deviation, compared to 1850-1899 average
( o C)
Temperature deviation, compared to 1850-1899 average (o C)
1.75
1.25
1.00
0.75
0.50
0.25
0.00
-0.25
-0.50
1.50
1.25
1.00
0.75
0.50
0.25
0.00
-0.25
-0.50
-0.75
-1.00
-0.75
-1.25
-1.00
-1.50
2000
1990
1980
1970
1960
1950
1940
1930
1920
1910
1900
1890
1880
1870
1860
1850
2000
1990
1980
1970
1960
1950
1940
1930
1920
1910
1900
1890
1880
1870
1860
1850
year
year
Figure 5.2.2.1 Observed global (left) and European1 (right) annual average temperature
deviations, 1850-2007, relative to the 1850-1899 average (in oC). The lines refer to 10-year
moving average, the bars to the annual 'land only' annual average temperature (source
Haylock et al, 2008, Land only data are based on HadCRUT3 data, land & ocean data based
on CRUTEMP3 data; European data obtained with climexp.knmi.nl)
1
Europe is defined as the area between 35 o to 70o Northern latitude, -25 o to 30 o Eastern longitude,
plus Turkey (=35 o to 40 o North, 30 o to 45 o East)
46
Relevance
Of all parameters used in monitoring and projecting climate change, air temperature is
the closest to the human perception of “climate”. Fortunately, air temperature data are
also the most reliable climate data. There is a sufficiently dense network stations in
the world -including Europe- with standardised measurements and with sophisticated
quality control systems and homogeneity procedures. This allows for high resolution
temperature information with a high level of significance. Daily and monthly data are
available in long time series (standardised data from 1850 onwards). Furthermore,
climate models have become increasingly sophisticated over the past years with the
uncertainty in the longer-term temperature projections being especially related to
future emissions of greenhouse gases rather than to uncertainties in the climate
system.
Temperature changes affect practically all indicators included in this report, directly
or indirectly. The projected further temperature rise may have some positive impacts
in the northern part of Europe (at least for a certain period), but the impacts in most
parts of Europe are and will be adverse. In relation to climate change policy, the
global temperature rise is relevant because of the EU objective of limiting global
average temperature increase to a maximum of 2 °C, as described in section 5.1.
Monitoring temperature change is thus also relevant for comparing actual
developments with this EU’s long-term goal. Some other studies have proposed an
additional ‘sustainable’ target of limiting the rate of temperature change, ranging from
0.1oC to 0.2oC per decade, based on the limited capability of ecosystems to adapt
(Rijsberman and Swart, 1990, WBGU, 2003, Van Vliet and Leemans, 2006). Within
limits and at a cost, for many of the consequences of temperature changes adaptation
options are available. These will be mentioned in the sections for the individual
indicators. Furthermore, it is important to notice that land warms faster then the
oceans, due to the heat balance (water absorbs heat).
Past trends
The IPCC (2007) has very high confidence that global warming is accelerating very
likely due to the observed increase in anthropogenic greenhouse gas concentrations
(IPCC, 2007). The increase in global average temperature (land and ocean) up to 2007
has been 0.76 °C, compared to pre-industrial (defined as 1850-1899 average, IPCC,
2007). There is also very high confidence that the observed increase over land only
has been 0.98 °C. Eleven of the last 12 years (1996 -2007) rank among the 12
warmest years (the exception being 1996). 2005 and 1998 were the warmest two years
on record (see Figure 5.2.21). The rate of change in the global average temperature
has accelerated from 0.08 °C per decade over the last 100 years, to 0.13 °C per decade
over the past 50 years and up to 0.20°C per decade over the last 10 years (all values
represent land & ocean area) (IPCC, 2007).
Europe has warmed slightly more than the global average (high confidence). The
increase of the annual average temperature for the European land area and combined
land and ocean area has been 1.22°C and 0.98°C, respectively, comparing the trend
towards 2007 with pre-industrial times. Geographically, particularly significant
warming has been observed in the past 50 years (Figure 5.2.2.2) over the Iberian
Peninsula, and central & north-eastern Europe and in mountainous regions (Böhm et
al., 2001; Klein Tank, 2004) (high confidence). The warming in the period 1979 –
2005 was strongest over northern Europe in winter. Europe as a whole warmed up
especially in spring and summer. Remarkably, autumn saw almost no warming.
47
Projections
The global average temperature is projected to continue to increase with high
confidence. The projected increase this century is in a range between 1.8 to 4.0°C,
and is considered likely between 1.1 – 6.4°C for the six IPCC SRES scenarios (see
Chapter 4), comparing the 2080-2100 average with the 1980-1999 average. Note that
these scenarios do not assume that additional policies are implemented to limit
greenhouse gas emissions (IPCC 2007). The range is caused by the uncertainties in
future socio-economic development and in climate models. The EU ‘sustainable’
target of limiting global average warming to not more than 2.0°C above pre-industrial
level likely to be exceeded between 2040 and 2060, for the IPCC scenarios.
The annual average temperature for Europe is projected to increase by 1.0 - 5.5°C
(comparing 2080-2100 with the 1961-1990 average). This range takes into account
the uncertainties in future socio-economic development by including two of the
IPCC-SRES scenarios (the high emissions A2 and the medium emissions A1b) and in
climate models (IPCC, 2007b; Christensen & Christensen, 2007). The warming is
projected to be greatest over eastern Europe, Scandinavia and the arctic in winter
(December to February), and over south-western and Mediterranean Europe in
summer (June to August) (Giorgi et al., 2004; Christensen & Christensen, 2007). The
temperature rise in parts of France and the Iberian Peninsula may exceed 6°C,
whereas the Arctic could become on average 6°C and possibly 8°C warmer than the
1961-1990 average (IPCC, 2007a,b, ACIA 2004, 2007).
Figure 5.2.2.2 Observed temperature change over Europe during the last 50 years. Left:
Annual, mid: Winter; right: summer, unit: oC/decade, (Source: http://eca.knmi.nl/download/)
Figure 5.2.2.3. Modelled change in mean annual, summer (JJA) and winter (DJF)
temperature over Europe from the period between 1980-1999 and 2080-2099, averaged over
21 models for the IPCC-SRES A1B scenario (Source: IPCC, 2007a).
48
5.2.3 European precipitation
Key Messages
• Annual precipitation trends in the 20th century showed an increase in northern
Europe (10 – 40%) and a decrease in some parts of southern Europe (up to 20%).
• Mean winter precipitation has increased in most of the Atlantic- and northern
Europe (20 – 40%).
• Models project an increase in winter precipitation in northern Europe and a decrease
in summer precipitation in southern Europe. But there are uncertainties on the
magnitude and geographical details of precipitation change.
Figure 5.2.3.1. Trend of annual precipitation for 1946 to 2006 (mm per decade); blue colour
mean increase, green means no change, yellow decrease trend (Source KNMI, CRU 2007) (to
be adapted in grid form and graphs with precipitation trends in Northern Europe and
Mediterranean)
Relevance
Precipitation is a major component of the hydrological cycle. Most of precipitation
over Europe is connected with advection of maritime air masses from the Atlantic and
Mediterranean. The amount and spatial distribution of European precipitation is
strongly influenced by circulation patterns (see Box in 5.2.1). Combination of the
changes in precipitation regime and increases in air temperature can lead to extreme
hydrological evens such a flooding and droughts (see e.g. Section 5.2.5, 5.6.4 and
5.7.3). Some systems or sectors, connected closely with the hydrological cycle, are
very sensitive to the combined effects of higher temperatures and changed
precipitation characteristics. Within limits and at a cost, adaptation to many of the
impacts is possible. These options will be briefly mentioned in the individual
indicator sections.
Homogenous time series of monthly precipitation data and interpolation and gridding
methods enable the analysis of various periods from 1901 at various temporal and
spatial scales. However, differences between climate models for future precipitation
49
projections indicate higher uncertainty in regional and seasonal results than for
temperature projections and observed precipitation trends.
Past trends
Precipitation in Europe has generally increased over the 20th century (high
confidence). The European average increase has been 6 - 8% between 1901 and 2005.
Geographically, in NW European regions the increase was generally significantly
higher. This is partly due to stronger advection of wet Atlantic air masses to this part
of the continent. Drying has been observed in the Mediterranean and eastern Europe.
In western Europe no clear trends have been observed (Norrant and Douguédroit,
2006). Similar results were found for the period 1946 – 2006 (see Figure 5.2.3.1).
Mean winter (December-February) precipitation is increasing in most of Atlantic- and
northern Europe (Klein Tank et al., 2002) (high confidence), because western
circulation was stronger in the cold part of the year. Conversely, southern Europe and
parts of central Europe were characterized by a drier winter. Trends in spring and
autumn were not significant.
Projections
Climate models project changes in precipitation that vary considerably from season to
season and across regions. Geographically, projections indicate a general precipitation
increase in northern Europe and a decrease in southern Europe . The annual areamean change from 1961-1990 to 2071-2100 for the high emissions IPCC SRES A2
projections varies from 0 to 20% in northern Europe and from –5% to –40% in
southern Europe and Mediterranean (Figure 5.2.3.2). Many impact studies (see other
indicators) use this scenario. Under the intermediate emissions IPCC SRES A1B
scenario the projected changes are smaller.
Seasonally, models project a large-scale increase in winter precipitation in mid and
northern Europe. Many parts of Europe are projected to experience dryer summers
(Figure 5.2.3.2). Relatively small precipitation changes were found for spring and
autumn (Räisänen et al., 2004; Kjellström, 2004).
Figure 5.2.3.2. Modelled change in precipitation over Europe from period 1961-1990 to
2071-2100 and seasonal changes for summer (JJA) and winter (DJF) seasons. (Source:
Dankers and R. Hiederer , 2008, using the HIRHAM4 regional climate model with boundary
conditions of the HadCM3 model and the IPCC-SRES A2 GHG emission scenario)
50
5.2.4 Temperature extremes in Europe
Key Messages
• Cold extremes have become less frequent in Europe, while the frequency of hot
days has almost tripled between 1880 and 2005 and the number of warm extremes
doubled.
• For Europe as a whole heat waves and droughts are projected to increase in
frequency, intensity and duration, whereas winter temperature variability and the
number of cold and frost extremes will further decrease. European regions
projected to be most affected are the Iberian Peninsula, central Europe including
the Alps, the eastern Adriatic seaboard, and southern Greece.
Figure 5.2.4.1: Observed changes in duration of warm spells in summer (left) and
frequency of frost days (right), in the period 1976-2006 (in days per decade)
Source: eca.knmi.nl and Haylock et al., 2008; data available from
http://ecad.knmi.nl/download/ensembles/download.php
Figure 5.2.4.2: Modelled number of tropical nights (minimum temperature >20º C)
over Europe from control period (1961-1990) and scenario period (2071-2100) during
summer seasons (JJA) and change between periods (right panel).
Data source: PRUDENCE, Dankers and R. Hiederer , 2008, using the HIRHAM4
regional climate model with boundary conditions of the HadCM3 model and the
IPCC-SRES A2 GHG emission scenario)
51
Relevance
In the public eye, climate change may manifest itself most clearly through changes in
the frequency of weather extremes and their impacts. Extreme temperature events
may lead to heat waves and intensive and long-lasting droughts, having, in turn,
numerous impacts on the environment (like natural ecosystems) and society (e.g.
agriculture, public health).
The time series for studying extremes are based on daily data. These data are more
variable than mean monthly data. However, there are more than 50-years European
time series data available on temperature and precipitation, which allow for detailed
assessment of extreme events. Within limits and at a cost, for many of the
consequences of temperature extremes adaptation options are available. These will be
mentioned in the sections for the individual indicators.
Past trends
Warm extremes like hot days, tropical nights, and heat waves have become more
frequent, while cold extremes like cold days, cold nights and frost have become less
frequent (IPCC 2007, Klein Tank et al, 2007, Fig. 5.4.2.1). The length of summer heat
waves over Western Europe has doubled over the period 1880 to 2005 and the
frequency of hot days has almost tripled (Della-Marta et al, 2007). It is more likely
than not that anthropogenic climate change has contributed to this increase (IPCC,
2007).
Projections
For Europe as a whole it is very likely that extreme high temperature events across
Europe, along with the overall warming, will become more frequent, intense and
longer (Schär et al., 2004, Tebaldi et al., 2006, IPCC, 2007a,b; Beniston et al., 2007).
Likewise, the night time temperature is projected to increase considerably (Fig.
5.2.4.2), possibly leading to additional health problems and even mortality (Halsnæs
et al, 2007, Sillman and Roekner, 2008), at least partly compensated by a reduced
mortality in winter (see Section 5.9.2).
Geographically, the maximum temperature during summer is projected to increase
much more in Southern and Central Europe than in Northern Europe, whereas the
largest reduction in cold extremes occurrence is projected for Northern Europe
(Kjellström et al., 2006; Beniston et al., 2007, Sillman and Roekner, 2008). Under the
A2 scenario, Central Europe, for example, is projected to experience by the end of the
twenty first century the same number of hot days as are currently experienced in
Spain and Sicily (Beniston et al., 2007).Also extreme cold days in winter are
projected to warm more strongly than mild winter days, which indicates a decrease in
winter temperature variability.
52
Box: The heat waves of the summer of 2003
Much of Europe was affected by heat waves during the summer 2003 (June, July and
August). Seasonal temperatures were highest on record in Germany, Switzerland,
France and Spain (Fig. 5.2.4.3). It is estimated that 2003 has been the hottest summer
since at least 1500 (Luterbacher et al., 2004). Average summer (June-August)
temperatures were far above the long-term mean by up to 5 standard deviations,
implying that this was an extremely unlikely event under current climatic conditions
(Schär and Jendritzky, 2004; Chase et al, 2006). The 2003 summer may, however,
become much more frequent during the second part of the 21st century (Beniston,
2007).
The 2003 heat waves were associated with a certain air pressure field pattern over
Europe, leading to an advection of heated air from the south which reinforced the
strength and persistence of the heat wave. Nearly all radiation from the Sun was
converted to heat because of the soil and vegetation dryness. At many locations,
temperatures rose above 40°C. In the European Alps, the average thickness loss of
glaciers reached about three metres water equivalent, nearly twice as much as during
the previous record year of 1998 (WMO, 2004; see Section 5.3.2). Annual
precipitation deficits up to 300 mm caused drought with consequences in reduced
agricultural production (Section 5.6.2), larger fires (Portugal, Section 5.6.6), record
low levels of many major rivers (e.g. Po, Rhine, Loire and Danube; Section 5.7.2). In
all the affected countries together over 75 000 additional deaths were related to the
2003 heat wave (Section 5.9.2).
Figure 5.2.4.3: Summer 2003 (J,J,A) daily maximum temperature anomaly with respect to
1961–90 (based on http://eca.knmi.nl/download/).
53
5.2.5 Precipitation extremes in Europe
Key Messages
•
•
•
•
For Europe as a whole, the intensity of precipitation extremes like heavy rain
events has increased in the past 50 years, even for areas with a decrease in mean
precipitation such as Central Europe and the Mediterranean.
The fraction of Europe experiencing extreme and/or moderate drought conditions
has not changed significantly during the 20th century.
For Europe as whole heavy precipitation events will continue to become more
frequent.
Dry periods are projected to increase especially in the different parts of Southern
Europe.
Figure 5.2.5.1: Trends in precipitation fraction due to very wet days (upper 5%)
between 1976 and 2006 (to be adapted in grid form)
Source: eca.knmi.nl and Haylock et al., 2008; data available from
http://ecad.knmi.nl/download/ensembles/download.php
Relevance
Both high and low precipitation extremes (high intensity or long-lasting rains and
droughts, respectively) can lead to periods with precipitation deficit even with a high
amount of total precipitation. The periods can vary from minutes (e.g. in case of
intense showers) to days or even weeks (with long-lasting rain events or absence of
precipitation). Regarding precipitation deficits, low precipitation extremes can lead to
drought occurrence (see Box). In the case of precipitation surplus, extremes can result
in fast flash floods, sewerage system failure and land-slides, or devastating floods,
54
affecting large catchments and having longer duration. As such, precipitation
extremes have a series of impacts. Within limits and at a cost, for many of the
consequences of temperature changes adaptation options are available. These will be
mentioned in the sections for the relevant impacts indicators.
Precipitation extremes can be assessed in different ways. Regarding precipitation
deficits, often the number and duration of dry periods is used (e.g. expressed as the
number of consecutive dry days). High precipitation events can, for example, be
assessed through number of wet days, the number of consecutive wet days, and the
frequency and intensity of heavy precipitation events (see Klein Tank et al, 2007).
The time series for studying precipitation extremes are based on daily data. These data
are more variable than mean monthly data. However, there is more than 50-years
European time series data available of precipitation, which allow for detailed
assessment of extreme events.
Past trends
The number of extreme precipitation events has increased over most of the European
land area, linked to warming and increases of atmospheric water vapour. For Europe
as a whole, the intensity of extreme precipitation has increased in the past 50 years,
even for areas with a decrease in mean precipitation, such as central Europe and the
Mediterranean. In particular, the contribution of extreme precipitation to the
precipitation totals has increased (Figure 5.2.5.1).
The fraction of Europe that has experienced extreme and/or moderate drought
conditions has not changed significantly during the 20th century (Lloyd-Hughes and
Saunders, 2002). Summer droughts do not show statistically significant trends in the
period 1901 – 2002 (Robock et al., 2005; van der Schrier at all, 2007).
Projections
For Europe as whole it is likely that heavy precipitation events will continue to
become more frequent (IPCC, 2007). The frequency of wet days is projected to
decrease in summer, but the intensity of extreme rain showers may increase. In
addition, frequency of several-day precipitation episodes is projected to increase.
Geographically, there is a considerable regional differentiation in the projections
(medium confidence). Extreme precipitation events are projected to decrease
especially in southern Europe, whereas northern and central Europe may experience
considerable increase (for the ECHAM 4 climate model, A1B scenario, 17% and
13%, respectively) (Fig. 5.2.5.2., J. Sillmann & E. Roeckner, 2008).
The combination of higher temperatures and reduced mean summer precipitation is
projected to enhance the frequency and intensity of droughts across Europe. This can,
for example, be illustrated by the projected number of the consecutive dry days (Fig.
5.2.5.3, for an intermediate emissions scenario IPCC SRES A1B and a low emissions
scenario IPCC SRES B1). The maximum number of these days is projected to
increase substantially in southern Europe during the 21st century. The longest dry
period within a year is prolonged by 1 month at the end of 21st century. In central
Europe, prolongation of longest dry period is by 1 week, but no trend is projected in
northern Europe. As such, regions in Europe that are dry now will stay more
vulnerable in the future.
55
Figure 5.2.5.2: Precipitation time series (1860–2100) of European land average for
maximum 5-day precipitation sum (mm) in Europe; 20th Century (black), A1B
(red) and B1 (blue) model simulations. (Displayed are the respective ensemble
means. The spread in the ensemble members is indicated by light (B1) and dark
(A1B) shading, respectively. Data are smoothed by 10-year running means) (J.
Sillmann & E. Roeckner, 2008).
Figure 5.2.5.3: Precipitation time series (1860–2100) of European land average
maximum number for consecutive dry days (g, h, i; days); 20th Century (black),
A1B (red) and B1 (blue) model simulations. (Displayed are the respective
ensemble means. The spread in the ensemble members is indicated by light (B1)
and dark (A1B) shading, respectively. Data are smoothed by 10-year running
means)* (J. Sillmann & E. Roeckner, 2008).
56
Drought box
Drought is a natural phenomenon, defined as sustained and regionally extensive
occurrence of below-average water availability. Drought should not be confused with
aridity, which is a long-term average feature of a dry climate. Nevertheless, the most
severe human consequences of drought can be found in arid regions, where water
availability is naturally lower. Likewise, drought should not be confused with water
scarcity, which reflects conditions of long-term imbalances between water supply and
demands.
Droughts can affect both high and low rainfall areas of Europe and can develop over
short periods of weeks and months or much longer periods of several seasons, years
and even decades. In many cases drought develops gradually, making it difficult to
identify and predict.
The mostly used definitions and types of the drought are:
• Meteorological drought: precipitation’s departure from normal values for an
extended period of time, the primary cause of the other types of drought.
• Hydrological drought: deficiencies in surface and subsurface water supplies,
reflecting effects and impacts of meteorological droughts.
• Agricultural drought: a deficit of soil moisture affecting a particular crop at a
particular time.
• Socio-economic drought: imbalance between supply of and demand for an
economic good, capturing both drought condition and human activities.
Drought duration
Weeks……………..Months……………..….Years………….………...Decades
Meteorological drought
Hydrological drought
Agricultural drought
Socio-economic drought
Precipitation is the primary factor controlling the origin and persistence of drought
conditions for all types of drought. Deficiency of precipitation results in water
shortage for some activity or for some group. The impacts of droughts on people and
the environment arise due to a combination of the intensity and duration of drought
events and the vulnerability of agricultural or water resources systems, including
water management policies, the characteristics of regional and local water
infrastructure and social responses to drought situations. Drought is a phenomenon
that is not constrained by international boundaries and can therefore grow to afflict
many countries simultaneously.
Projections of the future climate indicate that in warmer conditions droughts may
become longer lasting and more severe in current drought-prone regions because of
enhanced evaporation.
57
5.2.6 Storms and storm surges in Europe
Key Messages
•
•
•
•
•
Storminess in Europe has been relatively high during the late nineteenth and the
early twentieth century; then decreased again in central Europe and northern
Europe. The subsequent rise in the late 20th century was most pronounced in
north-western Europe, while slow and steady in central Europe.
The water level rise aboard most vulnerable European coastlines of the North Sea
and Mediterranean Sea had not shown an increase due to trend of storminess.
Extra-tropical storm tracks are projected to move pole wards, with consequent
changes in wind, precipitation, and temperature patterns, continuing the broad
pattern of observed trends over the past half-century.
Climate models indicate a slight decrease in the number of storms and an increase
of the strength of the heaviest storms.
For the end of 21st century the projections show a significant increase of storm
surge elevation for the continental North Sea and South East England.
Figure 5.2.6.1. Storm index time series for the period 1881 – 2004 for north-western,
north and central Europe, positive values of the index are connected with higher
storminess (from Matulla et al., 2007). (Comm. Left graph will be used only)
Relevance
Storms in Europe are extreme, near-surface damage-causing wind speeds, associated
with passage of intense extra-tropical cyclones (Pinto, et al., 2007). Storms can occur
in north or north-western Europe all year, whereas they occur in central Europe
mainly between November and February. Storm surges are temporary increases in
sea-level, above the level of the tide, causing often coastal flooding.
Storm events can have in general large impacts on many vulnerable systems such as
transport, forestry and energy infrastructures, but also on human safety.
The storm activity for Europe and the neighbouring part of the Atlantic is closely
connected with the intensity of the atmospheric circulation (North Atlantic Oscillation
- NAO). But the correlation of the NAO index with storminess across Europe varies
in space and time. Direct wind observation data of sufficient quality are often lacking.
58
Storms intensity and frequency can indirectly be assessed by the sea level pressure
field.
Note that modelling the future wind conditions (and changes in it) is highly uncertain,
mainly due to uncertainty in atmospheric circulation projections.
Storm surges result from the combined action of atmospheric pressure and strong
wind on the sea surface and mostly occur in shallow water. An increase in mean sea
level will directly affect extreme levels. Changes in water depth can also influence the
tidal component, modifying the extent of flooded areas. Adaptation options are
available, such as improved coastal management systems and spatial planning.
Past trends
Storminess in Europe has not significantly changed over the past century. Since the
mid 1960s to at least the mid-1990s, the western circulation above the Northern part
of the Atlantic Ocean has become stronger and more northerly than it was before (Yin
2005), causing a strengthening of the mid-latitude westerly winds in the Northern
Hemisphere (IPCC 2007). After the mid-1990s, these winds, however, have returned
to the pre-1960 values. These results are in agreement with the storm-index time
series in fig. 5.2.6.1 - based on air pressure data- which show that storminess in northwestern, northern and central Europe has been relatively high during the late
nineteenth and the early twentieth century; then decreased again in central Europe and
northern Europe. The subsequent rise in the late 20th century was most pronounced in
north-western Europe, while slow and steady in central Europe. Most recent years
feature average or calm conditions (Matulla et al., 2007).
The evaluation of the high tide levels along the North Sea in the past century found
clear signals in mean levels (the sea level rise) but not in storm-related variation (von
Storch at al., 2002). Also a study of storm surges occurrence in Northern Adriatic Sea
showed that high sea level trend is not associated to a trend of storminess (Lionello
2005).
Projections
Extra-tropical storm tracks are projected to move poleward, with consequent changes
in wind, precipitation, and temperature patterns, continuing the observed trends over
the last half-century (IPCC, 2007). The total number of storms is projected to
decrease, whereas the strength of the heaviest storms may increase (see Figure 5.2.6.2
showing the difference of a regional maximum wind distribution between different
model outputs). Note that these projections are still very uncertain and model
dependent.
During historical times, storminess and large-scale temperature variations were
mostly decoupled, but the scenarios show different conditions. Some projections,
based on the high emissions IPCC SRES A2 scenario, show an increase in the
frequency of heavy storms for the North Atlantic Ocean. In this region both higher
temperature and high storm frequency change, where the trend in storms is smaller
than for temperature. The future storminess in this region depends on projections of
sea surface temperature, retreat of arctic ice and changes of the air pressure field
(Fischer-Bruns et al., 2005).
Projections of storm surges are closely connected with the future storminess
behaviour. For the end of 21st century the projections point to a significant increase of
storm surge elevations for the continental North Sea coast by between 15 and almost
25 cm (Woth, 2005). Also for the UK coastline a large increase in relative surge
height has been projected for the high IPCC SRES scenario A2 and the intermediate
59
scenario B2, especially along the southeast coast of England, where the changes in
storminess will have their largest effect and where the land is sinking most rapidly
(Lowe and Gregory, 2005, Fig. 5.2.6.3).
Figure 5.2.6.2. Relative change (%) in the annual maximum daily mean wind speed at 10 m
height between the +2C scenario for 2050 and the reference climate (1961-2000) from three
similar GCMs (left) and one different GCM, MIROCHi (right) (Source: KNMI Climate
Change Scenarios 2006 for the Netherlands).
Figure 5.2.6.3. Change in the height (m) of an extreme water level event (with probability of
occurrence 2 times in 100 years) measured relative to the present day tide, due to changes in
atmospheric storminess, an increase in mean sea level and vertical land movements. Results
are shown for the (a) A2 scenario (b) B2 scenario for the end of 21st century (Source. Lowe,
J.A. and Gregory, J.M. 2005) (Note: will be specified used model)
60
5.2.7 Air pollution by ozone
Key Messages
•
•
•
Climate change has contributed to an increase in ozone concentration in
central and South-Western Europe (1-2% per decade).
During the summer of 2003, exceptionally long-lasting and spatially extensive
episodes of high ozone concentrations occurred, mainly in the first half of
August. These episodes appear to be associated with the extraordinarily hot
temperatures over wide areas of Europe.
Due to projected climate change, more frequent exceedances of the ozone
information threshold could be expected, even at the current emission levels.
Figure 5.2.7.1: Trend in tropospheric ozone concentrations (unit: % per decade) due
to climate change. White areas have no significant trend. For the left panel the time
period is 1958–2001. For the right panel, the time period is 1979–2001. (Source
Andersson et al., 2007)
Relevance
Tropospheric ozone is one of the air pollutants of most concern in Europe. Ozone is
estimated to cause about 20,000 annual acute mortalities (European Commission,
Clean Air for Europe impact assessment, 2006) and an economic damage due to crop
loss of €4625 million per year (Holland et al, 2006). Ozone is formed in the lower
troposphere from complex chemical reactions between volatile organic compounds
and nitrogen oxides, in the presence of sunlight. EU legislation has established ozone
exceedance thresholds and national emission ceilings for ozone precursor emissions in
order to protect human health and prevent damage to ecosystems, agricultural crops
and materials.
Episodes, i.e. periods with elevated ozone levels, mainly occur during periods of
warm sunny weather. An increasing likelihood of hot extremes, heat waves in Europe
is expected to cause a climate-driven impact on ozone episodes which may call for
more vigorous emission reduction measures. Also adaptation measures are available,
such as improved public information and health care services.
61
Past trends
A modelling study from 1958 to 2001 (Andersson et al., 2007) shows that an increase
in temperature contributed to increased ozone concentrations during the period 19792001 over south-central and south-western Europe. Temperature affects the emission
of biogenic organic compounds (isoprene) which are precursors for ozone formation,
and the photo-dissociation of NO2 (Figure 5.2.7.1).
A statistical analysis of ozone and temperature measurements in Europe for the period
1990 – 2004 shows that in Central-Western Europe and the Mediterranean area, a
change in the daily maximum temperature in the period 1999 – 2004 compared to
1990-94 has contributed to extra ozone exceedences (Figure 5.2.7.2). In South and
central Europe, the temperature trend was responsible for 8 extra exceedence days of
120 µg/m³, i.e. 17% of the total number of exceedences observed in that region.
An analysis of trends over the past 12 years indicates that in the European Union the
average number of hours per station when ozone concentration exceeded the
information threshold of 180 μg/m3 was higher in summer 2003 than in all previous
years (EEA, 2003).
Projections
The expected trends for ozone and other air pollutants are closely linked to
projections for radiation, temperature, cloudiness, and precipitation. On a global scale,
the effect of climate change only on tropospheric ozone is expected to be small, due to
a reduction in the ozone lifetime as a consequence of higher humidities (Stevenson et
al, 2006). However, regional differences can be large. A model study based on the
IPCC92a scenario shows a strong increase in tropospheric ozone over southern and
central Europe, and a decrease over northern Europe (Langner et al., 2005), in line
with observed trends form the past. These changes are due to climate change only,
and are likely to exacerbate air quality in regions with high NOx emissions.
62
Figure 5.2.7.2: Change in number of ozone exceedance days per year2, from period
1990-1994 to the period 1999-2004 (excluding the year 2003). Green bars indicate the
net change (increase or decrease) in the number of exceedences. Orange bars indicate
the contribution due to the temperature change during that period, the blue bars
indicate the contribution of other causes (e.g. changing emissions). Source: Van
Dingenen et al., manuscript submitted.
2
days where the maximal 8hr average ozone concentration exceeds 120 µg per m³
63
5.3 Cryosphere
5.3.1 Introduction
The cryosphere is the frozen part of the world. It includes all permanent or seasonal snow and
ice deposits on land, in the seas, in rivers and lakes and in the ground (permafrost). It is the
second largest component of the climate system after the oceans with regard to mass and heat
capacity. Snow and ice play a key role in the earth’s energy budget by reflecting heat from the
sun due to its light surfaces. As melting replaces white surfaces with darker, more heat is
being absorbed (the albedo effect). Three quarters of the world’s freshwater resources are
frozen. Snow and ice play a key role in the water cycle and are essential for storing fresh
water for the hotter and often dryer seasons.
The cryosphere is important for the exchange of gases between the ground and the
atmosphere – among these several greenhouse gases, e.g. methane. Finally, ice and snow are
defining components of ecosystems in the northern parts of the Northern hemisphere and in
high mountain areas. Many plants and animals have evolved to live under these conditions
and can not live without. The cryosphere hence plays a major role in various dimensions of
the climate system: it is affected if the climate changes, but its own changes affect the climate
system in turn. Therefore, monitoring these changes provides crucial knowledge about
climate change.
Selection of indicators
The various components play strong but different roles within the climate system:

Due to their large volumes and areas, the two continental ice sheets of Greenland and
Antarctica actively influence the global climate over very long time scales. Sea level
can however be affected more rapidly.
 Snow covers a large area too, but has relatively small volume. It is important for key
global interactions and feedbacks like albedo.
 Sea ice also covers a large area. It is important due to its albedo and its impacts on
ocean circulation, which transports heat from equator to the poles.
 Melting permafrost release the strong greenhouse gas methane from frozen organic
material. Together with seasonal snow, it influences the water content in the soil and
the vegetation.
 Glaciers, ice caps and seasonal lake ice, with their smaller areas and volumes, react
relatively quickly to changes in climate, influencing ecosystems and human activities
on a local scale. They are good indicators of climate change.
The selected indicators cover strategic information from all these compartments of the
cryosphere: glaciers and snow cover in Europe, the Greenland ice sheet, the Arctic sea ice and
mountain permafrost in Central Europe. Lake and river ice conditions are presented in the
water chapter.
Indicators and vulnerability
The cryosphere is vulnerable to global warming. It is a very visible expression of climate
change. It integrates climate variations over a wide range of time scales, from millennia to
seasonal variations throughout the year. This can sometimes complicate interpretations of
why changes happen.
64
In Europe, the most vulnerable areas are the high mountain areas, and the Arctic. Both in the
Arctic and the European Alps, temperatures have increased more than global rate in the past
few decades; in the Arctic as a whole, twice as much. The amount of ice and snow, especially
in the Northern Hemisphere has decreased substantially over the last few decades due to
increased temperatures. European glaciers shrink, snow covered areas creep higher up and
further north, sea ice in the Arctic melts and gets thinner, permafrost starts to thaw, and the
Greenland ice sheet shows increasing signals of disintegration and thawing at its borders.
These trends will accelerate as climate change is projected to continue. However, there are
also large uncertainties to the fate of global key components of the cryosphere; it is not
possible to give reliable predictions about when the Arctic sea ice will melt completely in
summer. Neither is it possible now to predict the future of the Greenland ice sheets with any
confidence
Data and info sources
Data on the components of the cryosphere are quite different with regard to availability and
quality. Long-term data records on glaciers of all European glaciated areas are provided in a
considerable quality by the World Glacier Monitoring Service (WGMS) in Zürich. Data on
snow cover and Arctic sea ice from the past are mostly regionally available only, measured by
scientific expeditions or military submarines. More recent data covering great areas result
mostly from satellite missions backed by ground-truth measurements, available e.g. at the
Global Snow and Ice Data Centre (NSIDC) in Boulder/U.S.A. Particularly area-wide data on
the Greenland ice-sheet are often available not longer than for 15 years, as well as data on
mountain permafrost measured in boreholes which are drilled into frozen rock walls.
The gaps in the cryospheric data base are well recognized by the scientific community and
many efforts are taken to improve the knowledge, e.g. during the actual running International
Polar Year (IPY).
65
5.3.2 Glaciers
Key Messages
• The vast majority of glaciers in the European glacial regions are in retreat.
• Since 1850, glaciers in the European Alps have lost approximately two thirds of their
volume, with clear acceleration since 1985.
• Glacier retreat is projected to continue. A 3°C warming of summer air temperature could
reduce the currently existing glacier cover of the European Alps by some 80 %. With
ongoing climate change nearly all of the smaller glaciers and one third of the overall
glacier area in Norway are projected to disappear by 2100.
• Glacier retreat has serious consequences for river flow. It impacts freshwater supply, river
navigation, irrigation and power generation. It could cause natural hazards and damages
to infrastructure.
Presentation of the main indicator
Figure 5.3.2.1.: Cumulative ice loss (in mm) of glaciers from all European regions with
glaciers Source: World Glacier Monitoring Service (www.wgms.ch), 2006.
Relevance
Glacier changes are among the most visible indications of the effects of climate change.
Glaciers are particularly sensitive to changes in the global climate because their surface
temperature is close to the freezing/melting point (Zemp et al., 2007). Glacier fluctuations
show a strong relation to air temperature throughout the 20th century (Greene, 2005), and a
strong physical basis exists to explain why warming would cause mass loss (IPCC, 2007a).
Therefore their mass balance, resulting from accumulation (mainly snow fall) and ablation
(mainly melting and calving), is considered as a immediate signal for the early detection of
global warming trends. Glaciers are important freshwater resources and act as `water towers’
for the lower lying regions. With a diminished ice mass, the annual melt water and therefore
the contribution to river flow and sea level rise is decreasing in the long term. This will cause
serious consequences for freshwater supply, river-navigation, irrigation facilities, and power
generation. Strong retreat of glaciers can cause glacier instabilities with hazardous incidents
66
such as glacier lake outbursts, rock-ice avalanches and landslides (Pralong et al., 2005;
Huggel et al., 2007). This may cause more damage to infrastructure. Glacier retreat affects
tourism and winter sports in the mountains (OECD, 2007) and changes the appearance of
mountain landscapes.
Adaptation options, such as improved glacier monitoring, water management measures,
draining of glacier-lakes and construction of protective walls can reduce some of the risks and
negative consequences, but not all.
Past trends
According to the high-quality data-records of the WGMS in all of the European glacier
regions, a general loss of glacier mass has been observed (Fig.: 5.3.2.1). Glacier retreat in
Europe started after the maximum glacier-extent of the so-called ‘Little Ice Age’ in the
middle of the nineteenth century. In the Alps, glaciers lost one third of their surface and one
half of their volume between 1850 and 1970s. After 1985 an acceleration in glacial retreat has
been observed, which led to a loss of 25% of the remaining ice until 2000 (Zemp et al., 2006).
This culminated in a further ice loss of 5-10% in the extraordinary hot and dry summer of
2003 (Zemp et al., 2005), and resulted in a total loss of about 2/3 of ice mass until present.
This is illustrated by the shrinking of the Vernagtferner-glacier in Austria (Fig.: 5.3.2.2).
Also the Norwegian coastal glaciers, which were expanding and gaining mass due to
increased snowfall in winter up to the end of the 1990s, are retreating now by reduced winter
precipitation and greater summer melting (Nesje et al., 2007; Andreassen et al, 2005).
Glaciers in Svalbard are experiencing mass loss at lower elevations, and the fronts of nearly
all glaciers there are retreating (Haeberli et al., 2005; Nuth et al. 2007). Some ice
caps in north-eastern Svalbard seem to increase in thickness at higher elevations (Bamber et
al., 2004, Bevan et al., 2007). However, estimates for Svalbard as a whole show the total
balance is negative (Hagen et al. 2003), and there is a clear signal of accelerated melting at
least in western Svalbard (Kohler et al., 2007).
Very recent findings by the WGMS (UNEP, 2008) indicate a clearly increased annual
reduction of the global-mean ice-thickness of glaciers since the turn of millennium (0.5m)
compared to the 1980-1999 period (0.3 m). Some of the most dramatic shrinking has taken
place in Europe (Scandinavia, Alps, and Pyrenees).
The centennial retreat of European glaciers is mainly attributed to increased summer
temperatures. However, changes in winter precipitation, the decreased glacier albedo due to
lack of summer snow fall and various other feedback processes alter the pattern on regional
and decadal scale. The recent strong warming has made disintegration and down wasting
increasingly dominant causes of glacier decline in the European Alps during the most recent
past (Paul et al., 2004).
Projections
According to a recently published sensitivity study (Zemp, 2006) the European Alps could
lose about 80 % of their currently existing glacier cover, should summer air temperatures rise
by 3° C and a precipitation increase of 25 % for each 1°C would be needed to offset the
glacial loss. The modelled remains of Alpine glaciers as a consequence of warming is
presented in Fig. 5.3.2.3. Sugiyama et al. (2007) investigated the potential evolution of Rhone
Glacier, Switzerland, in the 21st century based on a model considering more the glacier flow
dynamics. They found increasing mass loss as well, but to a gradual lesser extent. However,
both modelling studies did not consider feedback processes such as the development of
glacier lakes, which might accelerate the glacier retreat dramatically. Recent climate scenarios
for Norway basing on model calculations of the British Headley Centre and the German Max
Planck Institute which follow the SRES B2 emission-scenario indicate a rise in summertemperature by 2.3°C and an increase in winter-precipitation by 16% by the end of 21st
century. As a result, nearly all of the smaller Norwegian glaciers are likely to disappear and
the overall glacier area may be reduced by about one third at 2100 (Nesje et al., 2007).
67
Fig.5.3.2.3: Modeled remains of the glacier cover in the European Alps (climatic
accumulation area) for an increase in summer air temperature of +1 to +5 °C.
(The total of 100% refers to the ice cover of the period 1971–1990).
Source: Zemp, M., 2006
Figure 5.3.2.2: Shrinking of the Vernagt-ferner glacier, Austria
Source: Commission of Glaciology, Bavarian Academy of Sciences; Munich, 2006
www.glaziologie.de
68
5.3.3. Snow cover
Key Messages
• The northern hemisphere's snow cover extent has decreased at a rate of 1.3% per decade
during the last 40 years. The greatest losses are in spring and summer.
• Model simulations project widespread reductions of extent and duration of snow cover in
Europe over the 21st century.
• Changes in snow cover affect the Earth surface reflectivity, river discharge, vegetation,
agriculture and animal husbandry, tourism, snow sports, transport and power generation.
Presentation of the main indicator:
Fig. 5.3.3.1: Differences in the distribution of
March-April average snow-cover in Europe
between earlier (1967-1987) and later
((1988-2004) portions of the satellite era (%
coverage).
Negative values (brown) indicate greater
extent of snow cover in the earlier portion.
Red curves show the 0°C and 5 °C isotherms
averaged for March and April 1967 to 2004
Source:
modified from Lemke et al. (2007)
Fig. 5.3.3.2: Northern Hemisphere snowcover extent departures from monthly
means from 1966 to 2005
Source: Brodzik, 2006 (NOAA-data)
Relevance
Snow covers more than 33 % of the land surface north of the equator from November to
April. It reaches a maximum of approximately 45.2 million km2 in January, and a minimum
of about 1.9 million km2 in August (Clark et al., 1999a).
Snow cover is an important feedback mechanism in the climate system. The extent of snow
cover depends on the climate e.g. on temperature, precipitation and solar radiation. But it also
influences the climate and climate-related systems by its high reflectivity, insulating
properties, its effects on water resources and ecosystems, and cooling of the atmosphere. Thus
a decrease in snow cover reduces the reflection of solar radiation, contributing to an
accelerated climate change. Changes in extent, duration, thickness and properties of snow
cover have an impact on water availability, as for domestic use, navigation and power
generation. Changes in snow cover affect human well-being through influences on
agriculture, infrastructure, and livelihoods of Arctic indigenous people, environmental
69
hazards and winter recreation. Snow cover retreat can reduce complications in winter road
and rail maintenance, affecting the exploitation and transport of oil and gas in cold regions
(UNEP, 2007; ACIA, 2004).
Shallow snow cover at low elevations in temperate regions is the most sensitive to
temperature fluctuations and hence most likely to decline with increasing temperature (IPCC,
2007).
For several of these impacts, adaptation can reduce the negative implications of snow cover
change. Some adaptation options, such as artificial snowmaking in the Alps to maintain
tourism as a main source of income, have to be balanced against their negative implications
for mitigation, due to increased energy use and greenhouse gas emissions.
Past trends
Data from satellite monitoring (NESDIS-database at NOAA) from 1966 to 2005 show that
monthly snow-cover extent in the Northern Hemisphere is decreasing at a rate of 1.3% per
decade (Fig. 5.3.3.2), with the strongest retreat in spring (Fig. 5.3.3.1) and summer (UNEP,
2007). It has declined in all months except November and December, with the most
significant decrease during May to August (Brodzik et al., 2006). This was accompanied by
lower springtime water content, earlier disappearance of continuous snow cover in spring (by
almost two weeks in the 1972-2000 period; Dye, 2002), less frequent frost days and shorter
frost-seasons.
The trends in duration and depths of Northern Hemispheric snow cover at higher latitudes
differ by region. In contrast to a reduced duration of snow cover over North America, a long
term increase in depth and duration has been observed over most of northern Eurasia (Kitaev
et al., 2005).
Snow cover trends in mountain regions of Europe vary largely with region and altitude.
Recent declines in snow cover have been documented in mountains of Switzerland (e.g.
Scherrer et al., 2004) and Slovakia (Vojtek et al., 2003), but no change was observed in
Bulgaria over the 1931 to 2000 period (Petkova et al.,2004). Declines, when observed, were
largest at lower elevations, and Scherrer et al. (2004) statistically attributed the declines in the
Swiss Alps to warming. Lowland areas of central Europe are characterised by recent
reductions in annual snow cover duration by about 1 day/year (Falarz, 2002). At Abisko in
subarctic Sweden, increases in snow depth have been recorded since 1913 (Kohler et al.,
2006), and trends towards greater maximum snow depth but shorter snow season have been
noted in Finland (Hyvärinen, 2003).
Projections
Model simulations project widespread reductions in snow cover over the 21st century (IPCC,
2007). Decreases of between 9 and 17% in the annual mean NH snow cover by the end of 21st
century are projected by individual models (ACIA, 2004). Although winter precipitation is
projected to increase in northern and central Europe (Christensen et al., 2007), less frequent
frost occurrence associated with higher temperatures are projected to reduce the number of
days with snow cover (Fig. 5.3.3.3). Decreases of more than 60 snow cover days are projected
to occur around the northern Baltic Sea, on the west slope of the Scandinavian mountains and
in the Alps (Jylhä et al., 2007).The beginning of the snow accumulation season (the end of the
snowmelt season) is projected to be later (earlier), and the snow coverage is projected to
decrease during the snow season (Hosaka et al., 2005).
For every 1 °C increase in winter-temperature, the snowline rises by about 150 metres in the
European Alps (Beniston, 2003a). Regional climate model runs, following the SRES emission
scenarios A1B, B1 and A2, project milder winters with more precipitation in this region,
increasingly falling as rain (Jacob et al, 2007). A recently published study of the Alpine
region by Hantel et al. (2007) reported an estimated reduction of about 30 days in snow
duration (snow cover of at least 5 cm) in winter at the height of maximum sensitivity (about
700m) for a 1°C increase in temperature over central Europe (5-25°E and 42.5-52.5°N).
70
Snowfall in lower mountain areas is likely to become increasingly unpredictable and
unreliable over the coming decades (Elsasser and Bürki, 2002), with consequences for natural
snow-reliability and therefore difficulties in attracting tourists and winter sport enthusiasts in
the future (OECD, 2007).
Snowfall in lower mountain areas is likely to become increasingly unpredictable and
unreliable over the coming decades (Elsasser and Bürki, 2002), with consequences for natural
snow-reliability and therefore difficulties in attracting tourists and winter sport enthusiasts in
the future (OECD, 2007).
Fig. 5.3.3.3 Annual number of days with snow covers over land areas
(a) Observed means in northern Europe, 1961-1990 (coloured dots) and modelled means for
Europe, 1961-1990, based on seven regional climate model simulations (contours) ;
(b) Projected multi-model mean changes for the period 2071-2100, relative to 1961-1990
Source: Jylhä et al., 2007
71
5.3.4 Greenland Ice Sheet
Key Messages
• The ice loss from the Greenland ice sheet more than doubled in the last decade of the 20th
century and may have doubled again by 2005 (UNEP 2005). Accelerated movement of outlet
glaciers towards the sea accounts for 2/3 of the ice loss (Rignot et al 2006).
• The corresponding contribution to global sea level rise has been suggested having increased
from 0,14 to 0,28 mm/year between 1993 and 2003. In the long run, melting ice sheets have
the largest potential to increase sea level.
* It is not possible to give reliable predictions of the future of the ice sheets yet; Processes
causing the faster movements of the glaciers are poorly understood and there is a lack of long
term observations.
Presentation of the main indicators
Figure 5.3.4.1: Calculations of the ice mass changes in Greenland demonstrate increased ice
loss over time, but also uncertainties in the most recent estimates since estimates from the
same period do not overlap.
The rectangles depict the time period of observations (horizontal) and the upper and lower
estimates of mass balance. Open bars refer to direct calculations from gravity measuring
satellites, filled bars are other techniques. (Source: Cazenave 2006)
72
2007
Figure5.3.4. 2: The area of Greenland that undergoes melting during a year has increased
over time; melting now reaches altitudes up to xxx masl. The melting areas however also
receive snow, so the resulting changes in mass balance is not negative all over the red areas.
(Konrad Steffen and Russell Huff, CIRES, University of Colorado at Boulder)
Relevance
The polar ice sheets in Greenland and Antarctica contain 98-99% of the freshwater ice on
Earth’s surface. The frozen water stored in Antarctica equals a 57 meters rise in global sea
level, whereas the water in the Greenland ice equals 7 meters. When setting their upper
estimate of 80 cm sea level rise at the end of this century, the IPCC did not take into account a
possible acceleration in ice loss from the ice sheets. The uncertainty about their future is
therefore a main reason for uncertainties in projections of sea level rise. The Greenland ice
sheet is the most susceptible to warming because of its vicinity to the Atlantic Ocean and
other continents. But also the Antarctica now seems to have a net loss of ice that may be
accelerating (UNEP 2007). (See indicator on sea level)
The speed of ice loss is important in addition to the magnitude because a faster increase in sea
level reduces the time to take appropriate adaptation measures.
The melt water from Greenland will contribute to reduce the salinity of the surrounding
ocean. An upper layer of fresher water will reduce the formation of dense deep water, one of
the mechanisms setting up global ocean circulation.
Past trends
The Greenland Ice Sheet is a huge inland glacier with several glacier tongues calving into the
sea. It covers roughly 80% of Greenland. The average ice thickness is 1600 meters, with the
highest summit reaching 3200 meters above sea level. It has a volume of approximately 3
million km3.
Until recent improvements in remote sensing, it was hard to measure whether the polar ice
sheets were growing or shrinking. Still most time series are short. There is however a general
consensus from different measurements that the melting of the Greenland Ice sheet has
accelerated; From a near balance in the early 90s, the loss of ice is estimated to have doubled
in the last decade of the 20th century to about 100 billion tonnes per year and may have
doubled again by 2005 (UNEP 2007). The estimated contribution to sea level rise increased
from 0,14 to 28 mm/year between 1993 – 2003 (IPCC WG 1, page TS-23). However, there is
still considerable discrepancy between different estimates on current ice loss (see figure
5.3.4.1) The year 2007 ….(to be updated when new estimates come)
73
The ice has thickened in the interior of the Greenland ice sheet above 2000 meters since it has
received more snowfall – in average ap. 4 cm/year since 2000. This gain has been more than
offset by the loss in lower-lying regions. The air temperatures have risen (K Steffens: 7
degrees F since 1991), and the area that undergoes melting has increased by xx% from 1997
to 2007 with 2007 as the year with the most widespread melting measured (figure 5.3.4.2).
Large amounts of meltwater form rivers and melting ponds at the surface and penetrate
through crevasses to the bottom of the glacier. Here the water probably lubricates the
bedrock-ice interface, making the glaciers move faster. The fastest flowing glacier, Illulisat
doubled its speed up to 14 km/year and also discharges twice as much ice into the ocean.
Individual glaciers are also reported to slow down, possibly around a new equilibrium. The
phenomena is widespread mostly on the southeast coast and has moved northwards to ap. 70o
N. The speeding up of the glaciers is associated with large retreats and thinning. The surface
elevation can subside with tens of metres near the outlets. The outlet glaciers act as “bathtube
drains” to the inner ice; ice is being transported into the melting zone, and calving into the
ocean increases. The speeding up of the outlet glaciers accounted for 2/3 of the ice loss from
Greenland in 2005 (Rignot and Kanagaratnam 2006). The rest is the difference between snow
accumulation and melting.
Projections
It is not possible to predict the future development of the Greenland ice sheet with
confidence. Glacier models account mostly for accumulation of snow during winter and
melting in summer (surface mass balance). The processes at the bottom of the glaciers
believed to cause their acceleration are not understood well and can therefore not be modelled
with confidence. The sensitivity of the Greenland ice sheet to global warming therefore
largely is unknown – also if there are tipping points that can accelerate the melting because of
positive feed-backs.
Deep ice core drillings can give some indications. The Eem period 120 000 years ago was
approximately 5 degrees warmer that today, but the Greenland ice did not melt down
completely. Sea level rose 5 meters above today’s level, with the melting Greenland ice
contributing 1-2 meters (Dahl Jensen, pers. com).
The ice sheets of Greenland and Antarctica have been associated with slow climate responses
over thousands of years. But the acceleration of ice movement has caused a rethinking of how
fast they respond to warming. Paleodata show periods of rapid melting of the large
continental ice sheets after the last ice age, resulting in average rise of sea level with 1
cm/year and peak rates up to 4 cm/year (UNEP 2007). However, the observation periods for
the recent changes are too short to draw clear conclusions on the nature of changes observed.
If current trends continue, it seems however likely that there will be a faster ice loss than
previously expected. Since the contributions to sea level rise from rapidly changing flow in
ice sheets were excluded from the IPCC 2007 projections, this could also lead to a higher rise
in sea level than their upper estimate of +80 cm. As to the ocean currents, the IPCC projects a
25% reduction in this century of the North Atlantic Current, but no abrupt changes.
74
5.3.5 Arctic Sea Ice
Key Messages
• Sea ice extent in the melt season has declined at an accelerating rate. The record low ice
cover September 2007 was roughly half the size of the normal minimum extent in the 1950s.
* The summer ice is projected to continue to shrink and may even disappear at the height of
the melt season in the upcoming decades. There will still be substantial ice in winter.
•
Reduced polar ice will speed up global warming and is expected to impact ocean
circulation and weather patterns. Species specialized for life in the ice are threatened and
some could even get extinct.
* Less ice will ease the access to the Arctic’s resources. Oil and gas exploration, shipping,
tourism and fisheries will offer new opportunities, but also increase pressures and risks to the
Arctic environment.
Presentation of the main indicator(s)
Figure 5.3.5.1: Time trends of arctic sea-ice extent
(Will be combined in one graph with absolute values of extent at y-axis)
(Source: Serreze et al 2007, must be updated to 2007 for summer and 2008 for winter).
75
Figure 5.3.5.2: Geography of arctic sea ice extent
The extent of the summer sea ice in September 2007 reached a historical minimum, 39%
below the climatic average for the first two decades of satellite observations, starting in
1979 (red line). The weather conditions that summer was dominated by clear sky.
Continuous warm winds blew the ice towards the coast of Canada-Greenland and southwards
through the Fram Strait east of Greenland and out in the NE Atlantic.
Relevance
Reduction in Arctic sea ice has several feedbacks to the climate system. Snow- covered ice
reflects 85% of the sunlight (high albedo), whereas open water reflects only 7% (low albedo).
Less ice and snow will therefore both accelerate the sea ice decline and global warming.
Reduced ice formation will also reduce formation of dense deep water which contributes to
drive ocean circulation. As the ice cover influences air temperature and circulation of air
masses, changes in weather patterns such as storm tracks and precipitation can be expected
even at mid-latitudes (Serreze et al 2007).
The sea ice is an ecosystem filled with life uniquely adapted to these conditions, from micro
organisms in channels and pores inside the ice, rich algae communities underneath, to fish,
seals, whales and polar bears. The diversity of life in the ice grows with the age of the ice
floes. These species will be reduced in numbers with a risk of extinction for some of them
when the ice gets younger and smaller. Arctic indigenous peoples adapted to fish and hunt
species associated with ice, will have to face large economic, social and cultural changes.
Less summer ice will ease the access to the Arctic Ocean’s resources, though remaining ice
still will pose a major challenge to operations most of the year. Expectations of large
undiscovered oil and gas resources already drive the focus of the petroleum industry and
governments northwards. As marine species move northwards with warmer sea and less ice,
so will the fishing fleet. It is however hard to predict whether the fisheries will become richer
or not; fish species react differently to changes in marine climate, and it is hard to predict
whether the timing of the annual plankton blooms will match the growth of larvae and young
fish also in the future. Shipping and tourism are likely to increase, though drift ice, short
sailing seasons and lack of infrastructure will impede a fast development of transcontinental
shipping of goods; it is more likely that traffic linked to extraction of Arctic resources in the
76
outskirts of the northern sea routes will grow first. These activities offer new economic
opportunities. At the same time they also represent new pressures and risks to an ocean that so
far has been closed to most economic activities by the ice. This should be met by better
international regulations.
A race for the Arctic’s resources may represent a challenge for security in an area where there
still are some unsettled marine borders and jurisdiction disputes. Several such disputes have
found their solution through negotiations between the Arctic coastal states.
Past trends
The extent of the minimum ice cover at the end of the melt season in September 2007 broke
all previous records. If older ship and aircraft observations are taken into account, sea ice
coverage may have been halved since the1950s. (NSIDC, 2007; Meier, 2007). Since the more
reliable satellite observations started in 1979, the summer ice has shrunk with 10,2 % per
decade. (Comiso 2007, NSIDC, 2007). The reduction in maximum extent in winter is smaller,
with a decrease of 2,9% per decade .(Figure 5.3.5.1) (Stroeve et al 2007). Both summer and
winter decline has accelerated (Comiso 2008).
The ice also gets thinner and younger since less ice survives summer and grows into
thicker multi-year floes. There is a remarkable shift in the composition of the sea ice
towards less multi-year ice and larger areas covered with first-year ice. This young ice
absorbs more heat, is weaker and has less biological life than the older ice. (Figure
5.3.5.3).
Figure 5.3.5.3. Time series of area of perennial sea ice extent in March of each year in the
period 1957 - 2007 (Nghiem et al 2007)
Observations of thickness are more scattered, and it is hard to calculate trends for the
whole ice cover. Submarine data have been considered to be most representative and
have demonstrated a decrease of 40% from an average of 3,1 meters in 1956-78 to 1,8
meters in the 1990s (Rothrock et al 1999; UNEP 2007). British submarine data show
continued thinning (Wadham, pers com). German observations from the area around
the North pole indicate that ice thickness here has decreased from 2 m in 2001 to 0.9
m in 2007, both due to a general thinning and a shift towards generally younger ice
(Nghiem et al,;2007; Haas, pers.com.) These results are in stark contrast with
77
observations between Ellesmere Island and 86N, where ice thickness was still above 4
m in 2006 (Haas et al., 2006).
The Arctic sea ice reacts very sensitively to changes in air and ocean temperatures as well as
winds, waves and ocean currents. There are strong imprints of natural variability in the
observed changes, e.g. due to regular shifts in the circulation patterns of the polar atmosphere.
However, the changes that can be attributed to increases in greenhouse gases seem to increase
over time (Stroeve et. al., 2007).
Projections
The summer ice is very likely to continue to shrink in extent and thickness, leaving larger
areas of open waters for an extended period of time. It is also very likely that conditions for
freezing in winter will persist so the winter sea ice still will cover large areas.
The speed of change is however uncertain. Several international assessments until recently
concluded that mostly ice free late-summers may occur by the end of this century (ACIA,
2004; IPCC, 2007; UNEP, 2007). But the actual melting has been faster than the average
trends simulated by the climate models used for these assessments (Figure 5.3.5.4). New
model studies suggest that ice-free summers may occur in a much nearer future. (Winton,
2006; Holland et. al., 2006; Stroeve et. al, 2007; Maslowski, 2007). Exactly when is
impossible to predict, both due to the limited understanding of the processes involved and the
large variability of the system.
Most studies emphasize that it is very likely that a thinner and more vulnerable ice will break
up so more heat from the sun will be absorbed in the open water. This can lead to abrupt
melting and a high susceptibility to strong winds when weather conditions are favourable, like
in the summer of 2007. An increased influx of warm Atlantic and Pacific water can also be an
important mechanism for further weakening of the sea ice. Unless followed by consecutive
years of cold winters, such events will produce a thinner, younger and even more vulnerable
ice cover that can melt more easily the next summer, and more easily be transported out of the
Polar Ocean.
Figure 5.3.5.4: Arctic September sea ice extent from observations (thick red line) and 13
IPCC AR4 climate models, together with the mean value (solid black line) and variation
(dotted black line) from several models run, each run for several scenarios (Stroeve et al
2007).
78
5.3.6. Mountain Permafrost
Key Messages
• A long-term regional warming of mountain permafrost of 0.5-1.0°C has been observed
during the recent decades.
• Present and projected atmospheric warming will lead to wide-spread thaw of mountain
permafrost.
• Warming and melting of permafrost is expected to contribute to increasing the
destabilization of mountain rock-walls, the frequency of rock fall, debris flow activity and
geotechnical and maintenance problems in high-mountain infrastructure.
Presentation of the main indicators
Figure 5.3.6.1: Temperature in a mountain range Figure 5.3.6.2: Temperatures measured in
containing permafrost
different boreholes at ca. 10.0 m depth
(blue colours bordered by black line)
(drilled in rock-glaciers and frozen rock-walls)
Source: Gruber (2007)
Source: PERMOS (2007)
[http://maps.grida.no/go/graphic/mountain-permafrost-patterns-and-temperature-gradients]
Relevance
Permafrost is permanently frozen ground and consists of rock or soil material that has
remained at or below 0°C continuously for more than 2 years.
Mountain permafrost is abundant at high elevations in mid-latitude mountains, where the
annual mean temperature is below -3°C. It contains variable amounts of ice and exists in
different forms: in steep bedrock, in rock glaciers, in debris deposited by glaciers or in
vegetated soil. Because vegetation and circulating groundwater in the mountain permafrost
area are mostly absent, the temperature in the deeper rock material is largely determined by
the history of temperature at its surface. Therefore mountain permafrost contains valuable
information on climate change. Even if temperature profiles from alpine boreholes are
difficult to interpret in terms of past trends due to the effects of the complex topography
(Fig.5.3.6.1) and the availability of insulating snow-cover (Gruber et al., 2004), monitoring of
temperature changes at depth however provides valuable data on the thermal response of
permafrost to climate change.
Permafrost influences the evolution of mountain landscapes and affects human infrastructure
and safety. Permafrost warming or thaw affects the potential for natural hazards, such as rock
falls (e.g. at the Matterhorn in summer 2003) and debris flows (Noetzli et al., 2003; Gruber
and Haeberli, 2007). At least four large events involving rock volumes over 1 million m3 took
place in the Alps during the last decade. Its effects on infrastructure have motivated the
development of technical solutions to improve design lifetime and safety (Philips et al.,
2007).
79
Past trends
Data from a north-south transect of boreholes, 100m or more deep, extending from Svalbard
to the Alps (European PACE-project) indicate a long-term regional warming of permafrost of
0.5-1.0°C during recent decades (Harris et al., 2003). In Scandinavia and Svalbard,
monitoring over 5-7 years show warming down to 60m depth and present warming rates at
the permafrost surface of 0.04-0.07°C/year (Isaksen et al., 2007). In Switzerland, a warming
trend and increased active-layer depths were observed in 2003, but results varied strongly
between borehole locations due to variations in snow cover and ground properties (PERMOS,
2007). At the Murtel-Corvatsch (rock-glacier) borehole in the Swiss Alps, the only long-term
data-record (20 years), permafrost temperatures in 2001, 2003 and 2004 were only slightly
below -1°C (Fig. 5.3.6.2) and were, apart from 1993 and 1994, the highest since
measurements began in 1987 (Vonder Mühll et al., 2007). Such data measured at rockglaciers are difficult to interpret because the subsurface thermodynamics in ice-rich frozen
debris are rather complex. Complementary and clearer signals on thawing permafrost are
expected from boreholes that were directly drilled into bedrock (e.g. Schilthorn and M. Barba
Peider; Fig.:5.3.6.2). Corresponding monitoring programs, such as PACE and PERMOS,
however, have started only less than a decade ago.
.
Projections
No specific projections on the behaviour of mountain permafrost are available yet, but
changes in mountain permafrost are likely to continue in the near future and the majority of
permafrost bodies will experience warming and/or melting. According to recent model
calculations basing on the regional climate model REMO and following the IPCC SRESScenarios A1B, A2 and B1, in the Alpine region a warming up to 4°C until 2100 is projected
(Jacob et al., 2007). The further rising temperatures and melting permafrost could
increasingly destabilize mountain walls and increase the frequency of rock falls, posing
problems to mountain infrastructure and communities (Gruber et al., 2004). The warming and
thaw of bedrock permafrost can sometimes be fast and failure along ice-filled joints can occur
even at temperatures below 0°C (Davies et al., 2001). Water flowing along linear structures
and the advection of heat along joint systems will further accelerate destabilisation (Gruber
and Haeberli, 2007).
Fig.: 5.3.6.4 Rock-Glacier Murtel-Corvatsch
Source: M.Phillips
Fig.: 5.3.6.3 TheMatterhorn
Source: M.Phillips
80
5.4 Marine systems
5.4.1 Introduction
The oceans play a key role in the regulation of climate by transporting heat northward
and by distributing energy from the atmosphere into the deep parts of the ocean. The
Gulf Stream and its extensions, the North Atlantic current and drift, influence
European weather patterns and storm tracks. The heat transported northward by the
oceanic circulation impacts precipitation and wind regimes over Europe. On the other
hand, oceans are also affected by the climatic conditions and the induced changes in
physical conditions affect the marine ecosystems. Reductions in sea ice coverage in
the arctic polar region are probably the most visible indicators of climate change and
the economic consequences of Arctic sea ice disappearance are likely to be of great
importance for Europe (see also Section 5.3).
In the following chapters, changes in sea level and sea surface temperature are
discussed, followed by examples of the chemical consequences (acidification of the
ocean) and the biological consequences (changes in physiology, distribution,
phenology and genetic composition) and the associated impacts on marine life in all
European seas. The examples chosen to document changes in the marine food-web in
this report are all well accepted in the scientific community as examples of a climatic
impact in the marine environment. In general changes related to the physical marine
environment are better documented than chemical or biological changes simply
because observations have been made for longer. For example, systematic
observations of both sea level and sea surface temperature were started around 1880
and are today complemented by observations from space that have high resolution in
time and geographical coverage. The longest available records of plankton are from
the Continuous Plankton Recorder (CPR), a sampler that is towed behind many
different merchant vessels, along fixed shipping routes. Sampling with the CPR was
started in the North Sea in the 1950’ies and today a network covering the entire North
Atlantic has been established. No other plankton time series of equivalent length and
geographical coverage exist for the European Regional Seas.
The primary physical impact of climate change in European regional seas is increased
sea surface temperature. However, due to different geographical constraints, climate
change is expected to impact the physical conditions differently in the European
regional seas, and consequently biological impacts also vary depending on the region
as demonstrated in the following three examples:
North East Atlantic: Projections indicate that mean warming in the ocean will extend
throughout the water column during the course of the 21st century (Meehl et al., 2007).
Surface temperature changes have already resulted in an increased duration of the marine
growing season and in northward movement of marine zooplankton. Some fish species are
shifting their distributions northward in response to increased temperatures.
Baltic Sea: Climate models predict a mean warming of 2-4°C in the sea-surface temperature
in the 21st century, and both increasing run-off and decreasing frequency of Atlantic inflow
patterns which will decrease salinity of the Baltic Sea. Consequently, sea-ice extent is
expected to decrease by 50-80% over the same period (Graham et al. 2006) and stratification
is expected to become stronger increasing the probability of a deficiency of oxygen (hypoxia)
that kills marine life in the region. Changes in stratification are expected to impact
commercially relevant regional cod fisheries because stratification appears to be an important
component for the reproductive success of cod in the Baltic Sea.
81
Mediterranean Sea: Temperature is projected to increase and run-off to the Mediterranean
Sea to decrease. In contrast to the Baltic Sea, the combination of these two effects is not
expected to greatly change stratification conditions because of the compensating effects of
increasing temperature and increasing salinity in the density of sea water. Alien species
invasion and survival in the Mediterranean Sea have been correlated to the general SST
warming trend, causing replacement of local fauna with new species. Such changes not only
impact local ecosystems, but also impact the activities of the international fishing fleet when
commercial species are affected (Marine Board Position Paper, 2007)
Box 5.4.1: Ocean acidification
In addition to increasing atmospheric temperature, greenhouse gasses (specifically
CO2) also affect marine systems more directly. The global ocean is the primary
storage medium for greenhouse gases such as carbon dioxide (CO2) and the amount
stored in the ocean depends on the concentration of CO2 in the atmosphere. CO2 is
soluble in the ocean and during this process, carbon dioxide reacts with water to form
carbonic acid which then dissociate into hydrogen ions (H+), bicarbonate ions (HCO3-) and, to
a lesser extent, carbonate ions (CO32-). The higher the concentration of CO2 is in the
atmosphere, the more CO2 that is dissolved in the ocean, and the more the concentration of
H+ ions will increase which would cause a drop in the pH of sea water, causing it to become
less alkaline (or more acidic). Since the industrial revolution pH has already decreased by 0.1
units and simulations for the next century are predicting further reduction of the pH from 0.3
to 0.5 units, depending on which IPCC scenario is adopted in the calculation (Orr et al., 2005,
Caldeira and Wickett, 2005). The increased dissolved CO2 concentration will lower
saturation levels of the carbonate minerals such as calcite, aragonite, and high-magnesium
calcite, which means that the availability of materials used to form supporting skeletal
structures in many major groups of marine organisms will decrease. The decrease in pH of
oceans is viewed as particularly severe because ocean pH has been relatively stable for the
past 300 million years (Caldeira and Wickett, 2003), it will take a very long time to reverse
the decreasing trend, and it has the potential of fundamentally altering the lowest levels of the
marine food-web with unpredictable consequences for higher trophic levels.
Time
Pre-industrial
Ph
8.2
Ph change
0
Present day (1994)
8.1
-0.1
2050
8.0
- 0.2
2100 (based on IPCC
scenario IS92a, SRES
scenarios)
7.7-7.9
-0.3 – -0.5
Source
Model
(Houghton et al., 1995)
Model
(GLODAP reference year,
Hey et al., )
Model
(Orr et al. 2005)
Models
(Orr et al. 2005,
Caldeira and Wickett, 2005)
Average ocean surface pH values based on Houghton et al., 1995, Orr et al., 2005, and
Caldeira and Wickett, 2005.
Implications for European Seas
In European Seas, the largest effects are expected in the Arctic where a scenario of the
consequences of doubling the atmospheric CO2 concentration suggests the possibility of a
complete undersaturation of first aragonite (by 2100) and later calcite (by 2150-2200) is
expected to become the largest (Orr et al., 2005). Under these conditions, key marine
organisms such as coccolithophores (a diatom), echinoderms (sea urchins), and cold water
corals along the northwest European continental margin will have difficulties to build and
maintain their external structure (Orr et al., 2005). These changes at the bottom of the foodweb may have serious knock-on effects in all European marine ecosystems (Pearson et al.,
1999).
82
5.4.2 Sea level rise (SLR)
Key messages
 Global average sea level rose around 17 cm (1.7mm/year) during the last century. In Europe
sea-level change ranges from -0.3 mm/year to 2.8 mm/year.
Recent results from satellites and tide-gauges indicated an accelerated average rate of global
SLR of about 3.1mm/year in the previous 15 years.
 Projections for the end of the 21st century suggest an additional 60 cm SLR above the
current level, the upper end of the range of estimates suggesting up to 140 cm.
 Sea level rise can cause flooding, coastal erosion and the loss of flat and low-lying coastal
regions. It increases the likelihood of storm surges, enforces landward intrusion of salt water
and endangers coastal ecosystems and wetlands. An additional 1.6 million people living in
Europe’s coastal zones might experience coastal flooding by 2080.
Presentation of the main indicator:
Figure 5.4.2.1: Sea level change at different European
tide-gauge stations from 1896 to 2004
(post-glacial processes considered).
Source: Novotny et al., 2007.
83
Figure 5.4.2.2: Trends in sea level 1870-2006
Source: Church, J.A. and White, N.J. (2006).
http://maps.grida.no/go/graphic/trends-in-sea-level1870-2006
Figure 5.4.2.3: Map of sea level variation trends in Europe (1992-2007), based on satellite
altimeter data (Source: CLS; INGV; 2007)
Relevance
Sea level rise (SLR) results from thermal expansion of the oceans (the increase in volume due
to rising ocean water temperature) and from inflow of melt-water from glaciers and ice-sheets
(in particular the Greenland and West-Antarctic ice sheets) due to increasing air temperature.
Thus sea level rise is an important indicator of climate change, with great relevance in Europe
for flooding, coastal erosion and the loss of flat coastal regions. Rising sea level increases the
likelihood of storm surges, enforces landward intrusion of salt water and endangers coastal
ecosystems and wetlands. Coastal areas in Europe often feature important natural ecosystems,
productive economic sectors, and major urban centres. A higher flood risk increases the threat
of loss of life and property as well as of damage to sea-dikes and infrastructure, and might
lead to an increased loss of tourism, recreation and transportation functions (Nicholls et al.,
2006; Nicholls et al. 2007; Devoy, 2007). Low-lying coastlines with high population densities
and small tidal ranges will be most vulnerable to sea-level rise (Kundzewicz, 2001). Thus
coastal flooding related to sea-level rise could affect a large population (Arnell, 2004,
Nicholls, 2004). Because of the inertia of the climate system, climate change mitigation will
not reduce these risks over the coming decades to any significant degree, but various options
for adaptation exist.
Furthermore, it is important to note that some models may stop at 2100, but SLR will not.
Even if atmospheric greenhouse gas concentrations would be stabilized, sea level will
continue to rise for some hundreds of years due to the thermal expansion of the oceans and to
the contribution of the melting Greenland and West-Antarctic ice sheets, which have the
potential to contribute 7m sea level rise respectively. After 500 years, sea level rise from the
thermal expansion of oceans may have reached only half its attainable level and the ice sheets
will continue to melt. The global ice sheets have the potential to make the largest contribution
to SLR, but they are also the greatest source of uncertainty.
84
Past trends
Tide-gauge based data as e.g. from the Permanent Service for Mean Sea Level (PSMSL)
show that long-term average sea level at European coasts has been changing regional
differently at a rate of -0.3 mm/year up to 2.8 mm/year during the previous century (Fig.:
5.4.2.1). Global sea level rose in this time by an average of 1.7 mm/year (Church et al., 2006).
Previous satellite and tide-gauge data-sets indicate an accelerated global trend in sea-level rise
by about 3.1 mm/year (Fig.: 5.4.2.2) in the last 15 years (Nerem et al., 2006; Church et al.,
2006; Rahmstorf et al., 2007). It is very likely that the observed trend in sea level rise over the
past 100 years is mainly attributable to an increase in the volume of ocean water as a
consequence of temperature rise, although inflow of water from melting glaciers and icesheets plays an increasing role (Tab.: 5.4.2.2).
Satellite observations indicate a large spatial variability of SLR trends in the European seas
(Fig.: 5.4.2.2; Tab.: 5.4.2.1). For instance in the Mediterranean Sea positive trends can be
observed in the Levantine Sea while negative trends can be observed in the northern Ionian
Sea . These local variations could be explained by variability of the North Atlantic Oscillation
(NAO), by inter-annual wind variability, by changes in global ocean circulation patterns, and
by specific local structures of the circulation (e.g. gyres) (Demirov and Pinardi, 2002).
Table 5.4.2.1: Average in sea level rise in some European Seas (satellite observations; time
period 1993-2006) (Source: CLS; INGV; 2007)
European seas
Sea level rise
[mm/year]
North Atlantic (50°N to 70°N)
3.4
Central North Atlantic (30°N to 50°N)
1.15
Mediterranean Sea
1.5
Black Sea
7.5
Table 5.4.2.2: Contribution of different processes to global sea level rise (time period 19932006)
(Source: IPCC, 2007)
Process
Contribution to global sea level rise
[mm/year]
Ocean thermal expansion
1.6 ± 0.5
Melting of glaciers and ice caps
0.8±0.2
Melting of the Greenland Ice Sheet
0.2±0.1
Melting of the Antarctic Ice sheet
0.2±0.4
Unaccounted for
0.3
Total global sea level rise
3.1± 0.4
Projections (future trends)
According to the IPCC, it is likely that the observed trend of increased high sea levels will
continue globally at least during the 21st century. Model projections based on the IPCCSRES scenarios give a global mean SLR of 0.18 -0.59 m by 2100, with rates up to 3 times
higher than those at the previous decade (Fig.:5.4.2.3; IPCC, 2007).
85
Figure 5.4.2.4: Projected global average sea level rise (m) from 1990 to 2100 for the six
SRES scenarios; Source: UNEP, 2007
In Europe, regional influences may result in SLR being up to 50% higher than these global
estimates (Woodworth et al., 2005), where the impact of the NAO on winter sea levels adds
an uncertainty of 0.1-0.2 m to these estimates (Hulme et al., 2002; Tsimplis et al. 2004). SLR
projections for the Baltic and Arctic coasts based on SRES scenarios indicate an increased
risk of flooding and coastal erosion after 2050 but always lower than the risk in the North Sea
and in the Mediterranean (Johansson et al., 2004; Meier et al. 2004, 2006, Nicholls, 2004).
The A1F1-scenario, which assumes a very high greenhouse gas emission from fossil fuel
combustion, would lead to a greater impact of SLR in the Northern Mediterranean, as well as
in Northern and Western Europe. Where it was highly unlikely that the populations in those
coastal areas would experience flooding in 1990, up to 1.6 million people might experience
coastal flooding each year by 2080 (Nicholls, 2004).
Various adaptation options are available to reduce such risks. But there are limits to
adaptation: Due to thermal inertia of the oceans sea level rise does not stop by 2100, not even
if greenhouse gas concentrations would be stabilized. A very large SLR would result from the
melting of the world’s major ice sheets at Greenland and the West Antarctic Peninsula, which
have a SLR potential of about 13 m (UNEP, 2007).
Box: Long-term sea-level rise: insights since IPCC AR4
Observed sea level from 1990 onwards is near the upper end of the range projected in IPCC’s
Third Assessment Report (Fig.: 5.4.2.4, Rahmstorf et al., 2007),. This indicates that one or
more of the model contributions to SLR is underestimated and implies that SLR may be
higher than projected by the IPCC: using a statistical model, a SLR of up to 1.4 m was
recently estimated (Rahmstorf, 2007). In addition to the large uncertainty about the behaviour
of the world’s major ice sheets, also ocean dynamics and the effect of gravity changes
induced by melting of land based ice-masses (as e.g. the Greenland ice-sheet) can have a large
effect (Katsman et al., 2007). A melting Greenland ice-sheet would result in a weakening of
SLR in the Northern latitudes and in a strengthening of SLR in the tropics in the longer term.
86
5.4.3 Sea Surface Temperature
Key Messages



Sea surface temperature (SST) is increasing at a faster rate in European seas
than in the global oceans. The rate of increase is larger in the northern
European seas and relatively smaller in the Mediterranean Sea.
In the past 25 years sea surface temperature in all European seas has been
increasing at a rate that is roughly 10 times faster than the average over more
than a century (high confidence).
The trend observed in the last 25 years is the largest ever measured in any
previous 25 year period.
Presentation of the main indicator
Table 1: Summary of historical and recent sea surface temperature trends in the Global Ocean
and the four European Regional Seas.
Sea
1871-2006 annual rate
1982-2006 annual rate
(past 136 years) [°C/yr]
(past 25 years) [°C/yr]
Global Ocean
0.004
0.01
North Atlantic Ocean
0.002
0.03
Baltic Sea
0.006
0.06
North Sea
0.004
0.05
Mediterranean Sea
0.004
0.03
Black Sea
0.003
0.03
Figure 5.4.3.1: Annual average SST difference from the 1982-2006 average in different seas.
Panel a: Global Ocean (red), North Atlantic (yellow), North Sea (Cyan), Baltic Sea (magenta).
Panel b: Global Ocean (red), Black Sea (black), Mediterranean Sea (blue). Source: Appendix
1 and INGV, 2007
87
Figure 5.4.3.2. Spatial distribution of linear trend of the last 25 years (1982-2006) for the
European Seas as calculated from HADISST1 dataset. The units are oC/yr. Source: INGV,
2007.
Relevance
Changes in sea surface temperatures of the global ocean and the regional seas of Europe are
consistent with the changes in atmospheric temperature (Levitus, 2000, Rayner, 2006). The
increased surface temperature of the North Atlantic has impacts upon the atmospheric
circulation, and thus impacts climate in Western Europe. There is also an accumulating body
of evidence suggesting that many marine ecosystems, both physically and biologically are
responding to changes in regional climate caused predominantly by the warming of air and
SST (Halpern et al., 2008).
Variations in SST has for example been linked to the NAO-index, (see Section 5.2 in this
report) (Frankignoul and Kestenare, 2005). The general atmospheric warming is likely to
influence the Atlantic Ocean circulation system that carries warm upper waters north and
returns cold deep waters south also known as the Atlantic Meridional Overturning Circulation
(Marshall et al., 2001) because it will cause increased freshening and warming in the subpolar
seas. Observations indicate that there has indeed been a freshening of the North Atlantic
(Curry and Mauritzen, 2005) and thus possibly a weakening of the Atlantic Meridional
Overturning Circulation, although no observations are available to document this.
88
Past trends
In the European regional seas the SST changes are stronger than in the global oceans (Table
1). The strongest trend in the last 25 years is in the Baltic Sea and the North Sea, while the
rates are lower in the Black Sea and Mediterranean Sea. The regional seas experienced
warming trends that are up to six times larger than the in global oceans in the past 25 years.
These changes have not been observed in any other 25-year period since systematic
observations started more than a century ago (Fig. 5.4.3.1).
The spatial distribution of trend over the European seas is shown in Fig. 5.4.3.2. It shows that
the positive temperature trend is more pronounced in the North Sea, Baltic Sea, the area south
of the Denmark Strait, the eastern part of the Mediterranean, and the Black Sea. Absolute
maxima are located in the North Atlantic around 50° N, in the North Sea and Baltic Sea with
values over 0.06-0.07 °C/yr. Negative trends are detected in the Greenland Sea. Here, the
estimates also depend on the ice extent.
Projections
IPCC (2007) reports global-scale SST patterns for the SRES-A1B scenario for the time
periods 2011-2030, 2046-2065, and 2080-2099. In these scenarios, the ocean warming
evolves more slowly than the warming of the atmosphere. Initially ocean warming will be
greatest in the upper 100 meters of the ocean (in the surface mixed layer), but later in the 21stcentury temperature will also increase at a faster rate in the deep ocean (IPCC, 2007;
Watterson, 2003; Stouffer, 2004).
The scenario projects ocean warming to be relatively large in the Arctic and along the equator
in the eastern Pacific, with less warming over the North Atlantic and the Southern Ocean
(e.g., Xu et al., 2005). Enhanced oceanic warming along the equator is also evident, and can
be associated with oceanic heat flux changes (Watterson, 2003) and temperature changes in
the atmosphere (Liu et al., 2005). The projection of SST indicators is not feasible for the
different geographical regions across Europe because the spatial resolution of the coupled
ocean-climate models is not high enough to evaluate trends on the European regional sea
scale
89
5.4.4 Marine Phenology
Key Messages

Many marine organisms in the European seas now appear earlier in their seasonal
cycles than in the past, i.e. they have changed their phenology as a consequence of
increased sea surface temperatures. For example, some species have moved forward
in their seasonal cycle by 4-6 weeks.
 These changes in the timing of seasonal cycles have important consequences foron
the way organisms within an ecosystem interact and ultimately foron the structure
and efficiency of marine food-webs.
 The consequences of changed marine phenology are can be severe for the marine
ecosystem and include with medium confidence:
•
Decoupling of species relationships and changes to food-web structures.
•
To worsen the decline in North Sea cod stocks created by over-fishing.
•
Changes to other fish populations leading to decline in seabird populations.
 Genetic adaptations within species populations may be required to cope with these
changes but these may be hampered because the current pace of climate warming is
too fast for genetic adaptations to take place.
Presentation of the main indicator(s)
Figure 5.4.4.1: a. Year vs. month plot of decapod abundance in the central North Sea
highlighting the mean seasonal peak in abundance. Data collected by the continuous Plankton
Recorder Survey (1950-2005). b. The month of seasonal peak of decapod larvae for each
year 1958-2005 (green line) plotted with sea surface temperature red line). Decapod larvae
data from the Continuous Plankton Recorder survey for the central North Sea. Sea surface
temperature from the Hadley Centre of the UK met office (HadISST).
90
Relevance
One of the key indicators of climate change impacts on biological populations is phenology
change. Phenology is the study of annually recurring life-cycle events such as the timing of
migrations and flowering of plants. In the marine environment such phenology indicators
would include the timing of the spring phytoplankton bloom and the peak in the abundance of
other marine organisms such as the earlier appearance of dinoflagellates associated with
summer stratified conditions.
One of the ways synchrony between predator and prey can be measured is by studying
phenology changes. In the North Sea, many species are moving forward in the season (i.e.
they are appearing earlier from their normal seasonal cycles) while other species are not
moving forward. This has led to a decoupling of species relationships and changes to foodweb structures (Edwards & Richardson, 2004). These changes in plankton (apart from over
fishing) have been strongly implicated in exacerbating the decline in North Sea cod stocks
(Beaugrand et al. 2003) and changes to other fish populations leading to changes in seabirds
(Frederiksen et al. 2006).
The southern North Sea has been identified as being particular vulnerable to phenology
changes (Edwards, Woo & Richardson, in prep). Phenology changes have been related to the
degree of and speed of regional climate change. For example, the southern North Sea is
warming faster than other regions in the North East Atlantic and here is where phenological
movement has been found to be much more pronounced. It is uncertain to what extent species
will adapt to the changes.
Past trends
In the North Sea, work on pelagic phenology has shown that plankton communities, including
fish larvae, are very sensitive to regional climate warming with the response to warming
varying between trophic levels and functional groups. The ability and speed with which fish
and planktonic communities adapt to regional climate warming is not yet known. In other
European regional areas, however, long-term data on marine phenology changes are quite
sparse. According to some preliminary studies, there has also been some phenological
movement in certain copepod species in the Mediterranean Sea over the last decade (JuanCarlos Molinero pers. comm).
Due to the sensitivity of the physiological development of plankton to temperature, decapod
larvae (Lindley, 1987) were selected as a representative of phenological changes in shelf sea
environments. The zooplankton growing season indicator shows the annual peak seasonal
abundance ‘centre of gravity index’ of decapod larvae from 1958–2005 in the central North
Sea (i.e the peak in seasonal appearance). It is clearly visible that there is a major trend
towards an earlier seasonal peak (Figure 1a). In particular, since 1988, with the exception of
1996 (a negative NAO year), the seasonal development of decapod larvae has occurred much
earlier than the long-term average (baseline mean: 1958–2005) which in the 1990’ies has
been up to 4-5 weeks earlier than the long-term mean. This trend towards an earlier seasonal
appearance of meroplanktonic larvae during the last decade is highly correlated to Sea
Surface Temperature (Figure 1b).
Projections
Projections of how individual species behave to future climate change have not yet been
made. Using niche models and species thermal envelopes in combination with available
climate change scenarios, what may happen to certain species over the coming decades could
be analysed in the future. But even without detailed scenarios being available, it is very likely
that phenological changes will continue to occur as climate warming continues to accelerate.
However, there is a finite window in which species can move their seasonal cycles in
response to climate warming. What is currently much less certain is to what degree genetic
adaptations within species populations can cope with these changes and whether the current
pace of climate warming is too fast for genetic adaptations to take place.
91
Box 5.4.3 Phytoplankton biomass and growing season
Over the past fifteen years, considerable increase in phytoplankton biomass and an extension of
its growing season has occurred in the North Sea and eastern North Atlantic. This change is
closely related to changes in sea surface temperature and the NAO index.
The oceans are thought to absorb one third (approximately 2 Gt C y-1) of anthropogenic
emissions of CO2 because phytoplankton use CO2 for their photosynthesis. These microscopic
algae are responsible for removing carbon dioxide from the atmosphere through photosynthesis
and transferring the carbon to other trophic levels. Phytoplankton are also the lowest trophic level
of the marine food-web and thus, any change has consequences for all other trophic levels (e.g.
zoo-plankton, fish, seabirds) through bottom-up control. Increased sea surface temperature has
been linked to extending growing season in the North Sea (see example below) but because
phytoplankton growth is also regulated by nutrient and light availability, it is an area of active
research to identify exactly how climate change will impact phytoplankton growth in other parts
of Europe.
According to PCI time series (Figure 5.4.4.2) there has been a considerable increase in
phytoplankton biomass over the last decade in certain areas of the north-east Atlantic and North
Sea. Particularly, an increase in biomass is observed after the mid-1980s in the North Sea and
west of Ireland (Reid et al., 1998; Edwards et al., 2001). In contrast, a decrease in phytoplankton
biomass was detected in the area north-west of the European Shelf. The mechanisms for this
change remains poorly understood but the different regional responses can be partly explained by
variations in NAO-index and in sea surface temperature. The NAO-index has positive
correlations with SST and phytoplankton biomass in the North and to the west of Ireland, and
negative correlations with SST and phytoplankton biomass north-west of the European Shelf
(Edwards
et
al.,2001).
Figure 5.4.4.2: Change in color index since the 1950’ies. A. Year vs. month plot of color-index
change in the southern North Sea. B. Geographical distribution of changes in average annual
color-index from the 1950’ies until 2000s. Both figures are based on data from the Continuous
Plankton Recorder.
92
5.4.5 Northward movement of marine species
Key Messages



Increase in regional sea temperatures has triggered a major northward movement of
warmer water plankton in the north-east Atlantic and a similar retreat of colder water
plankton to the north. This northerly movement is about10° latitude (1100 km) over
the past 40 years, and has appeared to have accelerated since 2000.This will have an
impact on distribution of fish in that region.
Many species of fish and plankton have shifted their distributions northward. Subtropical species occur with increasing frequency in European waters and sub-arctic
species are receding northwards. The rate of north-ward movement of a particular
species, the sailfin dory, has been estimated at about 50 km/year.
Changes in the geographical distribution of some species of fish have been observed
and they may affect the management of fisheries. Fisheries regulations in the EU
include allocations of quotas based on historic catch patterns, and these may need to
be revised.
Presentation of the main indicator(s)
Figure 5.4.5.1: Northward movement of zooplankton based on CPR between two time periods
(1958-1981) and (1982-1999). The northward movement of the warm-water species (warmtemperate pseudo-oceanic species) is associated with a decrease in the mean number of cold water
species (subarctic species). Note also the rapid northward movement along the continental shelf
edge (approx. 1000 km northward movement).
93
Figure 5.4.5.2: Map showing recordings of tropical fish and warm-water copepods in different
years. (This figure is a template/placeholder only- a more readable version will be compiled)
Relevance
Many species of fish and plankton have shifted their distribution northward and sub-tropical
species occur with increasing frequency in European waters, changing the composition of
local and regional marine ecosystems in a major way (Brander et al., 2003; Beare et al., 2004;
Beare et al., 2005; Perry et al., 2005; Stebbing et al., 2002).
The kinds of fish which are available for human consumption are not necessarily affected by
the distribution changes shown above, because fish are often transported long distances from
where they are caught to where they are marketed. However, people eating locally caught
fish may notice changes in the species they catch or buy. Changes in distribution may affect
the management of fisheries. Fisheries regulations in the EU include allocations of quotas
based on historic catch patterns, and these may be need to be revised.
94
In a few situations e.g. early retreat of sea ice in Arctic areas the effect of climate change may
be to increase fish catches (ACIA, 2005), but in general it is not possible to predict whether
northward shifts in distribution will have a positive or a negative effect on total fisheries
production (Brander, 2007).
Past trends
The increase in regional sea temperatures has triggered a major re-organisation in
zooplankton species composition and biodiversity over the whole North Atlantic basin
(Beaugrand et al. 2002), shown in Figure 5.4.5.1. During the last 40 years there has been a
northerly movement of warmer water plankton by 10° latitude in the north-east Atlantic and a
similar retreat of colder water plankton to the north. This northerly movement has continued
over the last few years and has appeared to have accelerated since 2000. The movement is
particularly pronounced along the European continental shelf edge and has been associated
with the Shelf Edge Current running north. Also the marine environment is generally thought
to have much fewer barriers to dispersal than terrestrial systems.
Some clear, well documented examples of fish species shifts are shown in Figure 5.4.6.2.
The sailfin dory (Zenopsis conchifer) was first recorded in European waters off the coast of
Portugal at 38oN in 1966 and has since been recorded progressively further north, to north of
55oN by the early 1990’s (Quero et al., 1998). It is probably transported northward in the
continental slope current and the rate of northward shift in distribution for this species is over
50 km per year. Other species which have become much more common further north, such as
sea bass (Dicentrarchus labrax), red mullet (Mullus surmulletus) and European anchovy
(Engraulis encrasicolus) are probably now able to overwinter and establish breeding
populations Brander et al., 2003).
The ratio of catches of two common Pleuronectiform flatfish species – European plaice
(Pleuronectes platessa) and Common sole (Solea solea) can be used as an indicator showing
the increase in relative abundance of a warm-water vs. a cold-water species of flatfish (Figure
5.4.6.3).This is a change in their distribution, as sole and other warm-water species become
relatively more abundant in northerly areas, while plaice and other cold-water species become
rare in southerly areas (Brander et al. 2003). These changes are due to changes in birth and
death rates. Some species of fish may also be migrating northward, or into deeper (cooler)
water, as temperatures rise. Climate is only one of many factors which affect distribution and
abundance, however the consistency of the response of this particular indicator to
temperature, both within particular areas (i.e. time trend) and across all areas (i.e. geographic
trend) suggest that the causal relationship is quite strong.
Other factors affecting abundance and distribution include fishing, biological interactions,
salinity, oxygen, the North Atlantic Oscillation (NAO) and pollution. In some cases changes
in distribution are probably due to geographic patterns of fishing and not to climate effects.
However, an index based on ratios of catches minimises the influence of fishing when fishing
acts on both species in a similar way, as is the case with these flatfish, which are caught in the
same kinds of gear and often in the same fishing operations.
Projections
Scenario projections of future movements of marine species have not yet been made. The
principal uncertainty in making projections of fish distribution changes over the next 20-50
years arise from the projections of ocean climate. The rate of warming of SST in European
waters shown in the Figure (Section 5.4.3) is more than twice the global rate. Part of the
reason for this may be the effect of the Atlantic Multidecadal Oscillation (AMO) which is
superimposed on the global trend. If this multidecadal component has been responsible for
the high regional rate of warming and if it moves into a declining phase, then the temperature
in European waters may change at a much slower rate over the next decade or two, before
accelerating again. This will affect the northward movement of marine species.
95
Box 5.3.4 Past and future changes in fish and plankton distribution and abundance in
the Baltic
The Baltic aquatic ecosystems are species-poor and are balanced between predominantly
marine species in the western parts near the Skaggerak and mainly brackish and freshwater
species in the northern and eastern parts. Quite a small reduction in the salinity, driven by
climate induced changes in precipitation and inflow from the North Sea already appears to tip
the balance and change the composition of the Baltic biota.
The effects of climate on distribution of fish and plankton in the Baltic are thus due mainly to
changes in salinity and also oxygen, rather than temperature. They do not fit into the general
pattern of northward shift due to temperature, although temperature rise will interact with
oxygen limitation by increasing the metabolic demand for oxygen. Salinity in the Baltic has
decreased steadily since the mid 1980’s due to increased freshwater input (precipitation) and a
reduction in the frequency of inflow events through the Kattegat, which bring in more saline,
oxygenated water. Of the three major fished species, cod (Gadus morhua), herring (Clupea
harengus) and sprat (Sprattus sprattus), cod is particularly sensitive to reduced salinity - at
levels below 11 the eggs lose their buoyancy and the sperm become inactive. The major
zooplankton prey species for cod lavae Pseudocalanus acuspes also decline when salinities
are low.
Projections for future ocean climate of the Baltic are for continuing increases precipitation
and decreases in inflow from the Skaggerak, therefore the distribution and abundance of cod
and other marine species is likely to continue to diminish. Their position in the ecosystem
may be taken over by more brackish and freshwater species, such as whitefish, pikeperch and
perch (MacKenzie et al., 2007).
Appendix 1
The SST anomalies and linear trends have been calculated for different datasets:
a) HADISST1 dataset (1° horizontal resolution) for the Global Ocean and all the regional seas
from 1871-2006. It is one of the Hadley Centre (UK) global monthly data sets of sea ice
concentration and SST (Rayner et al., 2003) SST anomalies were calculated with respect to
1982-2006 mean;
b) MOON SST dataset (1/16 ° horizontal resolution) for the Mediterranean Sea from 19852006. The MOON dataset is based upon the longest time series of infrared satellite
observations (AVHRR) starting from the beginning of the 80’s. SST anomalies were
calculated subtracting the 1985-2006 MOON dataset mean;
c) BSH SST for the Baltic Sea (BSH-Baltic-Sea, 20 Km horizontal resolution) from 19902006 derived from BSH’s weekly SST analyses (Loewe et al. 2006). SST anomalies were
calculated with respect to the 1990-2006 mean;
d) BSH SST for the North Sea (BSH-North-Sea 20 Km horizontal resolution) produced with
the satellite infrared observations (AVHRR) for the period 1990-2006. SST anomalies were
calculated with respect to the 1982-2006 mean.
HADISST1 dataset has been used for the global ocean and for the centennial trend estimates
(1871-2006). The regional datasets where used for the corresponding regional seas when
available because of their higher horizontal resolution.
96
5.5 Terrestrial ecosystems, biodiversity
5.5.1 Introduction
Climate (change) is an important driving force for the functioning and distribution of
natural systems (Parmesan and Yohe, 2003). Europe’s biodiversity has adapted
repeatedly to glacial and inter-glacial periods in the past, with many species
recolonising the continent from ancient refugia. However, Europe’s landscapes are
now so profoundly modified and fragmented by human activity that, during the
current period of more rapid climate change, species’ movement is being severely
restricted. Local and regional extinctions are therefore likely (McKinney and
Lockwood, 1999). Species at greatest threat are specialists, species at the top of the
food chain, species that are attitudinally restricted, and species with northerly
distributions that are at the southern edge of their ranges; those with poor dispersal
abilities are especially vulnerable.
Biodiversity contributes to ecosystem functioning and the maintenance of ecosystem
services such as nutrient cycling, carbon storage, regulation of hydrology and climate,
pollination, natural pest regulation, and quality of life (IPCC, 2007; Diaz et al., 2006).
Thus impoverishment of Europe’s flora and fauna affects the delivery of ecosystems
services with potentially catastrophic consequences (Lovejoy and Hannah, 2005).
Approximately 60% of the world’s ecosystems that have been evaluated are currently
utilised unsustainably and are showing increasing signs of degradation (Reid et al.,
2005). This alone is causing widespread biodiversity loss. It is likely that this loss will
be accelerated by climate change. Between one fifth and one third of species in
Europe are at increased risk of extinction if global mean temperatures rise more than 2
to 3ºC above pre-industrial levels (Lovejoy and Hannah, 2005; IPCC, 2007); the
ability of many species to adapt naturally (their resilience) will be exceeded by a
combination of climate change and other drivers of change, particularly land-use
change and over-exploitation of natural habitats. Therefore, reducing these other
stresses can enhance the adaptive capacity of ecosystems to climate change (IPCC,
2007). Furthermore, new areas for conservation are needed, together with measures to
facilitate species movement in fragmented landscapes. As such, efforts are needed to
boost the robustness of the European ecological network of Natura 2000 sites,
including more widespread implementation of Article 10 of the Habitats Directive
(which relates to the network’s coherence).
This subchapter outlines the impacts of climate change on biodiversity by showing
both observed and projected changes in the distribution and phenology of plants and
animals, and the implications for communities.
97
5.5.2 Distribution of plant species
Key Messages
•
•
•
Climate change, in particular milder winters, is responsible for observed
northward and uphill distribution shifts of many European plant species.
Mountain ecosystems in many parts of Europe are changing as cold adapted
species are driven out of their ranges and pioneer species expand uphill.
By the late 21st Century, distributions of European plant species are projected
to have shifted several hundred kilometres to the north and 60% of mountain
plant species may face extinction (high confidence). The rate of change will
exceed the ability of many species to adapt.
Presentation of the main indicator
Figure 5.5.2.1. Change in species richness on Swiss Alpine mountain summits. As
pioneer species expand uphill, cold-adapted species are being driven out of their
distribution ranges.
Source: Walther, Beissner and Burga, 2005
Caption: Legend: × = 1900s; ◆= 1980s; ● = 2003; open symbols (P. Trovat, P.
Languard) indicate a (temporary) decrease in species number.
98
Figure 5.5.2.2. Expansion of holly (Ilex aquifolium)- a climate limited species. Maps
show the 0° C January isoline pre-1970 and updated for 1981-2000.
Source: Walther, Berger and Sykes, 2005
(a) Former range of I. aquifolium (dark grey shading) based on Enquist (1924) and
Meusel et al. (1965); isoline based on Walter & Straka (1970); symbols based on
Iversen (1944).
(b) Modelled range of I. aquifolium in the recent past (1931–60: vertical shading);
isoline as in (a).
(c) Former range of I. aquifolium (dark grey shading) as in (a); isoline updated for
1981–2000 based on Mitchell et al. (2004).
(d) Former range of I. aquifolium (dark grey shading) complemented with the
simulated species’ distribution under a moderate climate change based on 1981–2000
climate data (diagonal shading); isoline as in (c); triangles represent locations with
new actual occurrences of I. aquifolium. (NB symbol legend taken from Figure 2 in
same publication).
99
Relevance
The rate of climate change in Europe is likely to exceed the adaptive capacity of some
wild plant species (IPCC, 2007), whilst others are expected to benefit from changing
environmental conditions (e.g. Sobrino Vesperinas et al., 2001). Consequently,
completely new assemblages of plants are appearing and the ecological implications
of these, including the emergence of invasive non-native species, are unpredictable.
The threat of extinction of species at the edge of their geographical ranges –
particularly poorly dispersing endemics – will be greatly increased, with challenging
consequences for their long-term conservation (Gitay et al., 2002). Furthermore,
changes in plant composition within ecosystems can affect ecosystem services such as
agricultural and forest production, nature conservation, water purification and climate
regulation.
The adaptive capacity of species is linked to genetic diversity. In some parts of
Europe, including the Iberian Peninsula, Italy and the Balkans, the genetic diversity of
plants is relatively high, as many species have persisted in these regions since glacial
times. Reducing the impacts of climate change and other anthropogenic pressures
will be particularly important in conserving these sensitive and valuable relic
populations.
Past trends
Warmer temperatures in the last 30 years have significantly influenced seasonal
patterns. As evidenced during glacial and inter-glacial periods, the predominant
adaptive response of climate-limited plant species has been to shift distribution,
resulting in altitudinal and poleward movements. One such climate limited species is
holly (Ilex aquifolium), which has expanded in southern Scandinavia in a manner
consistent and synchronous with recorded regional climate changes, linked in
particular with increasing winter temperatures (Walther, Berger and Sykes, 2005).
There has also been a general increase in mountain summit vegetation during the 20th
Century (Figure 5.5.2.1). The uphill shift of Alpine plants showed an accelerating
trend that is likely to be linked with the extraordinarily warm conditions of the 1990s
(Walther, Beissner and Burga, 2005). Evidence also emerged of warming-induced
declines as cold-adapted species are driven out of their distribution ranges by the
uphill expansion of pioneer species. In effect, Europe’s summit region is shrinking as
characteristic vegetation is displaced and as tree lines rise. In the Swedish Scandes,
for example, the tree line of the Scots pine (Pinus sylvestris) rose by 150-200 metres
during the past century as warmer winters significantly lowered mortality and
increased rates of establishment. Observations from other continents suggest that
uphill tree line migration is a global phenomenon (Kullman, 2006, 2007). This could
become the major threat to biodiversity in high mountains (Pauli et al., 2007).
Projections
Projections show that, by the late 21st Century, the potential range of European plant
species may shift several hundred kilometres in a northerly direction. This is several
times faster than past rates as estimated from the Quaternary record or from historical
data (Huntley, 2007).
Modelling of late 21st Century distributions of 1350 European plant species under a
range of scenarios concluded that more than half could become threatened by 2080,
100
with high risks of extinction. The greatest changes were projected for the
Mediterranean and Euro-Siberian regions and many mountain regions (Thuiller et al.,
2005). Mountain communities may face up to a 60% loss of plant species under high
emission scenarios, as there are no climate adaptation options for the characteristic
vegetation of these regions (IPCC, 2007). Bakkenes et al. (2006) obtained similar
results from modelling stable areas of plant species distribution for 2100 under
different climate change scenarios (Figure 5.5.2.3.). This study suggests that 10-50%
of plant species in European countries are likely to disappear from their current
location in the absence of climate mitigation. Again, species in southeastern and
southwestern Europe are projected to be worst affected. For Europe as a whole, about
1% of plant species may go extinct because they no longer have a suitable climate
niche. This figure will be higher if movement is restricted or if they are out-competed
by invasive species.
Figure 5.5.2.3. Projected changes in plant species (left: number of species considered;
middle: %stable fraction in 2050; right: %stable fraction in 2050).
Source:Based on Bakkenes et al., 2006
Caption: Results for stable area per grid cell, using the EuroMove model with
HadCM2 A2 climate scenario.
101
Figure 5.5.2.4: Total number of species projected to have potentially suitable climate
space within each European region over time, under different scenarios.
Source: Berry et al., 2007
Caption: In the BRANCH project, bioclimatic modelling was applied to 389 species
(both plants and animals) encompassing a range of dominant and threatened taxa
found in Europe. HA2: HadCM3 A2 scenario; HB1: HadCM3 B1 scenario; PCM:
Parallel Climate Model A2 scenario.
102
5.5.3 Plant phenology
Key Messages
•
•
•
The timing of seasonal events in plants is changing across Europe; 78% of leaf
unfolding and flowering records show advancing trends and only 3% significant
delays. Between 1971 and 2000, the average advance of spring and summer was
2.5 days per decade
As a consequence of climate-induced changes in plant phenology, the pollen
season starts on average 10 days earlier and is longer than 50 years ago.
Advancing trends in seasonal events are set to continue as climate warming
increases in the years and decades to come.
Presentation of the main indicator
Figure 5.5.3.1. Phenological sensitivity to temperature changes. In a study of 254
national records across nine countries, most phenological changes correlated
significantly with mean monthly temperatures of the previous two months. The earlier
a spring event occurred, the stronger the effect of temperature.
Caption: Countries included: Austria, Belarus/northern Russia, Estonia, Czech
Republic, Germany, Poland, Slovenia, Switzerland, Ukraine/ southern Russia.
Phenophase groups included: farmers’ activities (b0), spring and summer with
different leafing, shooting and flowering phases (b1), autumn fruit ripening (b2) and
leaf colouring of deciduous trees in fall (b3). Source: Menzel et al. (2006a)
103
Figure 5.5.3.2. Oak (Quercus sp) leafing date in Surrey (1950-2005).
Source: Nature’s Calendar UK,
http://www.naturescalendar.org/uk/climate+change/past.htm
Relevance
Phenology is the study of changes in the timing of seasonal events, such as flowering,
migration and hibernation. Some phenological responses are triggered principally by
temperature, while others are more responsive to day length (Menzel et al., 2006a,
figure 5.5.3.1). Changes in phenology are linked with the growing season and affect
ecosystem functioning and productivity. Farming, gardening and forestry, as well as
wildlife, are affected: the timing of tilling, sowing and harvesting require adjustments,
and fruit is ripening earlier due to warmer summers (Menzel et al., 2006).
Changes in flowering have implications for the timing and intensity of the pollen
season; this is showing an advancing trend as many species start to flower earlier.
Allied to this, the concentration of pollen in the air is increasing (Nordic Council,
2005).
Past trends
There is a clear signal across Europe of changing phenology over the past decades
(Parmesan & Yohe, 2003; Root et al., 2003; Menzel et al., 2006a), although early
phenological events (i.e. spring) are more variable than later events (Menzel et al.,
2006b). For example:
 78% of all leaf unfolding, flowering and fruiting records across Europe show
an advancing trend and only 3% significant delays. The average advance of
spring/summer phenological events is occurring at a rate of 2.5 days per
decade (Menzel et al., 2006a).
 The pollen season starts on average 10 days earlier and is of longer duration
than 50 years ago.
 In Britain, the first flowering date for 385 plant species has advanced by 4.5
days on average during the past decade in comparison with the previous four
decades (Fitter and Fitter, 2002); oak leafing has advanced three weeks in the
last 50 years (DEFRA, 2007) (Figure 5.5.3.2).
 In the Arctic environment, rapid climate-induced advancement of spring
phenomena (e.g. flowering, egg laying) has been observed during the last 10
104
years. The strong responses of Arctic ecosystems and large variability within
species illustrate how easily biological interactions can be disrupted by climate
change (Høye et al., 2007).
Projections
Phenological changes will alter growing seasons, ecosystem production, populationlevel interactions and community dynamics (Fitter and Fitter, 2002). Different species
show different phenological responses; for example, annuals and insect-pollinated
species are more likely to flower early than perennials and wind-pollinated species
(Fitter and Fitter, 2002). Ecological research is investigating these different response
thresholds to better understand what the wider effects might be.
105
5.5.4 Distribution of animal species
Key Messages
•
•
•
•
A combination of the rate of climate change and the obstacles to adaptive
movement may lead to a progressive decline in biodiversity in Europe. This
has challenging consequences for current conservation practices and effective
responses are likely to be costly.
Europe’s fauna is moving northwards and uphill in response to climate
change. Next to a farmland effect, there is a clear and strong signal of climatic
change on European bird populations.
Observed rates of distribution change are not keeping pace with changing
climate. In Britain, for example, range boundary adjustments of butterflies are
lagging behind temperature changes.
Suitable climatic conditions for Europe’s breeding birds are projected to shift
nearly 550 km northeast by the end of the century, with the average range size
shrinking by 20 %. Projections analysing 120 native European mammals
suggest that up to 9% (assuming no migration) risk extinction during the 21st
Century.
Presentation of the main indicator(s)
Figure 5.5.4.1. Latitudinal shifts in northern range margins in Britain for 16
taxonomic groups of animal species over the past 40 years.
Source: Hickling et al., 2006
Caption: Results are given for three levels of data sub-sampling (recorded, blue;
well-recorded, yellow; heavily recorded, red). Only species occupying more than
twenty 10km grid squares were included in the analysis.
Relevance
The northward shift in distribution of animal species has a range of potential
consequences for agriculture and health, as well as biodiversity and the conservation
106
of resident species (Sparks et al., 2007). Warmer conditions, particularly warmers
winters, are allowing the establishment of new pest species such as the European corn
borer (Ostrinia nubilalis), American bollworm (Heliothis armigera), gypsy moth
(Lymantria dispar) and some migratory moths and butterflies (Miroslav et al., 2005).
Health risks associated with vector-borne diseases are linked to the invasions of
species such as ticks and mosquitoes (Box 1, section 5.9).
Box 1: the spread of the Asian tiger mosquito (Aedes albopictus)
The Asian tiger mosquito (Aedes albopictus) is a vector for 22 viruses, including West Nile,
dengue and yellow fever. In August 2007, the mosquito was implicated in the first European
outbreak of chikungunya, a relative of dengue fever normally found around the Indian Ocean.
Over 100 residents of the Italian region of Castiglione di Cervia were affected by symptoms
of high fever, exhaustion and excruciating bone pain that were eventually traced to this
tropical disease.
Twenty years ago, A. albopictus was unknown in Europe, mainland Africa and the Americas,
but it has since extended its geographic range extensively and rapidly. The first report of the
mosquito in Europe was from Albania in 1979, followed by Italy in 1990. It has subsequently
spread to the Americas, some Pacific islands, Australia, Africa and other parts of Europe.
This was primarily through the international trade in new and used tyres that contained eggs
and/or larvae of the mosquito. However, climate change is also a factor: studies of cold
hardiness and distributional data from Asia and North America indicate that the natural
northerly limits of the species are set by the -5°C cold month isotherm. As Europe’s winters
become warmer, Britain, Ireland and the Atlantic and Channel coasts of France will become
susceptible to colonisation by the mosquito.
Sources:
Past
trends
TheRamsdale
northwards
and2000;
uphill
movement of a wide variety of animal species has been
and Snow,
http://www.uel.ac.uk/mosquito/issue4/climatechange.htm
observed over recent decades across Europe and this has been attributed to climate
http://www.ehj-online.com/archive/2000/november/november3.html
change.
However, distribution changes are not only taking place in the context of
climate warming, but are also triggered by land-use and other environmental changes.
Forhttp://www.nytimes.com/2007/12/23/world/europe/23virus.html?pagewanted=1&_r=1&ei=5088&en=e70c
example, three amphibian and reptile species at the northwestern edge of their
b7bda0bd53ca&ex=1357275600&partner=rssnyt&emc=rss
ranges in Britain should have benefited from recent warming. Instead, their
distributions have collapsed southwards due to fragmentation, with each species
surviving in remnant populations restricted to only a small fraction of their former
range.
In a study of 57 non-migratory European butterflies, 36 had shifted their ranges to the
north by 35–240 km and only two species had shifted to the south (Parmesan et al.,
1999). The sooty copper (Heodes tityrus), for example, had spread north from
Catalonia to the Pyrenees and by 2006 had established breeding populations on the
Baltic coast (Parmesan et al., 1999).
A recent study of 122 terrestrial bird species indicated that, from around 1985,
climatic changes have influenced population trends across Europe (Gregory et al.,
2008), with impacts becoming stronger through time. The study shows that 30 species
have generally increased its population because of climate warming, and 92 have
declined.
107
In Britain, out of a total of 329 animal species analysed over the last 25 years, 275
shifted their range margins northwards, 52 shifted southwards, and two did not move
(UKCIP, 2005; Hickling et al., 2006); there was an average northward shift across all
species of 31–60km (Figure 5.5.4.1). Comparable findings were obtained with respect
to elevation shifts. However, many species, including butterflies, are failing to shift
their ranges as quickly as might be expected under the current rate of climate change
(Warren et al., 2001).
In Germany, the scarlet darter dragonfly (Crocothemis erythraea) (c.f. Subchapter
5.8) has spread from south to north, paralleling observed changes in climate. Only a
few decades ago, the species was rare even in southern Germany, but now it is found
in every federal state. Northward expansions of this species and other southern
species have also been recorded in other countries (Ott, 2007). Similarly, the spread of
the comma butterfly in the Netherlands has been linked to recent climate change
patterns (Fig. 5.5.4.2).
The uphill movement of animal species can bring problems for endemic species. For
example, the habitat of 16 mountain-restricted butterflies in Spain has been reduced
by about one third over the last 30 years; lower elevation limits have risen on average
by 212 metres – comparable with a 1.3ºC rise in mean annual temperature (Wilson et
al., 2005).
Figure 5.5.4.2. The recorded occurrence of the comma butterfly (Polygonia c-album)
in the Netherlands from 1975 to 2000.
Photo source: Saxifraga-Frits Bink. Map source: Dutch Butterfly Conservation,
processed by Netherlands Environmental Assessment Agency for the Environmental
Data Compendium
Projections
Projections indicate that the observed northward and uphill movement of many
animal species will continue this century. Widespread species may become less
vulnerable, while threatened endemic species – already under pressure- will be at
greatest risk, although there will be spatial variation (Levinsky et al., 2007).
108
Reptile and amphibian species could, potentially, expand their ranges as northward
warming creates new opportunities for colonisation (IPCC, 2007). However, limited
dispersal ability is very likely to reduce the ranges of many reptiles and amphibians
(Hickling et al., 2006; Araújo et al., 2006). Species are projected to be most
vulnerable in the Iberian Peninsula and parts of Italy (Fig. 5.5.4.3).
A study of 120 native terrestrial mammals projected that species richness is likely to
reduce dramatically this century in the Mediterranean region, but increase towards the
northeast and in mountainous areas such as the Alps and Pyrenees.
Under a 3˚C climate warming scenario (above pre-industrial levels), the ranges of
European breeding birds are projected to shift by the end of the 21st Century by about
550 km to the northeast. The projected average range will be 20% smaller than
current range and range overlap will be around 40%. Arctic, sub-Arctic, and some
Iberian species are projected to suffer the greatest potential range losses (Huntley et
al., 2008).
In polar regions, projected reductions in sea ice will drastically reduce habitat for
polar bears, seals and other ice-dependent species (IPCC, 2007). In addition to climate
change, these top predators are also affected by declining fish stocks.
Figure 5.5.4.3. Potential changes in climate space of reptiles and amphibians in 2050.
Source: Araújo et al., 2006. Maps: CBD and MNP, 2007
Caption: Projected data based on the GLM map using the HadCM3 A2 scenario for the
2050s are compared with the current situation (left: current number of species; right:
percentage of stable species).
109
5.5.5 Animal phenology
Key Messages
•
•
•
•
Climatic warming has caused spring advancement in the life cycles of most
animal groups studied. Observed trends include earlier frog breeding, bird
nesting and the arrival of migrant birds and butterflies.
Breeding seasons are lengthening, allowing extra generations of temperaturesensitive insects such as butterflies and dragonflies to breed during the year.
Advancement is particularly strong and rapid in the high Arctic. The date of
snowmelt has advanced by on average 14.6 days since the mid 1990s, and
strong and rapid advancement of spring events across species has been
documented.
Advancing trends in seasonal events are set to continue as climate warming
increases in the years and decades to come.
Presentation of the main indicator(s)
Figure 5.5.5.1. Changes in egg laying dates of the flycatcher across Europe.
Source: Both et al., (2004)
Caption: Colours show the trend in temperature: blue, trend towards colder springs;
-1
yellow, mild warming (tr
); red, strong warming
-1
). Filled symbols: flycatcher populations with a significant
advancement of laying date; open symbols: no significant laying date trends. Inset:
population response of laying date in relation to temperature, with each symbol
representing one population. Filled symbols: populations for which data were
110
available between 1980/81 and 1999 or later; open symbols: populations that had a
later start in the study.
Relevance
Butterflies, dragonflies and damselflies are particularly sensitive to temperature.
Milder springs are allowing them to take to the wing earlier in the season. Earlier
onset of breeding can allow an extra generation to emerge during the year and
population size can grow as a result. This may pose problems in the case of pests.
Climate change can also cause problems for animal populations. Populations may
crash as vulnerable young emerge too early for their main food source or explode as
insects are released from their normal predation pressure. The effects of such
disruption to ecosystems are also discussed in Section 5.5.6.
Past trends
As spring temperatures increased in Europe over the past 20 years, many organisms
responded by advancing the timing of their growth and reproduction. The egg laying
date for flycatcher populations, for example, advanced significantly during the period
1990-2002 in nine out of 25 populations studied, and 20 out of 25 populations showed
a significant effect of local spring temperature (Figure 5.5.5.1), both strongly driven
by climatic warming (Both et al., 2004). A study in Britain reached a similar
conclusion (Crick and Sparks, 1999); an analysis of 74,258 records for 65 bird species
from 1971 to 1995 showed significant trends towards earlier (8.8 days on average)
laying dates for 20 species (31%), with only one species laying significantly later.
The date of snowmelt in northeast Greenland has advanced by an average of 14.6
days since the mid 1990s, bringing forward the date at which birds lay eggs and plants
flower by on average 14.5 days per decade. The large and rapid phenological
advancements in the high Arctic are a response to warming in the region occurring at
twice the global average (Høye et al., 2007).
Projections
The impacts of future climate on phenological changes are poorly understood, but are
likely to include increasing trophic mismatch and disturbance to ecosystem
functioning. The trend towards warmer springs is likely to continue to induce earlier
breeding and migration activity. Unpredictable cold snaps will cause high mortality
amongst early movers. Meanwhile, species whose life cycles are calibrated according
to day length and which do not respond so readily to changing temperatures will not
be able to exploit earlier spring resources.
111
5.5.6 Impacts on ecosystem functioning
Key Messages
•
•
•
The stability of ecosystem functioning, and therefore of ecosystem services,
will become increasingly prone to unpredictable perturbations. Species are
responding individualistically to climate change, resulting in the formation of
new assemblages and novel interactions.
Established biotic interactions are becoming disrupted, benefiting generalists
at the expense of specialists.
The impacts of climate change on ecosystem functioning will increase the
potential for pest outbreaks and place additional pressures on species of
conservation importance.
Presentation of the main indicator(s)
Figure 5.5.6.1. Current distribution range of the butterfly Titania fritillary (Boloria
titania) and its host plant American bistort (Polygonum bistorta).
Source: Schweiger et al., in press.
Caption: Green: niche space of P. bistorta; yellow: niche space of B. titania; red:
overlap of both, which is the butterfly’s current range.
Relevance
Suitable climate is only one of a number of factors that determine species
distributions. An important constraint to range expansion of many insects is the
presence of their host plant (Schweiger et al., in press). The enhanced flush of spring
activity provides food for animals higher up the food chain. Animals that are able to
breed earlier can exploit these food sources; this is good for generalists, but not for
specialised feeders that depend upon synchrony with their food source. Asynchrony
between food supplies and breeding may result in starvation of young that emerge too
early or too late. This so-called trophic mismatch has been demonstrated in birds
(Both et al., 2006) and is disrupting predator-prey relations amongst other animal
species too, influencing population dynamics and in some cases causing crashes or
explosions in numbers.
Research has established that insect pests are likely to become more abundant as
temperature increases (Cannon, 1998). As the impacts of climate change on
112
ecosystems favour generalists, and as warmer temperatures increase insect survival
and reproduction rates, more frequent, severe and unpredictable pest outbreaks are
likely (McKinney and Lockwood, 1999). In temperate regions, milder winters are
allowing increased rates of winter survival (Bale et al., 2002) and it has been
estimated that, with a 2ºC temperature increase, some insects can undergo one to five
additional life cycles per season (Yamamura and Kiritani, 1998). In agricultural
systems, cold-susceptible species such as the western corn root worm (Diabrotica
virgifera virgifera) are being favoured by milder winters. This species has shown
recent expansion in several European countries, particularly in continuous cornfields.
Past trends
Many butterfly species are moving northward (c.f. 5.5.3), but often with overall
declines in abundance and range size (Warren et al., 2001). An additional constraint
on successful colonization of new habitats is the availability of ecologically linked
species such as host plants. Many butterflies are host-specific, and the changing
ranges of their hosts are an important constraint to their range expansion (Schweiger
et al., in press). Biotic interactions are important factors in explaining the distributions
of species. For example, many parts of Europe are climatically suitable for the
butterfly Titania fritillary (Boloria titania) (Fig. 5.5.6.1). The species may even be
able to migrate quickly in response to climate change. However, an important
constraint to range expansion is the presence of its larval host plant American bistort
(Polygonum bistorta) (Schweiger et al., in press). Likewise, the current distribution of
the clouded Apollo (Parnassius mnemosyne) is not only explained by climate
suitability but also by the presence of its Corydalis host plant (Araujo and Luoto,
2007).
Climate change may have disruptive effects of on plant and animal communities and
their food webs. During 2004 and 2005, major population crashes of North Sea
seabirds were observed. In Shetland, over 1000 guillemot nests and 24,000 nests of
the Arctic tern were almost entirely deserted, and on the nearby island of Foula, the
world’s largest colony of great skuas saw only a few living chicks. The cause of this
significant loss was found to be a drastic reduction in the populations of sandeel, their
principal food source. The disappearance of the sandeel, in turn, was due to the
northward movement of the cold-water plankton on which these fish feed (c.f.
Subchapter 5.4); the plankton’s range had shifted because the waters between Britain
and Scandinavia had become too warm for it to survive there. Since 1984, some
seabird species around Scotland have decreased by 60-70 % (CEH, 2005).
Projections
The response to climate change of the butterfly Titania fritillary (Boloria titania) and
its host plant American bistort (Polygonum bistorta) is likely to lead to a reduction in
range overlap and, thus, an uncertain future for this specialist butterfly. Played out on
a larger scale, these trophic mismatches benefit generalists at the expense of
specialists, putting additional pressures on the capacity of ecosystems to provide
certain services and on species of conservation importance (McKinney and
Lockwood, 1999; Reid et al., 2005; Biesmeijer et al., 2006).
113
Unlimited dispersal of P. bistorta
No dispersal of P. bistorta
a) Moderate change
b) Moderate change
c) High change
d) High change
Figure 5.5.6.2. (Mis)Match of projected distribution space of the butterfly Titania
fritillary (Polygonum bistorta) and its host plant American bistort (Boloria titania) for
moderate (a, b) and high (c, d) climate change scenarios for 2080 under the
assumption of unlimited (a, c) and no (b, d) dispersal of its host plant.
Source: Schweiger et al., in press.
Caption: Green: niche space of P. bistorta; yellow: niche space of B. titania; red:
overlap of both, which is the butterfly’s potential future niche space. Global change
scenarios based on storylines developed within the EU-funded project ALARM
(Settele et al., 2005, Spangenberg 2007, www.alarmproject.net).
114
5.6 Agriculture and forestry
5.6.1 Introduction
The impacts of medium and long-term climate change on agriculture and forestry are often
difficult to analyse separately from non-climate influences related to the management of the
resources (Hafner, 2003). However, there is growing evidence that processes such as changes
in phenology, growing season length and northwards shift of crops and forest species can be
related to climatic change (IPCC-AR4, 2007). There are also increasing impacts due to an
augmented frequency of extreme events, some of which can be attributed to climate change.
Potential positive impacts of climate change on agriculture in general are related to, longer
growing seasons and new cropping opportunities in Northern Europe, more efficient use of
water as well as better fertilization of plants throughout Europe. These possible benefits are
counterbalanced by potentially negative impacts that include increased water demand and
periods of water deficit, loss of soil carbon content, increased pesticide requirements and crop
damages and less cropping opportunities in some regions in Southern Europe (Olesen and
Bindi, 2004; Maracchi et al, 2005, Chmielewski et al., 2004, Menzel, 2003). In general,
changes in atmospheric carbon dioxide levels and increases in temperature are changing the
quality and composition of crops and grasslands and also the range of native/alien pests and
diseases. The latter could affect not only crops but also livestock and ultimately humans.
The link of forestry with climate change is twofold. Forests play a fundamental role in
mitigating climate change because they may act as sinks for carbon dioxide. However, forests
are also very vulnerable to changes in temperature, precipitation and extreme weather events
which might have destructive impacts and reduce the carbon sequestration potential of the
forest. Events such as forest fires have an even more negative effect since destroying the
forest increases the amount of carbon dioxide in the atmosphere. The majority of forests in
central Europe are growing faster than in the past, partly because of regional warming. In
contrast, the extended heat-wave of 2003 caused a significant reduction in biomass production
of forests (Gabon, 2005).
Although the economic impacts of climate change on agriculture and forestry in Europe are
very difficult to be determined due to the effects of policies and market influences and due to
the continuous technological development in farming and silviculture techniques, there is
evidence of wider vulnerability for both sectors (see also Chapter 6). Management actions can
counteract but also exacerbate the effects of climatic changes and will play an important role
as measures for adaptation to climate change.
The indicators included in this section are related to agricultural production, phenology,
forestry growth and distribution, and to observed and projected impacts of forest fires. The
effect of climate change on soil organic content is also covered in a separate section.
Furthermore, all sections have comments on quality and availability of monitoring data.
115
5.6.2 Crop yield variability
Key Messages
• Over the last decades the yields of major agricultural crops, including cereals,
increased across Europe.
• Climate change is responsible for variations of crop suitability and productivity in
Europe.
• Since the beginning of the 21st century, the variability of crop yields increased as a
consequence of the extreme climatic events, e.g. the summer heat of 2003 or the
spring drought of 2007.
• As a consequence of climatic changes those extreme climatic events are projected
to increase in frequency and magnitude and crop yields are expected to become
more variable. Changes in farming practices and land management can act as riskmitigating measures.
Presentation of the main indicator
Figure 5.6.2.1: Modelled suitability change
for grain maize cultivation in the past (1961–
1990) and in the future (2071–2100). 24
scenarios from 6 GCMs for each of the IPCC
SRES A1FI, A2, B1 and B2 emissions
scenarios (see Chapter 4) have been taken into
account. Green areas show the suitable area
without climate change, red depicts the
expansion common under all scenarios and
blue the area where scenario results diverge.
Grey areas are unsuitable under all scenarios.
Source: Olesen et al., 2007
Relevance
Climate change introduces new uncertainties for the future of the agricultural sector.
Climatic conditions are projected to become more erratic with an increase in the
frequency of extreme events (floods, hurricanes, heat waves, severe droughts) (Parry,
2000). Biomass production of plants, and thus crop yields, are fundamentally
determined by climatic conditions, i.e. the stable availability of energy (radiation,
temperature) and water (rain) to support growth. Other environmental and
anthropogenic factors influence crop yields, such as soil fertility, crop varieties and
farming practices. These factors imply that in principle, many adaptation options are
available to adjust agricultural practices to the changing climate, but that opportunities
differ between regions.
Past trends
While the area under arable land decreased for most parts of Western Europe over the
last 40 years crop yields have almost continuously increased (source: Eurostat). This
trend persisted into the 21st century, although crop yield variability increased as a
consequence of several extreme meteorological events in short succession: In 2003 a
late frost followed by a severe drought reduced cereal yields over most parts of
116
Europe. In 2005 a drought severely affected Western Europe (Iberian Peninsula). In
2006 an early drought was followed by extreme rains during summer, resulting in
lower cereal production especially in Eastern Europe. (EC, Mars Bulletins,
http://mars.jrc.it/marsstat/Bulletins/2008.htm) Alexander et al (2006) found a general
increase of the intensity of precipitation events. For the Mediterranean area, where the
climatologic vulnerability is high, several studies found an increasing trend towards
more intense precipitation and a decrease in total precipitation (Alpert et al., 2002;
Maheras et al., 2004, Brunetti et al., 2004). In general, it is difficult to separate the
climate effect from the effect of improved agricultural techniques in the development
of historic crop yield. Also in the future, adaptive management can help reducing the
risks of climate change for agricultural yields, and make better use of opportunities.
Projections
The effect of increasing mean daily temperatures on agricultural yields depends on the
magnitude and geographic extent. The production area of some crops could expand to
northern Europe, e.g for maize. For an increase in mean annual temperature of 2°C,
cereal yields are expected to increase not least due to the fertilisation effect of the
increase in CO2 (Parry et al, 2004). However, an increase in mean annual temperature
of 4°C or more will shorten the crop cycle and the CO2 effect will not compensate for
the resulting yield loss. Crop yields are also at risk from more intensive precipitation
and prolonged periods of drought in particular in areas bordering the Mediterranean
basin.
The figures below show the sensitivity of maize and wheat yields to climate change,
as derived from results of 69 published studies. The studies span a range of
precipitation changes and CO2 concentrations, and vary in how they represent future
changes in climate variability. Responses include cases without adaptation (red dots)
and with adaptation (dark green dots). Adaptation represented in these studies
included changes in planting dates and crop varieties, and shifts from rain-fed to
irrigated conditions.
Figure5.6.2.2: Yield variation due to temperature increase. A small increase of temperature
has a positive impact on cereals yield, while a high increase (3-5°C) has a negative impact.
Lines are best-fit polynomials and are used here to summarize results across studies rather
than as a predictive tool. Source: IPCC Report WGII (2007)
117
5.6.3 Timing of the cycle of agricultural crops
(agrophenology)
Key Message
•
•
There is evidence that flowering and maturity of several species in Europe
now occur two or three weeks earlier than in the past.
The shortening of the phenological phases are expected to continue if the
temperature will keep increasing.
Figure 5.6.3.1: Rate of advance in the yearly date of flowering of winter wheat. The
day of the year of flowering has been simulated by using a crop growth model
(CGMS - Crop Growth Monitoring System) for the period 1975-2007.
Relevance
Changes in crop phenology provide important evidence of responses to recent
regional climate change (IPCC, 2007). Although phenological changes are often
influenced by management practices and new farming technologies, the recent
warming in Europe has clearly advanced a significant part of the agricultural calendar.
Specific agro-phenological stages are particularly sensitive (e.g.: flowering, grain
filling, etc.) to weather conditions and critical for final yield. The timing of the crop
cycle (agrophenology) determines the productive success of the crop. In general, a
longer crop cycle is strongly correlated to higher yields. In fact, a longer cycle permits
to maximize the thermal energy, the solar radiation and the water resources available.
The impact of unfavourable meteorological conditions and extreme events vary
largely according to the timing of occurrence and the development stage of the crops.
Agriculture has already adapted to the changing climate by selecting suitable varieties
or adapting the crop calendar, and can be expected to do so in the future.
Past trends
Several studies have collected data and observed changes in the phenological phases
of several perennial crops in Europe, such as the advance in the beginning of the
growing season of fruit trees (2.3 days/10 years), cherry tree blossom (2.0 days/10
118
years), apple tree blossom (2.2 days/10 years) in agreement with 1.4°C annual air
temperature increase in Germany (Chmielewski et al., 2004) and the advance of fruit
tree flowering of 1-3 weeks over the last 30 years for apricot and peach trees in
France (Seguin et al., 2004).
An advance of sowing or planting dates has been observed for several agricultural
crops, ranging from 5 days for potatoes in Finland to 10 days for maize and sugar beet
in Germany and 20 days for maize in France (IPCC, 2007)
Grapevine phenology
Wine composition is determined by various conditions: grape variety, rootstock, soil type, cultivation techniques,
and climatic characteristics. The first three conditions are generally constant, while cultivation techniques are most
often responsible for long-term variability. Climate influences year-to-year variability and is responsible for
variations in the amount and quality of produced wines.
Wine production areas, and particularly those for premium wines, are limited to regions climatically conducive to
growing grapes with balanced composition and degree to which they reflect their origin (“varietal typicity”). Three
conditions are required: (i) adequate heat accumulation; (ii) low risk of severe frost damage; and (iii) the absence
of extreme heat. Moreover, vines are resistant to limited water availability in summer and it is essential to have no
rainfall during harvest time, in order to increase sugar concentration and reduce disease development.
Observed climate change during recent years determined a general increase of wine quality, mainly due to the
increase of heat and reduction of rainfall,
particularly during the last part of the
ripening period, with a gradual increase in
the potential alcohol levels (Duchen et al.,
2005).
Future possible impacts:
• Seasonal shift: moving forward in
time of all the phenological phases
with an increase of frost risk and a
shortening of the ripening period. As
a possible effect, the harvest time
can now fall during periods with
high temperatures, with negative
effect on wine quality.
• Shifting of wine production areas, to
north and more elevated regions.
• Water stress due to a reduction of
available water.
Figure 5.6.3.2 Evolution of potential alcohol levels at harvest
• Modification of pest and disease
for Riesling in Alsace (F). Source: Duchen et al., 2005
development.
• Increase of sugar concentration
determining wine with high alcohol and low acidity. The consequence is the reduced possibility of wine
ageing and the poorer phenolic ripening.
• Modification of natural yeast composition.
Projections
Assuming the warming trend will continue, further reductions of the number of days
required for reaching the flower opening (anthesis) and maturity phases may be
expected for areas in Western Europe, where the phenology is currently strongly
accelerated. However, the rate of the reductions can gradually decrease with a further
increase of temperature (+4°C and +6°C hypotheses) due to a reduced efficiency of
the photosynthetic process at high temperatures.
119
5.6.4 Irrigation demand
Key Message

Between 1975 and 2006 clear trends, both positive and negative, were evident in
irrigation demand across Europe, with marked spatial variability. A significant
increase of irrigation demand (50-70%) occurred mainly in Mediterranean areas;
large decreases were recorded mainly in northern and central European regions.

Current trends and future scenarios depict an increase in the demand for water in
agriculture, potentially reducing the amount available for other sectors.
Presentation of the main indicator
Figure 5.6.4.1: Variation in the annual meteorological water balance between April
and October (m3/ha/yr) between 1975 and 2007. In areas where the rainfall amount
has grown faster than the increase in temperature (for ex. in NL and DK in the period
1997-99), the figure shows a reduction of water deficit. The rate of variation of the
“meteorological (or climatic) water balance”, expressed in m3 ha-1 y-1 is calculated as
the difference between the volume of rainfall and the loss by evapotranspiration. A
negative “meteorological water balance” indicates that evapotranspiration exceeds
rainfall, and the amount by which it does so, represents the additional volume of water
required from irrigation to compensate and to maintain the potential crops
productivity. In other words, this indicator provides an estimation of the maximum
irrigation volume which is necessary to ensure that crop growth is not limited by
water availability.
Relevance
Climate change may affect agriculture primarily through increasing atmospheric CO2,
rising temperatures and changing rainfall. Where rainfall does not limit crop growth,
these conditions allow for earlier sowing dates and enhanced crop growth and yield
120
(see previous indicators). Where reduced rainfall is predicted, however, the increased
requirement for irrigation water can have an overall negative impact in economic and
environmental terms. In these areas, increased water shortages in the future will
enhance the competition for water between sectors (tourism, agriculture, energy, etc.),
particularly in Southern Europe where the agricultural demand for water is greatest.
Several adaptation options are available to mitigate the risks of water shortage.
Increased irrigation can further burden surface and groundwater resources and
increase greenhouse gas emissions, adding to the mitigation challenge.
Past trends
Systematic observation of water
demand for agriculture do not exist
at European scale, however local
trends cab ne reconstructed by
using meteorological data. On
average, the annual rate of increase
of the irrigation demand is around
50 m3/ha/yr, but in some cases
(Italy, Greece, Maghreb, Central
Spain, Southern France and
Germany) it is above 150-200
m3/ha/yr. Areas with upward trends
in the water balance (mainly due to
an increase in rainfall), have been
observed in the Balkan Peninsula,
the Alpine region, Scandinavia,
Scotland, Benelux, the Czech
Republic, Slovakia, Poland and
Hungary, as well as in many
Turkish areas.
Figure 5.6.4.2: In the Mediterranean area, a worsening
climatic water deficit has been observed over the past
32 years (1975-2006)
Projections
No quantitative projections of future irrigation demand are available. Many climatic
projections for Europe (IPCC, 2007) foresee a very likely precipitation increase in the
North and a decrease in the South, especially during the summer. Also the extremes of
daily precipitation will increase in the north and the annual number of rainy days will
decrease in the Mediterranean. Therefore, the risk of summer drought is likely to
increase in central Europe and in the Mediterranean area. Agricultural crops will be
affected, among other factors, by changes in the length and timing of the vegetative
cycle. Adaptation in crop management will be necessary in order to try to avoid that
crucial development stages sensitive to water-stress (flowering, grain filling, etc.) will
occur during generally dry periods.
121
5.6.5. Forest Growth
•
In much of continental Europe, the majority of forests are growing faster now than in
the early 20th century.
•
A changing climate will favour certain species in some forest locations, while making
conditions worse for others, leading to substantial shifts in vegetation distribution.
•
The distribution and phenology of other plant and animal species (both pests and
pollinators) is likely to change, leading to further alterations in competition dynamics in
forests that will be difficult to predict.
Presentation of the main indicator
Figure 5.6.5.1: Modelled current (year 2000, left hand side) and future (year 2100, right hand
side) of the 10 most dominant European Forest Categories (EEA, 2006), modelled to evaluate
the change of habitat suitability. Source: Casalegno et al. 2007.
Relevance
Forests accumulate 77% of the global carbon pool in the vegetation biomass and hence play
an important role in the global carbon cycle (Dixon et al., 1994; IPCC, 2007). The carbon
sink of European forests is currently estimated to offset about 10% of European fossil fuel
emissions (Janssens et al. 2005, EEA, 2007). Forests and woodlands provide many things
that society values, including food, marketable products, medicines, biodiversity, carbon
reservoirs and opportunities for recreation. In addition, they regulate biogeochemical cycles
and contribute to soil and water conservation. Changes in global climate and atmospheric
composition are likely to have an impact on most of these goods and services, with significant
impacts on socioeconomic systems (Winnett, 1998).
Management has a significant influence on the development of the growing stock and forest
productivity. Adaptation measures include changes to plantation practices and forest
management, the planting of different species mixtures, better matching of the species to the
specific site, and the planting of similar species from their places of origin and non-native
species in anticipation of climate change (Broadmeadow et al., 2003).
122
Past trends
For many centuries, most European forests were overexploited. Growth rates were reduced
and biomass stocks were depleted until the middle of the twentieth century, when growth
rates started to recover (Spieker et al. 1996; ICP Forests, 2005). Much of this increase can be
attributed to advances in forest management practices, genetic improvement and, in central
Europe, the cessation of site-degrading practices such as litter collection for fuel. It is also
very likely that increasing temperatures and CO2 concentrations, nitrogen deposition, and
reduction of air pollution (SO2) have had a positive effect on forest growth. Trees have long
been known to respond to changes in climate: variations in tree-ring widths from one year to
the next are recognized as an important source of climatic information over time (see Chapter
2).
Although systematic monitoring of forest conditions is not existing in Europe, several studies
have already noted changes in dates of budburst and therefore longer growing season in
several species (see Section 5.5.3), shifts in tree-line and changes in species distribution (see
Section 5.5.2). A north-east shift of forest categories has already been observed for European
forest species (Bakkens 2002, Harrison et al. 2006).
Projections
Tree growth is controlled by complex interactions between climate and non-climate related
factors, with forest management also having a significant effect. Possible future responses of
forests to climate change include increased growth rates, tree-line movements, changes to
forest growth, phenology, species composition, increased fire incidence (see Chapter 5.5.6),
more severe droughts in some areas, increased storm damage, and increased insect and
pathogen damage (Eastaugh/IUFRO 2008). Taken together this is likely to lead to a changed
pattern of forest cover in the future. Simulation of the IPCC SRES A1B scenario for the
period 2070-2100 shows a general trend of a south-west to north-east shift in suitable forest
categories habitat (Casalegno et al. 2007).
Although climate change is anticipated to have an overall positive effect on growing stocks in
Northern Europe, negative effects are also projected in some regions (e.g. drought and fire
pose an increasing risk to Mediterranean forests), making overall projections difficult.
Warmer winter weather is likely to increase
productivity by extending the length of the
growing season (Cannell et al., 1998).
Reduced summer rainfall may reduce tree growth and
severe droughts may kill increasing numbers of trees.
Elevated atmospheric CO2 concentrations can
have a fertilising effect.
Elevated atmospheric ozone concentrations may have a
negative impact on growth (Sitch et al., 2007; Karnosky
et al., 2005).
Cold and snow-related damage are likely to
become less common.
Violent storms may occur more often, and more trees
are likely to be damaged or blown down.
Possible increase in spring frost damage as trees
become more susceptible through earlier leafing.
Damage by forest fires and insect pests is projected to
increase.
123
5.6.6. Forest fire danger
Key Message
• Under a warmer climate more severe fire weather is expected and, as a
consequence, more area burned, more ignitions and longer fire seasons.
• Climate change will increase the fire potential during summer months, especially
in southern and central Europe.
• The period in which fire danger exists will be longer in the future due to climate
change, with a likely increase of the frequency of extreme fire danger days in
spring and autumn.
Figure 5.6.6.1: Past trends of fire danger level from 1958-2006 using the Seasonal Severity
Rating (SSR).
Relevance
Wildfires are a serious threat to forests and ecosystems in Europe and climate is the
most important driving force affecting fire potential changes over time (Flannigan et
al. 2000). Although it is generally recognized that forest fire occurrence in Europe is
mostly due to causes of an anthropogenic nature, the total burned area changes
significantly from year to year largely because of weather conditions. Changes in fire
regimes may have strong impacts on natural resources and ecosystems stability, with
consequent direct and indirect economic losses. On other hand active forest
management and fire management practices can counteract the impacts of a changing
climate to some extent.
Past trends
Fire risk depends on many factors of a different nature that change over time (such as
e.g., weather, fuel load, fuel type and condition, forest management practices, socioeconomic context…). Historical fire series can be used to support statements on trend
but unfortunately long and consistent time series of fire events are rarely available in
Europe. In addition, by looking at the historical fire series alone, it is difficult to get a
clear picture and recognize the effect of climate on fire potential. In contrast
meteorological fire danger indices, which are designed to rate the component of fire
124
risk dependent on weather conditions, can be usefully employed to analyze fire trend
in a consistent way throughout longer time series. These indices, normally applied on
a daily bases, can be summarized on a seasonal basis to rate the overall fire potential
of a given year (seasonal fire severity) due to meteorological conditions. The index of
Seasonal Severity Rating (SSR) has been derived from daily values of Van Wagner’s
Fire Weather Index (FWI, Van Wagner 1987) the fire danger assessment method most
widely applied all over the world (San Miguel-Ayanz et al. 2003). Results of a recent
study on SSR development are shown in Figure 5.6.6.1. The average trend 1958-2006
was computed for all the grid cells, but it resulted to be statistically significant only
for 21% of the cases (15% positive and 6% negative), which appear to be
concentrated in specific geographical areas (see Figure 5.6.6.1.).
Projections
Future projections were derived for the IPCC SRES scenario A2, processing data from the
PRUDENCE data archive, namely the daily high resolution data (12 km) from HIRHAM
model run by DMI, for the time periods 1960-90 (control) and 2070-2100 (projections) (see
Figure 5.6.6.2). Results confirm in Europe, projections assessed for North America
(Flannigan et al. 2005) with a significant increase of fire potential, an enlargement of the fire
prone area and a lengthening of the fire season.
Figure 5.6.6.2. Projected (2071-2100) and control (1961-1990) three-monthly fire danger levels in
Europe for the IPCC SRES high emissions A2 climate change scenario. Fire danger in winter months
(DJF) is not shown because negligible.
125
5.6.7 Soil Organic Carbon
Key Messages
o Soil in the EU contains around 71*109 tons (or 71 gigatons, Gt) of organic carbon,
nearly 10% of the carbon accumulated in the atmosphere) An increase in
temperature and a reduction in moisture tend to accelerate decomposition of
organic material and subsequently lead to a decline in soil organic carbon (SOC)
stocks in Europe.
o The projected changes in the climate during the 21st century will make soils a
source of CO2 in most areas of the EU. To counterbalance the climate-induced
decline of carbon levels in soil adapted land use and management practices can be
implemented.
o Changes in SOC have already been observed in measurements in various
European regions over the last 25 years.
Figure 5.6.7.1 Changes in soil
organic carbon contents across
England and Wales between 1978
and 2003. a) Carbon contents in
the original samplings (1978-83),
and b) rates of change calculated
from the changes over the
different sampling intervals
(1994-2003).
Source: Bellamy et al., 2005
Relevance
Organic carbon in the soil is not a static component, but part of the carbon cycle,
which includes the atmosphere, water and constituents of the above and below-ground
biosphere. The main source of organic carbon is provided by organisms that
synthesize their food from inorganic substances (autotropic), such as
photosynthesising plants. In this process atmospheric carbon is used to build organic
materials and enters the soil layers through decomposition and the formation of
humus.
Climatic conditions strongly influence both the trends and rates of the accumulation
and transformation of organic substances in the soil. Increases in temperature and
aridity lead to a decrease of the amount of organic carbon in soil in areas thus
affected. Lower levels of organic carbon in the soil are generally detrimental to soil
fertility and tend to increase soil compaction, which subsequently leads to increases in
surface water runoff and erosion. Other effects of lower organic carbon levels are a
depletion of biodiversity and an increased susceptibility to acid or alkaline conditions.
The projected changes to the climate cause accelerated rates in the release of CO2
from the soil, which contributes to higher GHG concentrations in the atmosphere.
126
The main measures to reduce the detrimental effect of higher temperatures combined
with lower soil moisture on the amount of soil organic carbon concern changes in land
cover and adapted practices of land management. (Liski et al., 2002; Janssens et al.,
2004; Smith et el., 2005, 2006). Under the same climatic conditions grassland and
forests tend to have higher stocks of organic carbon than arable land and are seen as
net sinks for carbon (Vleeshouwers and Verhagen, 2002). Land management practices
aim at increasing the net primary production and reducing losses of above-ground
biomass from decomposition. Adaptive measure on agricultural land are changes in
farming practices, such as a reduction in tilling or retaining crop residues after
harvesting.
Past trends
In the past changes in organic carbon in the soil were largely driven by conversion of
land for the production of agricultural crops and long-term monitoring activities
concentrate on agricultural land, such as the Rothamstead long-term experiments
(Rothamstead Research, 2006). In Europe drained peat lands the annual loss in carbon
is in the range of 0 to 47 gCm−2 (Lappalainen, 1996). A survey of Belgian croplands
(210,000 soil samples taken between 1989 and 1999) indicates a mean annual loss in
organic carbon of 76 gCm−2 (Sleutel et al., 2003). A large-scale inventory in Austria
estimated that croplands were losing 24 gCm−2 annually (Dersch and Boehm, 1997).
A general intensification of farming in the past is likely to exceed the effect of
changes in the climate on soil organic carbon on agricultural land.
Projections
The amount of organic carbon in the soil is largely determined by balancing the net
primary production (NPP) from vegetation with the rate of decomposition of the
organic material. Without an increase in NPP soil carbon for cropland may decrease
by 8.8 to 12.3 t C ha-1. When including changes in NPP and technological advances
the amount of organic carbon on cropland could increasedby 1-7 t C ha-1 (Smith, et
al., 2005).
Figure 5.6.7.2: Projected changes of soil organic carbon in the EU for cropland for the IPCC
SRES A2 scenario up to 2080. The map (left) shows that climate change can cause loss (redcolour) of SOC for most areas in Europe. This decline can be reversed (blue-colour) if
measures enhancing soil carbon are implemented. As these are modelled data the projected
developments should be regarded with caution. Source: Smith et al., 2005.
127
5.6.8 Growing season for agricultural crops
Key Message
•
There is evidence that growing season length has varied in Europe for several
•
•
agricultural crops.
The longer growing season increases the productivity of crop yields and insect
population and favours the introduction of new species in areas which were not
suitable for these species before. These opportunities are particularly important
for the northern latitudes.
At southern latitudes, the trend is towards shortening of the growing season with
consequent higher risk of frost damages for the crops due to delayed spring frost
events.
Figure 5.6.8.2:
Variation of
growing season
length.
Relevance
Increasing air temperatures significantly impact on the duration of the growing season
over large areas in European (Scheifinger et al, 2003). The number of consecutive
days with temperatures above 0°C can be assumed as the period favourable for
growth, also called “frost-free period”. The timing and length of the frost-free period
is of interest to naturalists, farmers and gardeners amongst others. The impact on
plants and animals are largely reported with a clear trend towards earlier appearance
dates of spring starting growth and its prolongation during autumn (Menzel et al,
1999). A longer growing season permits the proliferation of those species facing
optimal conditions for their development and increase of their productive
performances (e.g.: crop yields, insect population, etc.); the introduction of new
species (very sensitive to frost) in areas before limited by unfavourable thermal
conditions. Changes of management practices, e.g. changes in species grown,
different varieties or adaptations of the crop calendar, can counteract the negative
effects of a changing growing season (pests) and capture the benefits (agricultural
crops).
Past trends
Many studies report the lengthening of the period between the dates of occurrence of
last spring frost and first autumn frost. This occurred in the last decades in several
areas in Europe and more generally in the Northern Hemisphere (Keeling et al., 1996;
128
Myneni et al., 1997; Magnuson et al., 2000; McCarthy et al., 2001; Menzel and
Estrella, 2001; Tucker et al., 2001; Zhou et al., 2001; Walther et al., 2002; Root et al.,
2003 Tait et al, 2003, Yan et al., 2002, Robeson 2002, Way et al., 1997). An analysis
of the growing season lengths in Europe between 1975 and 2006 shows a general and
clear trend in lengthening of the growing season. However, in the Mediterranean
countries, in the Black Sea area and in parts of Russia a decline in lengths have been
observed.
Frost-free period in
Frost-free period in
Danmark (DK)
HIGHLANDS AND ISLANDS (UK)
300
400
350
250
Nr. of days
Nr. of days
300
200
150
250
200
150
100
100
2002
2004
2006
2008
2008
1996
1994
1992
1990
1988
1998
1998
1996
1994
1992
1990
1988
1986
1984
1982
1980
1978
2008
2006
2004
2002
2000
1998
1996
1994
50
1992
100
50
1990
150
100
1976
200
150
1988
1986
250
1974
200
1986
2006
300
250
1984
2000
300
1982
2004
350
1980
2002
350
Nr. of days
400
1978
1984
EXTREMADURA (ES)
400
1976
1982
Frost-free period in
THESSALIA (GR)
Nr. of days
2000
Frost-free period in
1980
1978
1976
1974
2008
2006
2004
2002
2000
1998
1996
1994
1992
1990
1988
1986
1984
1982
1980
1978
1976
50
1974
50
1974
The trend is not uniformly spread over
Europe. The highest rates of variation
(about 0.5-0.7 days per year) have been
recorded in central and Southern Spain,
Central Italy, along the Atlantic shores,
on the British Isles, in Denmark and in
the central part of the EU. The
extension of the growing season
occurred either due to a reduction of
spring frost events or due to a
progressive delay of the occurrence of
fall frost. By contrast, in areas where a
decrease of the length occurred, in
particular in Southern Europe, the
plants are more at risk to frost damages
due to a delay in the last winter-spring
frost.
Figure 5.6.8.2: Development of frost-free periods in
selected areas
Projections
Following the observed trends (even more accelerated in the last decade) and
according to the future projections for temperatures increase, a further lengthening of
the growing season (both for earlier onset of spring and delayed of autumn) as well as
a northward shift of species is projected. The latter development is already widely
reported (Aerts et al., 2006). The length of growing season will mainly be influenced
by the increase of temperatures in autumn and spring (Ainsworth and Long, 2005;
Norby et al., 2003; Kimball et al.,28 2002; Jablonski et al., 2002).
According to the IPCC analysis Europe undergoes a warming in all seasons in all
scenarios, but the warming will be greater over Western and Southern Europe in
summer and over Northern ad Eastern Europe in winter. Therefore, in these areas a
larger lengthening of growing season is expected, whilst in southern and western
Europe the limited water availability and the high temperatures stress during summer
will affect the real plant growth negatively.
129
5.7 Water quantity, droughts, floods
5.7.1 Introduction
Water is essential to any form of life and constitutes an indispensable resource for nearly all
human activity. It is intricately linked with climate through a large number of connections and
feedback cycles, so that any alteration in the climate system will induce changes in the
hydrological cycle. Global warming due to the anthropogenic rise in greenhouse gasses not
only results in widespread melting of snow and ice, but also augments the water holding
capacity of the air and amplifies evaporation. This leads to larger amounts of moisture in the
air, an increased intensity of water cycling, as well as to changes in the distribution, frequency
and intensity of precipitation (see also Section 5.2.3 and 5.2.5). Consequently, the distribution
in time and space of freshwater resources, as well as any socio-economic activity depending
thereon, is affected by climate variability and climate change.
There is growing evidence of changes in the global hydrological cycle over the last 50 years
that may be linked with changes in climate, such as an increasing continental runoff, a wetter
northern Europe and a drier Mediterranean, an increase in the intensity of extreme
precipitation events over many land regions, as well as changes in the seasonality of river
flow where winter precipitation dominantly falls as snow (see also Section 5.3).
Long-term trends in hydrological variables, however, are often masked by the significant
inter-annual to decadal variability. Compared to historic data availability about
meteorological variables, hydrometric records are more sparse and limited in time due to the
lack of dense observation networks of long term hydrological variables. Also, confounding
factors such as land-use change, water management practices or extensive water withdrawals
have considerably changed the natural flow of water, making it more difficult to detect
climate change induced trends in hydrological variables. It may therefore require substantially
more time before statistically significant changes can be observed, especially in the frequency
of extreme events such as floods and droughts, because of their long-term return intervals and
the random nature of their occurrence.
For the coming decades, global warming is projected to further intensify the hydrological
cycle, with impacts that will likely be more severe than those observed to date. Climate
change will lead to strong changes in the yearly and seasonal water availability across Europe.
Water availability will generally increase in northern parts of Europe, although summer flows
may decrease. Southern and south-eastern regions of Europe, which already suffer most from
water stress, will be particularly exposed to reductions in water resources and see an increase
in the frequency and intensity of droughts. On the other hand, an increase in extreme high
river flows is projected for large parts of Europe due to the increase of heavy rain events,
even in regions where it will get drier on average. Due to limitations of climate models, as
well as scaling issues between climate and hydrological models, quantitative projections of
changes in precipitation and river flows at the river basin scale remain, however, highly
uncertain.
The projected climate-induced changes will aggravate the impact of other stresses, such as
land use, demographic or socio-economic changes, on water availability, freshwater
ecosystems, energy production, navigation, irrigation, tourism, as well as on several other
sectors. In the face of these uncertain changes, adaptation procedures need to be designed that
can be altered or that are robust to changes. Such measures include, for example, stimulating
awareness, improving water efficiency and encouraging water conservation to mitigate water
stress, directing spatial planning and watershed management to enhance retention and reduce
flood risk, as well as effective monitoring, detection and early warning of hazards or changes
in water availability.
130
5.7.2 River flow
Key Messages
o
o
o
o
Over the 20th century, annual river flows have shown an increasing trend in northern parts
of Europe, with increases mainly in winter, and a slightly decreasing trend in southern
parts of Europe. These changes are linked with observed changes in precipitation patterns
and temperature.
Annual river flow is projected to decrease in southern and south-eastern parts of Europe
and to increase in northern parts of Europe, but absolute changes remain uncertain.
Climate change is projected to result in strong changes in the seasonality of river flows
across Europe. Summer flows are projected to decrease in most parts of Europe, also in
regions where annual flows will increase.
Regions in southern Europe, which already suffer most from water stress, are projected to
be particularly exposed to reductions in water resources due to climate change. This will
result in increased competition for available water resources.
Presentation of the main indicator(s)
Figure 5.7.2.1. Change in annual river
flow (in percent) for the period 1971-98
relative to 1900-70. Map is based on an
ensemble of 12 climate models driven by
estimated historical variations of solar
irradiance and radiatively active
atmospheric gases and aerosols, and
validated against observed river flows
(from Milly et al., 2005).
Relevance
Water is an indispensable resource for human health, ecosystems and socio-economic
activity. From a resource perspective, river flow is a measure of sustainable fresh water
availability in a basin. Variations in river flow are determined mainly by the seasonality of
precipitation and temperature, as well as by catchment characteristics such as geology, soils
and land cover. River flow can be used as an indicator because changes in temperature and
precipitation patterns due to global warming modify the distribution of water at the land
surface, and consequently the annual water budget of river basins as well as the timing or
seasonality of river flows. The consequent changes in water availability may adversely affect
ecosystems and several socio-economic sectors such as water management, energy
production, navigation, irrigation and tourism.
In view of global warming and the associated changes in water availability it will become
increasingly important to balance competing societal, industrial, agricultural and
environmental demands. Sustainable options to mitigate the effects of changes in water
131
availability include improved water efficiency, the re-use of water, or metering and water
pricing to stimulate awareness and encourage water conservation.
Past trends
In accordance with the observed changes in precipitation and temperature (see Section 5.2.2
and 5.2.3), there is some evidence of climate induced changes in annual river flow, as well as
in the seasonality of flow, in Europe during the 20th century. However, anthropogenic
interventions in the catchment, such as groundwater abstraction, irrigation, river regulation,
land use changes and urbanization, have considerably altered river flow regimes in large parts
of Europe, confounding climate change detection studies.
In northern parts of Europe, the mean annual river flow has in general increased (Lindström
and Bergström, 2004, Milly et al., 2005). Increases occurred mainly in winter and spring
season (Hisdal et al., 2007), likely caused by a general temperature increase during the last
decades (see Section 5.2.2) in combination with increased winter precipitation (see Section
5.2.3) in the northern regions. Significant increases in river flow have been observed also in
Scotland for one third of the sites in the last three decades (Werritty et al., 2002), as well as in
winter and autumn in western Britain, consistent with recent increases in winter rainfall and a
positive NAO index (Dixon et al., 2006). However, some of these changes could be part of
natural variability (Wade et al., 2005). In western and central Europe, annual and monthly
mean river flow series appear to be stationary over the 20th century (Wang et al., 2005). In
mountainous regions of central Europe, however, the main identified trends are an increase in
annual river flow due to increases in winter, spring and autumn river flow. In summer, both
upward and downward trends have been detected (Birsan et al., 2005). In southern parts of
Europe, a slightly decreasing trend in annual river flow has been observed (Milly et al.,
2005).
Projections
Annual river flow is projected to decrease in southern and south-eastern parts of Europe and
to increase in northern and north-eastern parts of Europe (Arnell, 2004, Milly et al., 2005;
Alcamo et al., 2007). Strong changes are also projected in the seasonality of river flows, with
large differences across Europe. Winter and spring river flows are projected to increase in
most parts of Europe, except for the most southern and south-eastern regions of Europe. In
summer and autumn, river flows are projected to decrease in most parts of Europe, except for
northern and north-eastern regions of Europe where autumn flows are projected to increase
(Dankers and Feyen, 2008). In snow dominated regions, such as the Alps, Scandinavia and
the Baltic, the drop in winter retention as snow, earlier snowmelt and reduced summer
precipitation will reduce river flows in summer (Andréasson, et al., 2004; Jasper et al., 2004;
Barnett et al., 2005), when demand is typically highest.
132
Figure 5.7.2.2. Relative change in mean annual and seasonal river flow between scenario
(2071-2100) and control period (1961-1990) for IPCC SRES scenario A2 (Dankers and
Feyen, 2008).
Figure 5.7.2.3. Change in daily average river flow between 2071-2100 (blue line) and 19611990 (black line) for IPCC SRES scenario A2 (Dankers and Feyen, 2008).
133
5.7.3 River floods
Key Messages
o
o
o
o
Although a significant trend in extreme river flows has not yet been observed to date, in
Europe twice as much river flow maxima occurred between 1981-2000 compared to
1961-1980.
Since 1990, 259 major river floods have been reported in Europe, of which 165 have been
reported since 2000 (EM-DAT). The rise in reported number of flood events over the last
decades is mainly caused by better reporting and land-use changes.
Nevertheless, it is projected that global warming will intensify the hydrological cycle and
increase the occurrence and frequency of flood events in large parts of Europe, even
though that estimates of changes in flood frequency and magnitude remain highly
uncertain.
Due to warming, projections suggest less snow accumulation during winter and therefore
a lower risk of early-spring flooding.
Presentation of the main indicator(s)
Figure 5.7.3.1. Recurrence
of flood events in Europe
for the period 1998-2005.
Source: EEA (2006), based
on data from Darthmouth
Flood Observatory.
Relevance
A flood can be defined as the temporary covering of land by water outside its normal
confines. There exist different types of floods, such as large scale river floods, flash floods,
ice-jam or snowmelt induced floods, and coastal floods due to sea level rise (see Section
5.4.2). Inland river floods are mainly linked with prolonged or heavy precipitation events as
well as with snowmelt, hence are suited as an indicator of climate change.
River floods are the most common natural disaster in Europe. They result in huge economic
losses due to damages to infrastructure, property and agricultural land, as well as due to
indirect losses in or outside the flooded areas, such as production loss caused by damaged
stocks or roads, or by the interruption of power generation plants and navigation. River floods
lead to loss of life, especially in the case of flash floods, displacement of people and have
adverse effects on human health and the environment.
134
To account for the projected changes in extreme precipitation and river flows current
procedures for designing flood-control infrastructures have to be revised. Flood management
policy will also have to shift from defensive action towards the management of risk and
enhancing societies’ ability to live with floods. This can be achieved via the use of nonstructural flood protection measures such as spatial planning, early warning, relief and postflood recovery systems, as well as flood insurance (Kundzewicz et al., 2002).
Past trends
Despite that there has been a considerable rise in the number of reported major flood events
and in economic losses caused by floods in Europe over the last decades (see Section 6.3), no
general climate related trend has yet been detected in extreme high river flows that induce
floods (Becker and Grunwald, 2003; Glaser and Stangl, 2003; Mudelsee et al., 2003;
Kundzewicz et al., 2005; Pinter et al., 2006; Hisdal et al., 2007; Macklin and Rumsby, 2007).
Some changes, however, have been reported that may be linked with climate change. For
example, in Europe twice as much river flow maxima occurred between 1981-2000 compared
to 1961-1980 (Kundzewicz, 2005), whereas globally there has likely been a increase in the
frequency of extreme rare flood events in very large catchments (Milly et al., 2002). On the
other hand, in Central Europe the frequency and severity of snowmelt and ice-jam floods
decreased over the last decades due to warming of European winters in combination with less
abundant snow cover (e.g., Mudelsee et al., 2003; Brázdil et al., 2006; Cyberski et al., 2006).
In the Nordic countries, snowmelt floods have occurred earlier due to warmer winters (Hisdal
et al., 2007). In Portugal, changed precipitation patterns resulted in an intensification of
floods during autumn but a decline in floods in winter and spring (Ramos and Reis, 2002). In
the UK, positive trends in high flows have been observed over the last 30-50 years (Robson,
2002; Dixon et al., 2006), some of which are consistent with observed changes in the North
Atlantic Oscillation (see Section 5.4.2). Comparison with historical climate variability and
flood records suggests however that many of the detected changes over the last decades could
reasonably be part of natural climatic variation. Changes in the terrestrial system, such as
urbanisation, deforestation, loss of natural floodplain storage, as well as river and flood
management have also strongly affected flood generation (Barnolas and Llasat, 2007).
Projections
Although there is yet no proof that the extreme flood events in recent years are a direct
consequence of climate change, they may be an indication of what can be expected in the
future. Namely, it is projected that the frequency and intensity of flood events is projected to
increase in large parts of Europe (Lehner et al., 2006; Dankers and Feyen, 2008). Especially
flash and urban floods, triggered by local intense precipitation events, are likely to be more
frequent throughout Europe (Christensen and Christensen, 2003; Kundzewicz et al., 2006).
Flood hazard will also likely rise during wetter and warmer winters, with increasingly more
frequent rain and less frequent snow (Palmer and Räisänen, 2002). Even in regions where
mean river flows will significantly drop, such as in the Iberian Peninsula, the projected
increase in precipitation intensity and variability may cause more floods. In snow-dominated
regions, such as the Alps, the Carpathian Mountains and northern parts of Europe, spring
snowmelt floods are projected to decrease due to a shorter snow season and less snow
accumulation in warmer winters (Kay et al., 2006; Dankers and Feyen, 2008).
135
Figure 5.7.3.2. Relative change
in 100-year return level of river
discharge between scenario
(2071-2100) and control period
(1961-1990) for the IPCC
SRES scenario A2 (Dankers
and Feyen, 2008).
136
5.7.4 River flow drought
Key Messages
o
o
o
o
Europe has been affected by several major droughts in the last decades, such as the
catastrophic 2003 drought associated with the summer heatwave in central parts of
Europe and the 2005 drought in the Iberian Peninsula.
Notwithstanding the absence of a general trend in Europe as a whole, in some regions
climate change has likely increased the frequency and/or severity of droughts.
Climate change is projected to increase the frequency and intensity of droughts in many
regions in Europe due to higher temperatures, decreased summer precipitation, as well as
more and longer dry spells.
Regions most prone to an increase in drought hazard are southern and south-eastern
Europe, but minimum river flows will also decrease significantly in many other parts of
Europe, especially in summer.
Presentation of the main indicator(s)
Figure 5.7.4.1. Change in
the severity of river flow
drought for the period
1962-1990. Red indicates
more severe, blue
indicates less severe
(from Hisdal et al, 2001).
Figure 5.7.4.2. Change in
the severity of river flow
drought in France for the
period 1960-2000. Red
indicates more severe,
blue indicates less severe
(from Lang et al., 2006).
137
Relevance
Drought may refer to meteorological drought (precipitation well below average, Section
5.2.3), hydrological drought (low river flows, lake and groundwater levels), agricultural
drought (soil moisture deficit, Section 5.6), environmental drought (impact on ecosystems)
and socio-economic drought (impact on economic goods and services). The focus here is on
hydrological drought, more specifically on river flow drought, as river flow is a measure of
sustainable fresh water availability in a basin and it is impacted by climate change. River flow
data are also more abundantly available compared to other hydrometric information such as
on groundwater recharge, surface water storage or soil moisture. Climate-induced trends in
extreme low river flows are however often masked by land use changes, water management
practices and extensive water withdrawals.
Prolonged droughts have considerable economical, societal and environmental impacts. They
affect several sectors, such as energy production, both in terms of water availability for
hydropower and cooling water in electricity generation, river navigation, agriculture, as well
as public water supply.
Adverse effects of droughts and low river flow conditions can be mitigated at the supply side
through the conjunctive use of surface and groundwater, desalination of sea water as well as
the storage and transfer of water. Demand side measures include improving water efficiency,
metering and water pricing. Shortages in water can be anticipated through effective
monitoring and forecasting of future river flows and storage in reservoirs.
Past trends
Over the past 30 years, Europe has been affected by a number of major drought events, most
notably in 1976, 1989, 1991, and more recently, the prolonged event over large parts of
Europe associated with the summer heatwave in 2003. The most serious drought in the
Iberian Peninsula in 60 years occurred in 2005, reducing overall EU cereal yields by an
estimated ten percent. The drought also triggered forest fires, killing 15 people and destroying
180,000 hectares of forest and farmland in Portugal alone (UNEP, 2006). However, to date,
there is no evidence that river flow droughts have become more severe or frequent in general
over Europe in recent decades (Hisdal et al., 2001), nor is there conclusive proof of a general
summer desiccation in Europe over the last 50 years due to reduced summer moisture
availability (van der Schrier et al., 2006).
Notwithstanding the absence of a general trend in Europe, distinct regional differences have
been found. In particular, more severe river flow droughts were observed in Spain, eastern
part of Eastern Europe and in large parts of the UK (Hisdal et al., 2001). However, in much of
the UK there is no evidence of a significant increase in the frequency of occurrence of low
river flows (Hanneford and Marsh, 2006). In large parts of Central Europe and in the western
parts of Eastern Europe droughts became less severe (Hisdal et al., 2001). In France, a
majority of stations showed a decreasing trend in the annual minima of 30 days mean river
flows over the last 40 years, but no such trend was found for drought severity or duration
(Lang et al., 2006). In southern and eastern Norway there has been a tendency towards more
severe summer droughts (Hisdal et al., 2006). On the other hand, several stations in Europe
have shown trends towards less severe low flows over the 20th century, consistent with an
increasing number of reservoirs becoming operational in the catchments over the period of
record (Svensson et al., 2004).
Projections
For the coming decades river flow droughts are projected to increase in frequency and
severity in southern and south-eastern Europe, the UK, France, Benelux, and in western parts
of Germany. In snow-dominated regions, where droughts typically occur in winter, droughts
138
are projected to be less severe because a lower fraction of precipitation will fall as snow in
warmer winters. In most parts of Europe, the projected decrease in summer precipitation,
accompanied by rising temperatures, which enhances evaporative demand, may lead to more
frequent and intense summer droughts (Douville et al., 2002; Lehner et al., 2006; Feyen and
Dankers, 2008). Due to both climate change and increasing water withdrawals more river
basins will be affected by severe water stress, resulting in increased competition for available
water resources. Regions most prone to an increase in drought risk are the Mediterranean and
south-eastern parts of Europe, which already suffer most from water stress (Alcamo et al.,
2003, Schröter, 2005).
Figure 5.7.4.3. Relative change in mean annual and summer minimum 7-day river flow
between 2071-2100 and 1961-1990 based on IPCC SRES scenario A2. Red indicates more
severe droughts, blue indicates less severe droughts (Feyen and Dankers, 2008).
139
Text box: Groundwater
The main pressures on the groundwater system due to climate change are sea level rise,
shrinking land ice and permafrost areas, declining groundwater recharge especially in
southern European countries, more extreme peak flows and more prolonged low flows of
rivers, and increased groundwater abstraction.
The resulting effects on groundwater quantity consist of shrinking fresh groundwater
resources especially in coastal areas and in southern European countries, while brackish and
salt groundwater bodies will expand. In addition, the fresh groundwater bodies will become
more vulnerable to pollution by reduced turnover times and accelerated groundwater flow.
Saline intrusion in coastal aquifers, making the water unsuitable for drinking, may be
exacerbated by future sea level rise Other effects on groundwater quality are more difficulty
to predict as they strongly depend on changes in land-use. Nevertheless it is already clear that
groundwater temperature has increased on average by 1 C since the 1970s (Stuyfzand et al.,
2007). Further increases will raise the salinity of groundwater due to increased
evapotranspiration losses, increased soil CO2 pressures and increased water – rock
interaction.
140
5.8 Water quality and fresh water ecology
5.8.1 Introduction
Climate change can result in significant changes in the variables that affect the quality of
water. These impacts come from a variety of alterations to the temperature regimes and
hydrology of water bodies, their physico-chemical and biological attributes, and changes in
anthropogenic pressures. Changes include:
 physical changes such as water temperature (see indicator on water temperature), river
and lake ice-cover (see Section 5.8.3), stratification of water masses in lakes and water
discharge including water level and retention time;
 chemical changes, in particular oxygen content, nutrient loading and water colour;
 biological changes affecting the freshwater ecology.
Changes in these variables lead to impacts on all the socio-economic and environmental
goods and services that depend on these variables directly or indirectly.
A rise in water temperature will affect the rate of biogeochemical processes which determine
water quality. This may result in:
 Reduced oxygen content. Increases in water temperature in streams and lakes reduce
oxygen content and increase biological respiration rates and thus may result in lower
dissolved oxygen concentrations, particularly in summer low-flow periods and in lake
hypolimnia. Higher temperature and lower oxygen concentration will cause stress and
may reduce the habitats for cold-water species such as salmonids in lakes and rivers.
 Less ice formation. Earlier ice break-up and longer annual ice-free period in rivers and
lakes (see further in the indicator on river and lake ice cover);
 a more stable vertical stratification and less mixing of water of deep water lakes, which in
turn affect deep water oxygen conditions, nutrient cycling and plankton community;
 Eutrophication. A warmer climate will generally enhance pollution load of nutrients to
surface and groundwater. Higher temperature will increase mineralization and releases of
nitrogen and carbon from soil organic matter and increased run-off and erosion that will
result in increased pollution transport. Also internal phosphorus load is expected to
increase in stratified lakes, due to declining oxygen concentrations in bottom waters.
 Change in timing of algal blooms and increase of harmful algal blooms (see further in the
indicator on freshwater ecology, Section 5.8.4);
 Alterations to habitats and distribution of aquatic organisms. For example, a number of
aquatic organisms conform to temperature preferences which determine their spatial
distribution. Higher water temperatures lead to changes in distribution (more northwards
in Europe) and may even lead to extinction of some aquatic species (see further in the
indicator on freshwater ecology).
Also climate change factors other than temperature can affect water quality. In areas where
amounts of surface water and ground water recharge will decrease, water quality may also
decrease due to lower dilution of pollutants. Higher intensity and frequency of floods and
higher intensity precipitation events is expected to increase the load of pollutants (organic
matter, nutrients, hazardous substances) being washed from soils and from overflows of
sewage systems to water bodies. Many lakes and rivers in areas with peat or forests have
brown-coloured water. An unprecedented increase in water colour over the last 10-15 years
has been reported in the Nordic countries, the UK, the foothills of central Europe, and northeastern USA. Several climate related mechanisms may be responsible for the observed
increases, but the reduction of acid rain in the last decades may be an important explanation
for the observed increase in water colour in these areas (Monteith et al. 2007).
141
Many of the diverse aspects of climate change (e.g. temperature increase, variations in rainfall
and runoff) affect the distribution and mobility of hazardous substances in freshwater
systems. Loading of hazardous substances may increase due to sewage overflow, as well as
higher pesticide use and run-off due to heavy rains, while higher temperature enhance the
degradation rate of pesticides and organic pollutants, which may reduce the concentration in
rivers and lakes. Thus the net effect of climate change on hazardous substances is uncertain.
European freshwaters are already affected by many human activities resulting in changes in
land-use, pollution with nutrients and hazardous substances, and acid deposition. Because of
difficulties in disentangling the effects of climatic factors from other pressures, there is
limited empirical evidence to demonstrate unequivocally the impact of climate change on
water quality and freshwater ecology. On the other hand, there are many indications that
freshwaters that are already under stress from human activities are highly susceptible to
climate change impacts. Currently many national and European research activities are
producing relevant and valuable results on climate change impacts on Europe’s freshwater;
see for example, Euro-limpacs (http://www.eurolimpacs.ucl.ac.uk/index.php) and CLIME:
Climate and Lake Impacts in Europe (http://clime.tkk.fi/)
142
5.8.2 Water temperature
Key Messages


During the last century the water temperature of European rivers and lakes increased by
1–3°C mainly due to air temperature increase, but also locally due to increased inputs of
heated cooling water from power plants.
As water temperature is closely linked to the change in air temperature, an increase in air
temperature due to climate change will be reflected in an increase of surface water
temperature.
Presentation of the main indicator
Figure 5.8.2.1: Trend in annual water temperature in river Rhine (1909–2006), Danube
(1901–1998) and average water temperature in August in Lake Saimaa, Finland (1924–2000)
Source: River Rhine: Rijkswaterstaat; River Danube: Hohensinner, 2006; Lake Saimaa: Korhonen,
2002; and L. Võrtsjärv 1947-2006 [missing reference].
Relevance
Since water temperature largely is determined by heat exchange with the atmosphere higher
air temperatures lead to higher surface water temperatures. Particularly in standing waters and
low flow situations in rivers, higher water temperatures will bring about changes in the
physico-chemical condition of water bodies with subsequent impacts on biological conditions.
This may have severe consequences for ecosystem structure and function as well as for water
use and ecosystem services.
Impacts of increased water temperatures also include more stable vertical stratification of
deep lakes and oxygen depletion; more frequent harmful algal blooms, reduced habitats for
cold-water aquatic species and increased incidence of the temperature-dependent illnesses
(see also indicator on freshwater ecology, Section 5.8.4).
Man can help freshwater ecosystems to adapt to increasing water temperature only in a
limited way, such as reducing the pressures from other human activities e.g. pollution by
nutrients and hazardous substances and hydromorphological modifications.
143
Past trends
From long time series covering the last 100 years it can be concluded that the surface water
temperature of some of the major rivers in Europe has increased by 1–3°C over the last
century (Figure 5.8.2.1). The temperature of the River Rhine increased by 3°C from 1910 to
2006. Two-thirds of this temperature rise is estimated to be due to the increased use of
cooling water in Germany and one-third to the increase in temperature as a result of climate
change (MNP, 2006). In the river Danube the annual average temperature increased by
around by 1°C during the last century. A similar temperature increase was found in some
large lakes: Lake Võrtsjärv in Estonia had a 0.7°C increase from 1947 to 2006 and the
summer (August) water temperature of Lake Saimaa, Finland more than 1°C over the last
century.
There are many shorter time series of water temperature covering the last 30-50 years and the
general trend has been that temperature has increased in European freshwater systems.
Generally the positive trend varies from 0.05 to 0.8oC decade-1.
 George et al. 2005 found that the temperature of Lake Windermere (England) and Lough
Feeagh (Ireland) increased by 0.7-1.4 °C from 1960 to 2000. The water temperature of
Lake Veluwe (the Netherlands) has increased by more than one degree since 1960 (MNP
2006);
 Marked increase in water temperature was found in eight Lithuanian lakes
(Pernaravičiūtė, 2004) and six Polish lakes (Dabrowski et al,. 2004);
 Since 1950, water temperatures in rivers and in lake surface water in Switzerland have in
some cases increased by more than 2°C (BUWAL 2004). In the large lakes in the Alps
the water temperature has generally increased 0.1 - 0.3°C per decade: Lake Maggiore and
other large Italian lakes (Ambrosetti & Barbanti, 1999); Lake Zürich, (Livingstone,
2003); Lake Constance and Lake Geneva (Anneville et al., 2005).
Dokulil et al. (2006) studied the trend in hypolimnion (the bottom water) temperature in 12
deep European lakes and found generally a temperature increase in the hypolimnion of 0.1–
0.2°C per decade. This temperature increase may have significant effects on thermal
stratification and mixing of water in lakes, which in turn affect deep water oxygen conditions
and nutrient cycling.
Projections
As water temperature is closely linked to change in air temperature, the predicted increase in
air temperature due to climate change will be reflected in increased surface water
temperature. This comes on top of possible temperature changes by other factors, such as
changes in cooling water releases. Forecasted increases in surface water temperatures are
often 50 to 70% of the forecasted increase in air temperature. In line with the projected
increase in air temperature (see section 5.2.2) lake surface water temperature maybe around
2°C higher by 2070, but with a clear seasonal dependency and depending on lake properties
(Malmaeus et al., 2006; George et al., 2007).
144
Figure 5.8.2.2: Observed changes in annual average deepwater temperatures. Time series and
regression lines for: (A) Windermere north basin 60 m and the first 10-week period (Q1), (B) Lake
Geneva, (C) Zürichsee, (D) Walensee, (E) Lake Constance, (F) Ammersee, (G) Lake Vänern, (H) Lake
Vättern, (I) Hallstättersee, (J) Traunsee, (K) Mondsee, and (L) Attersee for the depths indicated. Tincrease in all lakes was 0.1-0.2oC decade-1. Source: Dokulil et al., 2006.
145
5.8.3 Lake and river ice cover
Key Messages



In the Northern Hemisphere, the duration of ice cover has shortened at a mean rate of
12 days per 100 years accounted for by an average 5.7 days later ice-on and 6.3 days
earlier ice-off per century.
The strongest trends in northern Europe occur in the timing of ice break-up which is
consistently with the fastest warming in winter and spring.
Ice cover of lakes with mean winter temperature close to zero react much stronger on
change in temperature compared to lakes in colder regions such as northern
Scandinavia.
Presentation of the main indicator
Figure 5.8.3.1. Time series of ice break-up dates from selected European lakes and rivers and
the North Atlantic Oscillation (NAO) index for winter (Dec.- Feb.). Data smoothed with a 7year moving average. Se Section 5.4.3 for more information about the NAO.
Source: Benson & Magnuson 2000, (Updated to 2006 by J. Korhonen and D. Livingstone).
Relevance
The appearance of ice cover is indicative of a longer period with temperatures below 0 °C.
The deeper the lake, the more cold is needed to cool down the lake so that ice forms. Increase
in water temperature will result in less ice formation. Higher water temperature will affect the
duration of ice cover, the freezing and thawing up dates and the thickness of ice cover.
Changes in ice cover are of critical ecological importance for lakes because of their effect on
the underwater light climate (Leppäranta et al., 2003), nutrient recycling (Järvinen et al.,
2002) and oxygen conditions (Stewart, 1976; Livingstone, 1993), which influence the
production and biodiversity of phytoplankton (Rodhe, 1955; Phillips and Fawley, 2002;
Weyhenmeyer et al., 1999) and the occurrence of winter fish kills (Greenbank, 1945; Barica
and Mathias, 1979). Variations in lake and river ice are relevant in terms of freshwater
hydrology and human activities such as winter transportation, bridge and pipeline crossings,
and winter sports but no quantitative evidence for observed effects exists yet (IPCC, 2007). In
Europe there is some evidence for a reduction of ice-jam floods due to reduced freshwater
freezing during the last century (Svensson et al., 2006). There are only limited options for
helping the freshwater ecosystem to adapt to changes in ice cover, however, reducing other
human pressures will generally make the ecosystems more robust to cope with changing
climate conditions.
146
Past trends
An analysis of long (150-yr) ice records from lakes and rivers throughout the Northern
Hemisphere by Magnuson et al. (2000) indicated that for a 100 year period, ice-on has been
occurring on average 5.7 ± 2.4 days later (± 95% confidence interval), while ice-off has been
occurring on average 6.3 ± 1.6 days earlier, implying an overall decrease in the duration of ice
cover at a mean rate of 12 days per 100 years.
Change in ice parameters show mostly trends that are in agreement with an observed local
temperature increase. Air temperature is the key variable determining the timing of ice breakup (Palecki and Barry, 1986; Livingstone, 1997). As the fastest warming in the Northern
Hemisphere has occurred in winter and spring, the trends in break-up dates are generally
steeper than those in freeze-up dates.
A few longer time series reveal reduced ice cover (a warming trend) beginning as early as the
16th century, with increasing rates of change after about 1850 (see Figure 5.8.3.1). The early
and long term decreasing trend in the ice break-up dates is the result of the ceasing of the
Little Ice Age, which lasted about from 1400 to 1900 (Kerr, 1999). In the 20th century, the
effects of the North Atlantic Oscillation on the ice regime of European inland waters appears
to be stronger than the effect of increasing temperatures.
Studying ice cover information from 11 Swiss lakes over the last Century Franssen and
Scherrer (in press) found is found that in the past 40 years, and especially during the last two
decades, ice cover was significantly reduced.
Ice cover of lakes at lower latitude for example in southern Sweden are more sensitively to
climate change than lakes in the north, where mean winter temperatures are lower than zero
most of the winter. A study on 196 Swedish lakes along a latitudinal temperature gradient
revealed that a 1°C air temperature increase caused an up to 35 days earlier ice break-up in
Sweden’s warmest southern regions with annual mean air temperatures around 7°C. It only
caused an about 5 days earlier break-up in Sweden’s coldest northern regions with annual
mean air temperatures around −2°C (Weyhenmeyer et al. 2004; 2007). Also in Finland ice
break-up has become significantly earlier from the late 19th century to the present time,
except in the very north (Korhonen, 2006).
Projections
Future increases in air temperature associated with climate change are likely to result in
generally shorter periods of ice cover on lakes and rivers. The most rapid decrease in the
duration of ice cover will occur in the temperate region where the ice season is already short
or only occurring in cold winters (Weyhenmeyer et al., 2004). As a result, one part of the
present ice-covered lakes that mix from top to bottom during two mixing periods each year
(dimictic lakes) will potentially change into monomictic, open-water lakes with consequences
for vertical mixing (mixing only once), deep-water oxygenation, nutrient recycling and algal
productivity. This may lead to an alteration in the ecological status of ice-covered lakes in
temperate regions.
Regional climate model projections for northern Germany, based on the IPCC high emissions
SRES A2 and intermediate emissions B2 climate scenarios, imply that for the Müggelsee, the
percentage of ice-free winters will increase from ~2% now to over 60% by the end of the
current century (Livingstone & Adrian, submitted). By contrast, increases in mean annual air
temperature are likely to have much smaller effect on lakes in very cold regions (e.g. northern
Scandinavia) until these also reach the threshold of having winter temperature close to zero.
147
5.8.4 Freshwater ecology
Key Messages




Several freshwater species have shifted their ranges to higher latitudes (northward
movement) and altitudes in response to climate warming and other factors.
There are European examples on changes in life cycle events (phenology) such as
earlier spring phytoplankton bloom, clear water phase, first day of flight and spawning
of fish.
In several European lakes phytoplankton and zooplankton blooms occur one month
earlier compared to 30-40 years ago.
Climate change can cause enhanced phytoplankton blooms, favouring and stabilizing
the dominance of harmful cyanobacteria in phytoplankton communities, resulting in
increased threat of these bacteria and enhanced health risks, particularly in water
bodies used for public water supply and bathing.
Presentation of the main indicator(s)
Figure 5.8.4.1.: A) Northward shift of range margins of British Odonata, dragonflies and
damselflies, between 1960–1970 and 1985–1995. Northern species (white), ubiquitous
species (hatched) and southern species (black) are shown. B) Observed occurrence of seven
types of southern dragonflies in Belgium, 1980-2005
A
B
Source: A) Hickling et al. 2005 and B) Biodiversity Indicators, 2006
Relevance
Species and habitat dynamics in the face of climate change are complex and have many
aspects. Increased temperatures and increased CO2 concentrations will have an effect on the
different processes such as photosynthesis, respiration and decomposition and generally speed
up the different processes. Climate induced changes in ice cover period, thermal stratification
and nutrient availability and longer growing season affects the species composition and food
web structures.
Water temperature is one of the parameters that determines the overall health of aquatic
ecosystems. Most aquatic organisms (e.g. salmonids) have specific range of temperatures that
they can tolerate, which determine their spatial distribution along a river or at a regional scale.
Climate change could lead to the extinction of some aquatic species or at least can modify
their distribution in a river system or move their distribution northwards in Europe. Several
indications of climate impact on freshwater ecology functioning and biodiversity have already
been observed, such as northward movement; phenology changes and invasive alien species.
148
Reducing the stress/pressure to freshwater ecosystems due to other human pressures such as
pollution will make the ecosystems more robust to cope with the increasing stress due to
changing climate conditions.
Past trends
Northward and upward movement
There are European examples of aquatic species (dragonflies, brown trout) that shift their
ranges to higher latitudes (northward movement) and altitudes in response to climate
warming. Thermophilic fish and invertebrate taxa will to a certain extent replace cold-water
taxa. Examples include the brown trout in Alpine rivers (Hari et al., 2006), non-migratory
British dragonflies and damselflies (Hickling et al., 2005, see Figure 5.8.4.1.A), and SouthEuropean Dragonflies in Belgium (Biodiversity Indicators, 2006, see Figure 5.8.4.1.B).
Change in species composition and abundance
Climate change will generally have a eutrophication-like effect (e.g. Schindler, 2001), with
enhanced phytoplankton blooms (Wilhelm and Adrian 2007), and increased dominance of
cyanobacteria in phytoplankton communities, resulting in increased threat of harmful
cyanobacteria and enhanced health risks, particularly in water bodies used for public water
supply and bathing (Jöhnk et al. in press; Mooij et al. 2005;). Trends in temperature have
already had profound impacts on species composition of macrozoobenthos (fauna that are
partially or fully buried for the majority of their lives in water bodies) in Northern Europe
lakes (Burgmeer et al., 2007)). Fish and invertebrate communities have been found to respond
to increases in water temperature in the upper Rhône river in France (Daufresne et al (2004,
2007).
Phenology changes
Changes in growth season, such as the ice free period or periods above a certain temperature
will change life cycle events, such as an earlier spring phytoplankton bloom, the clear water
phase, the first day of flight of aquatic insects and the time of spawning of fish. Prolongation
of the growing season can have major effect on population abundances with increased number
of cell divisions, or generations per year. In several European lakes phytoplankton and
zooplankton blooms occur one month earlier compared to 30-40 years ago from
phytoplankton and zooplankton blooms (Weyhenmeyer 1999, 2001; Adrian et al 2006; Noges
et al. 2007) Evidence is available on change in phenology of British Odonata species (Hassall
et al., 2007).
Invasive freshwater species
Invasive species are species that arrive at locations where they were never recorded before.
Climate change is expected to support biological invasions of species that originate in warmer
regions. For example, the subtropical filamentous highly toxic cyanobacterium
Cylindrospermopsis raciborskii thrives in waters that have high temperatures, a stable water
column and high nutrient concentrations: it has recently spread rapidly in temperate regions
and is now commonly encountered throughout Europe (Dyble et al. 2002). The spreading to
drinking and recreational water supplies has caused international, public health concerns due
to its potential production of toxins. Fish species adapted to warmer waters, such as carp, may
replace native fish species, such as yellow perch and trout in many regions (Kolar and Lodge
2000)
The above changes are mostly related to increased water temperature. As discussed in Section
5.8.2, this warming is at least partly related to climate change, while also other factors play a
role (cooling water from power generation, sewage discharges).
149
Projections
Many species are predicted to shift their ranges to higher latitudes and altitudes in response to
climate warming. Southern species will move further north due to further increase of
temperatures. Species of colder regions will move north and towards higher altitudes or will
disappear when their migration is hampered (e.g. due to habitat fragmentation). Arctic and
mountainous species may disappear.
 A comparison of a large set of Danish shallow lakes with a corresponding one located in
colder climate in Canada (Jackson et al. 2007) suggest that warming will decrease winter
fish-kills and enhance overwintering success of planktivorous fish which, in turn,
suppress Daphnia development. As a result of decreased zooplankton grazing pressure,
there will be more phytoplankton biomass built up per unit total phosphorus in warmer
climate.
 Where river discharges decrease seasonally, negative impacts on Atlantic salmon may
occur. Walsh and Kilsby, 2006 found that salmon in northwest England will be affected
negatively by climate change because suitable flow depths during spawning time, which
occur all the time now, will only exist in 94% of the time in the 2080s (SRES A2
scenario)
 In the ongoing European research project Eurolimpacs there has been an evaluation of
Trichoptera taxa (aquatic insects) potentially endangered by climate change (see Figure
below). The main results are that more than 20% of the Trichoptera species are projected
to be endangered due to climate change in Southern Europe (droughts) and in the Alpine
region (too high temperatures), whereas in other parts of Europe, the impacts would be
less pronounced (Hering et al. in press).
Figure 5.8.4.2. The share of Trichoptera taxa sensitive to climate change (endemism, crenal preference
and cold stenothermy) in the European ecoregions. A distinct South-East – North-West gradient is
revealed: in all ecoregions of North-West Europe the share of potentially endangered taxa is lower than
10%, while the share is 51.7% on the Iberian Peninsula and 42.3% in Italy. Also the Balkan ecoregions
and the high mountain ranges (Alps, Pyrenees, and Carpathians) are characterized by more than 25%
potentially endangered taxa. Source: Hering et al. 2006: Evaluation of Trichoptera data in relation to
climatic gradients. Deliverable No. 190 from the Eurolimpacs European Research Project available at
http://www.eurolimpacs.ucl.ac.uk/oldsite/docstore/Deliverable_190.pdf
150
5.9. Human health
5.9.1 Introduction
Climate change currently contributes to the global burden of disease and premature
deaths (IPCC, 2007). Human beings are exposed to climate change through changing
weather patterns (temperature, precipitation, sea-level rise and more frequent extreme
events) and indirectly through changes in water, air and food quality and changes in
ecosystems, agriculture, industry and settlements and the economy. At this early stage
the effects are small but are projected to progressively increase in all countries and
regions (Confalonieri et al, 2007).
There is now emerging evidence of observed climate change effects on human health.
For example, climate warming in recent decades has altered the distribution of some
infectious disease vectors, altered the seasonal distribution of some allergenic pollen
species and increased the frequency and intensity of heat-waves. In the longer term,
many serious impacts on health may occur, including an increase in the number of
people suffering from death, disease and injury from heat-waves, floods, storms, fires
and droughts; a change in the range of some infectious disease vectors; increase of the
burden of diarrhoeal diseases from changes in water quality and quantity (IPCC,
2007). In parts of Europe, there may be some benefits of climate change to health,
including fewer deaths from cold. It is expected that the benefits will be outweighed
by the negative effects of rising temperatures worldwide.
This chapter describes some of the climate sensitive health outcomes in Europe, in
particular for those where action is necessary. A limited amount of studies, data and
information are currently available. Epidemiological studies have been undertaken on
the health impacts of individual extreme events (e.g., heatwaves, floods); spatial
studies where climate is an explanatory variable in the distribution of the disease or
the disease vector; and temporal studies assessing the health effects of climate
variability, including daily changes in temperature or rainfall. A very limited number
of intervention studies have been undertaken to investigate the effectiveness of
public-health measures to protect people from climate hazards.
151
5.9.2 Heat and health
Key Messages



High temperatures increase the risk of summer mortality, particularly in the
elderly. Studies have shown that mortality risk increases by 2% in northern
cities, and 3% in southern cities for every 1°C increase in apparent maximum
temperature above a cut point.
Heat wave events can have severe impacts on health. Longer heat waves have
an impact 1.5 to 5 times higher than shorter events. The hot summer of 2003
(June-September) was associated with more than 70,000 excess deaths, in 12
European countries.
86,000 net extra deaths per year are projected for the EU countries for a high
emissions scenario with a global mean temperature increase of 3°C in 20712100 relative to 1961-1990.
Presentation of the main indicator:
Figure 5.9.2.1: Number of excess death in summer 2003 and under a high emissions
scenario for 2070
Relevance
Heat directly affects the human physiology: Thermoregulation during heat stress
requires a healthy cardiovascular system. When environmental heat overwhelms the
heat coping mechanism, the body’s core temperature increases. This can lead to heat
illness, or deaths from heat–stroke, heart failure and many other causes.
152
Figure 5.9.2.2: Factors affecting human thermoregulation and the risk of heat illness
(adapted from Bouchama, 2007) in technical Summary EuroHEAT
Overall, the impact of hot weather and heat-waves depends upon the level of exposure
(timing, frequency, intensity and duration of the heat-wave); the size and structure of
the exposed population (demographic profile; aging, health status, children, etc); the
population sensitivity (prevalence of chronic diseases); preparedness of health
systems and the prevention measures in place.
Several medical factors can increase the risk of heatwave mortality, including
dehydration, drugs, ageing, and having a chronic disease that affects cardiac output
and the skin blood flow, Social factors, such as social isolation, may also be important
factors – although there has been little research in Europe (see Figure 5.9.2.2,
Bouchama, 2007)
Major heat-wave events are also associated with other health hazards such as air
pollution, wild fires, water, food and electricity supply failures—which also have
implications for public health action. The effects of heat-wave on mortality may be
larger during high ozone days, highlighting the interaction between climate change
and air pollution.
Future heat-related mortality may be reduced, by acclimatisation and adaptation. On
the other hand, increasing numbers of older adults in the population will increase the
proportion of the population at risk (Confalonieri et al, 2007). Adaptation measures
include national heat health action plans, heat health warning, health system
preparedness and response, advice on what to do, and particular care for those most
vulnerable.
153
Past trends
Many epidemiological studies have quantified the impact of temperature on daily
mortality. In all cities in Europe the temperature – mortality relationship in long time
series shows an increase in mortality over a certain temperature threshold. The point
at which temperatures increase (the threshold) is related to the local climate. The
estimated change in mortality risk by degree increase in temperature ranges from 0.7
to 3.6%. The research project “Assessment and Prevention of Acute Health Effects of
Weather Conditions in Europe (PHEWE)”, estimated an increase in mortality for
every 1°C increase in apparent temperature above thresholds of 2% in northern cities
and 3% in southern cities. (Michelozzi et al, forthcoming)
In 2003, more than 70,000 excess deaths were recorded in 12 European countries, as
compared to averages between 1998 and 2002,. In August alone, nearly 45,000
additional deaths were recorded in the 12 countries, including 15,251 in France
(+37%), 9713 in Italy (+21.8%), 7295 in Germany (+11%), 6461 in Spain (+22.9%)
and 1987 in England and Wales (+4.9%) (Robine et al, 2008). The timing, intensity
and duration of heat-waves have been shown to influence the amount of mortality.
Impacts of heat waves characterized by longer duration were from 1.5 to 5 times
higher than for short heat waves (EuroHEAT, 2008).
Projections
Heat-related morbidity and mortality is projected to increase. Within several National
Assessments, estimates on heat mortality have been carried out, using different
climate scenarios, population and adaptation assumptions. In the UK, annual heatrelated deaths are expected to increase from 798 in 1990s to 2,793 in 2050s and 3,519
in the 2080s under the medium-high scenario. Annual cold-related deaths decrease
from 80,313 in 1990s to 60,021 in 2050s and 51,243 in 2080s under the medium-high
scenario (Donaldson et al, 2001). In Germany, a 20% increase in heat-related
mortality is expected. This increase is estimated to be not likely to be compensated by
reductions in cold-related mortality (Koppe, 2003). In Portugal, an increase in heatrelated mortality from a baseline of 5.4 to 6 deaths/100,000 to a range of 19.5 to 248.4
deaths/100,000 by 2080s is expected (Dessai, 2003).
Almost 86,000 net extra deaths per year under a high emissions scenario (IPCC SRES
A2, see Chapter 4) with a global mean temperature increase of 3°C in 2071- 2100
relative to 1961-1990 in EU25 (PESETA, 2007). These results are preliminary and do
not assume physiological adjustment and do not separate out the impact of nonclimate changes (socio-economic changes in age structure or population movements).
The study is based on assumptions of a mortality-temperature relationship that does
not take into account the differences between the Mediterranean and northern
European countries.
154
5.9.3 Vector borne diseases
Key messages




Changes in bird migration, could shift geographic distribution of vector
borne diseases and change the life cycles of bird-associated pathogens.
The tiger mosquito, a transmittor a number of viruses, has extended its
range in Europe substantially over the last 15 years and is projected to
substantially further expand.
Changes in the geographical distribution of the sandfly vector have been
reported in several European countries.
For the UK it was estimated that projected temperature increases would
increase the risk of local malaria transmission by 8 to 15%; in Portugal a
significant increase in the number of days suitable for survival of malaria
vectors is projected.
Presentation of the main indicator:
Figure 5.9.3.1. Recent (January 2007) and projected distribution of Aedes albopictus
in Europe (Source: Scholte and Schaeffner, 2007).
Relevance
Climate change is likely to cause changes in ecological systems that will affect the
risk of infectious diseases in Europe, including the seasonal activity of local vectors,
and the establishment of tropical and semi-tropical species. Shifts in the global and
regional distribution and behaviour of insect and bird species are early signs that
biological systems are already responding to climate change (see also Section 5.5.5).
The IPCC (2007) projects that climate change will lead to significant changes in
infectious disease transmission by vectors (such as mosquitoes and ticks) as a result of
changes in the geographical range, seasonality, disease transmission and absolute
number of cases.
Patterns of infectious disease in Europe are and will be affected by movement of
people and goods, changes in hosts and land use. There are fears in Europe that new
infectious diseases could be triggering population health problems and/or previously
155
existing diseases could re-emerge. At the moment very limited information is
available and current ongoing research projects, like EDEN will provide in future
years more information on vector distribution.
Various adaptation options are available to reduce climate change risks for vectorborne diseases. For example early detection of any outbreak and improved public
health surveillance, as required by the International Health Regulation. However
locally some integrated vector-control measures will need to be revised or
strengthened.
Past trends
In Europe, shifting of life-cycle events, pole ward shift of boundaries, changes in
migratory patterns have been observed in some animals and plant species (see Box 1).
Box 1: Bird migration
Climate change has been implicated in changes in the migratory and
reproductive phenology (advancement in breeding and migration dates) of
several bird species, their abundance and population dynamics, as well as a
northward expansion of their geographical range in Europe. There are two
potential consequences: a) shifts in the geographical distribution of the vectors
and pathogens due to altered distributions or changed migratory patterns of bird
populations; and b) changes in the life cycles of bird-associated pathogens due to
the mistiming between bird breeding and the breeding of vectors, such as
mosquitoes. For example, the impact on bird migration patterns may play a role
in the distribution of West Nile fever.
in the distribution
of of
West
Nile fever.
Transmission
patterns
malaria
are intricately connected to meteorological
conditions such as temperature and precipitation. For example, conditions for
transmission in Europe have remained favorable as documented by repeated rare
autochthonous transmission of a tropical malaria strain by local vectors to a
susceptible person. At present, autochthonous malaria continues to pose a challenge in
Turkey.
Aedes aegypti (tiger mosquito)– a vector which between many others transmits
dengue - closely follows the 10 °C winter isotherm and is extending its range. Dengue
is frequently introduced into Europe by travellers returning from dengue-endemic
countries. Aedes albopictus has extended its range in Europe substantially over the
last 15 years (Scholte and Schaeffner, 2007). A traveler bitten by Ae. albopictus could
well be the source of transmission of a number of viruses in Europe. A recent cluster
of cases Chikungunya (a virus that is highly infective and disabling but is not
transmissible between people) has been observed in the Emilia-Romagna Region of
Italy. This is the first example that an imported human disease case was followed by a
sustained local mosquito transmission in continental Europe.
Under climate change, a shift towards milder winter temperatures may enable
expansion of the range of Lyme disease into higher latitudes and altitudes, but only if
all of the vertebrate host species required by the tick vector are equally able to expand
their population distribution. In contrast, droughts and severe floods will negatively
affect the distribution, at least temporarily. There is some observational evidence of
northern or altitudinal shifts in tick distribution from Sweden and the Czech Republic.
156
However, climate change alone is unlikely to explain recent increases in the incidence
of tick-borne diseases in Europe, as there is considerable spatial heterogeneity in the
degree of increase of tick-borne encephalitis. (see Figure 5) In Eastern Europe,
milder weather conditions, favoring tick reproduction, may influence the distribution
of Crimean Congo Haemorragic Fever.
Cutaneous leishmaniasis has been reported in dogs (reservoir hosts) further north in
Europe, although the possibility of previous under-reporting cannot be excluded.
Changes in the geographical distribution of the sandfly vector have been reported in
several European countries.
It is believed that at least some of these observed changes can be attributed to climate
change, although also other factors play an important, often dominant role, such as
human behaviour (outdoor activities, international travel) and the status of health care
services.
Projections
Projections of climate-change-related vector borne diseases use different approaches
to classify the risk of climate-sensitive health determinants and outcomes. For malaria
and dengue, results from projections are commonly presented as maps of potential
shifts in distribution (see Figure 5.9.3.1). Health-impact models are typically based on
climatic constraints on the development of the vector and/or parasite, and include
limited population projections and non-climate assumptions. Models with incomplete
parameterisation of biological relationships between temperature, vector and parasite
often over-emphasise relative changes in risk, even when the absolute risk is small.
Several modelling studies used the IPCC SRES climate scenarios, a few applied
population scenarios, and none incorporated economic scenarios. Few studies
incorporate adequate assumptions about adaptive capacity. The main approaches used
are inclusion of current ‘control capacity‘ in the observed climate–health function and
categorisation of the model output by adaptive capacity, thereby separating the effects
of climate change from the effects of improvements in public health (Confalonieri et
al, 2007).
Several climate change related models estimate an increase of malaria risk: For
example, in the United Kingdom it was estimated that, with temperature increases, the
risk of local malaria transmission could increase by 8–15% by 2050. In Portugal, the
number of days suitable for survival of malaria vectors is projected to increase.
Nevertheless, there is agreement that the risk of transmission of malaria related to
localized climate change is very small. Risks are greater in the countries where
importation of malaria coincides with socioeconomic degradation, the disintegration
of health and social services, uncontrolled cross-border migration and lack of
environmental management for mosquito control
Dengue
An empirical model estimated that, in the 2080s, 5-6 billion people would be at risk of
dengue as a result of climate change and population increase, compared with 3.5
billion people if the climate remained unchanged (Hales, 2002). This projections
include a risk for the Mediterranean countries for dengue.
157
5.9.4 Water and food borne diseases
Key messages




Changing frequency and intensity of precipitation events (and temperature)
from climate change may affect outbreaks of waterborne diseases.
Increasing risk of flooding due to climate change has the potential to mobilise
pathogens or chemicals from stores.
Climate change is likely to affect water quality in rivers due to chemical
contamination.
There is a linear increase in reported cases of some food borne diseases with
each degree increase in weekly or monthly temperature over a certain location
specific threshold.
Main indicator
Use access to safe waters – based on ‘Inland waters’ chapter from
the Belgrade report
Use maps / graphs from the report – state and projections, if
available
Relevance
Access to safe water remains an extremely important global health issue. The risk of
outbreaks of waterborne diseases increases where standards of water, sanitation and
personal hygiene are low. In Europe 95% of urban populations and 71% of rural
populations have access to public sources of drinking-water in 2004. However, in
many cases the supply is discontinuous and the water supplied does not meet the
microbial and/or chemical guidelines set by the WHO.
Intestinal infectious diseases that are transmitted through food or water are sensitive
to climate and weather factors. Such diseases are the main causes of infectious
diarrhoea and cause significant amounts of illness each year in Europe.
Approximately 20% of the population in western Europe are affected by episodes of
diarrhoea each year (IID study team, 2002;van Pelt et al., 2003). Such infections have
a significant economic impact through treatment costs and loss of working time
(Roberts et al., 2003).
Climate change is also likely to affect the quality of coastal waters, either by changing
the natural ecosystem or by changing the quality of the waters draining in the coastal
zone. This poses specific risks for the recreational use of bathing waters particularly
for the transient tourist population which may not have built-up resistance against
endemic water-related diseases or which may be faced with water quality that does
not meet stringent conditions imposed in the home country. The quality and safety of
seafood produced in this manner is directly linked to the quality of the water in the
coastal zone.
158
Droughts or extended dry spells can reduce the volume of river flow which may
increase the concentration of effluent pathogens posing a problem for the clearance
capacity of treatment plants
Various adaptation options are available, which include ensuring access to safe
drinking-water, provide sanitation services and establish common standards for
surveillance systems and contingency plans to detect and prevent waterborne disease
outbreaks. Water-safety plans are suggested to be revised for changing climate
conditions. These plans will need to include ensuring safe drinking-water from source
to tap through enhanced risk assessment and management. Improved management of
water demand in the context of fully integrated planning for river-basin management
will become imperative as a first coping mechanism, but it is unlikely to satisfy all the
needs created by demographic growth, rising living standards and economic
development. Alternative strategies will need to be explored, including reusing treated
wastewater, using grey water, harvesting rainwater and, where economically viable,
desalinization.
Past trends
Key foodborne and waterborne infections are monitored in Europe. The incidence of
Salmonella has been declining in many countries, but the incidence of other
pathogens is increasing. New pathogens have also emerged in recent years. In Europe
the risk of infectious disease outbreaks is relatively rare due to the standard of water
treatment and distribution infrastructure. Nevertheless, water-borne outbreaks have
the potential to be rather large but the actual disease burden in Europe is difficult to
estimate and most likely underestimated. Examples of an increased risk of infectious
disease outbreaks have been found in the UK (Reacher et al, 2004), Finland
(Miettinen et al, 2001), Czech Republic (Kriz et al, 1998) and Sweden (Lindgren,
2006). Heavy precipitation has been linked to a number of drinking water outbreaks
of Cryptosporidium (a pathogen causing a diarrheal illness) in Europe, due to spores
infiltrating drinking water reservoirs from springs and lakes and persisting in the
water distribution system (Lake et al, 2005, Semenza and Nichols, 2007). In
Germany, bacteriological and parasitic parameters spiked considerably during
extreme runoff events (Kistemann et al, 2002).
Harmful algal blooms (HABs) produce toxins that can cause human diseases, mainly
via consumption of contaminated shellfish. Warmer seas may thus contribute to
increased cases of human shellfish and reef-fish poisoning (ciguatera) and poleward
expansions of these disease distributions (Lehane and Lewis, 2000)(Hall et al, 2002,
Hunter, 2003, Korenberg, 2004). Several studies have confirmed and quantified the
effects of high temperatures on common forms of food poisoning, such as
salmonellosis (D'Souza et al, 2004)(Kovats et al, 2004)(Fleury et al, 2006). (see table
5..9.3.1) These studies found an approximately linear increase in reported cases with
each degree increase in weekly or monthly temperature. Temperature is much less
important for the transmission of Campylobacter (Kovats et al, 2005, Louis et al,
2005; Tam et al, 2006). Contact between food and pest species, especially flies,
rodents and cockroaches, is also temperature-sensitive. Fly activity is largely driven
by temperature rather than by biotic factors (Goulson et al, 2005).
159
Table 5.9.3.1 shows the increase of Number of cases of salmonella by degree increase
of temperature over a certain threshold
Country
% change (95% CI)
Poland
Scotland
Denmark
England and
Wales
Estonia
Netherlands
Czech Republic
Switzerland
Slovak Republic
Spain
8.7 (4.7, 12.9)
5.0 (2.2, 7.9)
0.3 (- 1.1, 1.8)
12.5 (11.6, 13.4)
9.2 (- 0.9, 20.2)
8.8 (8.0, 9.5)
9.2 (7.8, 10.7)
9.1 (7.9, 10.4)
2.5 (- 2.6, 7.8)
4.9 (3.4, 6.4)
Projections
With each degree increase of temperature 0.2% of increase of number of diarrhoea
cases are expected (REFERENCE).Water stress is projected to increase over central,
southern Europe. In EU27, the percentage area under high water stress is likely to
increase from 19% today to 35% by the 2070s, and the number of additional people
affected by the 2070s is expected to be between 16 millions and 44 millions.
Furthermore, in southern Europe and some parts of central and Eastern Europe,
summer water flows may be reduced by up to 80%. By 2025 it is estimated that an
additional 31 million people will be residing in the coastal zone of the Mediterranean,
and that 130 million more will visit the region.
160
6. Economic Consequences of Climate Change
6.1 Introduction
There are a wide range of economic effects that will arise from climate change in
Europe. These include effects on the natural environment and associated services
(including forests and fisheries), coastal zones, agriculture, tourism, energy, human
health and the built environment.
The observed and projected effects of climate change in Europe are different across
regions and sectors. Many of these impacts are projected to be adverse and to lead to
economic costs (‘losses’), though there will also be economic benefits (gains). There
is a strong distributional pattern for future economic effects predicted across Europe,
with a significant trend towards more potential negative impacts in South-eastern
Europe and the Mediterranean (e.g. in relation to energy demand, agricultural
productivity, water availability, health effects, summer tourism, ecosystems, etc). In
northern and Western Europe a more complex balance between negative and positive
impacts is projected for moderate levels of climate change in the coming decades,
potential benefits mainly derived from new farming and tourism opportunities. As
climate change continues, eventually the negative impacts are projected to dominate.
It is also evident that even if emissions of greenhouse gases stop today, changes in
climate will continue for many decades. Therefore, in addition to mitigation, it is
essential to develop proportionate adaptive responses (adaptation) as a means of
moderating damages or realising opportunities associated with climate change. There
is therefore a need to also consider the economic aspects of adaptation. While there is
a wide and increasing amount of information on the impacts of climate change, and
on adaptation, there is limited information on economic costs. This section
summarises some of the available information, and considers economic indicators
with a focus on Europe.
The economic costs of climate change impacts if no adaptation would take place are
known as the ‘costs of inaction’. They relate to both direct and indirect climate
change impacts, including the associated socio-economic developments. Estimates of
these costs and the costs of adaptation are increasingly helping to inform the policy
debate, in particular in discussing the level of mitigation efforts (reduction of
emissions of greenhouse gases) that is needed globally.
As a first indicator, direct losses from weather-related natural disasters are analysed.
Past trends indicate that economic losses due to weather-related natural disasters have
increased considerably, particularly in recent years. Since no statistically significant
increase in frequency of events like floods has as yet been observed, the increase of
economic losses is probably mainly determined by other factors, such as a possible
increase in the intensity of flood events, the overall increase in wealth and
possibilities for insurance, and increased amount and distribution of infrastructure
vulnerable to such disasters. Therefore we have also included a separate indicator on
economic losses from floods (which comprise the largest share of weather-related
natural disasters in Europe) for which such socio-economic effects have been
161
removed (“normalised”) in order to understand the actual weather/climate-related
trend better (6.3). It can be seen that by using such a normalisation method the losses
are generally lower3. Additional information and analysis are then presented for
energy (heating and cooling demand, hydropower), coastal areas, agriculture and
forestry, tourism and recreation, health, public water supply, nature and biodiversity
and the society as a whole.
A brief overview of the economic effects from climate change across Europe is
presented in the map below.
Figure 6.1.1 Summary of Economic Effects across Europe
3
For past trends in losses due to all weather-related natural disasters (6.2), the most recent Munich Re
dataset has been used, which is not normalised. It is important to be aware of these differences in the
indicators, and the normalised trends of economic losses provide more information on these issues.
162
6.2 Direct losses from weather disasters
Key messages




About 90% of all natural disasters in Europe occurred since 1980 are directly or
indirectly attributable to weather and climate. About 95 % of economic losses caused
by catastrophic events4 result from these weather and climate-related disasters.
The average number of annual disastrous weather and climate-related events in
Europe increased by about 65 % over the period 1998-2007 compared with the 1980s
decade, while non-weather events such as earthquakes remained stable. An unknown
share of this increase can be attributed to climate change, the rest to changes in
sensitivity of societal systems.
Overall losses resulting from weather and climate-related events have increased
clearly during the last 25 years. Even though social change and economic
development are mainly responsible for increasing losses, there is evidence that
changing patterns of weather disasters are drivers too. However, it is still not possible
to determine the proportion of increase in damages that might be attributed to
anthropogenic climate change.
While in the immediate future disaster losses are projected to increase predominantly
as a result of societal change and economic development, the most severe effects of
human-caused climate change on economic assets are expected in the second half of
the century.
Presentation of the main indicator(s)
200
Geophysical (earthquake, tsunami, volcanic activity)
180
Meteorological (storm)
160
Climatological (temperature extremes, drought, wildfire)
140
Hydrological (flood, mass movement)
120
100
80
60
40
20
0
1980
1982
1984
1986
1988
1990
1992
1994
1996
1998
2000
2002
2004
2006
Figure 6.2.1. Natural disasters in Europe 1980-2007 (Number of events)
Source: Münchener Rückversicherungs-Gesellschaft (Munich Re), Geo Risks
Research, NatCatSERVICE, (2008).
The following definitions apply (Munich Re): (1) A ‘major catastrophe’ is defined as a 100+ fatalities
event with overall losses in excess of US$ 200m; (2) A ‘devastating catastrophe’ is defined as a 500+
fatalities event with overall losses in excess of US$ 500m; (3) A ‘great natural catastrophe’ or ‘GREAT
disaster’ is defined as leading to thousands of fatalities with the economy being severely affected and
extreme insured losses (UN definition); interregional or international assistance is necessary, hundreds
of thousands are made homeless.
4
163
Number of events: 3,500
2%
Deaths: 88,500
11 %
22 %
25 %
39 %
3%
3%
4%
68 %**
Geophysical
(earthquake, tsunami, volcanic activity)
23 %
15 %
Meteorological (storm)
Insured losses:
Euro 59bn
Overall losses:
Euro 197bn
Hydrological
(flood, mass movement)
5% 2 %
5%
Climatological without heat wave
(temperature extremes, drought, wildfire)
27 %
6%
Heat wave
36 %
66 %
38 %
Figure 6.2.2. Natural disasters in Europe 1980-2007 (Percentage distribution)
Source: Münchener Rückversicherungs-Gesellschaft (Munich Re), Geo Risks Research,
NatCatSERVICE, (2008).
**: Most of all casualties were elderly people who died in the 2003 heat wave/hot summer
(surmortality)
Relevance
Changes in the frequency and intensity of storms, floods and extreme temperatures
due to climate change affect the financial sector, including the insurance sector, via
the amount of compensation payments. This indicator can also help to identify which
sectors, e.g. agriculture, forestry, infrastructure, industry or private households, are
most affected by damage and/or could be most affected in future.
A recently published report from the United Nation Environment Programmes
Finance Initiative (UNEP FI, 2006) estimated that losses from extreme weather events
are doubling globally every 12 years. Even though the observed increase in losses is
dominated by socio-economic factors (such as population growth, increased
habitation in vulnerable areas, increased wealth, increased amount/value of vulnerable
infrastructure), there is evidence that changing patterns of natural disasters are drivers
too (Figure 6.2.1). It is however not known how much of this increase in losses can be
attributed to anthropogenic climate change (Hoeppe et al., 2006).
Insurance mechanisms can play an important role in adapting to climate change by
covering the residual risks and providing incentives for risk reduction. Insurance
companies have inherent interests in minimizing the impact of climate change.
Through their asset management, finance services have great influence on company’s
investment decisions. They can therefore warrant that investments made are more
climate-resilient and channel investments into projects related to adaptation and
mitigation of climate change. On the other hand industries with greatest exposures
164
will have to respond increasingly with innovative products, e.g. catastrophe bonds
(Bouwer et al., 2007).
Figure 6.2.3: Overall and insured losses by weather disasters in Europe, 1980-2007
(in bn Euro).
Source: Münchener Rückversicherungs-Gesellschaft (Munich Re), Geo Risks
Research, NatCatSERVICE
35
Overall losses (2007 values)
Insured losses (2007 values)
30
Trend overall losses
Trend insured losses
25
20
15
10
5
06
20
04
20
02
20
00
20
98
19
96
19
94
19
92
19
90
19
88
19
86
19
84
19
82
19
19
80
0
As at March 2008
© 2008 Münchener Rückversicherungs-Gesellschaft, Geo Risks Research,
NatCatSERVICE
Past trends
In Europe, 64 % of all loss events since 1980 are directly attributable to weather and
climate events (storms, floods and heat-waves) and 25 % are attributable to wild
fires, cold spells, landslides and avalanches, which are also linked to weather and
climate. 95 % of the overall losses and 78 % of all deaths caused by disastrous events
result from these weather and climate-related events (Figure 6.2.2). The annual
average number of these weather and climate-related events in Europe increased
during the period 1997-2007 by about 65 % compared with the 1980s decade, while
non-climatic events, such as earthquakes remained stable (Figure 6.2.1).
In Europe, overall losses caused by weather and climate-related events have increased
during the period 1980-2007 from a decadal average of less than 7.2 billion € (19801989) to about 13.7 billion € (1998-2007). Six of the nine years with the largest
overall losses in this period have occurred since 1999 (Figure 6.2.3). The insured
portion of the losses generally rose, but the increase varied between the years.
Particularly disastrous extreme events in Europe in recent years include the severe
flooding in central Europe in August 2002 and the extended heat wave in Europe in
2003. The 2002 flooding of Austria, the Czech Republic, Germany, Slovakia and
165
Hungary resulted in overall losses of about € 16.8 billion and insured losses of about
€ 3.4 billion (Münchener Rück, 2008). The 2003 heat-wave (Schaer et al., 2004)
resulted in ten-thousands of deaths in central Europe additional to the normal numbers
(Kovats and Jendritzky, 2006; Robine at al., 2007) and caused significant losses in the
agricultural and energy producing sectors. As an example, the total loss of the hot
summer 2003 in France (including from power generation, stress on the transport
system, stress on forests and other ecosystems including fires, reduced wine
production and decreased agricultural productivity) has been estimated at 0.1/0.2 % of
GDP equivalent to 15-30 billion euros (Gillet, 2006). The 2003 summer was also
estimated to have increased building subsidence claims by 20 % in the UK, with
estimated impacts of £30 to £120 million (Hunt and Taylor 2006) and damage to
transport infrastructure (rail buckling and road subsidence) of £40 million (Watkiss et
all, 2006).
Projections
Extreme weather as heat waves, droughts and heavy precipitation is projected to
increase in frequency and intensity in Europe and the number of people at risk is
projected to grow, too. However, within the next 20 years projected changes in the
intensity and frequency of extreme events- depending on the time scale and hazardremain uncertain. The most severe effects of human-caused climate change are
expected in the second half of the century (IPCC, 2007a).
Predicting the future effects of extreme events remains difficult, particularly due to
increasing exposure caused by changes in economic development, which increases the
value and density of human and physical capital. Therefore disaster losses would rise
at a speed more rapidly than average economic growth, stressing the importance of
risk reduction (Bouwer et al, 2007).
Nonetheless, the Swiss Re has estimated that in Europe the costs of a 100-year storm
event could double by the 2080s with climate change ($50/€40 billion in the future
compared with $25/€20 billion today), while average storm losses are estimated to
increase by only 16 – 68 % over the same period (Heck et al., 2006).
The possible future increases in damage can enhance the vulnerability of the
insurance sector and will have important implications for the role of financial services
under climate change (IPCC, 2007b). In high-risk areas people will experience
increasing difficulty or costs to get adequate insurance. These are likely to lead to
greater levels of uninsured assets, particularly to socially deprived groups, i.e.
exacerbating inequalities. Thus governments may need to consider new ways of
ensuring that especially poorer people, that may be affected most, still will be able to
have insurance and/or may be compensated for possibly increasing losses in future,
e.g. through public-private insurance constructions as previously introduced in
Belgium and has been proposed in the Netherlands (Bouwer et.al., 2007).
166
6.3 Normalised losses from river flood disasters
Key Messages
• Economic losses as consequence of extreme flood events have been dramatic in
recent years and economic losses from flooding disasters have significantly increased
in Europe during the 1990s and 2000s. The estimated losses in Central Europe in
2002 were EUR 17.4 billion. This amount is more than the GDP of Bulgaria in that
year. Recently, in the summer 2007 the cost of the floods in UK was estimated at
around EUR 4.3 billion.
• Although there is scientific evidence of an ongoing intensification of the water cycle
there is no homogeneous trend in extreme river flows in Europe. Analyses of longterm records of flood looses indicate that societal and economic factors clearly play
an important role in the observed upward trends in flood-related losses.
Presentation of the main indicator(s)
Figure 6.3.1. EU flood losses per thousand of GDP (in grey decadal average).
Source: Barredo (2007).
167
Figure 6.3.2. Casualties caused by flood disasters in the EU. Source: Barredo (2007).
Relevance
There is good reason to be concerned about the growth of flood losses in Europe even
without taking climate change into account. Economic losses from flood disasters
have increased in Europe from the 1970s to the 2000s (Barredo, 2007). In addition to
the rising trend in flood damage the effects of unusually severe floods during the
1990s and 2000s increased awareness of the economic consequences of flooding. The
1997 floods in Poland and Czech Republic are responsible for losses of about EUR
5.2 billion. In 2000 Italy, France and Switzerland experienced losses of EUR 9.2
billion. And in 2002 the material flood damage recorded in Germany, Czech Republic
and Austria of EUR 17.4 billion has been higher than in any single year before
(Kundzewicz et al., 2005). Recently, in summer 2007 the cost of the floods in UK was
of about EUR 4.3 billion.
There is no clear evidence for a climate-related trend for floods during the last
decades in Europe (Mudelsee et al., 2003; Kundzewicz, 2005). Even if there is
scientific evidence of an ongoing intensification of the water cycle (Huntington, 2006)
there is no homogeneous trend in extreme river flows at European or regional scale.
Analyses of long-term records of flood looses indicate that societal and economic
factors play an important role in the observed upward loss trends (Pielke Jr and
Downton, 2000; Mills, 2005; Barredo, 2007).
Past trends
Flood disasters in Europe have increased in number and amount of loss from the
1970s to 2000s. The number of major flood disasters during the last 16 years
(between 1990 and 2005) is more than twice as large as between 1970 and 1989
(Barredo, 2007, see also Chapter 5). Figure 6.3.1 shows the normalised flood losses in
Europe. For the assessment of flood losses it is important to compensate for changes
in asset values and exposure over time. Failing to adjust for economic factors yields
loss amounts that are not directly comparable over time and a pronounced ever168
increasing trend for purely economic reasons (Höppe and Pielke Jr, 2006; Muir Wood
et al., 2006). In figure 6.3.1 a continuous increase is observed in the decadal average
of flood damage normalised by GDP. GDP has been used as surrogate measure of
exposure since other direct measures are not available for all the assessed countries.
Current evidence indicates that the growth of flood losses in recent decades is related
to both societal factors and climatic factors, but the shares are unclear (Pielke Jr and
Downton, 2000; Barredo, 2007). In fact in the period 1970-1999 the trend was not
statistically significant, and the increase registered in the last sub-period is a
consequence of one single event, the floods in Central Europe of the summer of 2002.
Indeed the significance of the observed upward trend is influenced by this large flood
disaster. However, because of a series of issues it is still not possible to determine the
proportion of the increase in damage that might be attributed either to climate change
or societal change and economic development (Höppe and Pielke Jr, 2006). There is
agreement in the fact that climate change cannot be understood as the dominant factor
for increasing flood losses. In fact there are no conclusive studies that confirm the
hypothesis of changes in the occurrence of extreme river flows in Europe. In a
hypothetical scenario without climate change total flood losses will continue to
increase as consequence of societal and economic factors (Pielke Jr and Downton,
2000).
Figure 6.3.2 shows the yearly number of casualties produced by floods in Europe for
the period 1970-2005. It can be stated that there is not a clear trend concerning this
indicator. Indeed, the number of deaths is very much dependent on single events, as it
is the case for the events of 1970 in Romania and Hungary, 1973 in Spain, and 1998
in Italy. In the last few decades early warning systems and prevention measures have
improved evacuation mechanisms in the areas exposed to floods.
Projections
Some preliminary estimates (ABI, 2005) indicate that annual flood losses in Europe
could rise to €100 – 120 billion (tenfold) by the end of the century (though flood
management could reduce this). Hall et al. (2005) presented a national-scale
assessment for England and Wales. Their results predicted an up to 20-fold increase in
losses by the 2080s in the scenario with the highest economic growth, and no
adaptation. These results include changes in sea level rise, increasing precipitation
and increasing economic vulnerability. More detailed disaggregated work under the
PESETA project (Feyen et al., 2007) has modelled changes in river flows in a
changing climate in Europe, studying two river catchments (Upper Danube and
Meuse) in detail.

For the Upper Danube the estimated total damage of a 100-year flood is projected
to rise by around 40 % of the current damage estimate (an increase of €18.5
billion) for the high emission scenario (A2) and around 19 % for the intermediate
emission scenario (B2) by 2100. The number of people affected in the Upper
Danube is projected to increase by 242,000 (around 11 %) for the A2, and
135,000 (around 6 %) for the B2 scenario.

For the Meuse, the potential damage of a 100-year flood is projected to rise by
about 14% for the A2 scenario and about 11% for the B2 scenario.
169
These regional studies have been expanded with European wide studies on river
flooding. The figure below shows the change in damage (averaged over nuts3 level)
for floods with 100 year return period for the SRES A2 scenario.
Figure 6.6.3. Change in damage (in €, averaged over nuts3 level) for river floods with
100 year return period for the A2 scenario.
A number of uncertainties should, however, be highlighted with these river catchment
and Europe-wide results. First, the numbers are the combined effect of the climate and
socio-economic effects, and second, they do not include existing or any future flood
protection measures, so strictly speaking they are a measure of potential exposure, not
impacts (though they may underestimate potential losses by not incorporating changes
in exposure). This highlights a broader issue of climate and socio-economic analysis
for future flood risk. Research into flood risks in The Netherlands indicates that
potential economic losses from flooding (river and sea) under socio-economic change
could increase by 22/45% in 2040 (WL Delft Hydraulics, 2007). The particular role of
climate change was not taken into account, because of unknown effects on flood
severity and frequency. Moreover, socio-economic factors are expected to dominate
future loss records, and will continue to complicate normalisation studies, because of
the large inaccuracies associated with actual loss estimates, compared to geophysical
data on extreme weather itself (Pielke Jr, 2007).
170
6.4. Energy
Key Messages
 Climate change will alter the relative balance between energy used for space heating and
cooling, while energy use for space heating is currently dominating the use for cooling in
Europe (particularly in the North).
 Historical data on heating degree days shows a fall in recent years in Europe, indicating a
benefit from reduced space heating. Actual energy demand from these changes is also
determined by technical and socio-economic factors. At present, no data is available on
cooling degree days across Europe, though country specific data shows some increases in
cooling degree days over the same period, consistent with greater space cooling demand.
 Future projections of climate change suggest reductions in heating degree days in Europe,
but increases in cooling degree days. The net change in energy demand is difficult to
predict, but there will be strong distributional patterns, with significantly reduced space
heating demand in northern Europe against increased space cooling demand in Southern
Europe with associated costs and benefits. There may also be increases in energy demand
associated with adaptation to climate change, e.g. for water supply.
 The projected change in river runoff due to climate change will result in an increase of
hydropower production by about 5 % and more in northern Europe and a decrease of about
25 % or more in the south.
 Climate change could have a negative impact on the thermal power production as most
studies show that summer droughts will be more severe, limiting the availability of cooling
water in terms of amount and adequacy of the temperature and power plant efficiency.
 Dam safety may be affected under changed climatic conditions with more frequent
extreme flows and possibly natural hazards.
Heating and Cooling Demand
Figure 6.4.1: Development of Heating Degree Days (HDD) in Europe, providing an
indication of the number of days when heating might be required, based around thresholds of
temperature. The relative degree days are weighted by population or area. The HDD figures
for individual member states vary considerably, with much higher HDD values for
Scandinavian countries, and much lower ones for Southern European countries, though there
is a downward trend across both regions.
Source: Eurostat and Joint Research Center. Eurostat Unit G4 'Energy Statistics' and JRC
IPSC/Agrifish Unit/MARS-STAT Action".
171
Relevance
Energy industries are the single most important source for greenhouse gas emissions
in Europe. The energy sector will also be affected by climate change. Numerous
studies have demonstrated that energy demand is linked to climatic conditions, with
changing demand for winter heating and summer cooling according to outside
temperature (particularly in the domestic sector, but also in service and industry
sectors). Data show higher energy consumption (gas and oil) in severe winters in the
domestic sector, and higher electricity consumption in high summer temperatures,
especially in southern European countries (Eurostat, 2007). The changing climate in
Europe is likely to lead to a decrease in the demand for winter heating, but an increase
in summer cooling (which can be described as either an impact or an adaptation that
in turn reinforces climate change impacts).
Note there are also other factors that affect the apparent temperature, and so the actual
energy demand, including wind chill, illumination and cloud cover, and other factors
that affect demand including precipitation.
Past trends
Energy demand has risen very strongly in Europe over recent years, due to technical,
socio-economic and economic factors (Eurostat, 2007). Actual final energy
consumption for heating since 1997 has been persistently below the projected
temperature-corrected consumption. This suggests warmer-than-average years at EUlevel, which is confirmed by the information in heating degree-days. The HDD data
shows that recent years (since 1996) are all lower than the long-term average. Note at
present, the net energy demand in Europe is dominated by space heating rather than
cooling. However, space heating is strongly influenced by technical, socio-economic
and economic factors, and it is difficult to separate (normalise) the specific effect of
temperature from these data5. There is currently less data available on space cooling
demand at an EU level. Space cooling relates to human comfort levels, but also
cooling for appliances.
The heating degree day indicator above shows a falling trend reflecting the recent
warmer years. This translates into a lower winter heating burden (a benefit) due to
climate change. However, some care is needed in interpreting these changes as actual
changes to energy demand and the economic consequences, because of technical,
socio-economic and economic factors.
Projections
HDD and CDD (cooling degree days) are parameters that can be derived from results
of global climate models. The forecast trends in Europe suggest further reductions in
heating degree days, and further increases in cooling degree days, due to mean
average temperature increases over future years. For cooling, there may be additional
peaks associated with heat-waves.
For Europe, the overall changes in energy and economic costs (at a net level) are
predicted to be modest in the short-medium term, due to the aggregated effects of
decrease winter heating demand vs. increased summer cooling demand. However,
5
For example, the effects of population, housing density, housing stock, insulation levels, technology,
equipment penetration level, efficiency of heating or cooling units, behaviour, perceived comfort
levels, energy prices, income
172
strong distributional patterns are expected across Europe – with rising cooling
(electricity) demand in summer in Southern Europe, compared to reduced heating
(energy) demand in winter in Northern Europe (Alcamo et al, 2007) – summarised in
the figure below. This translates in likely net benefit to Northern Europe because of
the high current space heating levels, but net negative impacts for Southern Europe.
Figure 6.4.2. Projections of energy demand for various time horizons in Europe
Source: Information from Alcamo et al, 2007. IPCC AR4
The actual net economic costs are more complex to estimate, due to differences in
energy sources6. Winter heating demand is primarily from fossil fuel use, whilst
summer cooling from electricity, and there may be additional issues of peak demand
levels in Southern Europe in the summer7. Adaptation has a role to play here –
particularly through alternatives to mechanical air conditioning, e.g. through passive
ventilation, building design, planning, etc. and is a key aspect to mitigation –
adaptation linkages.
6
There are complex issues in predicting future energy demand and prices, because of the need to
predict future energy and electricity prices (under future mitigation scenarios), and because of the
complex relationships between technology, socio-economic trends, etc.
7
Whilst the overall energy balance may not change that greatly in Europe from climate change, e.g.
with the reduction in energy for winter heating vs. the increase in energy for summer cooling, there
could still be important economic effects. Winter space heating is provided by fuels (coal, oil, gas) that
can be stored. Summer cooling is provided by electricity, which cannot be stored very easily, and so
electricity generation must match electricity demand. A rise in peak summer electricity demand,
associated with cooling in southern Europe, could increase the plant peak capacity needed, which
would be expected to lead to higher marginal prices at times of peak demand, e.g. during peak summer
temperatures and heat-waves.
173
Finally, there may also be an emerging issue of energy use increasing from water
supply (pumping, desalination, recycling, irrigation, water transfers). Again, these are
likely to be greater in Southern Europe where overall precipitation levels are predicted
to fall. There is also the potential for extreme weather events (e.g. storms) which also
increase the risk of energy infrastructure failure.
Hydropower
The production of electricity is strongly dependent on water, both for cooling in
power plants and for hydropower. In some areas, hydropower may benefit from
increased hydropower potential, while in other countries this potential will decrease
due to reduced river runoff. The generation of electric power in thermal (in particular
coal-fired and nuclear) power stations often relies on large volumes of water for
cooling. During heat waves and drought periods the use of cooling water may be
restricted if limit values for temperature are exceeded, which may force plant
operators to work at reduced capacity or even temporarily close plants, with
potentially serious consequences for supply.
In 2005 hydropower contributed 16.3 % of the electricity production in the European
countries, 10.2 % in EU-25 (Eurelectric, 2007). The share of hydropower in the
electricity production is generally high in the northern countries and countries in the
Alps. In 2001, the EU agreed that in 2010 21 % of the total electricity production
should come from renewable resources (EU, 2006). In 2005, the share of renewable
energy sources was 15 %, out of this hydropower represented 2/3 (Eurelectric, 2007).
Dam and reservoir safety may be affected under changed climatic conditions by more
frequent extreme flows. Evaluating changes in reservoir safety is complex; a flash
flood may be critical to a small reservoir or river weir, but might not influence a large
reservoir at all, as it would be more vulnerable to longer episodes. Regional studies of
climate change impacts on reservoir safety show varying results (Veijalainen and
Vehviläinen 2006; Andréasson et al. 2006).
Over the 20th century, annual river flow have shown an increasing trend in northern
parts of Europe, with increases mainly in winter, and a slightly decreasing trend in
southern parts of Europe (Section 5.7.1 River flow). These changes are linked with
observed changes in precipitation patterns and temperature. To the extent climate
change affects river flow, it will also affect hydropower production.
Since the 1970s, annual energy production of some existing hydropower stations in
Europe has decreased, in particular in Portugal, Spain and other Southern European
countries (UCTE 1999). This reduction has been attributed to changes in average
discharge; but whether this is due to temporary fluctuations or already the
consequences of long-term changing climate conditions is not yet known (Lehner et
al., 2001). .
The two most relevant European studies directly addressing hydropower are the
EuroWasser study from 2001, covering all Europe (Lehner et al. 2001, 2005), and the
Nordic Climate and Energy study, covering Scandinavia, Iceland and the Baltic states
(Bergström et al. 2007). Both studies use results from both the ECHAM4 and
HadCM3 models. EuroWasser analyses one “business-as-usual” emission scenario
(IPCC-IS92a), while the Climate and Energy study analyses both the IPCC SRES A2
and B2 scenarios. Another regional study is GLOWA-Danube, aimed at establishing
174
an integrated decision support system for water management in the upper Danube
(upstream Passau) under global climate change, including the hydropower sector
(Mauser and Strasser 2007).
The EuroWasser study demonstrates a clear north-south gradient, generally giving
increased hydropower production in northern Europe and reduced production in the
southern part. Although there are large local differences between the outcomes for the
two models, especially in the Alps and part of the Mediterranean region, both show
increases up to 25 % or more in north Europe, and reductions by 25 % or more in
southern parts by 2070.
The Climate and Energy simulations also project increases in hydropower production
in Scandinavia for both emissions scenarios and both climate models, but with more
detail and another regional distribution due to the use of RCMs (Regional Circulation
Models) for downscaling. Generally the increase is largest in the western coastal
regions. The figure below shows the hydropower production by regions for the
reference period (1961-1990) and for 2070-2100 based on the two GCMs (Global
Circulation Models).
Figure 6.4.3. Changes in hydropower production in Scandinavia. In addition to
climate change impacts, this indicator will reflect changes in hydropower system
configuration and changes in demand pattern.( Source: Mo et al (2006))
Cooling water for power plants
Decreased precipitation is expected to have a negative impact on the electricity
generation sector where rivers provide the cooling water. Power stations have to be
shut down when water temperatures exceed certain thresholds. Electricity production
has already been reduced in various locations in Europe during very warm summers
(e.g. 2003, 2005 and 2006) (BMU, 2007; Lehner et al., 2005).
During the 2003 hot summer the cooling capacity of several power stations in the
Netherlands and France was threatened as a result of the high water temperatures and
175
Figure 6.4.4. Projected changes in hydropower production potential in Europe, for one
scenario (XXXX) and two models. Blue denotes increase, red reduction. (Source:
Lehner et al (2001; 2005)).
low river level. The requirement that cooling water discharge may be no warmer than
30C meant that several companies could only satisfy this criterion by reducing their
production capacity. The three hydroelectric power stations on the Meuse, Nederijn
and Vecht also had to run on a very limited capacity for several weeks (10–25 % of
normal). The combined result was a threatened shortage of electricity in the
Netherlands. A water temperature of 23°C applies as the critical limit for the intake of
cooling water. Due to the temperature rise that has already occurred in recent
decennia, the number of days in the year that the water temperature is above 23°C has
also increased. In 2003, a tight situation even arose for a period of almost 40 days
when the water temperature was above 24°C.
It is highly likely that electricity companies will experience greater problems with
their cooling water systems due to the rise in temperature and more frequent low
discharges.
Figure 6.4.5. Trend in the number of days per year when the temperature of the water
in the Rhine was higher than 23°C during the period 1909–2003 (Source: MNP, 2006)
176
6.5. Coastal Areas
Key Messages
• Coastal flooding can lead to important climate-related losses. By 2100, exposed population
in the main coastal European cities to the combined effects of sea level rise and storm surge is
expected to be about 4 million people and the exposed assets more than 2 trillion € (without
adaptation).
• Future projections of sea level rise and storm surge show potentially large increases in the
risk of coastal flooding. These could have potentially large economic costs, with recent
estimates of 18 billion Euro/year for Europe in 2080 under the IPCC SRES A2 scenario. The
same estimates indicate that adaptation could significantly reduce this risk to around 1 billion
Euro.
Relevance
Coastal zones in Europe contain large human populations and significant
socioeconomic activities. They also support diverse ecosystems that provide
important habitats and sources of food. One third of the European Union (EU)
population is estimated to live within 50 km of the coast, and some 140,000 km2 of
land is currently within 1 m of sea level. Significantly inhabited coastal areas in
countries such as the Netherlands, England, Denmark, Germany and Italy are already
below normal high-tide levels, and more extensive areas are prone to flooding from
storm surges.
Climate change is an additional pressure and is likely to have significant impacts on
coastal zones, particularly via sea level rise and changes in the frequency and/or
intensity of extreme weather events, such as storms and associated surges.
Projections
There are estimates of the physical impacts and economic costs to coasts in Europe
from sea level rise and flooding storm events. Results using the DIVA database and
model produced from the DINAS-COASTS DG research project ((DINAS-COAST
Consortium, 2006; Hinkel and Klein, 2007; Nicholls et al., 2007a; Vafeidis et al.,
2004; 2007) have been developed for Europe in the PESETA project (Richards and
Nicholls, 2007). They show impacts increasing dramatically without adaptation: in the
2080s under the A2 SRES scenario and estimate around 19,000 km2 of land in Europe
could be permanently lost, leading to some 1.4 million people in Europe experiencing
coastal flooding each year. The economic costs of these events are estimated at 18
billion Euro/year (current prices)8. Large areas of coastal wetlands are also threatened,
with highest relative losses on the Mediterranean and Baltic Coasts.
However, adaptation has significant benefits. These strategies include (Nicholls et al.,
2007b): coastal defences (e.g. physical barriers to flooding and coastal erosion such as
dikes and flood barriers); realignment of coastal defences landwards; abandonment
(managed or unmanaged); measures to reduce the energy of near-shore waves and
currents; coastal morphological management; and resilience-building strategies.
Despite some difficulties in estimation, there is an extensive literature reporting the
direct cost of adaptation to sea level rise and even estimating the optimal levels of
protection (based on cost benefit analysis. (e. Tol, 2004; Anthoff et al., 2006)
8
This includes the combined effect of climate and socio-economic scenarios.
177
Figure 6.5.1. People flooded (thousands/year) across Europe (coastal areas)
Top Baseline. Bottom. for the IPCC SRES A2 scenario, 2080s (ECHAM4), without
adaptation (Richards and Nicholls et al, 2007 from PESETA)
Richards and Nicholls, 2007). Using the same climate and sea level projection as
above (A2 scenario in the 2080s), with adaptation included and optimised, the
DINAS-COAST Consortium (2006) suggests that the land loss falls to less than 1,000
km2 (and the economic costs fall from 18 billion to around 1 billion Euro/year). There
are costs of adaptation (coastal protection), estimated at some 1 billion Euro/year also
– but these achieve significant reductions in the residual damages.
Recent work for the OECD (2008) has also looked at current and future major coastal
cities with sea level rise (0.5 metres global average) and storm surge, and assessed
exposure to a 1 in 100 year flood event, looking at population and asset value exposed
now and with sea level rise in 2100 for Amsterdam, Rotterdam, Hamburg, London,
Copenhagen, Helsinki, Provence, Athens, Napoli, Lisbon, Porto, Barcelona,
Stockholm, and Glasgow. For these cities, the exposed population rise from 2.3
million to 4.0 million, and the exposed assets from $360 to $2220 billion (though the
values are dominated by London, Amsterdam, and Rotterdam).
178
6.6. Agriculture and Forestry
Key Messages
• The hot summer of 2003 in Europe is estimated to have led to $15 billion in
economic losses to farming, livestock and forestry from the combined effects of
drought, heat stress and fire. In the future, extreme events as a consequence of
climate change are expected to increase..
• Climate-related increases in crop yields are expected mainly in northern Europe, but
with reductions in the Mediterranean and the south-west Balkans. These show that
south and west Europe could experience a decrease of yields of 10% or more, though
there is an equivalent improvement of yields in Nordic countries.
• There are likely to be future changes in forest growth – and so economic
consequences - with climate change, though projections of future net changes in
Europe are uncertain. There are also observed changes in forest fires, and these lead
to economic costs through losses and through the costs of fire fighting. Forest fires
are likely to increase under a warmer climate, with increased economic costs.
Agriculture
Agriculture accounts for only a small part of gross domestic production (GDP) in
Europe, and it is considered that the overall vulnerability of the European economy to
changes that affect agriculture is low (EEA, 2006). Nonetheless, the agriculture
sector has a strong influence on other sectors, and moreover, effects of climate change
may still be substantial at a European level because of the spatial distribution of
changes. Agricultural indicators were included in 5.6, and historic trends are well
captured by Eurostat. The overall economic indicators are partly related to total yield
and market prices, though complicated by many other factors9.
Climate model studies indicate greater stresses will become apparent in southern
European (Mediterranean) and southerly eastern European countries with a changing
climate, due to the larger climate signals that these areas receive (with higher than
average increases in temperature for Europe), and also greater reductions in summer
water availability (and perhaps increases in drought), leading to lower yields.
Agriculture is a more significant sector for these countries in terms of employment
and GDP, which could compound these effects10. In contrast, the agricultural systems
in Western Europe are considered to have lower sensitivity to climate change, and the
modelling predictions show likely opportunities (yield increases and wider
agricultural crops) for Northern Europe. The recent IPCC 4th assessment report
(2007b) concludes that in Northern Europe, climate change is initially projected to
bring mixed effects, including some benefits such as increased crop yields and
increased forest growth. However, as climate change continues, its negative impacts
are likely to outweigh its benefits.
Most of the analyses now build in (autonomous) adaptation, reflecting a likely trend
of producers altering practices and even crop types by region as climate changes.
Several studies show the likely spatial patterns outlined above, with a strong
distribution of yield changes across Europe, as found in the recent PESETA project,
9
For example, agricultural subsidies, labour and production costs, global price changes, production
improvements, consumer demand, etc.
However, there will be socio-economic development at the same time as Europe’s climate changes. This highlights the need to
consider a changing climate alongside future projections of socio-economic development. Moreover, agricultural efficiency and
productivity are likely to increase with technological development.
10
179
(Iglesias and Garrote, 2008), which has forecasts for regional yield changes for 2020
to 2080. The analysis gives indications of the general spatial pattern of changes in
agriculture yields across Europe using two different models. These show that south
and west Europe could experience a decrease of yields of 10% or more, though there
is an equivalent improvement of yields in Nordic countries11.
Figure 6.6.1. Simulated crop yield changes by 2080s relative to the period 1961-1990
according to a high emission scenario (IPCC A2) and two different climate models:
(left) HadCM3/HIRHAM, (right) ECHAM4/RCA3
Source: PESETA project. http://peseta.jrc.es/docs/Agriculture.html. A. Iglesias/L. Garrote
The IPCC (Alcamo et al, 2007) summarises that climate-related increases in crop
yields are expected mainly in northern Europe, e.g., wheat: +2 to +9% by 2020, +8 to
+25% by 2050, +10 to +30% by 2080, while the largest reductions of all crops are
expected in the Mediterranean, the south-west Balkans and in the south of European
Russia. In southern Europe, general decreases in yield and increases in water demand
are expected for spring sown crops. The impacts on autumn sown crops are more
geographically variable; yield is expected to strongly decrease in most southern areas,
and increase in northern or cooler areas (e.g., wheat: +3 to +4% by 2020, -8 to +22%
by 2050, -15 to +32% by 2080).
Changes in agricultural productivity result also in changes in subsidy dependency.
Currently, however, there is no detailed data on subsidy distribution by crop and
region available which would allow analysis of the relationship between climate
change, yield and subsidies.
Recent valuation studies in the UK predict increases in yield and also revenue in the
2020s, but with these declining by the 2050s and with revenue changes becoming
11
The results show the increase in growing season, but also higher minimum temperatures in winter
lead to higher yields in higher latitudes, ranging from +3% to +70%. On the downside are the
Mediterranean countries with a shortening of the growing season and decreases between -2% and 22%, all findings subject to considerable uncertainty.
180
negative in nearly all regions by the 2080s with expected economic losses up to £24
million/year (Hamilton et al, 2006), particularly in more southern areas where water
becomes increasingly limiting. The study also indicated that similar agricultural losses
could occur as a result of flooding (without adaptation).
However, while these models generally consider the effects of projected changes in
temperature and CO2 fertilisation, they do not fully consider issues of water
availability, and rarely consider extreme events. The latter could be important for
Europe in relation to heat extremes and floods. As an example, the droughts of 1999
caused losses of more than Euro 3 billion in Spain (EEA, 2004) and the hot summer
of 2003 in Europe is estimated to have led to $15 billion in economic losses to
farming, livestock and forestry from the combined effects of drought, heat stress and
fire (Munich Re, 2004)12.
Finally, the role of autonomous and planned adaptation is extremely important for
agriculture – and has been studied more intensively for this sector than any other
(with the possible exception of coastal defences). While most analysis considers
short-term autonomous adaptation (to optimise production) there are also potential
long-term adaptations in the form of major structural changes to overcome adversity
caused by climate change13. These are usually the result of a planned strategy. There
are a number of studies that show the benefits of adaptation to farmers in reducing
negative impacts by at least 20 %, and even turning losses into gains (though it is
highlighted that such studies rarely explicit provide a cost for adaptation).
Forestry
Forestry is also a small part of European GDP, though in a large part of Europe,
forestry represents an important economic sector and also provides potential for
carbon sequestration. Forests in Europe are likely to be affected by climate change,
both in terms of distribution and species composition. Alcamo et al (2007) expect
that forest area will expand in the north, but contract in the south, so there will be
positive and negative effects (see 5.6 (Agriculture and Forestry).
Eurostat measures a number of indicators for forests14, some of which can at least
partly be linked to climate change, e.g. physical impact indicators of forest yield and
condition, and windstorm damage and forest fires. While the economic effects of
timber production can be captured using market prices, forests (natural and managed)
play a much greater role than timber alone, and there is a need to progress towards the
total economic value of forestry including full ecosystem goods and services and user
and non-user values.
In addition to yield changes, there will be potential economic consequences from
forest fires (see 5.6). Forest fires are likely to increase under a warmer climate, with
an enlargement of the fire prone area and a lengthening of the fire season. As well as
the economic losses (lost production) from forest fires, there are also direct costs of
fire fighting. These were evident in the summer 2003 heat-wave in France, which
increased the costs of fighting forest fires (for the Ministry of interior) to 179 Million
Euros, from 83 Million Euros in a normal year (Gillet, 2006).
12
Though overall positive effects on the UK agricultural, fruit and viticulture industries are also estimated to have occurred
(Metroeconomica, 2005), with estimated economic benefits of £64 million, though this included a mix of positive and negative
effects – though the authors note that it is not possible to conclude with any confidence that these gains / losses are wholly
attributable to the weather conditions that prevailed in the summer of 2003.
13
Adaptation can also be undertaken at different scales i.e. farm level, regional level and national level. Note there are
differences between models in the way that adaptation is included, e.g. between a spatial / Ricardian, or a structural approach.
14
Eurostat Pocketbooks – Forestry statistics 2007 edition
181
6.7. Tourism and recreation
Key Messages
• Changes in climate are starting to impact upon the attractiveness of many of the
Mediterranean’s major resorts, whilst improving at the same time in other regions.
• Future projections of climate change suggest that the Mediterranean tourism
suitability will decline during the key summer months, though there will be an
increase during other seasons (Spring and Autumn). This can produce shifts in the
major flows of tourism within the EU, which will be important in regions where
tourism is a dominant economic sector, though adaptation responses such as
economic diversification will be critical.
• The tourism industry will face significant adaptation costs such as those incurred for
producing artificial snow or implementing economic diversification. Adaptation
measures will be driven by climate change and socio-economic-related factors and
their sustainability will also have to be assessed in view of the associated
environmental impacts.
Summer tourism
Mass tourism is closely associated with climate, for both the source of tourists and
their destination. At present, the predominant (summer) tourist flows in Europe are
from north to south, to the coastal zone, which helps to transfer capital. However,
coastal and mountain tourism are the segments that are most vulnerable to climate
change in Europe, and the Mediterranean region is the world’s most popular holiday
region: it attracts some 120 million visitors from Northern Europe each year, the
largest international flow of tourists on the globe and their spending is in excess of
100 billion Euros.
The changes in and differences between the suitability of regions (or specific
locations) for summer-time tourism can be assessed by using the Tourism Comfort
Index (TCI). The TCI is an index, ranging from 0 to 100, based upon a range of
climate variables that reflects the suitability related to an individual’s bioclimatic
comfort. It can be used to analyse the variability in tourism attractiveness between
locations.
There are large differences within Europe and between seasons as to the attractiveness
for tourism. During the key summer months (June, July and August) the
Mediterranean has a close-to-ideal climate for tourism, with TCI values close to the
maximum value of 100. This ideal climate is one of the important drivers for the
current holiday market, next to cultural, social, landscape and other factors.
With growing income and increasing leisure time, the tourism industry in Europe is
expected to continue to grow. However, temperature rise is likely to change summer
destination preferences in Europe. Seasonality is a key issue in tourism, and the
summer months are the dominant period for the Mediterranean region and Europe as a
whole. The effect of climate change is likely to make outdoor activities in northern
Europe more attractive, while summer temperatures and heat waves in the
Mediterranean, potentially exacerbated by water supply issues, may lead to a
redistribution or a seasonal shift in tourism away from the current summer peak.
182
2071-2100
1961-1990
Figure 6.7.1. Simulated Conditions for Summer Tourism in Europe for 1961-1990
(left) and 2071-2100 (right) according to a High-Emissions Scenario (IPCC A2)
Source: PESETA project. http://peseta.jrc.es/docs/Tourism.html. (P. Martens/B.
Amelung/A. Moreno, 2008).
Data from climate change model experiments show a shift northward of tourism
during the 21st Century and the increasing bi-modal distribution of tourism over the
seasons in the Mediterranean (i.e. either side of a significant dip in summer). At the
same time northern European locations (the current source area for most tourism)
shows increasing attractiveness for tourism. The PESETA analysis shows the
direction of potential tourism shifting towards the end of the century (2071-2100)
under the IPCC SRES A2 high emission scenario. The maps indicate significant
potential shifts in the climatic suitability for tourism, with the belt of excellent
summer conditions moving from the Mediterranean towards northern Europe.
However, the reduction in attractiveness of current summer resorts is likely to be at
least partially offset by increased opportunities for tourism in northern Europe. In the
shoulder seasons (Spring and Autumn, not shown here), TCI scores are generally
projected to increase throughout Europe and particularly in Southern Mediterranean
countries, which could compensate for some losses experienced in summer.
The assessments above reflect the tourism suitability. Projections of the actual
changes that are likely to occur in the future, and the economic implications, are much
harder to assess. Much will depend on the flexibility of tourists and institutions such
as school holidays. If summer remains the predominant season for tourism activities
in Europe, major shifts of tourist flows may eventually occur from the Mediterranean
to more northern areas in Europe. However, shifts in the holiday season may be the
dominant form of adaptation. If these, as well as other societal changes (e.g. ageing
population), allow for a more flexible timing of holidays among a large share of the
population, then some of these effects may be offset, i.e. climate change may be
beneficial for the Mediterranean tourist industry if it levels out demand, reducing the
183
summer peak, while increasing occupancy in the shoulder seasons. In the absence of
such adjustments, the Mediterranean tourist industry will be among the losers. Other
work (Hamilton and Tol, 2006) has investigated the potential economic effects from
climate change on tourism. Their studies shows an increase in the number of inbound
tourists due to population and economic growth in the rest of the world, and that the
influence of climate change may be rather to change the rate of relative growth in
Northern regions of Europe compared with the Mediterranean. The study also shows
a potential shift towards a greater level of domestic tourism in regions with increasing
attractiveness (e.g. within the UK).
There are some other climate-related factors that could be important for tourism.
Water shortages due to extended droughts could potentially affect tourism flows,
especially in southeast Mediterranean where the maximum demand coincides with the
minimum availability of water resources.
Winter Snow Industry
There is also a major winter sports tourism industry in Europe, with the ski industry in
the European Alps and Pyrenees attracting millions of tourists each year. This
industry is a significant contribution to the economy of Alpine countries (OECD,
2007), generating close to EUR 50 billion in annual turnover. Studies project
widespread reductions in snow-cover over the 21 century (IPCC, 2007), see Chapter
5.3. These changes will affect the winter sports industry in Europe and its financial
viability, because of the availability of natural snow or conditions suitable for making
snow.
A recent study for the Organisation for Economic Cooperation and Development
(Abegg et al, 2007), reports that the number of snow reliable ski areas in Austria,
France, Germany, Italy, and Switzerland are projected to drop from approximately
600 to 500 if temperatures rise by 1.2°C, to approximately 400 if temperatures rise by
2°C, and to approximately 200 in a +4°C scenario.
There are already responses in place (e.g. artificial snow making, although adaptation
options also pose sustainability and environmental problems), and these increased in
recent years. For example, in France almost half a billion Euros were spent between
1990 and 2004 on artificial snow-making installations, while in Austria,
approximately EUR 800 millions were spent between 1995 and 2003. The
introduction of these machines is also driven by other socio-economic factors
(increasing the reliability of resorts to increase revenues and expand their ski areas
beyond previous natural limits). These measures have limits and the costs of their use
are likely to rise non-linearly as temperatures increases. The energy use and
associated greenhouse gas emissions add to the mitigation challenge. Furthermore, the
water use of snow-machine is increasing and already negatively affects current water
resources, which could be exacerbated in the future due to climate-induced decreases
in water availability. Other adaptation measures exist, including economic
diversification within or outside the tourism sector, e.g. diversification from winter
sports to other recreational or seasonal activities.
184
6.8.Health
Key Messages
• Climate change is likely to affect human health, either directly related to the
physiological effects of heat and cold, or indirectly, for example, through the
increased transmission of food-borne / vector-borne pathogens, or through the
flooding. These changes will have economic consequences, e.g. through medical
costs, lost time at work, and additional spending to avoid pain and suffering.
• Adaptation measures such as surveillance and outbreak control already exist, and
these build on well-establish public health approaches. Most adaptation measures
appear to be low cost, though further work is needed to assess the full cost of
adaptation.
The potential health effects of climate, such as from the 2003 heat wave, have been
well documented. There is some uncertainty over the net effects (the sum of heat and
cold effects) for Europe under a changing climate, and especially the distribution of
benefits across Europe. There are literature estimates (e.g. the WHO global burden of
disease (McMichael, 2004, IPCC, 2007, and the recent PESETA project: AEA, 2007):
the latter estimated that the economic effects of climate change in Europe could be
significant with potentially large economic costs (billions of Euro/year) from summer
mortality by the 2080s, though these will be offset by potentially larger economic
benefits from the reduction in winter mortality. However, populations may partly
acclimatise to future temperatures, and there is also the potential for adaptation
beyond this, for example with the current heat alert systems across many EEA
member countries (as implemented in France after the 2003 heat-wave, and also in
other member states). It is also highlighted, however, that heat related health
concerns are likely to lead to higher use of energy and air conditioning.
Climate sensitive infectious diseases, such as Salmonella, have the potential to
increase under a changing climate. Some emerging work (AEA, 2007, based on
Kovats, 2004) shows that the disease burden in Europe could be significant, and have
a potentially high cost (potentially several billion Euro a year by the period 20702100 through medical costs, lost time at work, willingness to pay to avoid pain and
suffering, and through the small number of cases of food poisoning that are fatal),
though adaptation offers a low cost means to reduce these.
Waterborne diseases may rise with increase in extreme rainfall due to sewage water
overflows. Higher water temperatures may also result in increased occurrence of
harmful algal blooms, and increase the faecal bacteria and incidence of pathogens.
This will affect drinking water intakes and water bodies used for recreation due to a
bad odour and taste of drinking water and occurrence of unsafe drinking and bathing
waters.
The increasing intensity of heavy rainfall is likely to make extreme floods more
frequent in some areas of Europe (see earlier section). While the number of deaths
and injuries from floods are relatively low in Europe, flood events do have wider
effects, notably in wider well being (mental health and stress and depression). There
is some emerging quantification and valuation of the latter well-being impacts (AEA,
2007, based on impact studies such as Reacher et al 2004), which show that without
adaptation, baseline costs from coastal flooding could be significant (several billion
per year). However, adaptation is likely to reduce these very significantly.
185
Climate-induced changes in the potential distribution of malaria are projected mainly
in poor and vulnerable regions globally. In Europe localised outbreaks are possible in
areas where the disease has been eradicated, but vectors are still present (Reiter et al.,
2004), though strengthening of effective surveillance and prevention programmes
(adaptation) should ensure these are minimised. There is emerging work assessing
climate change impacts on the distribution of vector-borne diseases in Europe
(malaria, leishmaniasis, West Nile virus) from the EDEN project.
Finally, there are a number of emerging health issues from climate change in Europe,
where quantification and valuation have not been explored. A warmer climate may
have important effects on air quality in Europe (for ozone formation). The seasonality
of allergic disorders may change with implications of direct costs in terms of over the
counter medications for allergic rhinitis, and wider economic costs to individuals.
Data on the costs of surveillance and outbreak control (adaptation costs) are starting
to emerge and there are adaptation strategies that can be implemented by health
sectors (e.g. see the cCASHh project), most of which are likely to build on wellestablished public health approaches, though further work is needed to fully assess the
costs of adaptation. Most adaptation measures appear to be low cost (e.g. provision of
information), but there is the potential for some to involve potentially costly largescale vaccination or other prevention programs against vector borne disease. Some
recent studies have considered the potential direct and indirect costs of health care
(e.g. Bosello et al, 2006) and show that these are likely to be relatively small for
Europe in terms of GDP. They also highlight that there are likely to be strong
distributional implications for climate change and health, with poorer countries being
either more exposed or more vulnerable.
186
6.9. Public water supply and drinking water
management
Key Messages
• Economic consequences of climate change impacts will be particularly pronounced
for those areas where increases in water stress are projected. The percentage area
under high water stress in Europe is likely to increase from 19% today to 35% by the
2070s, and the additional number of people affected by the 2070s is expected to be
between 16 millions and 44 millions.
• Evaluation studies of economic consequences of increasing water stress are now
emerging. They indicate that adaptation costs are generally significantly lower than
the economic losses incurred otherwise.
Changes in water demand strongly depend on economic and sectoral growth and
societal development. Household water demand is likely to increase with climate
change, with more water used for garden watering and personal hygiene, although a
clear separation exists between components that are sensitive to climate change
(showering, gardening, lawn sprinkling, and pool filling), and components that are
non-sensitive (e.g. dish washing, clothes washing). The general increase in wealth and
the generally hotter and longer summers may also increase the number of golf
courses, swimming pools and aqua parks, further increasing water demand. Problems
of water supply in tourist resorts are becoming increasingly common (see also
discussion in Tourism section earlier), in some cases tankers now have to transport
water to tourist islands, and the water deficits may be further increased by climate
change and increasing demand from other sectors.
Changes in the quantity and quality of river flows and groundwater recharge may
affect the drinking water supply systems (see section 5.7 Water quantity and 5.8
Water quality and freshwater ecology). Drinking water providers will be affected by
changed flow regimes and reduced annual water availability. Climate change may
alter the reliability of raw water sources by changing the frequency of low flows and
recharge of groundwater.
Hot summers such as 2003 may provide an indication of future climate impacts on
peak water demand. In the Netherlands public water supply increased by 15 % in
August compared with previous years. Other studies indicate that the increase of
household water demand may be rather small. Downing et al. (2003) concluded that
per capita domestic demand in England could rise by an extra 2 to 5 % during the
coming 20 to 50 years as a result of climate change.
During the summer 2007 Greece declared a state of emergency on the Cyclades
islands, including the popular holiday destinations of Mykonos and Santorini, because
of water shortages caused by a drought and heat-wave. Turkey's two major cities were
struggling with water shortages after record low levels of snow and rain in the winter
and searing summer temperatures. Reservoirs were down to less than 5% full in the
capital (Ankara, home to 4 million people) according to the country's water authority.
In March 2008 Cyprus was exploring the possibilities of transportation of water in
tankers from Lebanon after to survive chronic water shortage brought on by a twoyear drought and unseasonal warm weather.
187
High water temperature, low water flows and therefore less dilution of pollutants may
have severe consequences for the quality of drinking water and recreation activities
related to water. Saline intrusion in coastal aquifers making the water unsuitable for
drinking water may be exacerbated by future sea level rise.
These effects do have economic consequences, especially in areas where there are
predicted increases in water stress. Alcamo et al, 2007 (IPCC, WGII) predict that the
percentage area under high water stress in Europe is likely to increase from 19%
today to 35% by the 2070s, and the additional number of people affected by the 2070s
is expected to be between 16 millions and 44 millions.
Some studies on the economic consequences of increasing water stress are emerging.
Work in the UK has estimated the economic losses to households of foregone water
use due to the anticipated water deficit by 2100 in south-east England (Wade et al,
2006) at between £41 and 388 million a year (depending on scenario), but that the
costs of largely (but not entirely) eliminating these deficits would be only £6 to £39
million/year (effectively the costs of adaptation).
188
6.10. Nature and Biodiversity
Key Messages
• While methods for valuation of ecosystems are improving, it is as yet not possible to
cover the full range of ecosystem productivity, goods and services, and the economic
benefits to users and non-users.
The functioning of an ecosystem service provision from many natural and seminatural ecosystems in Europe are under threat from climate change and other
pressures (see Section 5.5). Such services include food and water supply, climate
regulation and species preservation. Particularly sensitive areas include the Arctic
region of Europe, mountain regions, and various coastal zones across Europe,
especially in the Baltic and parts of the Mediterranean.
These ecosystem services can be divided into supporting, provisioning, regulating and
cultural. Functions attributed to provisioning services have a direct market value. For
example, the ability of ecosystems to provide food (agriculture, fisheries), fibre
(timber) and fresh water to the population can be valued at market prices. Other
services, such as regulating and cultural services and the ability of an ecosystem to
provide natural habitat for flora and fauna and biodiversity loss, have however no
direct market price, though it is possible in some cases to approximate the value of
these functions.
Past and ongoing research tries to value ecosystem loss, reflecting ecosystem
productivity, goods and services, but also the wider use of ecosystems, increasingly
using the Millennium Ecosystem Assessment framework (MEA, 2005). MEA (2005)
uses the rate of extinction to illustrate some of the changes in ecosystem services.
This rate, though, cannot adequately monitor medium-term changes, as it tracks the
number of extinctions per thousand species per millennium. There is also a growing
body of more general economic studies on ecosystem and biodiversity. This includes
greater numbers of primary valuation studies (e.g. Eftec, 2002), and work studying
where biodiversity loss has led to the loss/degradation of ecosystem services and
consequently to economic costs (Kettunen and Brink, 2006). However, while
methods for valuation of ecosystems are improving, as yet they fail to cover the full
range of ecosystem productivity, goods and services, and direct and indirect economic
benefits to users and non-users. Nonetheless, there are some illustrative values
showing potentially very high estimates (e.g. IPCC, 2007). Hence, at this stage it is
extremely difficult to put forward accessible indicators for the economic effects on
ecosystems associated with climate change.
Following commitments made at the G8+5 meeting of Environment Ministers in
Potsdam in March 2007, preparatory work for a Review on the Economics of
Biodiversity Loss is taking place. The Review will evaluate the costs of the loss of
biodiversity and the associated decline in ecosystem services worldwide. It will
consider the failure to take protective measures, versus the costs of effective
conservation and sustainable use. The Review aims to achieve a better understanding
of why action to halt the loss of biodiversity would make economic sense.
189
6.11. The Costs of Climate Change for Society
Key Messages
• The total projected economic losses of the impacts of climate change are difficult to
assess, and the literature shows a very wide range of results. Due to the many
uncertainties involved, there is not one single “true” value that can be reported by
science to policy makers.
• Macro-economic and micro/sectoral economic assessments rely on different
methodologies for different levels of analysis and purposes. They provide
complementary estimates to better inform policy makers.
The costs of climate change will accrue to different individuals, in different sectors, in
different places, and at different times. Due to this complexity, the total projected
economic losses of the impacts of climate change (globally or in Europe alone) cannot
be easily assessed, and the literature shows a very wide range of results. In particular,
the quantification of economic damage is a difficult task. The transposition of
physical impacts into monetary terms is difficult and sometimes contentious step,
given that climate change impacts involve both market and non-market goods and
services, covering health, environmental and social effects, and potential large-scale
climatic events (potentially irreversible in nature). There are different ways that the
costs of inaction can be expressed. The most common ways of defining the cost of
inaction to climate change are either as ‘total costs’ or ‘marginal costs’.
Total costs are usually measured as the discounted aggregate of all future welfare
changes over some planning horizon. At the global level, there is an emerging
literature and studies have presented total costs of climate change impacts to the
world economy as a percentage change. For some regions, climate change could result
in economic benefit for some of the sectors in the short to medium term. However, the
evidence reported from the IPCC 4th Assessment Report is that the aggregated global
impacts of climate change will result in net costs into the future and these costs will
grow over time (IPCC, 2007b). At a global scale, previous economic estimations of
the costs of climate change impacts—as a result of rising sea levels, falls in
agricultural productivity and energy demand changes for instance—are up to around
2% of global GDP per year (EEA, 2007). But, recent studies and reviews have
indicated that these costs may be more significant (Ackerman and Stanton, 2006;
Stern Review, 2006). The Stern review, the British government’s prominent report
on the economics of climate change, takes a global perspective and provides
aggregate total costs from an integrated assessment model. It estimates that if
greenhouse gas emissions are not reduced, the total cost under a business as usual
scenario involves impacts and possible outcomes that will reduce welfare by an
amount equivalent to a reduction in consumption per head of between 5 and 20%.
A related and very useful concept is the marginal (social) costs of climate change.
Marginal costs of climate change are the additional damage costs of climate change
from a current emission to the atmosphere of one unit of greenhouse gas. The IPCC
(2007b) compiled the estimated marginal costs across some of the relevant studies in
the literature and it can be seen how wide the range of results can be. The estimates
ranges from -10 US$ to +350 US$ per tonne of carbon. Peer-reviewed estimates have
a mean value of US$43 per tonne of carbon with a standard deviation of US$83 per
tonne (IPCC, 2007b). It is also important to note that the marginal cost of climate
190
change is likely to increase over time, in line with the expected rising costs of climatechange-induced damage (Watkiss, 2006).
While this information is very valuable in informing climate change policy, it is clear
that there are many methodological issues involved in estimating the cost of inaction
(both in terms of total costs or marginal costs).
Climate change is comprised of numerous types of climatic parameters, which in turn
affect many sectors (market and non-market) in different ways. This leads to the
issue of coverage (or completeness). It is clear that different estimates of the costs of
climate change are based on different types of climate effects, and include different
impacts across varying sectors. The coverage issue has been framed in a risk matrix,
based on the different types of climate change and their uncertainty on the one hand
and the uncertainty in valuation, covering market effects, non-market effects and
socially contingent effects (a sub-category of non-market effects – defined as large
scale dynamics related to human values and equity) on the other hand (Watkiss,
2006). Mapping the literature estimates onto this matrix shows a large difference in
coverage between studies – and also reveals that most studies focus on the top left
area reflecting market damages from predictable events.
All current estimates of the costs of inaction are therefore incomplete, as they do not
cover all effects of climate change, across all impact categories, though we do not
know by how much (because the probability and consequences of many of the boxes
in the matrix are unknown). What is needed therefore is a recognition that the costs
have a wide range and policies should be designed so that they consider this
uncertainty. Also, it should be communicated clearly that there is not one single
“true” value out there which science could deliver to policy makers.
Figure 6.11.1. The coverage of marginal economic costs of climate change against the
risk matrix. (Source: Watkiss, 2006).
191
7. Adaptation
7.1 Adaptation is required even if global greenhouse gas
concentrations are stabilized
Until a few years ago, climate change impacts were something of the future, and more
relevant for other parts of the world than for Europe. Not anymore. They are being
observed for natural and social systems worldwide (IPCC, 2007a). This report
documents such impacts for vulnerable systems in Europe. The impacts are generally
projected to exacerbate in the future. In its latest, fourth assessment report, IPCC
considers the “reasons for concern”15 to be stronger than in its Third Assessment
Report in 2001. Many risks are now identified with higher confidence and some risks
are projected to be larger or to occur at lower increases in temperature. At the same
time, the understanding about the relationship between impacts (the basis for the
“reasons for concern”) and vulnerability (that includes the ability to adapt to impacts)
has improved. Some of the projected impacts can be abrupt or irreversible (IPCC,
2007b).
To limit the impacts, the European Union has agreed on a long-term goal to avoid that
the global mean temperature would exceed 2oC above pre-industrial levels. Even if
this goal would be achieved by very stringent world-wide mitigation action, some
impacts remain and vulnerability reduction and adaptation are imperative. Thus,
Europe has to adapt to climate change itself and has a moral obligation to assist
particularly the poor elsewhere (esp. in developing countries) –as they are most
vulnerable- in adapting., e.g. in the context of the Nairobi Five-year programme of
work on impacts, vulnerability and adaptation to climate change (UNFCCC, 2006),
the Bali Action Plan (UNFCCC, 2008) and the National Plans on Adaptation Action
(NAPAs). However, this chapter focuses on Europe itself.
Note that not always does climate change pose a threat. In some areas in the world,
for small temperature increases, without negative precipitation changes and
influenced by CO2 fertilization, agriculture can benefit. But also for other sectors
(e.g., tourism, energy supply, water management, construction and spatial planning,
shipping) it can also provide opportunities for innovations in areas or economic
structural development, technology and governance in many world regions. Capturing
such opportunities requires pro-active adaptation planning.
Many adaptation options are available, which are usually location and sector specific..
Adaptation is seldom undertaken for the sake of climate change alone. Mostly,
adaptation is integrated into broader policy action, such as disaster preparedness,
coastal zone management, rural development, health care services, spatial planning
and water management.
Not always does climate change pose a threat. In some areas in the world agriculture
can benefit, albeit only for small temperature increases without negative precipitation
15
Risks to unique and threatened ecosystems, risks of extreme weather events, distribution of impacts
and vulnerability, aggregate impacts, and risks of large-scale singularities.
192
changes and influenced by CO2 fertilization. But also for other sectors (e.g., tourism,
energy supply, water management, construction and spatial planning, shipping) it can
also provide opportunities for innovations in economic development, technology and
governance. Capturing such opportunities requires pro-active adaptation planning.
It is important to note that there are limits to adaptation as well. Natural systems often
have a lower adaptive capacity than human systems, especially when certain
thresholds – which are poorly but increasingly understood - are exceeded. But also for
human systems there will be limits. These are influenced by social, institutional and
economic constraints. With increasing climate change, adaptation costs will increase
and the options may decrease. Costs of adaptation can be significant, but are as yet
largely unknown (EEA, 2008).
7.2 Also Europe is vulnerable and will have to adapt
Because of a large body of research in Europe, much of the information underlying
the IPCC global assessment of impacts comes from Europe (see Figure 4.1). The
current report further elaborates and structures this knowledge on European impacts,
confirming that many European regions are vulnerable to climate change and that
impacts have already been observed for many vulnerable systems. Again, most of the
projected impacts are negative, certainly beyond a few decades. However, also in
Europe adaptation offers opportunities to make Europe more resilient to climate
change and as a side effect stimulating export of innovative adaptation technologies to
other regions.
To address the necessity of adapting to climate change in Europe, the EU has adopted
a Green Paper (EC, 2007) with four pillars: (a) early action by mainstreaming
adaptation into sector policies, (b) integrating adaptation into external relations,
including development cooperation, (c) strengthening research and applying the
knowledge basis, and (d) develop a participatory process to implement the strategy.
After a period of stakeholder consultations and an impact assessment of potential
options, a White Paper on a European adaptation strategy with a more concrete and
detailed set of actions is planned to be adopted in November 2008.
7.3 Different vulnerable systems at different geographic levels
require different approaches
European action on adaptation could take different directions, such as
 a formal legislative approach, requiring member states to develop and implement
strategies in a particular way, either as stand-alone adaptation policy or integration
in other policy areas;
 a supporting role, helping the member states in their adaptation actions only were
needed (e.g., transboundary issues, knowledge and tools development, structural
funds), or
 reporting requirements, only requiring member states to report about their
activities to encourage networking and information sharing.
In many cases a link with risk and disaster management will be appropriate. The
choice of the level of intervention and support will be different for different sectors,
193
taking into account the subsidiarity principle. Options will be further developed in the
context of a White Paper, expected in the fall of 2008.
In general, adaptation is a real crosscutting issue and integrating climate change
adaptation into sectoral policies is one of the key approaches to climate change
adaptation to be developed under the 1st pillar of the Green Paper. Integration of
climate change into other policy areas is aiming at protecting citizens and nature, and
making economic activities less vulnerable to climate change by appropriate
adaptation in sectors such as energy supply, transport, spatial planning, water
management, agriculture and nature protection. Figure 7.1 shows the differences in
emphasis between different regions in Europe. Most adaptation policies in Europe
currently focus on reduction of risk and vulnerability, with a smaller number of
policies aiming at improvements of adaptation capacity, increasing the capacity to
cope with extreme events, or to capture opportunities from climatic changes (Massey,
2007). Linkages between adaptation and mitigation have to be considered. Some
adaptation options can be developed in synergy with mitigation, e.g. in the land and
water management sectors. Other options can have negative impacts on mitigation,
because they can lead to additional greenhouse gas emissions.
Member states are at different stages of developing and implementing national
adaptation strategies, dependent on the magnitude and nature of observed impacts, the
assessment of current and future vulnerability, and the capacity to adapt. Several
countries are in the process of developing national adaptation strategies or have
developed one, including the Czech Republic, Denmark, Germany, Finland, France,
Hungary, the Netherlands, Portugal, the Slovak Republic, Spain, and the United
Kingdom.
194
Disease management
100%
90%
Financial mgmt.
80%
Development co-op
70%
Food production & security
60%
Biodiversity mgmt.
50%
40%
Energy/UPS
30%
Extreme temps (heat waves &
freezes)
20%
Water management
(quantity&quality)
10%
0%
W. Europe
S. Europe
N. Europe
Region
C. Europe
Landscape management (incl.
soil erosion, floods, fires)
Coastal zone management
Figure 7.1: Emphasis on different types of adaptation polices in different European
regions (Massey, 2007).
7.4 From European and national plans to regional and local
implementation
Such strategies provide the framework for adaptation actions many of which have to
be implemented at the regional and local level (regions, provinces, municipalities, or
even smaller scope). Various regionally-oriented initiatives are already underway in
Europe under the INTERREG programme that link research and policy development,
such as ASTRA (Developing Policies & Adaptation Strategies to Climate Change in
the Baltic Sea Region), AMICA (Adaptation and Mitigation – an Integrated Climate
Policy Approach), ADAGIO (Adaptation of Agriculture in European Regions at
Environmental Risk under Climate Change), BRANCH (Biodiversity Requires
Adaption in Northwest Europe under a CHanging climate) and ClimChAlp (Climate
Change, Impacts and Adaptation Strategies in the Alpine Space). National and
European action can provide or strengthen the enabling circumstances for regional
and local adaptation. National and European information sources such as this report
can provide the knowledge basis for identifying vulnerable areas and setting the
context for regional and local adaptation action. More particularly, the impact
indicators presented in the earlier chapters of this report can be related to the
economic sectors in the Green Paper that have to prioritize, develop and implement
adaptation strategies. The links vary between very weak and very strong (see Table
7.1).
195
Table 7.1: Linking the impacts indicators in this report to the Green Paper sector
adaptation
EEA impacts Atmosphere
GP
and
adaptation
climate
Cryosphere
Marine
and coast
Terr.
Ecosystems/
biodivers
ity
Agriculture and
forestry
Water
quantity
Water
quality
Human
health
Economic
loss
+++
o
+
++
+++
++
+
+
+
+
o
+
o
o
+
+
+
+++
+
o
o
o
+
+
o
o
+
+
+
o
o
o
+
o
o
+
Health
++
o
o
o
o
o
+
+++
+
Water
+++
+
+
+
++
+++
+++
+
+
o
+
o
o
o
+
Agriculture,
rural
development
Industry and
services
Energy
Transport
Marine and
+
+
+++
++
o
o
o
fisheries
Ecosystems and
+++
++
++
+++
++
++
++
biodiversity
Other resources
++
+
o
++
+++
++
+
(forests, soils)
+++: strong link, ++: moderate link, +: weak link, o: negligible link; link can be positive or negative
196
8. Uncertainties, data availability, gaps and future
needs
Numerous studies investigating climate change and its impacts on ecosystems and
society have been published since the release of the last EEA climate change indicator
report (2004). These have led to a significant improvement in our understanding of
climate change and its impacts as recently outlined in the fourth IPCC Report (IPCC,
2007). The number of indicators presented in this report also reflects the increase in
knowledge on climate change impacts in Europe and in capacity to analyse these
changes. But there are still uncertainties and information gaps as outlined in this
chapter.
8.1 Sources of uncertainty
Need to address uncertainty
Decision making has to rely on knowledge about climate change, its impacts and the
anthropogenic forcing. It requires awareness of the remaining uncertainties which
may affect investment decisions into mitigation and adaptation measures. Ignoring
uncertainty increases the risk of inappropriate action to tackle the challenge of climate
change and its impacts on environment, economy and human well being. Taking
climate change into account may have a risk of over–reaction, which would lead to
low cost efficiency of measures, or an insufficient response, which would mean too
low investments with the consequence of high costs by climate change impacts.
Decision makers furthermore need to take into account the precautionary principle
according to which absence of full scientific certainty shall not be used as a reason to
postpone measures where there is a risk of serious or irreversible harm to public
health or the environment.
Main types of uncertainty
Three basic types of uncertainty could be distinguished
1. Uncertainty in future socio-economic developments
Most important sources of uncertainty are human behaviour, evolution of political
systems, demographic, technological and socio-economic developments. Policies
to control greenhouse gas emissions affect the rate and intensity of future climate
warming. For projections of climate change and impacts, the unpredictable
component of the anthropogenic forcing is handled by using ensembles of
potential futures based on different storylines of socio-economic development
leading to a set of emission scenarios, such as the ones presented in the last IPCC
reports (IPCC 2001, 2007). These emission scenarios are used to analyse the
influence of human activities on the climate system in an “if – then” mode: if
emissions will increase or decrease to a certain degree then the anthropogenic
impacts on the climate system will change in a certain way. The IPCC scenarios
describe a range of possible futures.
Several decades of climate research and the series of IPCC reports allow for a first
comparison of early projections with observations over the last 20 years. The
results show that there is increasing evidence that emission of greenhouse gases
and increase in temperatures tend to be underestimated in projections. Even
‘business as usual’ or ‘worst case’ scenarios projected lower emissions and rises
197
in temperatures in the last 20 years compared to the actual measurements in the
same period (IPCC, 2007). Underestimation of economic growth, energy demand
and carbon intensity of energy supply might be one of the reasons.
It is interesting to note that the projected range of temperature has only slightly
changed for the end of the 21st century from 1.4-5.8°C in the third IPCC
assessment report (IPCC, 2001) to 1.1-6.4°C warming with a best estimate of 1.84.0°C in the fourth assessment report (IPCC, 2007). This is the current best
available framework for decision makers when they consider options for climate
policy measures.
2. Incomplete knowledge in models and assessments
Lack in understanding of the physical, chemical and biological processes and
attribution of climate change to anthropogenic and natural factors are other
sources of uncertainty. Also the attribution of impacts to climate and non-climate
factors is a source of uncertainty. There might be still completely unknown
processes in the Earth system we are still not aware off. However the physical
processes of the atmosphere and its interaction with land and sea surface are
reasonably well understood. This is shown by the fact that climate models are able
to reproduce the past and present climate rather well. However, there remain
uncertainties in the understanding of the climate system, of which the last IPCC
report (IPCC, 2007) gives a good overview. For example, one major source of
uncertainty is the impacts of cloud cover on the energy balance of the atmosphere,
which, for example, implies uncertainty in the climate sensitivity, a key variable
in climate change modelling.
Another process-related uncertainty for climate impact analysis is related to the
‘double attribution’: the attribution of climate change to anthropogenic causes and
the attribution of observed changes in the environment to climate change impacts.
The former attribution is approached indirectly by running models with and
without anthropogenic causes and comparing the results with observations. The
latter type of attribution is especially occurring in systems which are intensively
managed or affected by human activities. Generally, this type of attribution is
approached by analyzing consistency of observed indicator changes with observed
climate change, and taking into account understanding of the climate-dependency
of the indicator, e.g. derived from controlled experiments or model analysis.
Increase in crop yield for example may be caused by a combination of higher
temperatures, CO2 fertilization, growth of new crop varieties and/or improved
management. Responses in systems with high levels of complexity and selforganization like biological, social or economic systems are very difficult to
evaluate. Climate impacts can either be increased by other, non-climatic factors,
or be compensated by adaptation of the system, or be internally compensated until
a critical level of resilience is exceeded. Sensitivity analysis with computer
models can support a better understanding of these systems by analyzing the
different combinations of drivers.
3. Insufficient data and information
There may be high confidence in the understanding of particular processes of
climate change and impacts. But if there is insufficient data assessments only give
qualitative information which is of limited value for mitigation and adaptation
strategies. If data are too limited for appropriate modelling and assessments,
198
confidence in findings are normally low. Data might be completely missing or
insufficient in spatial or temporal resolution or coverage. Time series are often too
short for detecting trends and understanding causal links to either anthropogenic
climate change or natural variability. For instance, the question is still unresolved
whether frequency and intensity of extreme events like hurricanes and floods are
changing due to anthropogenic climate change or whether the observed changes
are part of the natural variability. This is partly because time series of observed
trends are still too short and partly because variability of these events is much
higher than the trend in climate. Scarcity in terms of temporal and spatial coverage
of data describing the so-called lower boundary conditions of the climate system,
like sea surface temperature, ice and snow cover, and permafrost, still limit the
reliability of climate modelling (GCOS, 2003).
Another source of underestimation when comparing projections done in the past
with actual temperature measurements, as mentioned above (IPCC, 2007) might
be the tendency to interpret lack in knowledge as evidence that the climate is not
changing. Since current climate conditions are very likely significantly warmer
than conditions of the last 1000 years, models validated on observed or estimated
data in current climate may underestimate future trends under accelerating climate
warming especially if validation of models require detailed measurements such as
physiological models for crop and forest growth.
Adaptation planning requirements
The need for regional assessments for better understanding of climate change and
impacts requires analysis at higher spatial resolutions including seasonal changes over
the year (IPCC, 2007). To adequately support decisions on adaptation measures, more
precise information at a regional and local level is required. Planning for winter
tourism for example requires very detailed information on local climate, especially on
changes in temperatures and snow fall during winter to analyze the cost-effectiveness
of managing ski resorts and the environmental impacts of running these facilities.
Taking climate change impacts into account for flood protection measures requires
very detailed information on changes in precipitation frequency and intensity for
appropriate planning of dams and dikes.
8.2 Uncertainties and data gaps for indicators
The number of indicators and the quality of the underlying information in terms of
pan-European coverage for describing climate change impacts have been significantly
increased since the last EEA climate change indicator report 2004. New indicators
have been developed especially for systems which are also influenced by non-climate
factors like ecosystems, biodiversity, forestry and agriculture (Table 8.2.1). Others
have been dropped either because they haven’t been considered to be important for
the communication of climate change impacts (e.g. greenhouse gas concentration) or
been replaced by a more Europe-wide view (e.g. plant species distribution in
mountains).
The current state in terms of uncertainties and data needs are summarized for each
category.
199
Table 8.2.1: Major changes in indicators 2004 – 2008
Sector
Atmosphere and climate
New indicator in 2008
Replaced or removed indicator
from 2004 report
Storms and storm surges in Europe
Air pollution by ozone
Greenhouse gas concentrations
Cryosphere
Greenland ice sheet
Mountain permafrost
Terrestrial ecosystems and
biodiversity
Distribution of animal species
Bird survival
Animal phenology
Impacts on communities
Agriculture and forestry
Agrophenology
Irrigation demand
Forest growth
Forest fire danger
Soil organic carbon
Terrestrial carbon uptake *
Growing season
Crop yield
Water quantity, droughts, floods
Crop yield losses in 2003
River floods
River flow drought
Water quality and
Water temperature
fresh water ecology
Lake and river ice coverage
Freshwater ecology
Human health
Water and food borne diseases
* from section ‘Terrestrial ecosystems and biodiversity’
Atmosphere and climate
Atmospheric and climate systems are the subject of long-term routine measurements
since many decades. Therefore data availability is relatively good. On global level
major gaps in data coverage are mostly identified for Africa, the oceans and the polar
regions (GCOS 2003). However, there is also still lack in data for regional and local
assessments at appropriate spatial resolution and quality for Europe. More detailed
and quantitative, tailor-made information is especially needed for regional climate
impact assessments and the development of cost effective adaptation strategies for
ecosystems, such as Natura2000 areas, economic sectors and human health in the
regions across Europe. For adaptation particularly information on extreme events are
most important, but changes in storms and storm surges in relation to climate change
are still uncertain since time series of observed data are too short to understand the
contributions of natural variability and anthropogenic forcing. To better analyse
changes in the frequency and intensity of other extreme events like heat waves and
heavy rain falls sufficient amounts of data are still lacking in parts of Europe and in
terms of length of observed time series.
200
Cryosphere
Ice and snow coverage are monitored using remote sensing techniques, since several
decades, delivering qualitative information on changes over time. Much more
important are changes in mass balances which are the key information for assessing
water availability and changes in sea level rise. The evidence for these changes is
rather robust for glaciers which are monitored intensively, but for large parts of the
cryosphere and especially for the Greenland Ice Sheet uncertainty in changes in massbalances is still very high. There is more and more evidence that melting rates for
Greenland Ice Sheet and Arctic Sea Ice are accelerating very fast and there might be
the risk that a tipping point for accelerated Arctic Sea Ice melting might be reached
rather soon. Only recently monitoring and research activities in Greenland have been
stepped up, and over time the understanding may improve. Melting of permafrost has
been observed but data and knowledge for quantitative assessments is still rather poor
due to too short time series.
Marine systems
Observed changes of physical conditions of the sea surface are linked to changes in
the atmosphere. Both satellite and ground based gauge stations are in place to monitor
changes in sea level. For Europe uncertainty in projected sea levels are relatively high
because global trend is overlaid by other processes such as land rise in Northern
Europe, land subsidence at parts of the North sea coast and changes in thermohaline
circulation in the North Atlantic which also affects regional sea level. Additional
monitoring and research is needed to improve the understanding of changes in the
thermohaline circulation (THC, now termed as Meridional Overturning Circulation,
MOC) in terms of impacts on regional climate and sea level.
Monitoring of physical parameters (mostly temperature, salinity) has been
significantly improved especially with the systems of ARGO floaters and remote
sensing techniques. Implementing the European Marine Strategy will improve the
sustainability of these monitoring programmes. Biological systems are observed
occasionally mostly by research campaigns and ship of opportunity programmes (e.g.
‘SAHFOS Continuous Plankton Recorder’). These programmes give a good overview
on changes in areas where ships are navigating. A more systematic evaluation of
changes in marine ecosystems is needed in terms of coverage of European seas and
comparability of applied methods.
Terrestrial ecosystems and biodiversity
Phenology and diversity of plants and animals show significant changes but observed
trends are still rather fragmentary in terms of European coverage and number of
observed species. A fundamental problem is the different sensitivity of each species to
changes in temperature, precipitation, humidity and other climate variables. Using the
observed distribution of species in relation to current climate conditions allows
projections of the natural habitat of these species under future climate. However, it
doesn’t take into account the individual resilience and adaptation capacity of the
species, positive and negative impacts of management and risk of disconnection of
predator - prey relations in the foodweb. Impacts on abundance, which is also very
important for the genetic variety and consequently for the survival, is considered only
to a limited extent. Monitoring of biodiversity is partly based on regular screening of
plants and species in protected areas but most of the pan-European information is
201
based on temporary research projects, voluntary networks and NGOs. There are more
information on observed changes in phenology mostly based on voluntary observation
networks, but there are only very general projections on phenological changes mostly
based on interpretations of changes in temperatures. More efforts are needed to
improve monitoring especially in areas where data sampling is still rather poor or data
are existing but are not accessible. A more systematic observation of species and their
abundance across Europe is needed to improve the still very fragmented knowledge
about climate change impacts on ecosystems and biodiversity.
Agriculture and forestry
Agriculture and forestry sector are affected by climate change and non climatic
impacts. Both sectors are clearly dominated by management, and it is therefore
difficult to attribute specific trends from field observations uniquely to a changing
climate.
Nevertheless, direct observations and model based reconstructions allow the
identification of specific climate related impacts on plant growth. Additionally very
extensive experiments e.g. FACE (free air CO2 enrichment experiments) are
performed to investigate impacts of climate change on crop and tree species including
interactions with increasing atmospheric CO2 concentrations and impacts of
management. Results are then extrapolated using models to simulate the physiological
response of plants to changing conditions. There is still significant lack in knowledge
and data on the individual responses of species to climate change and the subsequent
changes in competition in forests also because response of species depends on age and
time of exposure. More precise analyses and projections for agriculture and forestry
would further require more detailed information on management for each site and
forest stand. Projections on fire risks are rather robust but projections of impacts of
other extreme events like storms are very uncertain. A big gap in information exists
on the possible changes in pressures by pests and diseases on crops and forests under
a changing climate.
Pan-European assessments will certainly be more precise and accurate when specific
local conditions and physiological constrains of crops, trees and pests are fully taken
into account. But lack in knowledge on physiological responses of individual plant
species to climate change and still inaccurate projections of future climate on regional
level make projections rather uncertain especially in areas where precipitation is the
limiting factor for agriculture and forestry. Projections for areas where temperatures
are the most limiting factor are expected to deliver more robust results.
Water quantity
In European rivers and lakes there are already many observations of marked changes
in water temperatures, ice coverage, and stratification of lakes, which can be
attributed mainly to air temperature increase. Some long-term observations clearly
describe the changes and they are supported by a growing set of shorter term
observations (30-50 years) that provide additional evidence of changes in the
temperature regime of rivers and lakes.
Water quantity assessments and especially floods are linked to quantification of trends
in precipitation which is still uncertain especially for summer rainfall. Small
202
differences in projected changes of precipitation lead to very large differences in
water quantity especially in areas with low annual precipitation. In some areas of
Europe river discharge is also linked to changes in mass balances of snow and ice.
Additionally water extraction and water management across catchments and changes
in land-use and management make it very difficult to attribute changes in average
water discharge as well as in floods and droughts to climate change forcing.
Monitoring networks for river discharge and groundwater levels are relatively dense
and delivering robust numbers on average water flows. But due to non-climatic
factors uncertainties in climate change induced impacts are rather high. For
projections uncertainties in climate models are further enhancing uncertainties in
future changes of water quantities. Especially changes in floods and droughts require
much more detailed information on regional changes in precipitation and landmanagement. Reliable information on quantities is the key information for appropriate
water management and flood protection. A pan-European view on river discharges
and changes is still missing but will be implemented in the Water Information System
for Europe (WISE) in near future.
Water quality
European freshwaters are already affected by many human pressures related to landuse, pollution with nutrients and hazardous substances, and acid deposition. Because
of difficulties in disentangling the effects of climatic factors from other pressures,
there is limited empirical evidence to demonstrate unequivocally the impact of
climate change on water quality and freshwater ecology. On the other hand, there are
many indications that freshwaters that are already under stress from human activities
are highly susceptible to climate change impacts.
Changes in oxygen content and stratification of lakes which can be attributed to
climate change are derived from national and local research and monitoring activities.
It is obvious that especially higher temperatures and subsequent lower oxygen content
will increase pressure on freshwater ecosystems especially if there is already high
nutrient and pollutant load but quantification of these processes is still highly
uncertain. European water quality is regularly monitored and reported as part of the
Water Framework Directive and data is available via WISE.
Human health
Climate change impacts on human health include direct impacts mainly due to heat
waves storms and floods and indirect impacts by vector, water, and food borne
diseases but also lowers risk of deaths from low temperatures. However quantification
of these impacts is uncertain since monitoring systems in the health sector are not yet
taking climate change sufficiently into account. Regarding projections an important
source of uncertainty is the extent to which adaptation actions, such as preventive
measures and appropriate changes in the health systems, will be implemented.
Economy
Economic climate change effects include direct impacts by extreme events like storms
floods heat waves, and indirect impacts via changes in ecosystem services for
agriculture, forestry, fisheries or tourism. Changes in temperature are also affecting
energy demand leading to less heating in winter and more cooling in summer. Other
203
impacts e.g. on human health are also linked to the economic sector. Climate change
related economic effects are different across Europe and depend to a large extent on
the capability of societies to adapt to these changes.
For more detailed risk assessments more detailed data and studies are needed, on the
contribution of increases in damage costs caused by climate change and the
contribution caused by increases in wealth and infrastructure. Also more studies on
valuation of climatic impacts on economic sectors are needed. Furthermore data at
local level on costs of adaptation actions is only available to a limited extent and
much more data on good practices and their costs are needed.
8.3 Improvement of monitoring and reporting
There are many efforts to improve climate change related data availability in Europe
which are also linked to global activities. For improving the understanding of the
climate system, the Global Climate Observing System (GCOS) network identified a
data set as described in Table 8.3.1 requiring long term time series on atmospheric,
marine and terrestrial processes in appropriate temporal and spatial resolution (GCOS,
2003). Most of these are also part of EEA’s indicators on climate change impacts,
such as near surface atmospheric conditions, sea surface temperature, ice and snow
cover, permafrost etc.
Table 8.3.1: Essential climate variables as required by UNFCCC for detecting and
modelling of climate change Source: GCOS, 2003.
Domain
Essential climate variables
Atmospheric
Surface:
(over land, sea
and ice)
Upper-air:
Composition:
Oceanic
Surface:
Sub-surface:
Terrestrial
Air temperature, precipitation, air pressure, surface radiation
budget, wind speed and direction, water vapour.
Earth radiation budget (including solar irradiance), upper-air
temperature (including MSU – microwave sounding unit –
radiances), wind speed and direction, water vapour, cloud
properties.
Carbon dioxide, methane, ozone, other long-lived greenhouse gases,
aerosol properties.
Sea-surface temperature, sea-surface salinity, sea level, sea state,
sea ice, current, ocean colour (for biological activity), carbon
dioxide partial pressure.
Temperature, salinity, current, nutrients, carbon, ocean tracers,
phytoplankton.
River discharge, water use, ground water, lake levels, snow cover, glaciers and ice
caps, permafrost and seasonally frozen ground, albedo, land cover (including
vegetation type), fraction of absorbed photosynthetically active radiation (FAPAR),
leaf area index (LAI), biomass, fire disturbance.
The EU ‘Global Monitoring for Environment and Security’ (GMES) Programme (EC,
2004) aims to strengthen the monitoring capacity within Europe and the world by
making data available as ‘services’ from 2008 onwards. The combination of satellite
and ground based information to services and the envisaged long-term funding of the
services will significantly improve availability of data on changes in the environment.
Many of the services developed will also improve availability of essential climate
variables. E.g. information in ‘Service Marine’ will include sea surface temperatures
and salinity, ‘Service Land Monitoring’ will provide information on land cover
changes and ‘Service Atmosphere’ will monitor greenhouse gas concentrations,
204
aerosols and radiation. The services will be extended stepwise over the coming years
and will be fully established in 2013. GMES is also contributing to the global
activities currently coordinated and streamlined in the Global Observation System of
the Systems (GEOSS) programme (GEOSS, 2005). Another important programme
aiming on monitoring is the Date User Element (DUE) of the European Space Agency
(ESA) which include global and regional data on land-cover (GlobCover), ocean
colour (GlobColour), carbon (GlobCarbon) and cryosphere (GlobIce, GlobSnow,
Permafrost) and other climate relevant data.
European research projects improve our understanding of processes and impacts but
normally are lacking continuity in funding. Consequently, measurements performed
in these projects are only covering short time periods and/or only small areas which
often limit the value of information for climate change assessments. Some projects are
funded for the implementation of GMES monitoring services such as MERSEA and
MyOcean for marine, Geoland-1 and Geoland-2 for land and GEMS and MACC for
atmosphere services.
Climate related projects (e.g., PRUDENCE and ENSEMBLES) are improving the
availability of regional projections of climate change. Other ongoing or already
completed projects are dealing with specific sectors like ACCELERATES for
agriculture and ALARM for biodiversity or EUROLIMPACS and CLIME for
freshwater ecosystems. Some are focussed on regions like CIRCE for the
Mediterranean area or investigating climate change impacts and the social and
economic impacts like ATEAM on vulnerability of ecosystem services in Europe.
ADAM is focused on adaptation and mitigation strategies for supporting European
climate policy and PESETA is dealing with projections of economic impacts of
climate change in different sectors and regions like coastal systems, energy demand,
human health, agriculture, tourism, and floods. There are also numerous activities on
national and regional level including both public and non-governmental research and
monitoring programmes (e.g. Transboundary projects of the Interreg Programme).
The monitoring activities and results from research will further help to understand,
and maybe reduce, the uncertainties related to lack in data and knowledge about
climate change and impacts and will consequently improve the basis for decision
making. In summary, from the perspective of climate change impacts and adaptation,
future efforts should mostly aim at



better data and better models for projections for the next decades (in addition to
models for the period up to 2100) in terms of spatial density, quality and length of
time series (including historical data) for detecting environmental changes due to
climate change;
improved timely and free public access to data and tools by metadata description,
standardization, harmonization, improved architecture and infrastructure for data
management including distributed data centres and portals for standardized
access;
research to advance the understanding about the linkages between the behaviour
of vulnerable systems, climatic changes, and non-climatic drivers.
205
References
1. Introduction
ACIA, 2004. Impacts of a warming Arctic: Arctic Climate Impact assessment, Cambridge
University Press, Cambridge, UK
EEA, 2004: Impacts of Europe’s changing climate, An indicator-based assessment, EEA
Report No. 2/2004. EEA, Copenhagen
Human Development Report 2007/2008, Fighting climate change: Human solidarity in a
divided word, Palgrave Macmillan, 2007.( http://hdr.undp.org/).
IPCC, 2007: Summary for Policymakers. In: Climate Change 2007: The Physical Science
Basis. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA.
IPCC, SYR, 2007d. Climate Change 2007: Synthesis. Cambridge University Press,
Cambridge, United Kingdom and New York, NY, USA
UNEP, 2007: Global Outlook for Ice and Snow, (www.unep.org/geo/geo_ice)
2. The Earth, its Climate and Man
Petit et. al. 1999. Climate and atmospheric history of the past 420.000 years from the Vostok
ice core, Antarctica. Nature 399, 429-436
Scientific Expert Group on Climate Change (Rosina M. Bierbaum, John P. Holdren, Michael
C. MacCracken, Richard H. Moss, and Peter H. Raven, eds.), 2007. Confronting Climate
Change: Avoiding the Unmanageable and Managing the Unavoidable. Report prepared for
the United Nations Commission on Sustainable Development. Research Triangle Park, N.C.,
and Washington, D.C.: Sigma Xi and the United Nations Foundation
3. Observed impacts: a cascade of effects with feed-backs
Climate Change 2007, Fourth Assessment Report, Working Group reports I to III (IPCC,
Cambridge Univ. Press, Cambridge, 2007); www.ipcc.ch
Crutzen P.J. and E.F. Stoemer, 2000, The “Anthropocene” International Geosphere
Biosphere Programmme, Newsletter 41, pp17-18.
4. Climate change impacts: what the future has in store
Den Elzen, M and M. Meinshausen, 2005. Meeting the EU 2 degrees C climate target: global
and regional emission implications. Report no. 728001031, MNP, Bilthoven, Netherlands
(replace with some more formal publication?)
EC, 2008. 20 20 by 2020: Europe's climate change opportunity. Communication from the
Commission COM(2008) 13, 16, 17, 18, 19 final, Brussels
EEA, 2007.Climate change: the cost of inaction and the cost of adaptation. EEA Technical
Report No. 13/2007. EEA, Copenhagen
IPCC, 2001. Climate Change 2001: Mitigation. Cambridge University Press, Cambridge, United Kingdom.
IPCC, WG1: 2007a. Climate Change 2007: The Physical Science Basis. Contribution of
Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate
Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M.Tignor and
H.L. Miller (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New
York, NY, USA.
206
IPCC, WG2: 2007b. Climate Change 2007: Impacts, Adaptation and Vulnerability.
Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA.
IPCC, WG3, 2007c. Climate Change 2007: Mitigation. Cambridge University Press,
Cambridge, United Kingdom and New York, NY, USA
IPCC, SYR, 2007d. Climate Change 2007: Synthesis. Cambridge University Press,
Cambridge, United Kingdom and New York, NY, USA
UNFCCC, 2007. Decision -/CP.13, Bali Action Plan
van Vuuren, D.P., M.G.J. den Elzen, P.L. Lucas, B. Eickhout, B.J. Strengers, B. van Ruijven,
M.M. Berk, H.J.M. de Vries, M. Hoogwijk, M. Meinshausen, S.J. Wonink, R. van den Houdt,
and R. Oostenrijk, 2006. Stabilising greenhouse gas concentrations at low levels: an
assessment of options and costs, MNP Report 500114002/2006, MNP Bilthoven, Netherlands
5. An Indicator-based assessment
Introduction
Atmosphere and climate
ACIA overview report, 2004, Cambridge university press, Cambridge
Alexandersson, H., et al., 2000: Trends of storms in NW Europe derived from an updated
pressure data set. Climate Res., 14, 71–73.
Andersson, C., Langner, J., and Bergström, R., 2007, Interannual variationand trends in air
pollution over Europe due to Climate variability during 1958-2001 simulated with a regional
CTM coupled to the ERA40 reanalysis, Tellus, 59B, 77-98
Beniston, M., et al., 2007: Future extreme events in European climate: An exploration of
regional climate model projections. Clim. Change ; 81 ; 71-95
Fiala, J., Cernikovsky, L., de Leeuw, F., Kurfuerst, P., 2003, Air pollution by ozone in Europe
in summer 2003, EEA Topic Report 3/2003
Fink, A.H. et al., 2004: The 2003 European summer heatwaves and drought – synoptic
diagnostics and impacts. Weather, 59, 209-216.
Fischer-Bruns, I., H. von Storch, F. González-Rouco and E. Zorita, 2005: Modelling the
variability of midlatitude storm activity on decadal and century time scales, Clim. Dyn. 25:
461-476.
Gillett, N. P., Zwiers, F. W., Weaver, A. J. & Stott, P. A., 2003: Detection of human influence
on sea-level pressure, Nature, Vol 422, 292-294.
Gulev, S. K., O. Zolina, and S. Grigoriev, 2001: Extratropical cyclone variability in the
Northern Hemisphere winter from the NCEP/NCAR reanalysis data. Climate Dyn., 17, 795–
809.
Holland, M., Kinghorn, S., Emberson, L., Cinderby, S., Ashmore, M., Mills, G., and
Harmens, H.., 2006, Development of a framework for probabilistic assessment of the
economic losses caused by ozone damage to crops in Europe, CEH project No. C02309NEW.
Centre for Ecology and Hydrology, Natural Environment Research Council, Bangor, Wales,
49pp.
Hurk, B.J.J.M. van den et al., 2006: KNMI Climate Change Scenarios 2006 for the
Netherlands, KNMI Scientific Report: WR-2006-01, De Bilt 2006, 82 pp.
Hurrell, J.W., et al., 2003: An overview of the North Atlantic Oscillation. In: The North
Atlantic Oscillation: Climatic Significance and Environmental Impact [Hurrell, J.W., et al.
(eds.)]. Geophys. Monogr., 134, Amer. Geophys. U., Washington, DC, 1–35.
IPCC, 2001. Climate Change 2001: The scientific basis. Contribution of Working Group I to
the Third Assessment Report of the Intergovernmental Panel on Climate Change. J.T.
Houghton et al. (eds), Cambridge University Press.
207
IPCC, WG1: 2007. Climate Change 2007: The Physical Science Basis. Contribution of
Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate
Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M.Tignor and
H.L. Miller (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New
York, NY, USA.
Jones, P.D. and A. Moberg, 2003. Hemispheric and large-scale surface air temperature
variations: An extensive revision and update to 2001. J. Climate 16, 206–223.
Kjellström, E. et al., 2007: Variability in daily maximum and minimum temperatures: recent
and future changes over Europe. Clim. Change; 81, 249-265.
Klein Tank, A.M.G., and G.P. Können, 2003: Trends in indices of daily temperature and
precipitation extremes in Europe, 1946–1999. J. Climate, 16, 3665–3680.
Klein Tank, A.M.G., J.B. Wijngaard et al., 2002. Daily dataset of 20th-century surface air
temperature and precipitation series for the European Climate Assessment. Int. J. Climatology
22, 1441-1453.
Lambert, S.J. and J.C. Fyfe, 2006. Changes in winter cyclone frequencies and strengths
simulated in enhanced greenhouse warming experiments; results from the models
participating in the IPCC diagnostic exercise; Climate Dynamics; 26, No 7 – 8, 713-728.
Langner, J., Bergström, R., and Foltescu, V., 2005, Impact of climate change on surface
ozone and deposition od sulohur and nitrogen in Europe, Atm. Env., 39, 1129-1141.
Leckebusch, G.C. and U. Ulbrich, 2004: On the relationship between cyclones and extreme
windstorm events over Europe under climate change. Global Planet. change, 44, 181-193.
Lionello, P., 2005: Extreme surges in the Gulf of Venice. Present and future climate Venice
and its lagoon, State of Knowledge. C. Fletcher, and T. Spencer (Eds), Cambridge University
Press, Cambridge UK, pp. 59 -65.
Lloyd-Hughes, B., Saunders, M. A., 2002.: A drought climatology for Europe, International
Journal of Climatology, vol. 22, p 1571-1592.
Lowe, J.A. and Gregory, J.M. 2005: The effects of climate change on storm surges around the
United Kingdom. Proc. R. Soc. 363, 1313-1328.
Matulla, C., W. Schöner, H. Alexandersson, H. von Storch, and X. L. Wang, 2007: European
storminess: late nineteenth century to present. Climate Dynamics, DOI 10.1007/s00382-0070333-y)
McCabe, G.J., M.P. Clark, and M.C. Serreze, 2001: Trends in Northern Hemisphere surface
cyclone frequency and intensity. J. Climate, 14, 2763–2768.
Pinto, J. G., E. L. Fr¨ohlich, G. C. Leckebusch, and U. Ulbrich, 2007: Changing European
storm loss potentials under modified climate conditions according to ensemble simulations of
the ECHAM5/MPI-OM1 GCM Nat. Hazards Earth Syst. Sci., 7, 165–175, 2007
Räisänen, J. et al., 2004: European climate in the late 21st century: regional simulations with
two driving global models and two forcing scenarios. Clim. Dyn., 22, 13-31.
Robock, A., Mu, M., Vinnikov, K., Trofimova, I.V. and Adamenko, T.I. 2005.: Forty-five
years of observed soil moisture in the Ukraine: No summer desiccation (yet). Geophysical
Research Letters; Vol 32: 10.1029/2004GL021914.
Schär, C. et al., 2004: The role of increasing temperature variability in European summer
heatwaves. Nature, 427, 332-336.
Sillmann, J., Roeckner, E. 2008: Indices for extreme events in projections of anthropogenic
climate change, Climatic Change; 86, (1-2), 83-104.
Smits, A., A.M.G. Klein Tank en G.P. Können, 2005. Trends in storminess over the
Netherlands, 1962-2002. Int. J. Climatology 25, 1331-1344.
Stevenson, D.S.,et al., 2006, Multimodel ensemble simulations of present-day and near-future
tropospheric ozone, J. Geophys. Res., 111, D08301, doi:10.1029/2005JD006338
208
Thompson, D. W. J., Wallace, J. M. & Hegerl, G. C., 2000: Annular modes in the
extratropical circulation. Part II: Trends. J. Clim. 13, 1018–1036.
van der Schrier, G., Briffa, K.R., Jones, P.D. and Osborn, T.J. 2006: Summer moisture
variability across Europe. Journal of Climate; 19: 2818-2834.
Van Dingenen et al., Trends of ozone exceedances in Europe from 1990 to 2004: what is the
role of climate change? [manuscript in preparation]
Van Dingenen, R., Raes, F., Dentener, F., and Putaud, J.P., 2008: The relation between
temperature anomaly and ozone exceedance as an indication for the role of climate change on
ozone exceedance in Europe [manuscript submitted]
van Lannen, H. A. J. , Tallaksen, L. M., Rees, G. 2007: Drought and climate change (In:
Commission Staff Working Document Impact Assessment (SEC(2007) 993, (COM(2007)
414), Commission of the European Communities, Brussels, Belgium).
von Storch, H., F. Feser, A. Pfizenmayer, and R. Weisse, 2002: Changig coastal weather and
extremes. Proceedings WCSE/UKC-2002, Weather Risk Management, Seoul, 2002, 468-473.
WMO, 2004: World Meteorological Organization statement on the status of global climate in
2003. WMO publications, Geneva, 12 pp.
Woth, K., 2005: North Sea storm surge statistics based on projections in a warmer climate:
How important are the driving GCM and the chosen emission scenario? Geophys. Res. Lett.,
32, L22708, doi:10.1029/2005GL023762.
Woth, K., R, Weise, H. von Storch, 2006: Climate change and North Sea storm surge
extremes: an ensemble study of storm surge extremes expected in a changed climate projected
by four different regional climate models, Ocean Dynamics; 56, 3-15.
Cryosphere
ACIA, 2004. Impacts of a warming Arctic: Arctic Climate Impact assessment, Cambridge
University Press, Cambridge, UK
Andreassen, L.M., Elvehøy, H., Kjøllmoen, B., Engeset, R.V. and Haakensen, N.,2005.
Glacier mass balance and length variations in Norway. In Annals of Glaciology, 42, 317–325
Bamber, J.L., W. Krabill, V. Raper, and J. Dowdeswell, 2004. Anomalous recent growth of
part of a large Arctic ice cap: Austfonna, Svalbard. In Geophys. Res. Lett., 31(12), L12402
Beniston, M., 2003a. Climatic change in mountain regions: a review of possible impacts.
Climatic Change, 59, 5-31
Beniston, M., F. Keller and S. Goyette, 2003b. Snow pack in the Swiss Alps under changing
climatic conditions: an empirical approach for climate impact studies. Theor. Appl.
Climatol., 74, 19-31.
Bevan, S., A. Luckman, T. Murray, H. Sykes, & J. Kohler.,2007. Positive mass balance
during the late 20th century on Austfonna, Svalbard, revealed using satellite radar
interferometry. Ann. Glaciol. 46, 117-122
Brodzik, M.J., R.I. Armstrong, E.C. Weatherhead, M.H. Savoie, K.W.Knowles and D.A.
Robinson, 2006. Regional trend analysis of satellite derived snow extent and global
temperature anomaliesAmerican Geophysical Union Fall 2006, San Franzisco, U.S.A.
Carter, T. 2007. Presentation at stakeholder consultation workshop and round table on the
EU-GP on adaptation: Northern Europe; 27-28 September 2007, Finnish Environment
Institute (SYKE), Helsinki, Finland
Cazenave, A., 2006. How fast are the ice sheets melting? Science 24. November.2006, vol
314.
Christensen, J.H. and O.B. Christensen, 2007. A summary of the PRUDENCE model
projections of changes in European climate by the end of this century. Climatic Change
81(Supplement 1):7–30
209
Clark, M.P., M.C. Serreze, and D.A. Robinson, 1999. Atmospheric controls on Eurasian snow
extent. Int. J. Climatology, 19, 27-40
Comiso, Josefino C., Claire Parkinson, Robert Gersten and Larry Stock 2008: Accelerated
decline in the Arctic sea ice cover. Geophysical research letters, vol 35, L01703
Davies, M. C. R., O. Hamza u. C. Harris, 2001. The effect of rise in mean annual temperature
on the stability of rock slopes containing ice-filled discontinuities.Permafrost and Periglacial
Processes, 12(1): 137-144.
Dye, D.G., 2002. Variability and trends in the annual snow-cover cycle inNorthern
Hemisphere land areas, 1972–2000. Hydrolog. Process., 16, 3065–3077.
Elsasser, H. and R. Burki, 2002. Climate change as a threat to tourism in the Alps.,Clim. Res.,
20, 253-257
Falarz, M., 2002. Long-term variability in reconstructed and observed snow cover over the
last 100 winter seasons in Cracow and Zakopane (southern Poland). Clim. Res., 19(3), 247–
256
Greene, A.M., 2005. A time constant for hemispheric glacier mass balance. In J. Glaciol.,
51(174), 353–362.
Gruber, S. and Haeberli, W., 2007, Permafrost in steep bedrock slopes and its temperaturerelated destabilization following climate change. In print
Gruber, S., Hoelzle, M. and Haeberli, W., 2004b. Permafrost thaw and destabilization of
alpine rock walls in the hot summer of 2003.Geophys. Res. Lett., 31, L13504
Gruber, S., King, L., Kohl, T., Herz, T., Haeberli, W. and Hoelzle, M., 2004a. Interpretation
of geothermal profiles perturbed by topography: the Alpine permafrost boreholes at
Stockhorn Plateau, Switzerland. Permafrost and Periglacial Processes, 15(4), 349-357
Haas, C., 2004: Late-summer sea ice thickness variability in the Arctic Transpolar Drift
1991--2001 derived from ground-based electromagnetic sounding, Geophysical research
letters, 31, L09402, 5pp., doi:10.1029/2003GL019394.
Haas, C., S. Hendricks, and M. Doble, 2006: Comparison of the sea ice thickness distribution
in the Lincoln Sea and adjacent Arctic Ocean in 2004 and 2005. Annals of Glaciology, 44
247-252.
Haeberli, W., Noetzli, J., Zemp, M., Baumann, S., Frauenfelder R. and Hoelzle, M., 2005.
Glacier Mass Balance Bulletin No. 8, 2002–2003. IUGG(CCS)–UNEP–UNESCO–WMO,
World Glacier Monitoring Service, Zurich
Hagen, J. O., K. Melvold, J. Kohler, J.-G. Winther. 2003b. Glaciers in Svalbard: mass
balance, runoff and freshwater flux. Polar Research 22(2), 145-159
Hantel, M. and L.-M. Hirtl-Wielke, 2007. Sensitivity of Alpine snow cover to European
temperature. Int. J. Climatol., 27, 1265-1275
Harris, C., Vonder Mühll, D., Isaksen, K., Haeberli, W., Sollid, J.L., King, L., Holmlund, P.,
Dramis, F., Guglielmin, M. and Palacios, D., 2003. Warming permafrost in European
mountains. Global and Planetary Change, 39, 215-225
Holland, M., C. Blitz and B. Tremblay 2006. Future abrupt reductions in the summer Arctic
sea ice. Geophysical research letters, Vol 33.
Holloway and Su (J. Climate) 2002
Hosaka, M., D. Nohara, and A. Kitoh, 2005. Changes in snow coverage and snow water
equivalent due to global warming simulated by a 20km-mesh global atmospheric
model.Scientific Online Letters on the Atmosphere, 1, 93–96
Huggel, C., Haeberli, W. and Kääb, A., 2007. Glacial hazards: changing threats, response
and management in different high-mountain regions of the world. In: B. Orlove, B. Luckman,
E. Wiegandt (Eds.), The Darkening Peaks: Glacial Retreat in Scientific and Social Context,
University of California Press.
210
Hyvärinen, V., 2003. Trends and characteristics of hydrological time series in Finland. Nord.
Hydrol., 34(1–2), 71–90.
IPCC, 2007a Climate Change 2007: The Physical Science Basis. Cambridge University Press,
UK
IPCC, 2007b Climate Change 2007: Climate change impacts, Adaptation and Vulnerability.
CUP, UK
Isaksen, K., Sollid, J.L., Holmlund, P. and Harris C., 2007. Recent warming of mountain
permafrost in Svalbard and Scandinavia. J. Geophys. Res., 112, F02S04
Jacob, D., Göttel, H., Lorenz, P., 2007. Hochaufgelöste regionale Klimaszenarien für
Deutschland, Österreich und die Schweiz, DMG- Mitteilungen, 03/07, Berlin
Jylhä, K., Fronzek, S., Tuomenvirta, H., Carter, T.R. and Ruosteenoja, K., 2007. Changes in
frost and snow in Europe and Baltic sea ice by the end of the 21st century. Climatic Change,
DOI 10.1007/s10584-007-9310-z.
Kääb, A., Frauenfelder, R. and Roer, I., 2007. On the response of rockglacier creep to surface
temperature increases. Global and Planetary Change, 56, 172–187
Kitajev, L., Foerland, E., Razuvaev, V., Tveito, O.E. and Krueger, O., 2005. Distribution of
snow cover over Northern Eurasia. Nordic Hydrology, 36, 311-319
Kohler, J., O. Brandt, M. Johansson, T. Callaghan. 2006. A long-term Arctic snow depth
record from Abisko, northern Sweden, 1913-2004 Polar Research 25 (2), 91–113
Kohler, J., T. D. James, T. Murray, C. Nuth, O. Brandt, N. E. Barrand, H. F. Aas, & A.
Luckman. 2007. Acceleration in thinning rate on western Svalbard glaciers, Geophys. Res.
Lett., 34, L18502, doi:10.1029/2007GL030681
Maslowski 2007 (Gunnar: I have never seen a publication, just media articles)
Murray, T., 2006. Greenland’s ice on the scales. Nature, 21.September 2006, Volume 443.
National Snow and Ice Data Center 2007: Arctic Sea Ice Shatters All Previous Record Lows.
See http://www.nsidc.org/news/press/2007_seaiceminimum/20071001_pressrelease.html
Nesje, A., Bakke, J., Dahl, S.O., Lie, O. and Matthews, J.A., 2007. Norwegian mountain
glaciers in the past, present and future. In print
Nghiem, S.V., I.G. Rigor, D.K. Perovich, P.Clemente-Colon, J.W. Weatherly and G.
Neumann 2007: Rapid reduction of Arctic perennial sea ice. Geophysical research letters, vol
34, L19504.
Noetzli, J., Hoelzle, M., and Haeberli, W., 2003. Mountain permafrost and recent Alpine
rock-fall events: a GIS-based approach to determine critical factors. In: Phillips, M.,
Springman, S.M. and Arenson, L.U. eds. 8th International Conference on Permafrost, 20-25
July 2003, Zurich. Lisse, A.A. Balkema Publishers
Nuth, C., J. Kohler, H.F. Aas, O. Brandt, & J.O. Hagen, 2007. Glacier geometry and
elevation changes on Svalbard (1936–90). Ann. Glaciol. 46, 106-116
OECD, 2007. Climate Change in the European Alps, OECD publishing; Paris, France
Overland et al. 2007 (regional changes)
Paul, F., Kääb, A., Maisch, M., Kellenberger, T. and Haeberli, W., 2004. Rapid disintegration
of Alpine glaciers observed with satellite data. In Geophysical Research Letters, 31(L21402
PERMOS, 2007. Permafrost in Switzerland 2002/2003 and 2003/2004. Glaciological Report
(Permafrost) 4(5) of the Glaciological Commission of the Swiss Academy of Sciences (SAS)
and Department of Geography, University of Zurich. In preparation
Petkova, N., E. Koleva, and V. Alexandrov, 2004. Snow cover variability and change in
mountainous regions of Bulgaria, 1931-2000. Meteorol. Z., 13(1), 19–23.
Phillips, M., Ladner, F., Müller, M., Sambeth, U., Sorg, J. and Teysseire,P., 2007. Monitoring
and reconstruction of a chairlift midway station in creeping permafrost terrain, Grächen,
Swiss Alps. Cold Regions Science and Technology, 47, 32-42
211
Pralong, A. and Funk, M., 2005. On the instability of hanging glaciers. In Journal of
Glaciology, 52(176), 31–48
Rignot, Eric and Pannir Kanagaratnam, 2006. Changes in the Velocity Structure of the
Greenland Ice Sheet. Science 17. February 2006, vol 311.
Rothrock et al. 1999
Scherrer, S.C., C. Appenzeller, and M. Laternser, 2004. Trends in Swiss alpine snow days –
the role of local and large scale climate variability. Geophys. Res. Lett., 31, L13215
Serreze, Mark, Marika Holland and Julienne Stroeve, 2007. Perspectives on the Arctic’s
Shrinking Sea-Ice Cover. Science 16.3.07, vol 315
Stroeve, Julienne, Marika Holland, Walt Meier, Ted Scambos and Mark Serreze, 2007. Arctic
sea ice decline: Faster than forecast. Geophysical research letters, vol 34
Sugiyama, S., Bauder, A., Zahno, C. and Funk, M. (2007). Evolution of Rhonegletscher,
Switzerland, over the past 125 years and in the future: application of an improved flowline
model. Annals of Glaciology, 46: p. 268-274.
Truffer, M. and M. Fahnestock, 2007. Rethinking Ice Sheet Time Scales. Science 16 March
2007, vol 315.
UNEP 2007. Global outlook for snow and ice. UNEP Arendal/Nairobi 2007.
UNEP, 2008. Meltdown in the Mountains. UNEP Zürich/Nairobi 2008
Vaughan, David and Robert Arthern 2007. Why Is It Hard to Predict the Future of Ice Sheets?
Science 16 March 2007, vol 315.
Vojtek, M., P. Faško, and P. Šťastný, 2003. Some selected snow climate trends in Slovakia
with respect to altitude. Acta Meteorologica Universitatis Comenianae, 32, 17–27.
Vonder Mühll, D., Noetzli, J., Makowski, K. and Delaloye, R. 2007. Permafrost in
Switzerland 2002/2003 and 2003/2004, Glaciological Report (Permafrost) No. 4/5 of the
Cryospheric Commission (CC) of the Swiss Academy of Sciences (SCNAT) and Department
of Geography, University of Zurich, 106 p.
Wadhams and Davis (2000)
Walt Meier, et al. 2007: Whither Arctic sea ice? A clear signal of decline regionally,
seasonally and extending beyond the satellite record. Annals of Glaciology annals 46
Winton, Michael 2006: Does the Arctic sea ice have a tipping point? Geophysical research
letters, Vol. 33
Yu et al. (JGR) 2004
Zemp, M., Frauenfelder, R., Haeberli, W. and Hoelzle, M., 2005. Worldwide glacier mass
balance measurements: General trends and first results of the extraordinary year 2003 in
Central Europe.
Zemp, M., Haeberli, W., Hoelzle, M. and Paul, F.,2006. Alpine glaciers to disappear within
decades? In Geophysical Research Letters, 33(L13504)
Zhang, T., Barry, R.G., Knowles, K., Heginbottom, J.A. and Brown, J., 2000. Further
statistics on the distribution of permafrost and ground ice in the Northern Hemisphere. Polar
Geography, 24, 126-131
Marine systems
ACIA 2005. Arctic Climate Impact Assessment. Cambridge University Press.
Arnell, N.W., 2004. Climate change and global water resources: SRES emissions and socioeconomic scenarios. In Glob. Environ. Change, 14, 31-52.
Beare, D., Burns, F., Jones, E., Peach, K., and Reid, D. 2005. Red mullet migration into the
northern North Sea during late winte. Journal of Sea Research, 53: 205-212.
212
Beare, D. J., Burns, F., Greig, A., Jones, E. G., Peach, K., Kienzle, M., McKenzie, E., and
Reid, D. G. 2004. Long-term increases in prevalence of North Sea fishes having southern
biogeographic affinities. Marine Ecology Progress Series, 284: 269-278.
Beaugrand G, Brander KM, Lindley JA, Souissi S, Reid PC. 2003. Plankton effect on cod
recruitment in the North Sea. Nature 426:661-664
Brander, K. M. 2007. Global Fish Production and Climate Change. Proceedings of the
National Academy of Science, accepted.
Brander, K. M., Blom, G., Borges, M. F., Erzini, K., Henderson, G., MacKenzie, B. R.,
Mendes, H., Ribeiro, J., Santos, A. M. P., and Toresen, R. 2003. Changes in fish distribution
in the eastern North Atlantic: Are we seeing a coherent response to changing temperature?
219, 261-270.
Church, J.A. and White, N.J., 2006. A 20th century acceleration in global sea-level rise.
In Geophysical Research Letters, 33, L01602
Demirov E. and Pinardi N., 2002. Simulation of the Mediterranean Sea circulation from 1979
to 1993: Part I. The inter-annual variability, In J. Mar. Syst. 33-34, 23-50
Devoy, R.J.N., 2007. Coastal vulnerability and the implications of sea-level rise for Ireland.
In J. Coast. Res. (in press)
Edwards M. and Richardson A.J. 2004 Impact of climate change on marine pelagic phenology
and trophic mismatch. Nature 430: 881-884
Edwards, M., Reid P. C. and Planque B. 2001 Long-term and regional variability of
phytoplankton biomass in the Northeast Atlantic (1960-1995) ICES J Mar Sci 58: 39-49
Edwards, M., Woo, J, Richardson, A.J. (in prep). Spatial variability in marine pelagic
phenology.
Frederiksen, M., Edwards, M., Richardson, A.J., Halliday, N.C., Wanless, S. 2006. From
plankton to top predators: bottom-up control of a marine food web across four trophic levels.
Journal of Animal Ecology, 75: 1259-1268Frankignoul, C., and E. Kestenare, 2005: Observed
Atlantic SST Anomaly Impact on the NAO: An Update. J. Climate, 18, 4089–4094.
Gerald D. Bell, Michael S. Halpert, Chester F. Ropelewski, Vernon E. Kousky, Arthur V.
Douglas, Russell C. Schnell, and Melvyn E. Gelman. Climate Assessment for 1998. Bulletin
of the American Meteorological Society. Vol. 80, No. 5, May 1999
Greve W, Prinage S, Zidowitz H, Nast J, Reiners F (2005) On the phenology of North Sea
ichthyoplankton. ICES J Mar Sci 62:1216-1223
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, R. Watson, February 2008 Science
Issue, p. 948-952.
Hulme, M., G. Jenkins, X. Lu, J.R. Turnpenny, T.D. Mitchell, R.G. Jones, J. Lowe, J.M.
Murphy, D. Hassell, P. Boorman, R. McDonald and S. Hill, 2002. Climate Change Scenarios
for the United Kingdom: In The UKCIP02 Scientific Report. Tyndall Centre for Climate
Change Research, University of East Anglia, Norwich
Hurrell, J. W., 2003: An overview of the North Atlantic Oscillation. Geophys. Monogr.
Series, No. 134, Amer. Geophys. Union, 1–22.
ICES. 2007. OSPAR Request: assessment of changes in the distribution and abundance of
marine species in the OSPAR maritime area in relation to changes in hydrodynamics and sea
temperature. In ICES Advice., Book 1.
IPCC, 2007: Summary for Policymakers. In: Climate Change 2007: Impacts, Adaptation and
ulnerability. Contribution of Working Group II to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change, M.L. Parry, O.F. Canziani, J.P. Palutikof, P.J.
van der Linden and C.E. Hanson, Eds., Cambridge University Press, Cambridge, UK, 7-22.
213
Johansson, M.M., K.K. Kahma and H. Bowman, 2004. Scenarios for sea level on
the Finnish coast. In Boreal Environ. Res. 9, 153-166.
Katsman, C. A., W. Hazeleger, S. S. Drijfhout, G. J. van Oldenborgh and G. J. H. Burgers,
2007. Climate scenarios of sea level rise for the northeast Atlantic Ocean: a study including
the effects of ocean dynamics and gravity changes induced by ice melt Kluwer Academic
publishers, the Netherlands
Kundzewicz, Z.W., M. Parry, W. Cramer, J.I. Holten, Z. Kaczmarek, P. Martens, R.J.
Nicholls, M. Oquist, M.D.A Rounsevell and J. Szolgay, 2001. Climate Change 2001:
Impacts, Adaptation, and Vulnerability. Contribution of Working Group II to the Third
Assessment Report of the Intergovernmental Panel on Climate Change, University Press,
Cambridge, 641-692
Lindley, J.A. 1987. Continuous Plankton Records: the geographical distributions and seasonal
cycles of decapod crustacean larvae and pelagic post-larvae in the North-eastern Atlantic
Ocean and the North Sea, 1981-83. J. mar. biol. Ass. U.K. 67, 145-150
Loewe, P., G. Becker, U. Brockmann, S. Dick, A. Frohse, J. Herrmann, B. Klein, H. Klein, H.
Nies, S. Schmolke, D. Schrader, A. Schulz, N. Theobald, S. Weigelt: Nordseezustand 2004,
Berichte des BSH, Nr. 40, 221pp, 2006.
www.bsh.de/Produkte/Buecher/Berichte/Bericht40/index.jsp
MacKenzie, B. R., Gislason, H., Möllmann, C., and Köster, F. W. 2007. Impact of 21st
century climate change on the Baltic Sea fish community and fisheries. Global Change
Biology, 13: 1-20.
Marullo S., B. Buongiorno Nardelli, M. Guarracino, R. Santoleri: Observing The
Mediterranean Sea from space: 21 years of Pathfinder-AVHRR Sea Surface Temperatures
(1985 to 2005). Re-analysis and validation. Ocean Science, 3, 299-30, 2007. (a)
Marullo S., R. Santoleri, M. Guarracino, B. Buongiorno Nardelli , V. Artale: Sea surface
temperature trends in the Mediterranean Sea: from interannual to decadal variations.
Geophysical Research Abstracts, Vol. 9, 03578, 2007. (b)
Meehl, G.A., T.F. Stocker, W.D. Collins, P. Friedlingstein, A.T. Gaye, J.M. Gregory, A.
Kitoh, R. Knutti, J.M. Murphy, A. Noda, S.C.B. Raper,I.G. Watterson, A.J. Weaver and Z.-C.
Zhao, 2007: Global Climate Projections. In: Climate Change 2007: The Physical Science
Basis. Contribution of Working Group I to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change [Solomon, S.,D. Qin, M. Manning, Z. Chen, M.
Marquis, K.B. Averyt, M. Tignor and H.L. Miller (eds.)]. Cambridge University Press,
Cambridge, United Kingdom and New York, NY, USA.
Meier, H.E.M., B. Broman, H. Kallio and E. Kjellström, 2007. Projection of Future
Surface Winds, Sea Levels and Wind Waves in the Late 21st Century and their Application
for Impact Studies of Flood Prone Areas in the Baltic Sea Region.
In Geological Survey of Finland, Special Paper, Helsinki, (in press).
Meier, H.E.M., B. Broman and E. Kjellström 2004: Simulated sea level in past and future
climates of the Baltic Sea. Clim. Res., 27, 59-75.
Nerem, R.S., Leulette,E., and A. Cazenave, 2006. Present-day sea level change: A review
Comptes Rendus Geoscience, 338,1077-1083, Elsevier, Amsterdam
Nicholls, R. J., 2004. Coastal flooding and wetland loss in the 21st century: changes
under the SRES climate and socio-economic scenarios. In Glob. Environ. Change, 14, 69-86.
Nicholls, R.J. and A.C. de la Vega-Leinert, Eds., 2007. Implications of sea-level rise for
Europe’s coasts. In J. Coast. Res., Special Issue, (in press).
214
Nicholls, R.J. and Tol, R.S.J., 2006. Impacts and responses to sea-level rise: a global analysis
of the SRES scenarios over the twenty-first century. In Philos. Trans. R. Soc. Lond. A, 364,
1073-1095
Perry, A. L., Low, P. J., Ellis, J. R., and Reynolds, J. D. 2005. Climate Change and
Distribution Shifts in Marine Fishes. Science, 308: 1912-1915.
Quero, J.-C., Du Buit, M.-H., and Vayne, J.-J. 1998. Les observations de poissons tropicaux
et le rechauffement des eaux dans l'Atlantique europeen. Oceanologica Acta, 21: 345-351.
Rahmstorf, S., 2007. A semi-empirical approach to projecting future sea-level rise.
In Science, 315, 367-370
Rahmstorf, S., A. Cazenave, J. A. Church, J. E. Hansen, R. F. Keeling, D. E. Parker, R. C. J.
Somerville, 2007: Recent Climate Observations Compared to Projections, Science 4 May
2007: Vol. 316. no. 5825, p. 709
Raitsos D.E., Reid P.C., Lavender S.J., Edwards M., Richardson A.J. (2005). Extending the
SeaWiFS chlorophyll data set back 50 years in the northeast Atlantic. Geophysical Research
Letters. 32: L06603 [doi:10.1029/ 2005 GL 022484]
Rayner, N. A., D. E. Parker, E. B. Horton, C. K. Folland, L. V. Alexander, D. P. Rowell, E.
C. Kent, and A. Kaplan, 2003: Global analyses of SST, sea ice and night marine air
temperature since the late nineteenth century. J. Geophys. Res., 108, 4407,
doi:10.1029/2002JD002670.
Reid, P.C., Edwards, M., Hunt, H.G. and Warner, A.J. 1998. Phytoplankton change in the North
Atlantic. Nature, 391:546.
Rodionov, S. and J. E. Overland: Application of a sequential regime shift detection method to
the Bering Sea ecosystem, ICES J. Mar. Sci., 62, 328-332, 2005.
Stebbing, A. R. D., Turk, S. M. T., Wheeler, A., and Clarke, K. R. 2002. Immigration of
southern fish species to the south+west of England linked to warming of the North Atlantic
(1960-2001). Journal of the Marine Biological Association U. K., 82: 177-180.
Trenberth, K.E., P.D. Jones, P. Ambenje, R. Bojariu, D. Easterling, A. Klein Tank, D. Parker,
F. Rahimzadeh, J.A. Renwick, M. Rusticucci, B. Soden and P. Zhai, 2007: Observations:
Surface and Atmospheric Climate Change. In: Climate Change 2007: The Physical Science
Basis. Contribution of Working Group I to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M.
Marquis, K.B. Averyt, M. Tignor and H.L. Miller (eds.)]. Cambridge University Press,
Cambridge, United Kingdom and New York, NY, USA.
Tsimplis, M.N., D.K. Woolf, T.J. Osbourn, S. Wakelin, J. Wolf R. Flather, P. Woodworth,
A.G.P. Shaw, P. Challenor and Z. Yan, 2004a. Future changes of sea level and wave heights
at the northern European coasts. In Geophys. Res. Abs., 6, 00332
UNEP, 2007. Global Outlook for Ice& Snow ; UNEP/GRID-ARENDAL, 2007
Volpe, G., R. Santoleri, V. Vellucci, M. Ribera d'Alcalà, S. Marullo, F. D'Ortenzio. The colour of
the Mediterranean Sea: Global versus regional bio-optical algorithms evaluation and implication
for satellite chlorophyll estimates. Remote Sensing of Environment, 107 (2007), 625–638.
Woodworth, P.L., J.M. Gregory and R.J. Nicholls, 2005. Long term sea level changes and
their impacts. In The global coastal ocean: multiscale interdisciplinary processes, A.R.
Robinson, and K. H. Brink, Eds., Cambridge, Massachusetts, p. 715-753.
Terrestrial ecosystems, biodiversity
Araújo, M.B., Thuiller, W. & Pearson, R.G. (2006) Climate warming and the decline of
amphibians and reptiles in Europe. J. Biogeogr. 33: 1712-1728.
Araújo, M.B. & Luoto, M. (2007) The importance of biotic interactions for modelling species
distributions under climate change. Global ecology and biogeography 16: 743-753.
215
Bakkenes, M. et al. (2006) Impacts of different climate stabilisation scenarios on plants
species in Europe, Global Environmental Change.
Bale, J.S. et al. (2002) Herbivory in global climate change research: direct effects of rising
temperatures on insect herbivores. Global Change Biology 8: 1-16.
Berry, P.M., Thomson, C.L., O'Hanley, J.R. & Glucksman E. (2007) Assessing the
vulnerability of species in Europe and the biogeographic regions to climate change. In Berry,
P.M, Jones, A.P., Nicholls, R.J. & Vos, C.C. (eds) (2007) Assessment of the vulnerability of
terrestrial and coastal habitats and species in Europe to climate change, Annex 2 of Planning
for biodiversity in a changing climate – BRANCH project Final Report, Natural England,
UK.
Biesmeijer, J. C. et al. (2006) Parallel declines in pollinators and insect-pollinated plants in
Britain and the Netherlands. Science 313: 351-354.
Both, C. et al. (2004) Large-scale geographical variation confirms that climate change causes
birds to lay earlier. Proc. R. Soc. Lond. B 271: 1657-1662.
Both, C. et al. (2006) Climate change and population declines in a long-distance migratory
bird. Nature 441: 81-83.
Cannon, R.J.C. (1998) The implications of predicted climate change for insect pests in the
UK, with emphasis on non-indigenous species. Global Change Biology 4: 785-796.
CBD & MNP (2007) Cross-roads of life on earth. Exploring means to meet the 2010
Biodiversity Target. Solution-oriented scenarios for Global Biodiversity Outlook 2. CBD
Technical Series no. 31, Secretariat of the Convention on Biological Diversity (sCBD) and
Netherlands Environmental Assessment Agency (MNP), Montreal and Bilthoven.
CEC (2007) Green Paper on climate change adaptation. Annex, educational material.
CEH, Morton Frederiksen (2005) cited in ‘A warm unwelcome’, by Audrey Schulman
http://www.grist.org/news/maindish/2005/01/25/schulman-seabirds/
Crick, H.Q.P. & Sparks T.H. (1999) Climate change related to egg-laying trends. Nature 399,
423-424.
DEFRA (2007) Conserving biodiversity in a changing climate: guidance on building capacity
to adapt. DEFRA, UK.
Díaz, S., Fargione, J. Stuart Chapin III, F.& Tilman,D. (2006) Biodiversity loss threatens
human well-being. PLoS Biol 4(8): e277. DOI: 10.1371/journal. pbio.0040277.
EEA (2005) Vulnerability and adaptation to climate change.
Estrella, N. & Menzel, A. (2006) Responses of leaf colouring of four deciduous tree species
to climate and weather in Germany. Climate Research 32, 253-267.
Fitter, A. & Fitter, R. (2002) Rapid Changes in Flowering Time in British Plants. Science
296, 1689 (2002).
Gitay, H., Suarez, A., Watson, R.T. & Dokken, D.J. (2002) Climate change and biodiversity;
IPCC Technical paper 5, April 2002, 77p.
Hickling, R. et al. (2006) The distributions of a wide range of taxonomic groups are
expanding polewards. Global Change Biology (2006) 12, 450–455.
Høye, T.T. et al. (2007) Rapid advancement of spring in the high Arctic. Current Biology 17,
no. 12.
Huntley, B. et al. (2008). A Climatic Atlas of European Breeding Birds. Barcelona: Lynx
Edicions.
Huntley, B. (2007) Climatic change and the conservation of European biodiversity: Towards
the development of adaptation strategies. Bern Convention Standing Committee on Climate
Change. Council of Europe, Strasbourg.
IPCC (2007) Working Group II Fourth Assessment Report. Chapter 4: Ecosystems, their
Properties, Goods and Services, and Chapter 12: Europe.
216
Kullman, L. (2006) Long-term geobotanical observations of climate change impacts in the
Scandes of West-Central Sweden. Nordic Journal of Botany 24: 445-467. Copenhagen.
Kullman, L. (2007) Tree line population monitoring of Pinus sylvestris in the Swedish
Scandes, 1973-2005: implications for tree line theory and climate change ecology. Journal of
Ecology 2007 95: 41-52.
Levinsky, I., Skov, F., Svenning, J. & Rahbek, C. (2007) Potential impacts of climate change
on the distributions and diversity patterns of European mammals. Biodiversity and
Conservation. 0960-3115 (Print) 1572-9710 (Online).
Lovejoy, T. & Hannah, L. (2005) Climate change and biodiversity. London, Yale University
Press.
Luo, Z. et al. (2007). Phenological responses of plants to climate change in an urban
environment. Ecological Research 22: 507–514.
McKinney, M.L. & Lockwood, J.L. (1999) Biotic homogenization: a few winners replacing
many losers in the next mass extinction. Trends in Ecology and Evoloution 14: 450-453.
Menzel, A. et al. (2006) European phenological response to climate change matches the
warming pattern. Glob. Change Biol. 12 1969-1976.
Mitchell T., Carter, T.R., Jones P. & Hulme, M. (2004) A comprehensive set of highresolution grids of monthly climate for Europe and the globe: the observed record (19012000) and 16 scenarios (2001-2100): Tyndall Centre Working Paper 55.
Nature’s Calendar UK website http://www.naturescalendar.org.uk/climate+change/past.htm
Nordic Council (2005) Conservation of Nordic Nature in a Changing Climate. Nordic Council
of Ministers, Copenhagen.
Ott, J. (2007) The expansion of Crocothemis erythraea (Brullé, 1832) in Germany – an
indicator of climatic changes. Odonata: Biology of Dragonflies. Tyagi B.K. (Ed.) Scientific
Publishers (India) pp 201-222.
Parmesan, C. et al. (1999) Poleward shifts in geographical ranges of butterfly species
associated with regional warming. Nature 399:579–83.
Parmesan, C. (2006) Ecological and Evolutionary Responses to Recent Climate Change.
Annu. Rev. Ecol. Sust. 37: 637-69.
Parmesan, C. & Yohe, G. (2003) A globally coherent fingerprint of climate change impacts
across natural systems. Nature 421: 37-42.
Pauli, H. et al. (2007) Signals of range expansions and contractions of vascular plants in the
high Alps: observations (1994–2004) at the GLORIA master site Schrankogel, Tyrol, Austria
. Global Change Biology 13, 147–156.
Reid, W. V. et al. (2005) Ecosystems and human well-being: Synthesis, and Policy
Responses. Island Press, Washington, DC.
Root, T.L. et al. (2003) Fingerprints of global warming on wild animals and plants. Nature
421: 57-60.
Schweiger, O. et al. (in press) Climate change can cause spatial mismatch of trophically
interacting species.
Settele, J. et al. (2005) ALARM: Assessing large scale risks for biodiversity with tested
methods. GAIA – Ecological Perspectives in Science, Humanities and Economics 14: 96-72.
Sobrino Vesperinas, E., González Moreno, A., Sanz Elorza, M.,Dana Sánchez, E., Sánchez
Mata, D. & Gavilán, R. (2001) The expansion of thermophilic plants in the Iberian Peninsula
as a sign of climatic change. ‘Fingerprints’ of climate change — adapted behaviour and
shifting species ranges (ed. by G.-R. Walther, C.A. Burga and P.J. Edwards), pp. 163–184.
Kluwer Academic/Plenum Publishers, New York.
Spangenberg, J.H. (2007) Integrated scenarios for assessing biodiversity risks. Sustainable
Development, Online Early DOI: 10.1002/sd.320.
217
Sparks, T.H. et al. (2007) Increased migration of Lepidoptera linked to climate change Eur. J.
Entomol. 104: 139–143, 2007.
Thuiller, W. et al. (2005) Climate change threatens plant diversity in Europe. Proc. Natl.
Acad. Sci. U.S.A. 102: 8245-8250.
Trnka, M. et al. (2005) Modelling of the European Corn Borer climatic niche under expected
climate conditions. Rožnovský, J. & Litschmann, T. (eds): Bioklimatologie současnosti a
budoucnosti“, Křtiny 12. – 14.9.2005, ISBN 80-86 690–31-08.
UKCIP (2005) UK Climate Impacts Programme. http://www.ukcip.org.uk.
Visser, M. E. et al. (2003) Variable responses to large-scale climate change in European
Parus populations. Proc. R. Soc. Lond. B 270, 367–372.
Walther, G-R., Berger, S. & Sykes, M.T. (2005) An ecological “footprint” of climate change.
Proceedings of the Royal Society of London Series B – Biological Sciences 272: 1427-1432.
Walther, G-R., Beissner S & Burga, C.A. (2005) Trends in the uphill shift of alpine plants.
Journal of vegetation science 16: 541-548.
Warren, M. S. et al. (2001) Rapid response of British butterflies to opposing forces of climate
and habitat change. Nature 414, 65-69.
Wilson, J.W. et al. (2005) Changes to the elevational limits and extent of species ranges
associated with climate change. Ecol. Lett. 8:1138–46.
Yamamura, K. & Kiritani, K. (1998) A simple method to estimate the potential increase in the
number of generations under global warming in temperate zones. Appl. Ent. and Zool.
33:289-298.
Agriculture and forestry
R. Aerts, J.H.C. Cornelissen and E. Dorrepaal: Plant performance in a warmer world: general
responses of plants from cold, northern biomes and the importance of winter and spring
events. (2006) Plant Ecology. Volume 182, Numbers 1-2, pp. 65-77
Adams, R.M., R.A. Fleming, C.C. Chang, B.A. McCarl, and C. Rosenzweig, (1995): “A
reassessment of the economic effects of global climate change in on U.S. agriculture”.
Climate Change, 30, 147-167.
Alexander, L.V. et al. (2006). “Global observed changes in daily climate extremes of
temperature and precipitation”. Journal of Geophysical Research, 111(D5): D05109.
Allen, C.D., Breshears, D.D., 1998. Drought-induced shift of a forest–woodland ecotone:
Rapid landscape response to climate variation. Proc. Natl. Acad. Sci. USA 95, 14839-14842.
Alpert, P., et al. (2002). “The paradoxical increase of Mediterranean extreme daily rainfall in
spite of decrease in total values”. Geophysical Research Letters 29 (11): 31-1 - 31-4.
Bielza, M., Conte, C., Dittmann, C., Gallego, J., Stroblmair, J. (2008) Agricultural Insurance
Schemes. EUR report in press.
Brinkmann, W.A.R.: Growing season length as an indicator of climatic variations? (1979)
Climatic Change, 2 (2), pp. 127-138.
Broadmeadow, M., Ray, D., Sing, L. and Poulsom, E. 2003 Climate change and British
woodland: what does the future hold? In Forest Research Annual Reports and Accounts
2002–2003 . HMSO, Edinburgh, pp. 70 – 83.
Brooker, R.W. Plant-plant interactions and environmental change. 2006. New Phytologist
171, 271-284.
Brunetti, M. et al. (2004). “Changes in daily precipitation frequency and distribution in Italy
over the last 120 years”. Journal of Geophysical Research 109 (D05): D05102.
Caesar, J., L. Alexander, and R. Vose (2006). “Large-scale changes in observed daily
maximum and minimum temperatures: Creation and analysis of a new gridded data set”.
Journal of Geophysical Research 111 (D05): D05101.
218
Cannell, M.G.R., Thornley, J.H.M., Mobbs, D.C. and Friend, A.D. 1998 UK conifer forests
may be growing faster in response to increased N deposition, atmospheric CO 2 and
temperature. Forestry 71 , 277 – 296.
Casalegno S., Amatulli G., Bastrup-Birk A., Houston T. 2007, Modelling Current and Future
Distribution of European Forest Categories. Proceedings of the 6th European Conference on
Ecological Modelling: Challenges for ecological modelling in a changing world: Global
Changes, Sustainability and Ecosystem Based Management. November 27-30, 2007. Trieste –
Italy.
Chmielewski, F.-M., Rötzer, T. : Annual and spatial variability of the beginning of growing
season in Europe in relation to air temperature changes. (2002) Climate Research, 19 (3), pp.
257-264.
Commission Bulletin, Forestry commission, Edinburgh. No 125: 99-118.
Cox D.R., Stuart A., 1955. Some quick tests for trend in location and dispersion. Biometrika,
(42): 80-95.
Darwin, R., 2004: “Effects of greenhouse gas emissions on world agriculture, food
consumption, and economic welfare”. Climatic Change 66, 191–238.
Defila, C., Clot, B.: Phytophenological trends in Switzerland. (2001) International Journal of
Biometeorology, 45 (4), pp. 203-207.
Dersch, G. and Boehm, K., 1997. Bodenschutz in Österreich, edited by Blum, W. E. H.,
Klaghofer, E., Loechl, A., and Ruckenbauer, P., (Bundesamt und Forschungszentrum fÜr
Landwirtschaft, Österreich, Germany), pp. 411–432.
Dixon, R.K., Brown, S., Houghton, R.A., Solomon, A.M., Trexler, M.C., and Wisniewski, J.
1994. Carbon pools and flux of global forest ecosystems. Science (Washington, D.C.), 263:
185–190.
Duchêne, E. and C. Schneider (2005). Grapevine and climatic changes: a glance at the
situation in Alsace. Agron. Sustain. Dev. 24, 93-99.
Eastaugh, C. IUFRO Occasional Paper No. 21: Adaptations of Forests to Climate Change: A
Multidisciplinary Review ISSN 1024-414X (NB Before citing this document, please contact
IUFRO for updated details on the paper’s status).
EC (2005a). “Communication from the Commission to the Council on risk and crisis
management in agriculture” COM (2005) 74, Brussels. In:
EC (2005b). “Commission Staff Working Document accompanying the Communication on
risk and crisis management in agriculture”. In:
EC GHG Inventory 2007 Annual European Community greenhouse gas inventory 1990–2005
and inventory report 2007.
EC MARS Bulletin (2006) http://mars.jrc.it/Bulletins/Europe/2006/IP-06-1097_EN.pdf
EC, 2006a. Communication from the Commission to the Council, the European Parliament,
the European Economic and Social Committee and the Committee of the Regions, Thematic
Strategy for Soil Protection: Summary of the Impact Assessment, Sec(2006)1165.
EC, 2006b. Communication of 22 September 2006 from the Commission to the Council, the
European Parliament, the Economic and Social Committee and the Committee of the
Regions: Thematic Strategy for Soil Protection [COM (2006) 231 final].
EC, 2006c. Proposal for a Directive of the European Parliament and of the Council
establishing a framework for the protection of soil and amending Directive 2004/35/EC
[COM (2006) 232 final].
European Environment Agency Technical report No 7/2007: Annual European Community
greenhouse gas inventory 1990–2005 and inventory report 2007: Submission to the UNFCCC
Secretariat (Version 27 May 2007).
Evans, H., Straw, N. & Watt, A. 2002. Climate change: implications for insect pests. Forestry
219
Fischer, G., H. van Velthuizen, M. Shah, and F.O. Nachtergaele, (2002b): “Global Agroecological Assessment for Agriculture in the 21st Century: Methodology and Results”.
Research Report RR-02-02. ISBN 3-7045-0141-7, International Institute for Applied Systems
Analysis, Laxenburg, Austria, 119 and CD-Rom pp.
Fischer, G., M. Shah, and H. van Velthuizen, (2002a): “Climate Change and Agricultural
Vulnerability”, contribution to the UN World Summit on Sustainable Development,
Johannesburg 2002. International Institute for Applied Systems Analysis, Laxenburg, Austria.
Flannigan M.D., Amiro B.D., Logan K.A., Stocks B.J., Wotton B.M., 2005. Forest Fires and
Climate Change in the 21st Century. Mitigation and Adaptation Strategies for Global Change.
Springer (2005) 11: 847-859.
Flannigan M.D., Stocks B.J., Wotton B.M., 2000. Climate Change and Forest Fires. The
Science of the Total Environment. Elsevier. 262(2000) 221-229.
Frich, P. et al. (2002). “Observed coherent changes in climatic extremes during the second
half of the twentieth century”. Climatic Research 19: 193–212.
Genovese, G.; Lazar, C; Micale, F.; Orlandi, Stefania; Royer, A. 2004 “Climate changes for
Europe reflected in the phenology of wheat simulated with the CGMS Model”. VIII Congress
of European Society for Agronomy, Copenhagen, 11-15 July 2004, pp. 263-264
Groisman, P.Y. et al. (2004). “Contemporary changes of the hydrological cycle over the
contiguous United States: Trends derived from in situ observations”. Journal of
Hydrometeorology 5 (1): 64–85.
Groisman, P.Y., et al. (2005). “Trends in intense precipitation in the climate record”. Journal
of Climate 18 (9): 1326–1350.
H. Spiecker et al. (eds., 1996) Growth Trends in European Forests. Springer-Verlag Berlin,
372 pp.
Heim, R.R. (2002). “A review of twentieth-century drought indices used in the United
States”. Bulletin of the American Meteorological Society 83: 1149
Hoogenboom, G;, Jones, J.W.; Porter, C.H.; Wilkens, P.W.; Boote, K.J.; Batchelor, W.D.;
Hunt, L.A.; Tsuji, G.Y. (editors). 2003. Decision Support System for Agrotechnology
Transfer Version 4.0. Volume 1: Overview. University of Hawaii, Honolulu, HI.
http://europa.eu.int/comm/agriculture/publi/communications/risk/com74_en.pdf
http://europa.eu.int/comm/agriculture/publi/communications/risk/workdoc_en.pdf
IPCC 2007 Intergovernmental Panel on Climate Change. 2007. Fourth Assessment Report.
IPCC, 2000. Special Report on Land Use, Land-Use Change and Forestry. Cambridge
University Press, Cambridge, U.K.
IPCC, 2007: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of
Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on
Climate Change, M.L. Parry, O.F. Canziani, J.P. Palutikof, P.J. van der Linden and C.E.
Hanson, Eds., Cambridge University Press, Cambridge, UK, 976pp.
Janssens I A et al 2005 The carbon budget of terrestrial ecosystems at country-scale—a
European case study Biogeosciences 2 15–26
Janssens I.A., A. Freibaur, B. Schlamadinger, R. Ceulemans, P.Ciais, A. Dolman, M.
Heimann, G.-J. Nabuurs, P. Smith, R. Valentini and E.-D. Schulze, 2004. The carbon budget
of terrestrial ecosystems at the country-scale – a European case study. Biogeosciences
Discussions, 1, pp. 167–193, www.biogeosciences.net/bgd/1/167/SRef-ID: 18106285/bgd/2004-1-167 © European Geosciences Union 2004.
Jones, P.D., Briffa, K.R.: Growing season temperatures over the former Soviet Union. (1995)
International Journal of Climatology, 15 (9), pp. 943-959.
Jones, R.J.A., Hiederer, R., Rusco, E., Loveland, P.J. and Montanarella, L., 2004. The map of
organic carbon in topsoils in Europe, Version 1.2, September 2003: Special Publication
220
Ispra, No.72 (S.P.I.04.72). Office for Official Publications of the European Communities,
Luxembourg.
Keeling CD, Chin FJS, Whorf TP. 1996. Increased activity of northern vegetation inferred
from atmospheric CO2 measurements. Nature 382: 146–149.
Kellomäki, S., Strandman, H., Nuutinen, T., Peltola, H., Korhonen, K.T. and Väisänen, H.
2005. Adaptation of forest ecosystems, forests and forestry to climate change. FINADAPT
Working Paper 4, Finnish Environment Institute Mimeographs 334, Helsinki, 44 pp.
Klein Tank, A.M.G. and G.P. Können (2003). “Trends in indices of daily temperature and
precipitation extremes in Europe, 1946–1999”. Journal of Climate 16: 3665–3680.
Lappalainen, E., 1996. Global Peat Resources (International Peat Society, Jyskä), Finland.
Lazar, C. and G. Genovese (eds). 2004. Methodology of the MARS Crop Yield Forecasting
System. Volume 2. Agrometeorological data collection, processing and analysis. In: European
Communities (ISBN 92-894-8181-1).
Liski, J., Perruchoud, D., and Karjalainen, T., 2002. Increasing carbon stocks in the forest
soils of western Europe, For. Ecol. Manage., 169, pp. 159–175.
Magnani F et al 2007 The human footprint in the carbon cycle of temperate and boreal forests
Nature 447, 848–50
Magnuson JJ, Robertson DM, Benson BJ, Wynne RH, Livingstone DM, Arai T, Assel RA,
Barry RG, Card V, Kuusisto E, Granin NG, Prowse TD, Stewart KM, Vuglinski VS. 2000.
Historical trends in lake and river ice cover in the Northern Hemisphere. Science 289: 1743–
1746.
Maheras, P., et al., (2004): “On the relationships between circulation types and changes in
rainfall variability in Greece”. Int. J. Climatol., 24, 1695–1712
McCarthy JJ, Canziani OF, Leary NA, Dokken DJ, White KS (eds). 2001. Climate Change
2001: Impacts, Adaptation, and Vulnerability. Contribution of the Working Group II to the
Third Assessment Report of the Inter-governmental Panel on Climate Change. Cambridge
University Press: Cambridge, UK; 1000 pp.
Menzel A. 2000. Trends in phenological phases
Menzel, A., Estrella, N.: Plant phenological changes. (2001) Fingerprints of Climate Change
- Adapted Behaviour and Shifting Species Ranges, pp. 123-137. Walther GR, Burga CA,
Edwards PJ (eds). Kluwer Academic/Plenum: New York and London
Menzel, A., Fabian, P.: Growing season extended in Europe. (1999) Nature, 397 (6721), p.
659.
Menzel, A.: Trends in phenological phases in Europe between 1951 and 1996. (2000)
International Journal of Biometeorology, 44 (2), pp. 76-81.
Mikhail V. Kozlov and Natalia G. Berlina (2002): Decline in Length of the Summer Season
on the Kola Peninsula, Russia. Climatic Change, Volume 54, Number 4, pp 387-398.
Myneni RB, Keeling CD, Tucker CJ, Asrar G, Nemani RR. 1997. Increased plant growth in
the northern high latitudes from 1981 to 1991. Nature 386: 698–702.
Parry, M.L., C. Rosenzweig, A. Iglesias, G. Fischer, and M. Livermore, (1999): “Climate
change and world food security: a new assessment”. Global Environmental Change, 9, 51-67.
Parry, M.L., C. Rosenzweig, A. Iglesias, M. Livermore, and G. Fischer, (2004): “Effects of
climate change on global food production under SRES emissions and socio-economic
scenarios”. Global Environmental Change, 14(1), 53-67.
Robeson, S.M.: Increasing growing-season length in Illinois during the 20th century. (2002)
Climatic Change, 52 (1-2), pp. 219-238
Root TL, Price JT, Hall KR, Schneider SH, Rosenzweig C, Pounds A. 2003. Fingerprints of
global warming on wild animals and plants. Nature 421: 57–60
San Miguel-Ayanz, J., Carlson, J.D., Alexander, M., Tolhurst, K., Morgan, G., Sneeuwjagt,
R. and Dudley, M. 2003, Current Methods to Assess Fire Danger Potential. In: Wildland Fire
221
Danger Estimation and Mapping. The Role of Remote Sensing Data (E. Chuvieco, Ed.)
World Scientific Publishing, Singapore, pp. 21-61.
Scheifinger, H., Menzel, A., Koch, E., Peter, Ch.: Trends of spring time frost events and
phenological dates in Central Europe. (2003) Theoretical and Applied Climatology, 74 (1-2),
pp. 41-51.
Sitch, S., Cox, P.M., Collins, W.J., Huntingford, C., 2007. Indirect radiative forcing of
climate change through ozone effects on the land-carbon sink. Nature 448, 791-795.
Sleutel, S., De Neve, S., and Hofman, G., 2003. Estimates of carbon stock changes in Belgian
cropland, Soil Use & Manage., 19, pp. 166–171.
Smith J., P. Smith, M. Wattenbach, S. Zaehle, R. Hiederer, R.J.A. Jones, L. Montanarella,
M.D.A Rounsevell, I. Reginster, F. Ewert, 2005. Projected changes in mineral soil carbon of
European croplands and grasslands, 1990-2080. Global Change Biology. 2005 11(12). p.
2141
Tait, A., Zheng, X.: Mapping frost occurrence using satellite data. (2003) Journal of Applied
Meteorology, 42 (2), pp. 193-203.
Trenberth, K.E. et al. (2003). “The changing character of precipitation”. Bulletin of the
American Meteorological Society 84: 1205–1217.
Trenberth, K.E., P.D. Jones, P. Ambenje, R. Bojariu, D. Easterling, A. Klein Tank, D. Parker,
F. Rahimzadeh, J.A. Renwick, M. Rusticucci, B. Soden and P. Zhai (2007). “Observations:
Surface and Atmospheric Climate Change”. In: Climate Change 2007: The Physical Science
Basis. Contribution of Working Group I to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M.
Marquis, K.B. Averyt, M. Tignor and H.L. Miller (eds.)]. Cambridge University Press,
Cambridge, United Kingdom and New York, NY, USA.
Tucker CJ, Slayback DA, Pinzon JE, Los SO, Myneni RB, Taylor MG. 2001. Higher northern
latitude normalized difference vegetation index and growing season trends from 1982–1999.
International Journal of Biometeorology 45: 184–190
Van Wagner, C.E. 1987. Development and structure of the Canadian Forest Fire Weather
Index System. Canadian Forestry Service, Ottawa, Ontario. Forestry Technical Report 35. 37
p.
Vleeshouwers, L. M. and Verhagen, A., 2002. Carbon emissions and sequestration by
agricultural land use: a model study for Europe, Glob. Change Biol., 8, pp. 519–530.
Walther GR, Post E, Convey P, Menzel A, Parmesan C, Beebee TJC, Fromentin JM, HoeghGuldberg O, Bairlein F. 2002. Ecological responses to recent climate change. Nature 416:
389–395
Way, J. d , Zimmermann, R., Rignot, E., McDonald, K, Oren, R.: Winter and spring thaw as
observed with imaging radar at BOREAS. Journal of Geophysical Research D: Atmospheres.
Volume 102, Issue 24, 26 December 1997, Pages 29673-29684.
Winnett, S.M., 1998: Potential effects of climate change on U.S. forests: a review. Climate
Research, 11, 39–49.
Yan, Z., Jones, P.D., Davies, T.D., Moberg, A., Bergström, H., Camuffo, D., Cocheo, C., (...),
Yang, C.: Trends of extreme temperatures in Europe and China based on daily observations.
(2002) Climatic Change, 53 (1-3), pp. 355-392.
Zhang, X., F.W. Zwiers, and G. Li, 2004: Monte Carlo experiments on the detection of trends
in extreme values. J. Clim., 17, 1945–1952.
Zhou, L., Tucker, C.J., Kaufmann, R.K., Slayback, D., Shabanov, N.V., Myneni, R.B.:
Variations in northern vegetation activity inferred from satellite data of vegetation index
during 1981 to 1999. (2001) Journal of Geophysical Research D: Atmospheres, 106 (D17),
pp. 20069-20083
222
Water quantity, droughts, floods
River flow
Andréasson, J., S. Bergström, B. Carlsson, L.P. Graham, G. Lindström, 2004. Hydrological
change – climate change impact simulation for Sweden. Ambio, 33, 228-234.
Alcamo, J., M. Flörke, M. Märker, 2007. Future long-term changes in global water resources
driven by socio-economic and climatic change. Hydrological Sciences Journal, 52, 247-275.
Arnell, N.W., 2004. Climate change and global water resources: SRES emissions and socioeconomic scenarios. Global Environmental Change, 14, 31-52.
Barnett, T.P., Adam, J.C., Lettenmaier, D.P., 2005. Potential impacts of warming on water
availability in snow-dominated regions. Nature, 438, 303-309.
Birsan, M.V., P. Molnar, P. Burlando, M. Pfaundler, 2005. Streamflow trends in Switzerland.
Journal of Hydrology, 314, 312-329.
Dankers, R., L. Feyen, 2008. Climate change impacts on river flow in Europe. Manuscript in
preparation.
Dixon, H., D.M. Lawler, A.Y. Shamseldin, 2006. Streamflow trends in western Britain.
Geophysical Research Letters, 23, L19406, doi:10.1029/2006GL027325.
Hanneford J., T. Marsh, 2006. An assessment of trends in UK runoff and low flows using a
network of undisturbed catchments. International Journal of Climatology, 26, 1237-1253.
Hisdal, H., E. Holmqvist, E. Kuusisto, G. Lindström, L.A. Roald, 2007. Has streamflow
changed in the Nordic countries? Climate Research, submitted.
Jasper, K., Calanca, P., Gyalistras, D., Fuhrer, J., 2004. Differential impacts of climate change
on the hydrology of two alpine rivers. Climate Research, 26, 113-125.
Lindström, G., S. Bergström (2004). Runoff trends in Sweden 1807-2002. Hydrological
Sciences Journal, 49(1), 69-83.
Milly, P. C. D., K. A. Dunne, A. V. Vecchia, 2005. Global pattern of trends in streamflow and
water availability in a changing climate. Nature 438, 347-350.
Wade, S., J.-P. Vidal, C. Dabrowski, P. Young, R. Romanowicz, 2005. Effect of climate
change on river flows and groundwater recharge. A practical methodology. Trends in UK
river flows: 1970-2002. UKWIR Report 05/CL/04/5.
Wang, W., P.H.A.J.M. Van Gelder, J.K. Vrijling, 2005. Detection of changes in streamflow
series in western Europe over 1901-2000.
Werritty, A., 2002. Living with uncertainty: climate change, river flows and water resource
management in Scotland. The Science of the Total Environment, 294, 29-40.
River floods
Barnolas, M., M.C. Llasat, 2007. A flood geodatabase and its climatological implications: the
case of Catalonia for the last century. Natural Hazards and Earth System Sciences, 7, 271281.
Brázdil, R., O. Kotyza, P. Dobrovolný, 2006. July 1432 and August 2002 – two millennial
floods in Bohemia? Hydrological Sciences Journal, 51(5), 848-863.
Becker, A., U. Grunewald, 2003. Disaster magement – flood risk in Central Europe. Science,
300(5622): 1099-1099.
Christensen O.B., J.H., Christensen, 2003. Severe summertime flooding in Europe, Nature,
421, 805-806.
Cyberski J. M. Grześ, M. Gutry-Korycka, E. Nachlik, Z. Kundzewicz, 2006. History of floods
on the River Vistula. Hydrological Sciences Journal, 51(5), 799-817.
223
Dankers, R., L. Feyen, 2008. Climate change impact on flood hazard in Europe: An
assessment based on high resolution climate simulations. Manuscript submitted to Journal of
Geophysical Research Atmospheres.
Dixon, H., D.M. Lawler, A.Y. Shamseldin, 2006. Streamflow trends in western Britain.
Geophysical Research Letters, 33, L19406, doi: 10.129/2006GL027325.
Glaser, R., H, Stangl, 2003. Historical floods in the Dutch Rhine Delta. Natural Hazards and
Earth System Sciences, 3, 605-613.
Hisdal, H., E. Holmqvist, E. Kuusisto, G. Lindström, L.A. Roald, 2007. Has streamflow
changed in the Nordic countries? Climate Research, submitted.
Kay, A., Reynard, N.S., R.N. Jones, 2006. RCM rainfall for UK flood frequency estimation.
II. Climate change results. Journal of Hydrology, 318: 163-172.
Kundzewicz, Z.W., S. Budhakooncharoen, A. Bronstert, H. Hoff, D. Lettenmaier, L. Menzel,
R. Schulze, 2002. Coping with variability and change: floods and droughts. Natural Hazards
Forum, 26, 263-274.
Kundzewicz, Z.W., Graczyk D., Maurer T., Pińskwar I., Radziejewski M., Svensson C.,
Szwed M. (2005) Trend detection in river flow series: 1. Annual maximum flow.
Hydrological Sciences Journal, 50(5), 797-810.
Kundzewicz Z.W., M. Radziejewski, I. Pińskwar, 2006. Precipitation extremes in the
changing climate of Europe. Climate Research, 31, 51-58.
Lehner B., Döll P., Alcamo J., Henrichs T., Kaspar F., 2006. Estimating the impact of global
change on flood and drought risks in Europe: a continental integrated analysis. Climatic
Change, 75, 273-299.
Macklin, M.G., B.T. Rumsby, 2007. Changing climate and extreme floods in the British
Uplands. Transactions of the Institute of British Geographers, 32(2), 168-186.
Milly, P.C.D., Wetherald, R.T., Dunne, K.A. and Delworth, T.L., 2002, Increasing risk of
great floods in a changing climate. Nature, 415.
Mudelsee, M., M. Börngen, G. Tetzlaff, U. Grünewald, 2003. No upward trends in the
occurence of extreme floods in Central Europe. Nature, 425, 166-169.
Palmer, T.N., and J. Räisänen, 2002. Quantifying the risk of extreme seasonal precipitation
events in a changing climate. Nature, 415, 512-514.
Pinter, N., B.S. Ickes, J.H. Wlosinski, R.R. van der Ploeg, 2006. Trends in flood stages:
Contrasting results from the Mississippi and Rhine River systems. Journal of Hydrology, 331,
554-566.
Ramos, C., E. Reis, 2002. Floods in Southern Portugal: their physical and human causes,
impacts and human response. Mitigation and Adaptation Strategies for Global Change, 7,
267-284.
Robson, A., 2002. Evidence for trends in UK flooding. Phil. Trans. R. Soc. Lond. A, 360,
1327-1343.
River flow drought
Alcamo, J., Döll, P., Henrichs, T., Kaspar, F., Lehner, B., Rösch, T. and Siebert, S., 2003.
Global estimates of water withdrawals and availability under current and future business-asusual conditions. Hydrological Sciences Journal, 48, 339-348.
Barnett, T.P., J.C. Adam, D.P. Lettenmaier, 2005. Potential impacts of a warming climate on
water availability in snow-dominated regions. Nature, 438, 303-309.
Feyen, L., R. Dankers, 2008. River flow drought in Europe in a future climate. Manuscript
submitted to Global Environmental Change.
Döll, P., 2002. Impact of climate change and variability on irrigation requirements: a global
perspective. Climatic Change, 54, 269-293.
224
Douville, H., F. Chauvin, S. Planton, J.F. Royer, D. Salas-Melia, S. Tyteca, 2002. Sensitivity
of the hydrological cycle to increasing amounts of greenhouse gasses and aerosols. Climate
Dynamics, 20, 45-68.
Hanneford J., T. Marsh, 2006. An assessment of trends in UK runoff and low flows using a
network of undisturbed catchments. International Journal of Climatology, 26, 1237-1253.
Hisdal, H., Stahl, K., Tallaksen, L.M. & Demuth, S., 2001. Have droughts in Europe become
more severe or frequent? International Journal of Climatology, 21, 317-333.
Hisdal, H., E. Holmqvist, E. Kuusisto, G. Lindström, L.A. Roald, 2007. Has streamflow
changed in the Nordic countries? Climate Research, submitted.
Kundzewicz, Z.W., S. Budhakooncharoen, A. Bronstert, H. Hoff, D. Lettenmaier, L. Menzel,
R. Schulze, 2002. Coping with variability and change: floods and droughts. Natural Hazards
Forum, 26, 263-274.
Lang, M., B. Renard, E. Sauquet, P. Bois, A. Dupeyrat, C. Laurent, O. Mestre, H. Niel, L.
Neppel, J. Gailhard, 2006. Climate Variability and Change – Hydrological Impacts, IAHS
Publ., 308.
Lehner B., Döll P., Alcamo J., Henrichs T., Kaspar F., 2006. Estimating the impact of global
change on flood and drought risks in Europe: a continental integrated analysis. Climatic
Change, 75, 273-299.
Schröter, D., W. Cramer, R. Leemans, C. Prentice, M.B. Araújo, N.W. Arnell, A. Bondeau,
H. Bugmann, T.R. Carter, C.A. Gracia, A.C. de la Vega-Leinert, M. Erhard, F. Ewert, M.
Glendining, J.I. House, S. Kankaanpää, R.J.T. Klein, S. Lavorel, M. Lindner, M.J. Metzger, J.
Meyer, T.D. Mitchell, I. Reginster, M. Rounsevell, S. Sabaté, S. Sitch, B. Smith, J. Smith, P.
Smith, M.T. Sykes, K. Thonicke, W. Thuiller, G. Tuck, S. Zaehle, B. Zierl, 2005. Ecosystem
Service Supply and Vulnerability to Global Change in Europe. Science, 310(5752), 13331337.
UNEP, 2006. Geo Year Book 2006. An overview of our changing environment.
van der Schrier, G., K.R. Briffa, P.D. Jones, T.J. Osborn, 2006. Summer moisture availability
across Europe. Journal of Climate, 19, 2818-2834.
Boxes
Stuyfzand P. J., Ruediger R.L, and Kristensen P. 2007. Impact of climate change on
groundwater. A background note for EEA produced by the EEA ETC/Water.
Water quality and fresh water ecology
Adrian R, Wilhelm S. and Gerten D. 2006: Life-history traits of lake plankton species may
govern their phenological response to climate warming. Global Change Biology 12, 652–
661.
Ambrosetti W., and Barbanti, L. 1999. Deep water warming in lakes: an indicator of climate
change. Journal of Limnology 58: 1-9.
Anneville, O., Ginot, V. and Angeli, N. 2002. Restoration of Lake Geneva: Expected versus
observed responses of phytoplankton to decreases in phosphorus. Lakes & Reservoirs:
Research and Management, 7, 67–80.
Anneville, O.;Gammeter, S., and Straile, D. 2005. Phosphorus decrease and climate
variability: mediators of synchrony in phytoplankton changes among European peri-alpine
lakes. Freshwater Biology 50:1731-1746.
Barica, J., and J. A. Mathias (1979), Oxygen depletion and winterkill risk in small prairie
lakes under extended ice cover, J. Fish. Res. Board Can., 36, 980– 986.
Bauernfeind, E. and U.H. Humpesch 2001. Die Eintagsfliegen Zentraleuropas (Insecta:
Ephemeroptera) Bestimmung und Ökologie. Verlag des Naturhistorischen Museums, Wien
Austria.
225
Benson, B., and J. Magnuson. 2000, updated 2006. Global lake and river ice phenology
database. Boulder, CO: National Snow and Ice Data Center/World Data Center for
Glaciology. Digital media.
Biodiversity Indicators, 2006. Climate Change: Trend of Southern European dragonfly
species . Research Institute for Nature and Forest, Brussels. www.natuurindicatoren.be
(updated 08-05-2006). Available at
http://www.natuurindicatoren.be/indicatorenportal.cgi?lang=en&detail=404&id_structuur=25
Burgmer T., Hillebrand H. & Pfenninger M. 2007: Effects of climate-driven temperature
changes on the diversity of freshwater macroinvertebrates. Oecologia 151: 93–103
BUWAL, BWG, MeteoSchweiz. 2004. Auswirkungen des Hitzesommers 2003 auf die
Gewässer. Schriftenreihe Umwelt Nr. 369. Bern-Ittigen: Bundesamt für Umwelt, Wald und
Landschaft, 174 p.
Dabrowski, M.;Marszelewski, W., and Skowron, R. 2004. The trends and dependencies
between air and water temperatures in lakes in northern Poland from 1961-2000. Hydrology
and Earth System Sciences 8:79-87.
Daufresne, M.;Bady, P., and Fruget, J. F. 2007. Impacts of global changes and extreme
hydroclimatic events on macroinvertebrate community structures in the French Rhone River.
Oecologia 151:544-559.
Daufresne, M.;Roger, M. C.;Capra, H., and Lamouroux, N. 2004. Long-term changes within
the invertebrate and fish communities of the Upper Rhone River: effects of climatic factors.
Global Change Biology 10:124-140.
Dokulil, M. T.;Jagsch, A.;George, G. D.;Anneville, O.;Jankowski, T.;Wahl, B.;Lenhart,
B.;Blenckner, T., and Teubner, K. 2006. Twenty years of spatially coherent deepwater
warming in lakes across Europe related to the North Atlantic Oscillation. Limnology and
Oceanography 51:2787-2793.
Dyble J., Paerl H.W. and Neilan B.A. 2002. Genetic characterization of Cylindrospermopsis
raciborskii (Cyanobacteria) isolates from diverse geographic origins based on nifH
andcpcBA-IGS nucleotide sequence analysis. Appl. Environ. Microbiol. 68: 2567–2571.
Franssen H. J. Hendricks and Scherrer S. C. 2008: Freezing of lakes on the Swiss plateau in
the period 1901–2006. Int. J. Climatol. (in press)
George G., Hewitt D, Jennings E., Allott N. and McGinnity P. 2005: The impact of changes
in the weather on the surface temperatures of Windermere (UK) and Lough Feeagh (Ireland)
In Proceedings of the Fourth Inter-Celtic Colloquium on Hydrology and Management of
Water Resources, Guimaraes, Portugal, Lobo Ferreira JP, Vieira JMP (eds). Associa¸cˆao
Portuguesa dos Recursos H`ıdricos (APRH), Lisbon,
http://www.aprh.pt/celtico/TEXT/papers.html
George G., Hurley M. and Hewitt D. 2007: The impact of climate change on the physical
characteristics of the larger lakes in the English Lake District. Freshwater Biology 52, 1647–
1666.
George, D.G. and Hurley, M.A. (2004) The influence of sampling frequency on the detection
of long-term change in three lakes in the English Lake District. Aquat. Ecosyst. Health and
Management 7, 1-14.
Greenbank, J. (1945), Limnological conditions in ice-covered lakes, especially related to
winterkill of fish, Ecol. Monogr., 15, 343– 392.
Hari, R. E.;Livingstone, D. M.;Siber, R.;Burkhardt-Holm, P., and Guttinger, H. 2006.
Consequences of climatic change for water temperature and brown trout populations in
Alpine rivers and streams. Global Change Biology 12:10-26.
Hassall C., Thompson D. J., French G. C., &. Harvey I. F 2007: Historical changes in the
phenology of British Odonata are related to climate Global Change Biology
Hering D., Schmidt-Kloiber A, Murphy J., Lücke S., Zamora-Muñoz C., López-Rodríguez
M.J., Huber T. & Graf W. 2007/08: Potential impact of Climate Change on aquatic insects:
226
A sensitivity analysis for European caddisflies (Trichoptera) based on species traits. In
review.
Hickling R., Roy D. B., Hill J. K., & Thomas C. D.2005: A northward shift of range margins
in British Odonata . Global Change Biology 11 (3), 502–506
Hohensinner, S., FWF-Project Machland 1715-1991, Grant-No. P14959-B06
Jackson L. J., Lauridsen T. L., Søndergaard M. and Jeppesen E. 2007: A comparison of
shallow Danish and Canadian lakes and implications of climate change . Freshwater Biology
52, 1782-1792,
Järvinen, M., M. Rask, J. Ruuhija¨rvi, and L. Arvola (2002), Temporal coherence in water
temperature and chemistry under the ice of boreal lakes (Finland), Water Res., 36, 3949–
3956.
Jöhnk, K.D., Huisman, J., Sharples, J., Sommeijer, B., Visser, P.M. & Stroom, J.M. Summer
heatwaves promote blooms of harmful cyanobacteria. Global Change Biology (in press)
Kaste, Ø, Wright, RF, Barkved LJ, Bjerkeng B, Engen-Skaugen T, Magnusson J, Sælthun
NR. 2006. Linked models to assess the impacts of climate change on nitrogen in a Norwegian
river basin and fjord system. Science of the Total Environment 365;1-3:200-222.
Kerr, R. A. 1999. The Little Ice Age - only the latest big chill. Science 248, 2069. DOI:
10.1126/science.284.5423.2069.
Kolar, C.S. and Lodge, D.M. 2000. Freshwater Nonindigenous Species: Interactions with
Other Global Changes. In Mooney, H.A. and Hobbs, R. (eds.): Invasive Species in a
Changing World. pp. 3-30. Island Press, Washington. ISBN 1-55963-782-x
Korhonen J. 2006: Long-term changes in lake ice cover in Finland. Nordic Hydrology 37:
347–363
Korhonen J., 2002: SY566 Suomen vesistöjen lämpötilaolot 1900-luvulla. Part II
http://www.ymparisto.fi/download.asp?contentid=10498&lan=fi
Korhonen, J. 2002. Water temperature conditions of lakes and rivers in Finland in the 20th
century (in Finnish). Suomen Ymparistö 566:1-115.
Korhonen, J. 2005. Suomen vesistöjen jääolot. Suomen Ympäristö 751, Suomen
Ympäristökeskus, Helsinki. (English summary) Available at
http://www.ymparisto.fi/download.asp?contentid=34384&lan=fi
Leppäranta, M., A. Reinart, A. Erm, H. Arst, M. Hussainov, and L. Sipelgas (2003),
Investigation of ice and water properties and under-ice light fields in fresh and brackish water
bodies, Nord. Hydrol., 34, 245– 266.
Livingstone, D. M. (1993), Lake oxygenation: Application of a one-box model with ice cover,
Int. Rev. Ges. Hydrobiol., 78, 465– 480.
Livingstone, D. M. (1997) Break-up dates of Alpine lakes as proxy data for local and regional
mean surface air temperatures, Climatic Change 37, 407-439.
Livingstone, D. M. 2003. Impact of secular climate change on the thermal structure of a large
temperate central European lake. Climatic Change 57:205-225.
Livingstone, D. M. and R. Adrian (2007) Modeling the duration of intermittent ice cover on a
lake for climate-change studies. Limnology and Oceanography (submitted)
Magnuson, J. J., Robertson, D. M., Benson, B. J., Wynne, R. H., Livingstone, D. M., Arai, T.,
Assel, R. A., Barry, R. G., Card, V., Kuusisto, E., Granin, N. G., Prowse, T. D., Stewart K.
M. and Vuglinski V. S. (2000a) Historical trends in lake and river ice cover in the Northern
Hemisphere, Science 289, 1743-1746.
Malmaeus, J. M.;Blenckner, T.;Markensten, H., and Persson, I. 2006. Lake phosphorus dynamics and climate warming: A mechanistic model approach. Ecological Modelling 190:1-14
MNP 2006: The effects of climate change in the Netherlands. (Bresser et al. (Eds.) Report
from MNP available at
http://www.mnp.nl/en/publications/2006/TheeffectsofclimatechangeintheNetherlands.html
227
Monteith, D.T., Stoddard, J. L., Evans, C. D., de Wit, H., Forsius, M., Hoegaasen, T.,
Wilander, A., Skjelkvaale, B.L., Jeffries, D.S., Vuorenmaa, J., Keller, B, Kopacek, J., Vesely,
J., 2007. Dissolved organic carbon trends resulting from changes in atmospheric deposition
chemistry. Nature 450: 537-540.
Mooij WM, Hu¨lsmann S, Domis LND, Nolet BA, Bodelier PLE, Boers PCM, Pires LMD,
Gons HJ, Ibelings BW, Noordhuis R, Portielje R, Wolfstein K, Lammens EHRR (2005) The
impact of climate change on lakes in The Netherlands: a review. Aquat Ecol 39(4):381–400
Moore, K., 2007. Climate change impacts on the catchment contribution to lake water
quantity and quality. PhD thesis Uppsala University, Sweden
(http://publications.uu.se/theses/abstract.xsql?dbid=8236)
Noges et al. 2007: The impact of changes in the weather and climate on the seasonal
dynamics of phytoplankton. Journal???
Palecki, M. A. and Barry, R. G. (1986) Freeze-up and break-up of lakes as an index of
temperature changes during the transition seasons: a case study for Finland, Journal of
Climate and Applied Meteorology 25, 893-902.
Pedersen, N. L., and Sand-Jensen, K. 2007. Temperature in lowland Danish streams:
contemporary patterns, empirical models and future scenarios. Hydrological Processes
21:348-358.
Pernaravièiûtë, B. 2004. The impact of climate change on thermal regime of Lithuanian lakes.
Ekologija 2: 58-63.
Phillips, K.A. and Fawley, M.W. (2002) Winter phytoplankton blooms under ice associated
with elevated oxygen levels, Journal of Phycology 38, 1068-1073.
Rijkswaterstaat, measurements Rhine River at Lobith period 1908-2006 (non published). See
also MNP 2006
Rodhe, W. (1955), Can phytoplankton production proceed during winter darkness in subarctic
lakes?, Verh. Int. Ver. Limnol., 12, 117–122.
Schindler, D.W. 2001. The cumulative effects of climate warming and other human stresses
on Canadian freshwaters in the new millennium. Can J. Fish. Aquat. Sci., 58: 18-29.
Šporka, F., Livingstone, D. M., Stuchlík, E., Turek J. and Galas J. (2006) Water temperatures
and ice cover in the lakes of the Tatra Mountains, Biologia 61 (Suppl. 18), S77-S90.
Stewart, K. M. (1976), Oxygen deficits, clarity and eutrophication in some Madison lakes,
Int. Rev. Ges. Hydrobiol., 61, 563– 579.
Walsby, A.E, 2005. Stratification by cyanobacteria in lakes: a dynamic buoyancy model
indicates size limitations met by Planktothrix rubescens filaments, New Phytologist 168, 365–
376.
Walsh CL, Kilsby CG (2006) Potential impacts of climate change on Atlantic salmon: case
study in the Eden catchment, Cumbria, UK. Hydrol. Earth. Syst. Sci., in press (abstract)
Weyhenmeyer, G. A., 2007: Water chemical changes along a latitudinal gradient in relation
to climate and atmospheric deposition. Climate change
Weyhenmeyer, G. A., Meili, M. and Livingstone, D. M. (2004) Nonlinear temperature
response of lake ice breakup, Geophysical Research Letters 31(7), L07203,
doi:10.1029/2004GL019530.
Weyhenmeyer, G. A., Meili, M. and Livingstone, D. M. (2005) Systematic differences in the
trend toward earlier ice-out on Swedish lakes along a latitudinal temperature gradient,
Verhandlungen der internationalen Vereinigung der Limnologie 29(1), 257-260.
Weyhenmeyer, G. A., T. Blenckner, and K. Pettersson (1999), Changes of the plankton spring
outburst related to the North Atlantic oscillation, Limnol. Oceanogr., 44, 1788– 1792.
Weyhenmeyer, G.A. 2001: Warmer winters - are planktonic algal populations in Sweden's
largest lakes affected? Ambio 30, 565-571.
228
Weyhenmeyer, G.A., Blenckner, T.and Pettersson, K. 1999: Changes of the plankton spring
outburst related to the North Atlantic Oscillation, Limnology and Oceanography 44, 17881792.
Whitehead, P. G., Wilby, R. L., Butterfield, D. and Wade, A. J. 2006. Impacts of climate
change on nitrogen in a lowland Chalk stream: an appraisal of adaptation strategies Science of
the Total Environment, 365, 260-273.
Wilhelm, S., and Adrian, R. 2007. Impact of summer warming on the thermal characteristics
of a polymictic lake and consequences for oxygen, nutrients and phytoplankton. Freshwater
Biology. Doi: 10.1111/j.1365-2427.2007.01887.x
Human health
Bouchama A (2007). What should health professionals know? EuroHEAT final meeting,
Bonn, Germany, 22-23 March 2007.
Confalonieri U, Menne B, Akhtar R et al. (2007). Human Health. In: Parry ML et al., eds.
Impacts, Adaptation and Vulnerability. Working Group II contribution to the
Intergovernmental Panel on Climate Change Fourth Assessment Report. Cambridge, United
Kingdom and New York, NY, USA.: Cambridge University Press.
D'Souza R, Becker N, Hall G et al. (2004). Does ambient temperature affect foodborne
disease? Epidemiology, 15 86-92.
Dessai S (2003). Heat stress and mortality in Lisbon Part II. An assessment of the potential
impacts of climate change. Int J Biometeorol, 48 (1):37-44.
Donaldson G, Kovats RS, Keatinge WR et al. (2001). Heat-and-cold-related mortality and
morbidity and climate change. In: Health effects of climate change in the UK. London:
Department of Health; p. 70-80.
Ebi KL (2006). Floods and Human Health. In: Menne B, Ebi KL, eds. Climate Change and
Adaptation Strategies for Human Health. Darmstadt: WHO Regional Office for Europe,
Steinkopff Verlag; p. 99-121.
EC Communication (January 2007)Preliminary results of PESETA have been published in
the Staff Working Paper accompanying the EC Communication on "Limiting Global Climate
Change to 2 degrees Celsius. The way ahead for 2020 and beyond" (see DG Environment
related Web site http://ec.europa.eu/environment/climat/future_action.htm).
Fleury M, Charron DF, Holt JD et al. (2006). A time series analysis of the relationship of
ambient temperature and common bacterial enteric infections in two Canadian provinces.
International Journal of Biometeorology, 50 (6):385-391.
Goulson D, Derwent LC, Hanley M et al. (2005). Predicting calyptrate fly populations from
the weather, and the likely consequences of climate change. Journal of Applied Ecology And
Society, 42 795-804.
Hall GV, D'Souza RM, Kirk MD (2002). Foodborne disease in the new millennium: out of
the frying pan and into the fire. Medical Journal of Australia, 177 (2/16 Dec):614-618.
Hunter PR (2003). Climate change and waterborne and vectorborne disease. J
Appl.Microbiol., 94 37-46.
Kistemann T, Classen T, Koch C et al. (2002). Microbial Load of Drinking Water Reservoir
Tributaries during Extreme Rainfall and Runoff. 68 (5):2188-2197.
Koppe C, Jendritzky G, Pfaff G (2003). Die Auswirkungen der Hitzewelle 2003 auf die
Gesundheit. In: DWD, editor. Klimastatusbericht 2003; p. 152-162
Korenberg E (2004). Environmental causes for possible relationship between climate change
and changes of natural foci of diseases and their epidemiologic consequences. Climate change
and public health in Russia in the XXI Century. Proceeding of the international workshop,
Moscow.
229
Kovats RS, Edwards S, Hajat S et al. (2004). The effect of temperature on food poisoning:
time series analysis in 10 European countries. Epidemiology and Infection, 132 (3):443.
Kovats RS, Edwards SJ, Charron D et al. (2005). Climate variability and campylobacter
infection: an international study. Int J Biometeorol, 49 (4):207-214.
Kriz B, Benes C, Castkova J et al. (1998). Monitorování Epidemiologické‚ Situace V
Zaplavených Oblastech v České Republice v Roce 1997. [Monitoring the Epidemiological
situation in flooded areas of the Czech Republic in 1997.] In: Konference DDD '98;
Kongresove‚ Centrum L ze•sk Kolona da Podebrady, 11. -13. Kvetna 1998 [Proceedings of
the Conference DDD'98, 11-12th May, 1998, Prodebrady, Czech Republic.]. Prodebrady,
Czech Republic.
Lake IR, Bentham G, Kovats RS et al. (2005). Effects of weather and river flow on
cryptosporidiosis. J Water Health, 3 (4):469-474.
Lehane L, Lewis RJ (2000). Ciguatera: recent advances but the risk remains. International
Journal of Food Microbiology, 61 (2-3):91-125.
Louis VR, Gillespie IA, O'Brien SJ et al. (2005). Temperature-driven Campylobacter
seasonality in England and Wales. Appl Environ Microbiol, 71 (1):85-92.
Miettinen I, Zacheus O, von Bonsdorff C (2001). Waterborne
epidemics in Finland in 1998-1999. Water Sci Technol, 43 67-71.
Randolph SE, Rogers DJ (2000). Fragile transmission cycles of tick-borne encephalitis virus
may be distrupted by predicted climate change. Philosl Trans Roy Soc London - Series B, 267
1741-1744.
Reacher M, McKenzie K, Lane C et al. (2004). Health impacts of flooding in Lewes: a
comparison of reported gastrointestinal and other illness and mental health in flooded and
non-flooded households. Commun Dis Public Health, 7 (1):39-46.
Robine JM, Cheung SL, Le Roy S et al. (2007). Report on excess mortality in Europe during
summer 2003. EU Community Action Programme for Public Health, Grant Agreement
2005114. 2003 Heat wave project.Available under
http://ec.europa.eu/health/ph_projects/2005/action1/docs/action1_2005_a2_15_en.pdf
Semenza J, Nichols G (2007). Cryptosporidiosis surveillance and water-borne outbreaks in
Europe. Euro surveill, 12 (5):Available online:
http://www.eurosurveillance.org/em/v12n05/1205-1227.asp.
Tam C, Rodrigues L, O'Brien S et al. (2006). Temperature dependence of reported
Campylobacter infection in England, 1989-1999. Epidemiol Infect, 134 (1):119-125.
6. Economic consequences of Climate change
Abegg, B., Agrawala S., Crick F., and de Montfalcon, A. (2007). Climate change impacts and
adaptation in winter tourism. In: Authors (initials first) Climate change in the European Alps:
adapting winter tourism and natural hazards management. Paris: Organization for Economic
Co-operation and Development. 25-60.
ABI (2005). Financial risk of climate change. Association of British Insurers, London, UK.
Ackerman, F. and Stanton, E. (2006). ‘Climate Change – the Costs of Inaction, Report to
Friends of the Earth England, Wales and Northern Ireland, October , 2006.
Amelung, B., Nicholls, S., & Viner, D. (2007). Implications of Global Climate Change for
Tourism Flows and Seasonality. Journal of Travel Research, 45(3), 285-296.
Amelung B and Viner D. 2006 Mediterrannean Tourism: Exploring the future with the
Tourism Comfort Index Journal of Sustainable Tourism Vol. 14. Nos. 4 pp 349-366
Barredo, J.I., 2007, Major flood disasters in Europe: 1950-2005. Natural Hazards, 42, 125148.
230
Bouwer, L. M., R. P. Crompton, E. Faust, P.Höppe, R. A. Pielke Jr., 2007: DISASTER
MANAGEMENT: Confronting Disaster Losses; Science, 2. November 2007, 318. no. 5851,
p. 753
Crompton, R.P. and K.J. McAneney (in press). Normalised Australian insured losses from
meteorological hazards: 1967-2006. Environmental Science and Policy
EEA (2004): Mapping the impacts of recent natural disasters and technological accidents in
Europe, Environmental issue report No 35, European Environment Agency, Copenhagen,
Denmark.
EEA (2006). Working paper. Climate Change: the Cost of Inaction A review of assessment
studies with a focus on the methodologies used. Specific Agreement No
3570/B2005.EEA.52247. Final working paper, January 2006. Barbara Buchner, Michela
Catenacci, Alessandra Goria (FEEM) Onno Kuik, Richard Tol (IVM).
EEA (2007), ‘Climate Change: the Cost of Inaction and the Cost of Adaptation Costs of
Action and Inaction’, EEA Technical Paper, No. xx/2007, Denmark
Eurostat (2007). Panorama of Energy Energy statistics to support EU policies and solutions.
2007. http://epp.eurostat.ec.europa.eu/cache/ITY_OFFPUB/KS-76-06-604/EN/KS-76-06604-EN.PDF
Feyen, L., Dankers, R., Barredo, J.I., de Roo, A. and Lavalle, C., 2007, Climate change and
floods in Europe. In Climate Change Impacts on the Water Cycle, Resources and Quality Research-Policy Interface, M. Moren-Abat, P. Quevauviller, L. Feyen, A.-S. Heiskanen, P.
Noges, A.L. Solheim and E. Lipiatou (Eds.). Luxembourg: Office for Official Publications of
the European Communities, 37-43.
Fischlin, A., G.F. Midgley, J.T. Price, R. Leemans, B. Gopal, C. Turley, M.D.A. Rounsevell,
O.P. Dube, J. Tarazona, A.A. Velichko, 2007: Ecosystems, their properties, goods, and
services. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of
Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on
Climate Change, M.L. Parry, O.F. Canziani, J.P. Palutikof, P.J. van der Linden and C.E.
Hanson, Eds., Cambridge University Press, Cambridge, 211-272.
Hall, J.W., Sayers, P.B. and Dawson, R.J., 2005, National-scale Assessment of Current and
Future Flood Risk in England and Wales. Natural Hazards, 36, 147-164.
Höppe, P. and Pielke, R.A., Jr (Eds.), 2006, Workshop on Climate Change and Disaster
Losses - Understanding and Attributing Trends and Projections, Workshop Summary Report.
Hohenkammer, Germany.
P. Höppe, R. A. Pielke Jr., Eds., 2006: Climate Change and Disaster Losses:Understanding
and Attributing Trends and Projections, Report of a workshop,Hohenkammer, Germany, 25 to
26 May 2006 (University of Colorado, Boulder and Munich Re, Munich, 2006);
http://sciencepolicy.colorado.edu/sparc/research/projects/extreme_events/munich_workshop/.
Huntington, T.G., 2006, Evidence for intensification of the global water cycle: Review and
synthesis. Journal of Hydrology, 319, 83-95.
IPCC (2001b): Climate change 2001: Impacts, adaptation and vulnerability, IPCC WGII
report, Cambridge University Press, Cambridge, UK.
IPCC (2007b): Climate change 2007: Impacts, adaptation and vulnerability, IPCC WGII
report, Cambridge University Press, Cambridge, UK.
Kovats, R.S. and Jendritzky, G., (2006): Heat waves and human health, in: Climate Change
and Adaptation Strategies for Humans Health (eds. B. Menne and K.L. Ebi), Steinkopff,
Darmstadt, Germany.
Kundzewicz, Z.W., Graczyk, D., Maurer, T., Przymusińska, I., Radziejewski, M., Svensson,
C. and Szwed, M., 2004, Detection of change in world-wide hydrological time series of
maximum annual flow. Report 32, GRDC Report Series, Global Runoff Data Centre
(GRDC), Koblenz, Germany, p. 36.
231
Kundzewicz, Z.W., 2005, Intense precipitation and high river flows in Europe – Observations
and projections. Acta Geophysica Polonica, 53, 385-400.
Kundzewicz, Z.W., Ulbrich, U., Brücher, T., Graczyk, D., Krüger, A., Leckebusch, G.C.,
Menzel, L., Pińskwar, I., Radziejewski, M. and Szwed, M., 2005, Summer Floods in Central
Europe – Climate Change Track? Natural Hazards, 36, 165-189.
Lehner, B., Döll, P., Alcamo, J., Henrichs, T. and Kaspar, F., 2006, Estimating the Impact of
Global Change on Flood and Drought Risks in Europe: A Continental, Integrated Analysis.
Climatic Change, 75, 273-299.
Lise W. and Tol R.S.J 2002 Impact of climate on tourism demand. Climatic Change Vol 55
Nos 4 429-49.
Millennium Ecosystem Assessment, 2005. Ecosystems and Human Well-being: Synthesis.
Island Press, Washington, DC.
Mills, E., 2005, Insurance in a Climate of Change. Science, 309, 1040-1044.
Milly, P.C.D., Wetherald, R.T., Dunne, K.A. and Delworth, T.L., 2002, Increasing risk of
great floods in a changing climate. Nature, 415, 514-517.
Mudelsee, M., Borngen, M., Tetzlaff, G. and Grunewald, U., 2003, No upward trends in the
occurrence of extreme floods in central Europe. Nature, 425, 166-169.
Muir Wood, R., Miller, S. and Boissonnade, A., 2006, The search for trends in a global
catalogue of normalized weather-related catastrophe losses. In Workshop on Climate Change
and Disaster Losses - Understanding and Attributing Trends and Projections, P. Höppe and
R.A. Pielke, Jr (Eds.). Hohenkammer, Germany, 188-194.
Munich Re (2007a): Topics-annual Review of Natural Disasters 2006, Munich Reinsurance
Group, Munich, Germany.
Munich Re (2007b): Naturkatastrophen in Europa, unpublished data.
Nogués-Bravo, D., Araújo, M. B., Errea, M. P., & Martínez-Rica, J. P. (in press). Exposure of
global mountain systems to climate warming during the 21st Century Global Environmental
Change.
OECD (2007). Authors. Nicholls, R.J.(1), Hanson, S. (1), Herweijer, C.(2), Patmore, N. (2),
Hallegatte, S.(3), Corfee-Morlot, J.(4), Chateau, J.(4), and Muir-Wood, R. (2)Screening
Study: Ranking Port Cities With High Exposure And Vulnerability To Climate Extremes
Interim Analysis: Exposure Estimates. ENV/EPOC/GSP(2007)11
Pielke, R.A., Jr and Downton, M.W., 2000, Precipitation and Damaging Floods: Trends in the
United States, 1932–97. Journal of Climate, 13, 3625–3637.
Pielke Jr., R.A. (2007). Future economic damage from tropical cyclones: sensitivities to
societal and climate changes. Philosophical Transactions of the Royal Society, 365, 1-13.
Pielke Jr., R.A., J. Gratz, C.W. Landsea, D. Collins, M. Saunders and R. Musulin (2008).
Normalized hurricane damages in the United States: 1900-2005. Natural Hazards Review
9(1), 29-42.
Schär et al. (2006): ‘The role of increasing temperature variability in European summer
heatwaves’, Nature, 427, pp. 332-336.
Stern, N., (2006). The Economics of Climate Change. Cabinet Office – HM Treasury.
Cambridge University Press.
UNEP FI (2006): Adaptation and vulnerability to climate change: the role of the finance
sector, United Nations Environment Programme Finance Initiative, Geneva, Switzerland.
Also available at
http://www.unepfi.org/fileadmin/documents/CEO_briefing_adaptation_vulnerability_2006.pd
f
Viner D. 2006 Tourism and its interactions with Climate Change. Journal of Sustainable
Tourism Vol. 14. Nos. 4 pp 317-323
232
WL Delft Hydraulics (2007). Overstromingsrisico’s in Nederland in een veranderend klimaat
(Flood risks in the Netherlands under climate change), report Q4290.00, Delft, Netherlands;
in Dutch.
Watkiss, P. (2006) ‘The Social Costs of Carbon (SCC) Review – Methodological
Approaches for Using SCC Estimates in Policy Assessment’, Final Report to Defra.
Published January 2006.Paul Watkiss with contributions from David Anthoff, Tom Downing,
Cameron Hepburn, Chris Hope, Alistair Hunt, and Richard Tol.
http://www.defra.gov.uk/environment/climatechange/carboncost/aeat-scc.htm
Hydropower and cooling water for power plants
Aguiar, R., H. Goncalves, and M. Oliveira 2004. Energia. Principais Conclusões das Equipas
Sectoriais da Fase II, Sessão de apresentação pública dos resultados do SIAM II. Series
Energia. SIAM. http://www.siam.fc.ul.pt/siamII_pdf/Energia.pdf
Aguiar, R., H. Goncalves, M. Oliveira, and M. J. Reis 2002. Energy. In F. D. Santos, K.
Forbes and R. Moita [eds.], Climate Change in Portugal. Scenarios, Impacts and Adaptation
Measures - SIAM Project. Gradiva. http://www.siam.fc.ul.pt/SIAM_Book/9_Energy.pdf.
Andréasson, J., S. Bergström, M. Gardelin, and S.-S. Hellström 2006. Climate Change Effects
on Dam Safety – A Sensitivity Analysis of the Swedish Hydrological Dam Safety Guidelines.
European Conference on Impacts of Climate Change on Renewable Energy Sources. Series
Climate Change Effects on Dam Safety – A Sensitivity Analysis of the Swedish Hydrological
Dam Safety Guidelines.
http://verkefni.vatn.is/ce/files/Hydrological_models/johan_andreasson.pdf
Arnell, N., and M. Hulme 2000. Implications of Climate Change for Large Dams and their
Management, Thematic Review II.2 prepared as an input to the World Commisson on
Dams84 pp., Cape Town. http://www.wcainfonet.org/servlet/BinaryDownloaderServlet?filename=1066751858491_management.pdf
Bergström, S., T. Jóhannesson, G. Aðalgeirsdóttir, A. Ahlstrøm, L. M. Andreassen, J.
Andréasson, S. Beldring, H. Björnsson, B. Carlsson, P. Crochet, M. de Woul, B. Einarsson,
H. Elvehøy, G. E. Flowers, P. Graham, G. O. Gröndal, S. Guðmundsson, S.-S. Hellström, R.
Hock, P. Holmlund, J. F. Jónsdóttir, V. Radoc, N. Reeh, L. A. Roald, S. Rogozova, J.
Rosberg, Sigurðsson, Oddur, M. Suomalainen, T. Thorsteinsson, B. Vehviläinen, and N.
Veijalainen 2007. Impacts of climate change on river runoff, glaciers and hydropower in the
Nordic area. Joint final report from the CE Hydrological Models and Snow and Ice
Groups.CE Report CE-6. 40 pp. Norden - Nordic Energy Research, Reykjavik.
http://www.os.is/cefiles/hydro/final_report_high.pdf
BMU 2007. Time to Adapt Symposium: Climate Change and the European Water Dimension.
Discussion paper on Electricity. Series Time to Adapt Symposium: Climate Change and the
European Water Dimension. Discussion paper on Electricity. Bundesministerium für Umwelt,
Naturschutz und Reaktorsicherheit.
Christensen, J. H., T. R. Carter, M. Rummukainen, and G. Amantidis 2007. Evaluating the
performance and utility of regional climate models: the PRUDENCE project. Climate Change
81 supp 1 1-6.
Déqué, M., D. P. Rowell, D. Lüthi, J. H. Christensen, B. Rockel, D. Jacob, E. Kjellström, M.
d. Castro, and B. v. d. Hurk 2007. An intercomparison of regional climate simulations for
Europe: assessing uncertainties in model projections. Climate Change 81 supp 1 53-70.
EEA 2007. Climate change and water adaptation issues.EEA Technical report 2/2007. 114 pp.
European Environment Agency, Copenhagen.
http://reports.eea.europa.eu/technical_report_2007_2/en/eea_technical_report_2_2007.pdf
EU 2006. A European Strategy for Sustainable, Competitive and Secure Energy.Green paper
SEC (2006) 317. 20 pp. European Commission, Brussels. http://ec.europa.eu/energy/greenpaper-energy/doc/2006_03_08_gp_document_en.pdf
233
Eurelectric 2007. Latest Industry Statistics as at 31 December 2005. Series Latest Industry
Statistics as at 31 December 2005. Eurelectric.
http://www2.eurelectric.eu/Statistics/Latest2005.htm
Graham, L. P., S. Hagemann, S. Jaun, and M. Beniston 2007. On interpreting hydrological
change from regional climate models Climate Change 81 supp 1 97-122.
Hisdal, H., G. Lindström, E. Kuusisto, P. Jónsson, R. Barthelmie, J. Kriauciuniene, L.
Lizuma, and A. Reihan 2006. Climate and Energy 2003-2006, Statistical Analysis.CE flyer 6.
pp. Norden - Nordic Energy Research, Reykjavik. CE flyer 5.
http://www.os.is/cefiles/info/flyers/flyer_CE_Statistical.pdf
Horton, P., B. Schaefli, A. Mezghani, B. Hingray, and A. Musy 2006. Assessment of climatechange impacts on alpine discharge regimes with climate model uncertainty. Hydrological
Processes 20 2091-2109.
ICOLD 1992. Selection of Design Flood - Current Methods. International Commission on
Large Dams, Bull. 82.
IPCC 2007. Climate Change 2007 - 4th Assessment Report. WGII Impacts, Adaptation and
Vulnerabilitypp. International Panel on Climate Change, Geneva.
Kirkinen, J., A. Martikainen, H. Holttinen, I. Savolainen, O. Auvinen, and S. Syri 2005.
Impacts on the energy sector and adaptation of the electricity network business under a
changing climate in Finland.Mimeographs 340. 36 pp. Finnish Environment Institute,
Helsinki. http://www.environment.fi/download.asp?contentid=45340&lan=en
Lehner, B., G. Czisch, and S. Vassolo 2005: The impact of global change on the hydropower
potential of Europe: a model-based analysis. Energy Policy 33: 839–855
Lehner, B., G. Czisch, and S. Vassolo 2001. Europe’s hydropower potential today and in the
future. In B. Lehner, T. Henrichs, P. Döll and J. Alcamo [eds.], EuroWasser – Model-based
assessment of European water resources and hydrology in the face of global change. Kassel
World Water Series. Center for Environmental Systems Research, University of Kassel.
http://www.usf.uni-kassel.de/usf/archiv/dokumente/kwws/5/ew_8_hydropower.pdf
Lehner, B., G. Czisch, and S. Vassolo 2005. The impact of global change on the hydropower
potential of Europe: a model-based analysis. Energy Policy 33 839-855.
Mauser, W., and U. Strasser 2007. GLOWA – Danube, Integrative Techniken, Szenarien und
Strategien zur Zukunft des Wassers im Einzugsgebiet der Oberen Donau, Jahresbericht 2006
300 pp. Department für Geo- und Umweltwissenschaften, Ludwig-Maximilians-Universität
München, München. http://www.glowa-danube.de/frameset.htm
Mimikou, M. A., and E. A. Baltas 1997. Climate change impacts on the reliability of
hydroelectric energy production. Hydrol.Sci.J. 42 661-678.
Mo, B., G. Doorman, and B. Grinden 2006. Climate Change –Consequences for the
electricity system. Analysis of the Nord Pool system.Climate and Energy Report CE-5. 159
pp. Norden - Nordic Energy Research, Reykjavik.
http://www.os.is/cefiles/energy/Report_CE5.pdf
Nachtnebel, H.-P. 1999. CC-HYDRO. Impact of Climate Change on River Basin Hydrology
under Different Climatic Conditions, Final Reportpp., Vienna.
Nachtnebel, H.-P., and M. Fuchs 2001. HYD-KLIMA. Die Hydrologie Österreichs unter dem
Einfluß von Szenarien einer möglichen Klimaänderung. Endbereicht.pp. Institut für
Wasserwirtschaft, Hydrologie und konstruktiven Wasserbau, Universität fü Bodenkultur,
Wien.
Santos, F. D., and P. Miranda [eds.]. 2006. Alterações Climáticas em Portugal. Cenários,
Impactos e Medidas de Adaptação - Projecto SIAM II. Gravida.
Schaefli, B., B. Hingray, and A. Musy 2007. Climate change and hydropower production in
the Swiss Alps: quantification of potential impacts and realted modelling uncertainties. HESS
11 1191-1205.
234
Shablakova, M. V., W. P. A. van Deursen, and T. A. Buishand 2003. Assessing future
discharge of the river Rhine using regional climate model integrations and a hydrological
model. Climate Research 23 233-246.
Sælthun, N. R., P. Aittoniemi, S. Bergström, K. Einarsson, T. Jóhannesson, G. Lindström, P.
E. Ohlsson, T. Thomsen, B. Vehviläinen, and K. O. Aamodt 1998. Climate Change Impacts
on Runoff and Hydropower in the Nordic Countries. Final report from the project "Climate
Change and Energy Production".TemaNord 1998:552. 170 pp. Nordic Council of Ministers,
Copenhagen. http://www.norden.org/pub/sk/showpub.asp?pubnr=1998:552
UCTE 1999. Langfristige Entwicklung von Abflußreihen im UCPTE-Raum.Bericht der UG
Wasserwirtschaft 9 pp. Union for the Coordination of Transmission of Electricity, Wien.
Veiga da Cunha, L., R. Proença de Oliveira, L. Ribiero, and J. Nascimento 2004. Impactos
das Alterações Climáticas nos Recursos Hídricos Portugueses. Principais Conclusões das
Equipas Sectoriais da Fase II, Sessão de apresentação pública dos resultados do SIAM II.
Series Impactos das Alterações Climáticas nos Recursos Hídricos Portugueses. SIAM.
http://www.siam.fc.ul.pt/siamII_pdf/RecursosHidricos.pdf
Veijalainen, N., and B. Vehviläinen 2006. Climate change effects on dam safety in Finland.
European Conference on Impacts of Climate Change on Renewable Energy Sources. Series
Climate change effects on dam safety in Finland.
http://verkefni.vatn.is/ce/files/Hydrological_models/noora_veijalainen.pdf
Zierl, B. and H. Bugman, 2005. Global change impacts on hydrological processes in Alpine
catchments. Wat.Resourc.Res. 41(2)
Public water supply
Dyble J., Paerl H.W. and Neilan B.A. 2002. Genetic characterization of Cylindrospermopsis
raciborskii (Cyanobacteria) isolates from diverse geographic origins based on nifH
andcpcBA-IGS nucleotide sequence analysis. Appl. Environ. Microbiol. 68: 2567–2571.
Mooij WM, Hu¨lsmann S, Domis LND, Nolet BA, Bodelier PLE, Boers PCM, Pires LMD,
Gons HJ, Ibelings BW, Noordhuis R, Portielje R, Wolfstein K, Lammens EHRR (2005) The
impact of climate change on lakes in The Netherlands: a review. Aquat Ecol 39(4):381–400
Jöhnk, K.D., Huisman, J., Sharples, J., Sommeijer, B., Visser, P.M. & Stroom, J.M. Summer
heatwaves promote blooms of harmful cyanobacteria. Global Change Biology (in press)
Schindler, D.W. 2001. The cumulative effects of climate warming and other human stresses
on Canadian freshwaters in the new millennium. Can J. Fish. Aquat. Sci., 58: 18-29.
Wilhelm, S., and Adrian, R. 2007. Impact of summer warming on the thermal characteristics
of a polymictic lake and consequences for oxygen, nutrients and phytoplankton. Freshwater
Biology. Doi: 10.1111/j.1365-2427.2007.01887.x
Losses
ABI, 2005, Financial risk of climate change. Association of British Insurers, London, UK, p.
39.
Barredo, J.I., 2007, Major flood disasters in Europe: 1950-2005. Natural Hazards, 42, 125148.
Feyen, L., Dankers, R., Barredo, J.I., de Roo, A. and Lavalle, C., 2007, Climate change and
floods in Europe. In Climate Change Impacts on the Water Cycle, Resources and Quality Research-Policy Interface, M. Moren-Abat, P. Quevauviller, L. Feyen, A.-S. Heiskanen, P.
Noges, A.L. Solheim and E. Lipiatou (Eds.). Luxembourg: Office for Official Publications of
the European Communities, 37-43.
Hall, J.W., Sayers, P.B. and Dawson, R.J., 2005, National-scale Assessment of Current and
Future Flood Risk in England and Wales. Natural Hazards, 36, 147-164.
235
Höppe, P. and Pielke Jr, R.A., 2006, Workshop Summary Report. In Workshop on Climate
Change and Disaster Losses: Understanding and Attributing Trends and Projections, P. Höppe
and R.A. Pielke, Jr (Eds.). Hohenkammer, Germany, 4-12.
Huntington, T.G., 2006, Evidence for intensification of the global water cycle: Review and
synthesis. Journal of Hydrology, 319, 83-95.
IPCC, 2007, Climate change 2007: Impacts, adaptation and vulnerability, IPCC WGII report
(Cambridge, UK: Cambridge University Press).
Kundzewicz, Z.W., Graczyk, D., Maurer, T., Przymusińska, I., Radziejewski, M., Svensson,
C. and Szwed, M., 2004, Detection of change in world-wide hydrological time series of
maximum annual flow. Report 32, GRDC Report Series, Global Runoff Data Centre
(GRDC), Koblenz, Germany, p. 36.
Kundzewicz, Z.W., 2005, Intense precipitation and high river flows in Europe – Observations
and projections. Acta Geophysica Polonica, 53, 385-400.
Kundzewicz, Z.W., Ulbrich, U., Brücher, T., Graczyk, D., Krüger, A., Leckebusch, G.C.,
Menzel, L., Pińskwar, I., Radziejewski, M. and Szwed, M., 2005, Summer Floods in Central
Europe – Climate Change Track? Natural Hazards, 36, 165-189.
Kundzewicz, Z.W., Radziejewski, M. and Pinskwar, I., 2006, Precipitation extremes in the
changing climate of Europe. Climate Research, 31, 51-58.
Lehner, B., Döll, P., Alcamo, J., Henrichs, T. and Kaspar, F., 2006, Estimating the Impact of
Global Change on Flood and Drought Risks in Europe: A Continental, Integrated Analysis.
Climatic Change, 75, 273-299.
Mills, E., 2005, Insurance in a Climate of Change. Science, 309, 1040-1044.
Milly, P.C.D., Wetherald, R.T., Dunne, K.A. and Delworth, T.L., 2002, Increasing risk of
great floods in a changing climate. Nature, 415, 514-517.
Mudelsee, M., Borngen, M., Tetzlaff, G. and Grunewald, U., 2003, No upward trends in the
occurrence of extreme floods in central Europe. Nature, 425, 166-169.
Muir Wood, R., Miller, S. and Boissonnade, A., 2006, The search for trends in a global
catalogue of normalized weather-related catastrophe losses. In Workshop on Climate Change
and Disaster Losses - Understanding and Attributing Trends and Projections, P. Höppe and
R.A. Pielke, Jr (Eds.). Hohenkammer, Germany, 188-194.
Pielke Jr, R.A. and Downton, M.W., 2000, Precipitation and Damaging Floods: Trends in the
United States, 1932–97. Journal of Climate, 13, 3625–3637.
Pielke Jr, R.A., 2007, Future economic damage from tropical cyclones: sensitivities to
societal and climate changes. Philosophical Transactions of the Royal Society, 365, 1-13.
WL Delft Hydraulics, 2007, Overstromingsrisico’s in Nederland in een veranderend klimaat
(Flood risks in the Netherlands under climate change). report Q4290.00, WL Delft Hydraulics
Delft, Netherlands, p.
7. Adaptation
EC, 2007. Adapting to climate change in Europe – options for EU action. SEC(2007) 849,
European Commission, Brussels
EEA, 2005. Vulnerability and adaptation to climate change in Europe Technical report No
7/2005
EEA, 2007. Climate change and water adaptation issues. Technical report No 2/2007
EEA, 2008. Climate change: the cost of inaction and the cost of adaptation. EEA Technical
report No 13/2007
IPCC, 2007a. Climate Change 2007: Impacts, adaptation and vulnerability. Cambridge
University Press
IPCC, 2007b. Climate Change 2007: Synthesis Report. Cambridge University Press
236
Massey, E., 2007. Presentation during EIONET workshop on impacts, vulnerability and
adaptation; European Environment Agency, Copenhagen
UNFCCC, 2006. Five-year programme of work on impacts, vulnerability and adaptation to
climate change, in: Report of the Subsidiary Body for Scientific and Technological Advice on
its twenty-fifth session held at Nairobi from 6 to 14 November 2006
UNFCCC, 2007. Decision -/CP.13, Bali Action Plan
8. Uncertainties, data availability, gaps and future needs
EC (2004): Communication from the Commission to the European Parliament and the
Council: Global Monitoring for Environment and Security (GMES): Establishing a GMES
capacity by 2008 - (Action Plan (2004-2008)) COM(2004)65 final
http://www.gmes.info/library/files/
EC (2005): European Research on Climate Change Catalogue of FP6 Projects Volume 1,
Environment Directorate,
http://ec.europa.eu/research/environment/pdf/european_research_climate_change_en.pdf
EC (2006): Global Change and Ecosystems Catalogue of projects funded during the Sixth
Framework Programme, Environment Directorate,
http://ec.europa.eu/research/environment/pdf/global_change_ecosystem.pdf
EC (2006): European Research on Climate Change Catalogue of FP5 and FP6 Projects on
Carbon Cycle and Green House Gases, Environment Directorate,
http://ec.europa.eu/research/environment/pdf/catalogue_fp5fp6_carbon.pdf
EC (2006): International Symposium Climate Change Research Challenges, Environment
Directorate, http://ec.europa.eu/research/environment/pdf/climate_research_chall.pdf
EC (2007): European research on Polar Environment and Climate Results and information
from FP5 and FP6 projects, Environment Directorate,
http://ec.europa.eu/research/environment/pdf/polar_catalogue_final.pdf
EC (2007): Polar Environment and Climate The challenges, International Symposium,
European research in the context of the International Polar Year, Environment Directorate,
http://ec.europa.eu/research/environment/pdf/polar_env_and_climate_proceedings.pdf
GCOS (2003): The second report on the adequacy of global observing systems for climate in
support of the UNFCCC, report no. GCOS-82.
GEOSS (2005): The Global Earth Observation System of Systems (GEOSS) 10-Year
Implementation Plan http://www.earthobservations.org/docs/
IPCC (2001): Climate change 2001: The scientific basis, Cambridge University Press,
Cambridge, UK.
IPCC (2007): Climate Change 2007: The Physical Science Basis, Summary for Policymakers;
http://www.ipcc.ch/
237