FINAL DRAFT                       Chapter 13          IPCC WGII Sixth Assessment Report

1

2                                     Chapter 13: Europe

3

4 Coordinating Lead Authors: Birgit Bednar-Friedl (Austria), Robbert Biesbroek (The Netherlands), Daniela
5 N. Schmidt (United Kingdom/Germany)

 6

 7 Lead Authors: Peter Alexander (United Kingdom), Knut Yngve B鴕sheim (Norway), Jofre Carnicer
 8 (Spain), Elena Georgopoulou (Greece), Marjolijn Haasnoot (Netherlands), Gon閞i Le Cozannet (France),
 9 Piero Lionello (Italy), Oksana Lipka (Russian Federation), Christian M鰈lmann (Germany), Veruska
10 Muccione (Switzerland/Italy), Tero Mustonen (Finland), Dieter Piepenburg (Germany), Lorraine Whitmarsh
11 (United Kingdom)

12

13 Contributing Authors: Magnus Benzie (Sweden), Pam Berry (United Kingdom), Sara Burbi (United
14 Kingdom), Erika Coppola (Italy), Mladen Domazet (Croatia), Frank Ewert (Germany), Federica Gasbarro
15 (Italy), Matthias Gauly (Italy), Fran鏾is Gemenne (Belgium), Peter Greve (Austria/Germany), Ana Iglesias
16 (Spain), Elizabeth Kendon (United Kingdom), Heidi Kreibich (Germany), Nikos Koutsias (Greece), Anna
17 Laine-Pet鋔鋕angas (Finland), Dimitris Lalas (Greece), Cristina Linares Gil (Spain), Danijela Markovic
18 (Germany), Sadie McEvoy (Netherlands/Ireland), Ana Mijic (United Kingdom), Raya Muttarak
19 (Austria/Thailand), Rita Nogherotto (Italy), Hans Orru (Estonia), Mark Parrington (United Kingdom), Jeff
20 Price (United Kingdom), Kaisa Raitio (Sweden), Marta Guadalupe Rivera Ferre (Spain), Jan C. Semenza
21 (Switzerland), Rub閚 Valbuena (United Kingdom), Michelle van Vliet (The Netherlands), Heidi Webber
22 (Germany), Laura Wendling (Finland), Katherine Yates (United Kingdom), Monika Zurek (United
23 Kingdom).

24

25 Chapter Scientists: Sadie McEvoy (The Netherlands/Ireland), Phoebe O'Brien (United Kingdom/Sweden)

26

27 Review Editors: Georg Kaser (Austria/Italy), Jose Manuel Moreno (Spain)

28

29 Date of Draft: 1 October 2021

30

31 Notes: TSU Compiled Version

32

33

34 Table of Contents

35

36 Executive Summary..........................................................................................................................................3

37 13.1 Point of Departure ....................................................................................................................................7

38  13.1.1 Introduction and Geographical Scope............................................................................................7

39  13.1.2 Socio-Economic Boundary Conditions ...........................................................................................8

40  13.1.3 Impact Assessment of Climate Change based on Previous Reports ...............................................9

41  13.1.4 European Climate: Main Conclusions of WGI AR6 .......................................................................9

42 13.2 Water .......................................................................................................................................................12

43  13.2.1 Observed Impacts and Projected Risks.........................................................................................12

44 Box 13.1: Venice and its Lagoon ...................................................................................................................12

45  13.2.2 Solution Space and Adaptation Options .......................................................................................16

46  13.2.3 Knowledge Gaps ...........................................................................................................................20

47 13.3 Terrestrial and Freshwater Ecosystems and their Services ...............................................................20

48  13.3.1 Observed Impacts and Projected Risks.........................................................................................20

49  13.3.2 Solution Space and Adaptation Options .......................................................................................26

50  13.3.3 Knowledge gaps ............................................................................................................................27

51 13.4 Ocean and Coastal Ecosystems and their Services..............................................................................28

52  13.4.1 Observed Impacts and Projected Risks.........................................................................................28

53  13.4.2 Solution Space and Adaptation Options .......................................................................................31

54  13.4.3 Knowledge Gaps ...........................................................................................................................33

55 13.5 Food, Fibre, and Other Ecosystem Products .......................................................................................33

56  13.5.1 Observed Impacts and Projected Risks.........................................................................................33

57  13.5.2 Solution Space and Adaptation Options .......................................................................................37

    Do Not Cite, Quote or Distribute  13-1                                  Total pages: 143
    FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

1   13.5.3 Knowledge Gaps ...........................................................................................................................41

2 13.6 Cities, Settlements and Key Infrastructure..........................................................................................41

3   13.6.1 Observed Impacts and Projected Risks.........................................................................................41

4   13.6.2 Solution Space and Adaptation Options .......................................................................................47

5   13.6.3 Knowledge Gaps ...........................................................................................................................52

6 13.7 Health, Wellbeing and the Changing Structure of Communities ......................................................53

7   13.7.1 Observed Impacts and Projected Risks.........................................................................................53

8   13.7.2 Solution Space and Adaptation Options ......................................................................................57

9   13.7.3 Knowledge Gaps ...........................................................................................................................59

10 13.8 Vulnerable Livelihoods and Social Inequality .....................................................................................59

11  13.8.1 Observed Impacts and Projected Risks.........................................................................................60

12 Box 13.2: S醡i Reindeer Herding in Sweden ..............................................................................................63

13  13.8.2 Solution Space and Adaptation Options .......................................................................................65

14  13.8.3 Knowledge Gaps ...........................................................................................................................66

15 13.9 Interregional Impacts, Risks and Adaptation......................................................................................66

16  13.9.1 Consequences of Climate-change Driven Impacts, Risks and Adaptation Emerging in Other

17               Parts of the World for Europe ..............................................................................................66

18  13.9.2 Interregional Consequences of Climate Risks and Adaptation Emerging from Europe ..............67

19  13.9.3 European Territories Outside Europe ..........................................................................................68

20  13.9.4 Solution Space and Adaptation Options........................................................................................68

21 13.10         Detection and Attribution, Key Risks and Adaptation Pathways..................................69

22  13.10.1 Detection and Attribution of Impacts....................................................................................69

23  13.10.2 Key Risks Assessment for Europe .........................................................................................70

24  13.10.3 Consequences of Multiple Climate Risks for Europe ...........................................................77

25  13.10.4 Knowledge Gaps ...................................................................................................................78

26 13.11         Societal Adaptation to Climate Change across Regions, Sectors and Scales ................79

27  13.11.1 Policy Responses, Options and Pathways ............................................................................79

28  13.11.2 Societal Responses, Options and Pathways..........................................................................82

29  13.11.3 Adaptation, Transformation and Sustainable Development Goals ......................................84

30 Box 13.3: Climate Resilient Development Pathways in European Cities..................................................85

31 FAQ 13.1: How can climate change affect social inequality in Europe?...................................................87

32 FAQ 13.2: What are the limits of adaptation for ecosystems in Europe? .................................................88

33 FAQ 13.3: How can people adapt at individual and community level to heat waves in Europe? ..........89

34 FAQ 13.4: What opportunities does climate change generate for human and natural systems in

35  Europe? ...................................................................................................................................................90

36 References........................................................................................................................................................92

37

38

    Do Not Cite, Quote or Distribute  13-2        Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 Executive Summary

 2

 3 Where are we now?

 4

 5 Our current 1.1癈 warmer world is already affecting natural and human systems in Europe (very high
 6 confidence1). Since AR5, there has been a substantial increase in detected or attributed impacts of climate
 7 change in Europe, including extreme events (high confidence). Compound hazards of warming and
 8 precipitation have become more frequent (medium confidence). Climate change has resulted in losses of and
 9 damages to people, ecosystems, food systems, infrastructure, energy and water availability, public health,
10 and the economy (very high confidence) {13.1.4;13.2.1;13.3.1;13.4.1;13.5.1;13.6.1;13.7.1;13.8.1;13.10.1}.

11

12 As impacts vary both across and within European regions, sectors, and societal groups (high
13 confidence), inequalities have deepened (medium confidence). Southern regions tend to be more
14 negatively affected, while some benefits have been observed, alongside negative impacts in northern and
15 central regions. Traditional lifestyles, for example in the European Arctic, are threatened already (high
16 confidence). Poor households have lower capacity to adapt to, and recover from, impacts (medium
17 confidence) {13.5.1;13.6.1;13.7.1;13.8.1.;13.8.2;13.10.1;Box 13.2}.

18

19 The range of options available to deal with climate-change impacts has increased in most of Europe
20 since AR5 (high confidence). Growing public perception and adaptation knowledge in public and private
21 sectors, increasing number of policy and legal frameworks, and dedicated spending on adaptation are all
22 clear indications that the availability of options has expanded (high confidence). Information provision,
23 technical measures, and government policies are the most common adaptation actions implemented. Nature-
24 based solutions (NbS) that restore or recreate ecosystems, build resilience and produce synergies with
25 adaptation and mitigation are increasingly used. Many cities are taking adaptation action, but with large
26 differences in level of ambition and implementation (high confidence)
27 {13.2.2;13.3.2;13.4.2;13.5.2;13.6.2;13.7.2;13.8.2;13.10.2;13.11.1;13.11.2;13.11.3}.

28

29 Observed adaptation actions are largely incremental with only a few examples of local transformative
30 action; adaptation actions have demonstrated different degrees of effectiveness in reducing impacts
31 and feasibility of implementation (high confidence). For example, adaptation actions such as flood
32 defences and early warning systems have reduced flood damages and heat-related mortality in parts of
33 Europe. Despite progress on adaptation, impacts are observed. Adaptation actions in the private sector are
34 limited, with many businesses and regions remaining under-prepared. A gap remains between planning and
35 implementation of adaptation action (high confidence) {13.2.2;13.5.2;13.6.2;13.7.2;13.11}.

36

37 What are the future risks?

38

39 Warming in Europe will continue to rise faster than the global mean, widening risk disparities across
40 Europe in the 21st century (high confidence). Largely negative impacts are projected for southern regions
41 (e.g., increased cooling needs and water demand, losses in agricultural production, and water scarcity) and
42 some short-term benefits anticipated in the north (e.g. increased crop yields and forest growth)
43 {13.1.4;13.2.1;13.3.1;13.4.1;13.5.1;13.6;13.7.1;13.10.2}.

44

45 Four key risks (KR) have been identified for Europe, with most becoming more severe at 2癈 GWL
46 compared to 1.5癈 GWL in scenarios with low to medium adaptation (high confidence). From 3癈
47 GWL and even with high adaptation, severe risks remain for many sectors in Europe (high confidence). Key
48 risks are: mortality and morbidity of people and ecosystems disruptions due to heat (KR1: heat); loss in
49 agricultural production due to combined heat and droughts (KR2: agriculture); water scarcity across sectors
50 (KR3: water scarcity); impacts of floods on people, economies and infrastructure (KR4: flooding) {13.10.2}.

1 In this Report, the following summary terms are used to describe the available evidence: limited, medium, or robust;
and for the degree of agreement: low, medium, or high. A level of confidence is expressed using five qualifiers: very
low, low, medium, high, and very high, and typeset in italics, e.g., medium confidence. For a given evidence and
agreement statement, different confidence levels can be assigned, but increasing levels of evidence and degrees of
agreement are correlated with increasing confidence.

Do Not Cite, Quote or Distribute  13-3        Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 KR1: The number of deaths and people at risk of heat stress will increase two- to threefold at 3癈
 3 compared to 1.5癈 GWL (high confidence). Risk consequences will become severe more rapidly in
 4 Southern and Western Central Europe and urban areas (high confidence). Thermal comfort hours during
 5 summer will decrease significantly (high confidence), by as much as 74% in Southern Europe at 3癈 GWL.
 6 Above 3癈 GWL, there are limits to the adaptation potential of people and existing health systems,
 7 particularly in Southern Europe and Eastern Europe and areas where health systems are under pressure (high
 8 confidence) {13.6.1;13.6.2;13.7.1;13.7.2;13.8.1;13.10.2.1}.

 9

10 KR1: Warming will decrease suitable habitat space for current terrestrial and marine ecosystems and
11 irreversibly change their composition, increasing in severity above 2癈 GWL (very high confidence).
12 Fire-prone areas are projected to expand across Europe, threatening biodiversity and carbon sinks (medium
13 confidence). Adaptation actions, e.g. habitat restoration and protection, fire and forest management, and
14 agroecology, can increase the resilience of ecosystems and their services. Trade-offs between adaptation and
15 mitigation options (e.g., coastal infrastructure and Nature based Solutions) will result in risks for the integrity
16 and function of ecosystems (medium confidence) {13.3.1;13.3.2;13.4.1;13.4.2;13.10.2.1; Cross-Chapter Box
17 SLR in Chapter 3; Cross-Chapter Box NATURAL in Chapter 2}.

18

19 KR2: Due to a combination of heat and drought, substantive agricultural production losses are
20 projected for most European areas over the 21st century, which will not be offset by gains in Northern
21 Europe (high confidence). Yield losses for maize will reach 50% in response to 3癈 GWL, especially in
22 Southern Europe. Yields of some crops, e.g. wheat, may increase in Northern Europe when warming does
23 not exceed 2癈 (medium confidence). While irrigation is an effective adaptation option for agriculture, the
24 ability to adapt using irrigation will be increasingly limited by water availability, especially in response to
25 GWL above 3癈 (high confidence) {13.5.1;13.5.2;13.10.2.2}.

26

27 KR3: Risk of water scarcity will become high at 1.5癈 and very high at 3癈 GWL in Southern Europe
28 (high confidence) and increase from moderate to high in Western Central Europe (medium
29 confidence). In Southern Europe, more than a third of the population will be exposed to water scarcity at 2癈
30 GWL; under 3癈 GWL, this risk will double, and significant economic losses in water and energy dependent
31 sectors may arise (medium confidence). The risk of water scarcity is strongly increasing for Western Central
32 and Southern Europe, and many cities under 3癈 GWL. Adaptation becomes increasingly difficult at 3癈
33 GWL and above, due to geophysical and technological limits; hard limits are likely2 first reached in parts of
34 Southern Europe {13.2.1;13.2.2;13.6.1;13.10.2.3}.

35

36 KR4: Due to warming, changes in precipitation and sea level rise, risks to people and infrastructures
37 from coastal, riverine, and pluvial flooding will increase in Europe (high confidence). Risks of
38 inundation and extreme flooding will increase with accelerating pace of sea level rise along Europe's coasts
39 (high confidence). Above 3癈 GWL, damage costs and people affected by precipitation and river flooding
40 may double. Coastal flood damage is projected to increase at least 10-fold by the end of the 21st century, and
41 even more or earlier with current adaptation and mitigation (high confidence). Sea level rise represents an
42 existential threat for coastal communities and their cultural heritage, particularly beyond 2100
43 {13.2.1;13.2.2;13.6.2;13.10.2.4;Box 13.1; Cross-Chapter Box SLR in Chapter 3).

44

45 European cities are hotspots for multiple risks of increasing temperatures and extreme heat, floods,
46 and droughts (high confidence). Warming beyond 2癈 GWL is projected to result in widespread impacts
47 on infrastructure and businesses (high confidence). These include increased risks for energy supply (high
48 confidence) and transport infrastructure (medium confidence), increases in air conditioning needs (very high
49 confidence), and high water demand (high confidence) {13.2.2;13.6.1;13.7.1;13.10.2}.

2 In this Report, the following terms have been used to indicate the assessed likelihood of an outcome or a result:
Virtually certain 99�100% probability, Very likely 90�100%, Likely 66�100%, About as likely as not 33�66%,
Unlikely 0�33%, Very unlikely 0�10%, and Exceptionally unlikely 0�1%. Additional terms (Extremely likely: 95�
100%, More likely than not >50�100%, and Extremely unlikely 0�5%) may also be used when appropriate. Assessed
likelihood is typeset in italics, e.g., very likely). This Report also uses the term `likely range' to indicate that the assessed
likelihood of an outcome lies within the 17-83% probability range.

Do Not Cite, Quote or Distribute  13-4        Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 European regions are affected by multiple key risks, with more severe consequences in the south than
 3 in the north (high confidence). These risks may co-occur and amplify each other, but there is uncertainty
 4 about their interactions and their quantifications. There is high confidence that consequences for socio-
 5 economic and natural systems will be substantial; the number of people exposed to KRs and economic losses
 6 are projected to at least double at 3癈 GWL compared to 1.5癈 GWL (medium confidence); and increased
 7 risks are also projected for biodiversity and ecosystem services, such as carbon regulation. The risks
 8 resulting from changes in climatic and non-climatic drivers in many sectors is a key gap in knowledge (high
 9 confidence). This gap prevents the precise assessment of systemic risks, socio-ecological tipping points and
10 limits to adaptation {13.10.2;13.10.3;13.10.4}.

11

12 Climate risks from outside Europe are emerging due to a combination of the position of European
13 countries in the global supply chain and shared resources (high confidence). There is emerging evidence
14 that climate risks in Europe may also impact financial markets, food production and marine resources
15 beyond Europe. Exposure of European countries to interregional risks can be reduced by international
16 governance and collaboration on adaptation in other regions (medium confidence){13.5.2;13.9.1;
17 13.9.2;13.11; Cross-Chapter Box INTEREG in Chapter16}.

18

19 What are the solutions, limits and opportunities of adaptation?

20

21 There is a growing range of adaptation options available today to deal with future climate risks (high
22 confidence). Examples for adaptation to the key risks include: behavioural change combined with building
23 interventions, space cooling and urban planning to manage heat risks (KR1); restoration, expansion and
24 connection of protected areas for ecosystems, while generating adaptation and mitigation benefits for people
25 (KR1: heat); irrigation, vegetation cover, changes in farming practices, crop and animal species, and shifting
26 planting (KR2: agriculture); efficiency improvements, water storage, water reuse, early warning systems, and
27 land use change (KR3: water scarcity); early warning systems, reserving space for water and ecosystem
28 based adaptation, sediment or engineering based options, land use change and managed retreat (KR4:
29 flooding). Nature-based solutions for flood protection and heat alleviation are themselves under threat from
30 warming, extreme heat, drought and sea level rise (high confidence)
31 {13.2.2;13.3.2;13.4.2;13.5.2;13.6.2;13.7.2;13.8.2;13.9.4;13.10.2;13.11}.

32

33 In many parts of Europe, existing and planned adaptation measures are not sufficient to avoid the
34 residual risk, especially beyond 1.5癈 GWL (high confidence). Residual risk can result in losses of
35 habitat and ecosystem services, heat related deaths (KR1), crop failures (KR2), water rationing during
36 droughts in SEU (KR3), and loss of land (KR4) (medium confidence). At 3癈 GWL and beyond, a
37 combination of many, maybe even all, adaptation options are needed, including transformational changes, to
38 reduce residual risk (medium confidence).
39 {13.2.2;13.3.2;13.4.2;13.5.2;13.6.2;13.7.2;13.8.2;13.9.4;13.10.2;13.11}.

40

41 Although adaptation is happening across Europe, it is not implemented at the scale, depth and speed
42 needed to avoid the risks (high confidence). Many sectors and systems, such as flood risk management,
43 critical infrastructure and reforestation, are on self-reinforcing development paths that can result in lock-ins
44 and prevent changes needed to reduce risks in the long term and achieve adaptation targets. Forward-looking
45 and adaptive planning can avoid path-dependencies, maladaptation, and ensure timely action (high
46 confidence). Monitoring climate change, socio-economic developments and progress on implementation is
47 critical to assess if and when further actions are needed, and evaluating whether adaptation is successful
48 {13.2.2;13.10.2;13.11.1;13.11.2;13.11.3;Cross-Chapter Box DEEP in Chapter 17}.

49

50 Systemic barriers constrain the implementation of adaptation options in vulnerable sectors, regions
51 and societal groups (high confidence). Key barriers are limited resources, lack of private sector and citizens
52 engagement, insufficient mobilisation of finance, lack of political leadership, and low sense of urgency. Most
53 of the adaptation options to the key risks depend on limited water and land resources, creating competition
54 and trade-offs, also with mitigation options and socio-economic developments (high confidence). Europe
55 will face difficult decisions balancing these trade-offs. Novel adaptation options are pilot tested across
56 Europe, but upscaling remains challenging. Prioritisation of options and transitions from incremental to

Do Not Cite, Quote or Distribute  13-5        Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 transformational adaptation are limited due to vested interests, economic lock-ins, institutional path-
 2 dependencies, and prevalent practices, cultures, norms, and belief systems {13.11.1;13.11.2;13.11.3}.

 3

 4 Several windows of opportunity emerge to accelerate climate resilient development (CRD) (medium
 5 confidence). Such windows are either institutionalised (e.g. budget cycles, policy reforms and evaluations,
 6 infrastructure investment cycles), or open unexpectedly (e.g. extreme events, COVID-19 recovery
 7 programs). These windows can be used to accelerate action through mainstreaming and transformational
 8 actions (medium confidence). CRD is visible in European cities, particularly in green infrastructure, energy-
 9 efficient buildings and construction, and where co-benefits (e.g. to health, biodiversity) have been identified.
10 Private sector adaptation takes place mostly in response to extreme events or regulatory, shareholder, or
11 consumer pressures and incentives (medium confidence) {13.11.3; Box13.3;Cross-Chapter Box COVID in
12 Chapter 7}.

13

14 Closing the adaptation gap requires moving beyond short-term planning and ensuring timely and
15 adequate implementation (high confidence). Inclusive, equitable and just adaptation pathways are critical
16 for climate resilient development. Such pathways require consideration of SDGs, gender, and Indigenous
17 knowledge and local knowledge and practices. The success of adaptation will depend on our understanding
18 of which adaptation options are feasible and effective in their local context (high confidence). Long lead
19 times for nature-based and infrastructure solutions or planned relocation require implementation in the
20 coming decade to reduce risks in time. To close the adaptation gap, political commitment, persistence and
21 consistent action across scales of government, and upfront mobilization of human and financial capital is key
22 (high confidence), even when the benefits are not immediately visible {13.2.2;13.8;13.11;Cross-Chapter Box
23 GENDER in Chapter 18}.

24

Do Not Cite, Quote or Distribute  13-6        Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 13.1 Point of Departure

 2

 3 13.1.1 Introduction and Geographical Scope

 4

 5 This regional chapter on climate-change impacts, vulnerabilities and adaptations in Europe examines the
 6 impacts on the sectors, regions and vulnerable populations of Europe, assesses the causes of vulnerability,
 7 and analyses ways to adapt, thereby considering socio-economic developments, land use change, and other
 8 non-climatic drivers. Compared to AR5 and in the context of the Paris Agreement (2015), we have placed
 9 emphasis on the planned and implemented solutions, assessed their feasibility and effectiveness, considered
10 the Sustainable Development Goals (SDG) and shared socioeconomic pathways (SSPs). Global warming
11 level (GWL) refers to global climate change emissions relative to preindustrial levels, expressed as global
12 surface air temperature (Chen et al., 2021, 1.6.2).

13

14 The chapter generally follows the overall structure of AR6 WGII. We first present our point of departure
15 (13.1) followed by the key sectors, starting with water, as water is interconnected and of fundamental
16 importance to subsequent sections (13.2-13.8). For each section, we assess the observed impacts and
17 projected risks, solution space and adaptation options, and knowledge gaps. The solution space is defined as
18 the space within which opportunities and constraints determine why, how, when, and who adapts to climate
19 risks (Haasnoot et al., 2020a). Section 13.9 discusses impacts and adaptation beyond Europe, followed by the
20 key risks for Europe (13.10). The chapter ends with an assessment of adaptation solution space, climate
21 resilient development pathways, and SDGs (13.11), although recognizing that scientific literature on these
22 aspects is only slowly beginning to emerge.

23

24 With the rapidly growing body of scientific literature since WGII AR5 (Callaghan et al., 2020) our
25 assessment prioritized systematic reviews, meta-analyses, and synthesis papers and reports. Feasibility and
26 effectiveness assessments used revised methods developed for the Special Report of Global warming of 1.5�
27 (de Coninck et al., 2018; Singh et al., 2020). Protocols, as well as supporting material for figures and tables,
28 can be found in the Supplementary Material.

29

30

31

32 Figure 13.1: Geographical subdivision of land (a, b, c, d) and ocean (i, ii, iii) regions of Europe. The overlay represents

33 the WGI AR6 subdivisions for climate-change projections of land, while the colour coding indicates the European

Do Not Cite, Quote or Distribute  13-7        Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 countries (or, in case of the Russian Federation, the European part of the country (EEU) used for this chapter. Note that
 2 for WGI AR6 region MED includes both southern Europe and northern Africa while this chapter only includes the
 3 northern (European) part of the MED region.

 4

 5

 6 The geographical scope and subdivision of European land, coastal and ocean regions is largely the same as in
 7 WGII AR5 Chapter 23 (Kovats et al., 2014): Southern Europe (SEU), Western Central Europe (WCE),
 8 Eastern Europe (EEU) and Northern Europe (NEU). Note that WGI assesses a larger region for the
 9 Mediterranean (MED) which includes north Africa and the Middle East compared to the assessment in this
10 chapter (SEU). The European part of the Arctic region is not systematically assessed here as it is extensively
11 captured in Cross-Chapter Paper (CCP) 6. Information relevant to Europe is also synthesised in the CCPs,
12 including European biodiversity hotspots (CCP 1), coastal cities and settlements (CCP 2), Mediterranean
13 region (CCP 4) and mountains (CCP 5). European seas are broadly divided by latitude into (i) European
14 Arctic waters (NEUS), (ii) European Temperate Seas (TEUS) and (iii) Southern Seas with the Mediterranean
15 and Black Sea (SEUS) (Figure 13.1).

16

17 13.1.2 Socio-Economic Boundary Conditions

18

19 The adaptive capacity, as measured by GDP per capita, tends to be higher in northern and western parts of
20 Europe (Figure 13.2a). In the last decades, climate change has led to substantial losses and damages to
21 people and assets across Europe, mostly from riverine flooding, heatwaves, and storms (Figure 13.2b).
22 Public concern about climate change, which is an indicator of the intention to mitigate and adapt, is
23 particularly high in parts of Southern and Western Central Europe (Figure 13.2c). Current vulnerability to
24 extreme weather and climatic events in European countries is low to moderate compared to the rest of the
25 world (Figure 13.2d).

26

27

28

29 Figure 13.2: Indicators of reported damages to people and assets, vulnerability and adaptive capacity across European
30 countries: (a) GDP per capita (average 2013-2018), in constant 2011 international dollars (WorldBank, 2020); (b)
31 Exposure as measured by the global rank of the Climate Risk Index, which is based on economic damages and fatalities
32 due to climate-related extreme weather events between 1999 and 2018 (Germanwatch, 2020); (c) Level of climate
33 change concern among a representative weighted sample of residents 15 years and older in private households
34 (European Social Survey, 2020) and (d) Vulnerability to disasters and humanitarian crisis in 2021; the index is based on
35 socioeconomic factors (development, inequality, aid dependency) and vulnerable groups (DRMKC, 2020).

36

Do Not Cite, Quote or Distribute  13-8        Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 13.1.3 Impact Assessment of Climate Change based on Previous Reports

 3

 4 The main findings of previous reports, particularly the WGII AR5 (Kovats et al., 2014) and the Special
 5 Report of Global warming of 1.5癈 GWL (Hoegh-Guldberg et al., 2018), highlighted the impacts of
 6 warming and rainfall variations and their extremes on Europe, particularly southern Europe and mountainous
 7 areas. At 2癈 GWL, 9% of Europe's population was projected to be exposed to aggravated water scarcity,
 8 and 8% of the territory of Europe were characterized to have a high or very high sensitivity to desertification
 9 (UNEP/UNECE, 2016). These impacts are driven by changes in temperature, precipitation, irrigation
10 developments, population growth, agricultural policies, and markets (EEA, 2017a). Heat is a main hazard for
11 high-latitude ecosystems (Kovats et al., 2014; Jacob et al., 2018; Hock et al., 2019). The majority of
12 mountain glaciers lost mass during the last two decades, and permafrost in the European Alps and
13 Scandinavia is reducing (Hock et al., 2019). In central Europe, Scandinavia and Caucasus, mountain glaciers
14 were projected to lose 60% to 80% of their mass by the end of the 21st century (Hock et al., 2019). The
15 combined impacts on tourism, agriculture, forestry, energy, health and infrastructure were suggested to make
16 southern Europe highly vulnerable and increase the risks of failures and vulnerability for urban areas (Kovats
17 et al., 2014). Previous reports stated that the adaptive capacity in Europe is high compared to other regions of
18 the world, but that there are also limits to adaptation from physical, social, economic, and technological
19 factors. Evidence suggested that staying within 1.5癈 GWL would strongly increase Europe's ability to
20 adapt to climate change (de Conick and Revi, 2018).

21

22 13.1.4 European Climate: Main Conclusions of WGI AR6

23

24 Changes of several climatic impact-drivers have already emerged in all regions of Europe: increases in mean
25 temperature and extreme heat and decreases of cold spells (Ranasinghe et al., 2021; Seneviratne et al., 2021).
26 Lake and river ice has decreased in NEU, WCE and MED and sea ice in NEUS (Fox-Kemper et al., 2021;
27 Ranasinghe et al., 2021). With increasing warming, confidence in projections is increasing for more drivers
28 (Figure 13.3). Mean and maximum temperatures, frequencies of warm days and nights, and heat waves have
29 increased since 1950, while the corresponding cold indices have decreased (high confidence) (Ranasinghe et
30 al., 2021; Seneviratne et al., 2021). Average warming will be larger than the global mean in entire Europe,
31 with largest winter warming in NEU and EEU and largest summer warming in the MED (high confidence)
32 (Guti閞rez et al., 2021; Ranasinghe et al., 2021). An increase of hot days and a decrease of cold days are very
33 likely (Figure 13.4.a,b). Projections suggest a substantial reduction of European ice glacier volumes and of
34 snow cover below elevations of 1500-2000 m, as well as further permafrost thawing and degradation, during
35 the 21st century even at a low GWL (high confidence) (Ranasinghe et al., 2021).

36

37

Do Not Cite, Quote or Distribute  13-9        Total pages: 143
FINAL DRAFT                                 Chapter 13                       IPCC WGII Sixth Assessment Report

Changes in climate impact drivers
Observations from 1970-2019, Projected changes based on warming levels

(a) European land Areas NEU  WCE       EEU  MED         (b) European seas       NEUS  TEUS  SEUS

Mean    Observed 贩�              贩�  贩�  贩�         Relative sea Observed   贩穏  贩�   贩�
                                                                                贩穏  贩�   贩�
warming 1.5癈          贩�        贩�  贩�  贩�         level & coastal 1.5癈    贩�  贩�   贩�
                                                                                 贩�  贩�   贩�
        4癈            贩�        贩�  贩�  贩�         flooding        4癈      贩�  贩�   贩�

Heat    Observed 贩�              贩�  贩�  贩�         Acidification Observed

extremes 1.5癈         贩�        贩�  贩�  贩�                         1.5癈

        4癈            贩�        贩�  贩�  贩�                         4癈

Cold    Observed 贩�              贩�  贩�  贩�

extremes 1.5癈         贩�        贩�  贩�  贩�         Direction of            Increasing
                                                        change                  Decreasing
        4癈            贩�        贩�  贩�  贩�                                 Increasing & decreasing
                                                                                No change
Mean    Observed 贩               贩   贩   贩

precipitation 1.5癈    贩                   贩

        4癈            贩�                  贩�

Heavy   Observed 贩�              贩   贩�  �           Confidence in       � Low
                                                        assessment of      贩 Medium
precipitation 1.5癈    贩         贩   贩   穌          change             贩� High

        4癈            贩�        贩�  贩�  贩穌

Droughts Observed 贩              積   積   贩

(for all types) 1.5癈  積         積   積   贩                                  Not assessed
                                                                                Assessed, but low confidence in
        4癈            贩e        贩   �    贩�                                 direction of change

Severe wind Observed �            �    �    �

storms  1.5癈          �          �    �    穎

        4癈            贩         贩   贩   贩f

         (d) There are subregional differences with decreases or no change for the southern part of Europe, such as the southern Mediterranean
         (e) There are differences among types, areas, seasons and metrics.
         (f) Increased intensity is associated with decreased frequency
 1 (g) Future increase in NEUS does not apply to the northern Baltic Sea

 2 Figure 13.3: Observed and projected direction of change of climate impact drivers at 1.5癈 and 4癈 GWL for
 3 European sub-regions and European Seas (assessment from Guti閞rez et al., 2021; Ranasinghe et al., 2021; Seneviratne
 4 et al., 2021)

 5

 6

 7 The assessment of climate change in WGI AR 6 concludes that during recent decades mean precipitation has
 8 increased over NEU, WCE and EEU, while magnitude and sign of observed trends depend substantially on
 9 time period and study region in MED (medium confidence) (Douville et al., 2021; Guti閞rez et al., 2021;
10 Ranasinghe et al., 2021). Precipitation extremes have increased in NEU, and EEU (high confidence)
11 (Seneviratne et al., 2021), vary spatially in WCE (medium confidence), and have not changed in MED (low
12 confidence). For >2癈 GWL, of mean precipitation in NEU in winter is increasing and decreasing in MED in
13 summer (high confidence). A widespread increase of precipitation extremes is projected for > 2癈 GWL for
14 all subregions (high confidence), except for MED where no change or decrease is projected in some areas
15 (Figure 13.4c,d, Guti閞rez et al., 2021; Ranasinghe et al., 2021). WGI assessed projections for
16 meteorological, agricultural/ecological, hydrological drought (Ranasinghe et al., 2021) with low confidence
17 in the direction of change in NEU, WCE, EEU at 1.5癈 GWL. MED is projected to be most affected within
18 Europe with all types of droughts increasing for 1.5癈 (medium confidence) and 4癈 GWL (high
19 confidence). At 4癈 GWL, hydrological droughts in NEU, WCE and EEU will increase (medium
20 confidence). Projections for the 21st century show increases of storms across all Europe (medium confidence)
21 for >2癈 GWL with a decrease of their frequency in the MED (Ranasinghe et al., 2021).

22

23

Do Not Cite, Quote or Distribute                 13-10                                Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.4: Changes of climate hazards with respect to the CMIP6 baseline (Guti閞rez et al., 2021) for global warming
 3 levels of 1.5癈 and 3癈 GWL and present exposure or vulnerability: (a,b) number of days with temperature maximum
 4 above 35癈 (TX35) and population density (European Comission, 2019); (c,d) daily precipitation maximum (Rx1day)
 5 and built up area (JRCdatacatalogue, 2018); (e,f) consecutive dry days (CDD) and annual harvested rainfed area
 6 (Portmann et al., 2010); (g,h) sea surface temperature (TOS) and Marine Protected Areas (EEA, 2021b); and (k,l) sea
 7 level rise (SLR) and coastal population (Merkens et al., 2016). SLR data consider the long term period (2081-2100) and
 8 SSP1-2.6 for (i) and SSP3-7.0 for (j).

 9

10

11 Sea surface warming between 0.25癈 and 1癈 has been observed in all regions over the last decades (high
12 confidence) (Ranasinghe et al., 2021) and projected to continue increasing (high confidence), particularly in

Do Not Cite, Quote or Distribute  13-11       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 the SEUS and at the NEUS (Figure 13.4 g,h, Guti閞rez et al., 2021). Salinity has increased in the SEUS and
 2 decreased in northern European seas and projected to continue (medium confidence) (Fox-Kemper et al.,
 3 2021). European waters have been and will continue acidifying (virtually certain) (Eyring et al., 2021; Naik
 4 et al., 2021), resulting in a mean decrease of surface pH of about 0.1 and 0.3 pH units at 1.5癈 and 3.0癈
 5 GWL with largest changes at high latitudes (Guti閞rez et al., 2021).

 6

 7 Relative sea level has risen along the European coastlines (Ranasinghe et al., 2021), regionally mitigated by
 8 post-glacial rise of land masses in Scandinavia (Fox-Kemper et al., 2021). SLR will very likely continue to
 9 increase during the 21st century (Figure 13.4 k,l) (high confidence), with regional deviations from global
10 mean sea level rise (low confidence). Extreme water levels, coastal floods, and sandy coastline recession are
11 projected to increase along many European coastlines (high confidence) (Ranasinghe et al., 2021).

12

13

14 13.2 Water

15

16 13.2.1 Observed Impacts and Projected Risks

17

18 13.2.1.1 Risk of Coastal Flooding and Erosion

19

20 Almost 50 million European citizens live within 10 m above mean sea level (Vousdoukas et al., 2020;
21 McEvoy et al., 2021). Without further adaptation (section 13.2.2), flood risks along Europe's low-lying
22 coasts and estuaries will increase due to sea level rise (SLR) compounded by storm surges, rainfall and river
23 runoff (high confidence) (Mokrech et al., 2015; Arns et al., 2017; Sayol and Marcos, 2018; Vousdoukas et
24 al., 2018a; Bevacqua et al., 2019; Couasnon et al., 2020). The population at risk of a 100-year flood event
25 starts to rapidly increases beyond 2040 (Vousdoukas et al., 2018a) reaching 10 million people under RCP8.5
26 by 2100 but stays just below the 10 million under RCP2.6 by 2150 (Figure 13.5, Haasnoot et al., 2021b)
27 assuming present population and protection. The number of people at risk is projected to increase and risk to
28 materialise earlier particularly under SSP5 due to increasing population trends(Vousdoukas et al., 2018a;
29 Haasnoot et al., 2021b). Under high rates of SLR resulting from rapid ice-sheet loss from Antarctica, risks
30 may increase by a third by 2150 (Haasnoot et al., 2021b). Expected annual (direct) damages due to coastal
31 flooding are projected to rise from 1.3 billion today to 13�39 billion by 2050 between 2癈 and 2.5癈
32 GWL and 93�960 billion by 2100 between 2.5� and 4.4癈 GWL, largely depending on socio-economic
33 developments (Cross-Chapter Box SLR in Chapter 3, Vousdoukas et al., 2018a) (high confidence in the sign;
34 low confidence in the numbers). UNESCO World Heritage sites in the coastal zone are at risk due to SLR,
35 coastal erosion and flooding (CCP4, Section 13.8.1.3, Marzeion and Levermann, 2014; Reimann et al.,
36 2018b) as are coastal landfills and other key infrastructure in Europe (AR6/SROCC, Brand et al., 2018;
37 Beaven et al., 2020).

38

39 Observations indicate that soft cliffs and beaches are most affected by erosion in Europe, with e.g. 27 to 40%
40 of Europe's sandy coast eroding today, without climate change being identified as the main driver so far
41 (Pranzini et al.; Luijendijk et al., 2018; Mentaschi et al., 2018; Oppenheimer et al., 2019). SLR will increase
42 coastal erosion of sandy shorelines (high confidence) (Ranasinghe et al., 2021), but there is low confidence in
43 quantitative values assessment of erosion rates and amounts (Athanasiou et al., 2019; Le Cozannet et al.,
44 2019; Thieblemont et al., 2019). Without nourishment or other natural or artificial barriers to erosion, sandy
45 shorelines can retreat by about 100m in Europe at 4癈 GWL; limiting warming to 3癈 GWL can reduce this
46 value by one third (Vousdoukas et al., 2020).

47

48

49 [START BOX 13.1 HERE]

50

51 Box 13.1: Venice and its Lagoon

52

53 Venice and its lagoon are a UNESCO World Heritage Site. This socio-ecological system is the result of
54 millennia of interactions between people and the natural environment. It is exposed to climatic and non-
55 climatic hazards: more frequent floods, warming, pollution, invasive species, reduction of salt marshes,
56 hydrodynamic and bathymetric changes, and waves generated by cruise ships and boat traffic.

57

Do Not Cite, Quote or Distribute  13-12       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 The elevation of the average city pedestrian level and of its inner historic area are respectively 105cm and
 2 55cm above the present relative mean sea level (RMSL). Consequently, even small surges and compound
 3 events cause floods when they coincide with high tide (Lionello et al., 2021a). During the 20th century,
 4 RMSL has risen at about 2.5 mm/yr due to SLR and land subsidence (Zanchettin et al., 2021). The frequency
 5 of floods affecting the city has increased from once per decade in the first half of the 20th century to 40 times
 6 per decade in the period 2010-2019 (Figure Box 13.1.1a).

 7

 8 In 1973, the Italian government established a legal framework for safeguarding Venice and its lagoon.
 9 Construction of the flood protection system started in 2003 and were used for the first time in October 2020
10 to protect the city from floods (Lionello et al., 2021b). This system of mobile barriers (MoSE) closes the
11 lagoon inlets to avoid floods when needed, while under normal conditions they lay on the seabed, thus
12 allowing ship traffic and the exchange between the lagoon and the sea (Molinaroli et al., 2019). To prevent
13 the flooding of the central monument area, additional measures are proposed including inlets, expansion of
14 saltmarshes, and pumping seawater into deep brackish aquifers to raise the city's level (Umgiesser, 1999;
15 Umgiesser, 2004; Teatini et al., 2011).

16

17 Without adaptation, potential economic damages between 7 and 17 billion have been estimated for the
18 next 50 years (Caporin and Fontini, 2016). Additionally, the ecosystem is vulnerable to warming (Solidoro et
19 al., 2010) and sea level rise (Day Jr et al., 1999; Marani et al., 2007). The duration of the closure of the
20 lagoon inlets is expected to increase from 2 to 3 weeks per year for RMSL rises of 30 cm, to 2 months per
21 year for 50 cm and 6 months per year for 75cm (Umgiesser, 2020; Lionello et al., 2021b) (Figure Box
22 13.1.1b), resulting in disconnection from the sea for most of the time for RMSL rise exceeding 75 cm.
23 Frequent closures of the inlets would prevent ship traffic and in/outflow of water. For Venice, adaptation
24 pathways considering the full range of plausible RMSL (Figure Box 13.1.1c) levels are not available,
25 indicating a long-term adaptation gap. As planning and implementation of adaptation of this extent can take
26 several decades (Haasnoot et al., 2020b, Cross-Chapter Box SLR in Chapter 3) this increases the risk that the
27 city will not be prepared in case of rapid sea level rise.

28

29

Do Not Cite, Quote or Distribute  13-13       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 Figure Box 13.1.1: Venice sea level rise and coastal flooding. (a) Evolution of relative and mean sea level in Venice
 2 and decadal frequency of floods above the safeguard level in the city centre (Frederikse et al., 2020; Lionello et al.,
 3 2021a; Lionello et al., 2021b; Zanchettin et al., 2021). (b) Projected relative sea level rise at the Venetian coast (Fox-
 4 Kemper et al., 2021) and (c) timing when critical relative sea level thresholds will be reached depending on scenarios
 5 and confidence level (Lionello, 2012; Umgiesser, 2020; Lionello et al., 2021a). Panel d: landsat view of Venice, its
 6 lagoon with the three inlets connecting it to the Adriatic Sea.

 7

 8

 9 [END BOX 13.1 HERE]

10

11

12 13.2.1.2 Risks Related to Inland Water

13

Do Not Cite, Quote or Distribute  13-14       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 13.2.1.2.1 Riverine and Pluvial Flooding
 2 Precipitation has raised river flood hazards in WCE and UK by 11% per decade from 1960-2010 and
 3 decreased in EEU and SEU by 23% per decade (Douville et al., 2021; Ranasinghe et al., 2021). The most
 4 recent three decades had the highest number of floods in the past 500 years with increases in summer
 5 (Bl鰏chl et al., 2020). Economic flood damages increased strongly, reflecting increasing exposure of people
 6 and assets (Visser et al., 2014; Hoegh-Guldberg et al., 2018; Merz et al., 2021).

 7

 8 Projections indicate a continuation of the observed trends of river floods hazards in WCE (high confidence)
 9 of 10% at 2癈 GWL and 18% at 4.4癈 GWL and a decrease in NEU and SEU (medium confidence) with
10 respectively 5% and 11% in NEU and SEU for a 100-year peak flow, making Europe one of the regions with
11 the largest projected increase in flood risk (Di Sante et al., 2021; Ranasinghe et al., 2021). While there is
12 disagreement on the magnitude of economic losses and people affected, there is high agreement on direction
13 of change, particularly in WCE (Alfieri et al., 2018). New research increases confidence in AR5 statements
14 that without adaptation measures, increases in extreme rainfall will substantially increase direct flood
15 damages (e.g., Madsen et al., 2014; Alfieri et al., 2015a; Alfieri et al., 2015b; Bl鰏chl et al., 2017; Dottori et
16 al., 2020; Mentaschi et al., 2020). With low adaptation, damages from river flooding are projected to be 3
17 times higher at 1.5癈 GWL, 4 times at 2癈 GWL, and 6 times at 3癈 GWL (Alfieri et al., 2018; Dottori et
18 al., 2020). At 2癈 GWL, incidence of summer floods is expected to decrease across the whole alpine region,
19 whereas winter and spring floods will increase due to extreme precipitation (Gobiet et al., 2014) and snow
20 melt driven runoff (Coppola et al., 2018).

21

22 Pluvial flooding and flash floods due to intense rainfall constitute most flood events in SEU and a substantial
23 risk in other European regions (CCP4, Llasat et al., 2016; Rudd et al., 2020). The majority (56 %) of flood
24 events between 1860-2016 were flash floods (Paprotny et al., 2018a). These had considerable impacts
25 including danger to human lives, e.g. causing total economic damage of USD 1 billion in Copenhagen
26 (Denmark) 2011 (W骿cik et al., 2013), a damage to private households of more than 70 million in M黱ster
27 (Germany) 2014 (Spekkers et al., 2017), and during the 2021 floods in Belgium, Germany and the
28 Netherlands over 200 deaths, damage to thousands of homes and disrupted water and electricity supply
29 (Kreienkamp et al., 2021). The intensity and frequency of heavy rainfall events is projected to increase (high
30 confidence) (Figure 13.3, Ranasinghe et al., 2021). Combined with increasing urbanization, the risk of
31 pluvial flooding is projected to increase (Westra et al., 2014; Rosenzweig et al., 2018; Papalexiou and
32 Montanari, 2019). Small catchments, steep river channels and cities are particularly vulnerable due to large
33 areas of impermeable surfaces where water cannot penetrate (Section 13.6).

34

35 13.2.1.2.2 Low Flows and Water Scarcity
36 The frequency and severity of low flows are projected to increase, making streamflow drought and water
37 scarcity more severe and persistent in SEU and WCE (medium confidence) (Figure 13.3, Ranasinghe et al.,
38 2021), but decreases are projected in most of NEU except southern UK (Forzieri et al., 2014; Prudhomme et
39 al., 2014; Schewe et al., 2014; Roudier et al., 2016; Ranasinghe et al., 2021). In EEU, uncertainty about
40 changes in water scarcity pose distinct challenges for adaptation (Greve et al., 2018). At 1.5癈 GWL, the
41 number of days with water scarcity (water availability vs water demand) and drought will increase slightly in
42 SEU (Schleussner et al., 2016; Naumann et al., 2018), resulting in 18% of the population exposed to at least
43 moderate water scarcity, increasing to 54% at 2癈 GWL (Byers et al., 2018). Moderate water scarcity is
44 emerging in some parts of WCE (Bisselink et al., 2018) increasing to16% of the population under 2癈 GWL
45 and SSP2 (Byers et al., 2018). Under 4癈 GWL, areas in WCE experience water scarcity, especially in
46 summer and autumn. Future intensive water use can aggravate the situation, in particular in southern Europe
47 (13.5.1 and 13.10.3).

48

49 Groundwater abstraction rates reach up to 100 million m�/year across WCE and SEU, and exceed 100
50 million m�/year in parts of SEU (Wada, 2016). Low recharge rates lead to a depletion of groundwater
51 resources in parts of SEU and WCE (Doll et al., 2014; Wada, 2016; de Graaf et al., 2017), increasing the
52 impacts on water scarcity in SEU. Groundwater pumping and declines in groundwater discharge already
53 threaten environmental flow limits in many European catchments, especially in SEU, extending to almost all
54 basins and sub-basins within the next 30-50 years (de Graaf et al., 2019).

55

56 The combined effect of increasing water demand and successive dry climatic conditions further exacerbates
57 groundwater depletion and lowers groundwater levels in SEU but also WCE (Goderniaux et al., 2015).

Do Not Cite, Quote or Distribute  13-15       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 Declines in groundwater recharge of up to 30% further increase groundwater depletion (Aeschbach-Hertig
 2 and Gleeson, 2012) especially in SEU and semi-arid to arid regions (Moutahir et al., 2017). Even in WCE
 3 and NEU projected increases in groundwater abstraction will impact groundwater discharge, threatening
 4 sustaining environmental flows under dry conditions (de Graaf et al., 2019).

 5

 6 The risks for soil moisture drought are projected to increase in WCE and SEU for all climate scenarios
 7 (Grillakis, 2019; Tramblay et al., 2020; Ranasinghe et al., 2021). At 3癈 GWL compared to 1.5癈 GWL, the
 8 drought area will increase by 40% and the population under drought by up to 42%, especially affecting SEU,
 9 and to a lesser extent in WCE (Samaniego et al., 2018 ).

10

11 13.2.1.2.3 Water Temperature and Quality
12 Water temperatures in rivers and lakes have increased over the past century by ~1 to 3癈 in major European
13 rivers (CBS, 2014; EEA, 2017b; Woolway et al., 2017). Warming is accelerating for all European river
14 basins (Wanders et al., 2019) increasing by 0.8癈 in response to 1.5癈 GWL and 1.2癈 for 3癈 GWL
15 relative to 1971-2000 (van Vliet et al., 2016a) aggravated by declines in summer river flow.

16

17 (Ground)water extractions or drainage have caused saltwater intrusions (Rasmussen et al., 2013; Ketabchi et
18 al., 2016). During summer, seawater will also penetrate estuaries further upstream in response to reduced
19 river flow and SLR and result in more frequent closure of water inlets in the downstream part of the rivers in
20 a period when water is most needed (e.g., Haasnoot et al., 2020b) (high agreement, low evidence).

21

22 13.2.2 Solution Space and Adaptation Options

23

24 In recent decades water management in Europe has increasingly shifted towards integrated and adaptive
25 strategies, with most noticeable shifts in WCE (high confidence) (e.g., Kreibich et al., 2015; Bubeck et al.,
26 2017). While adaptive strategies are increasingly considered as an approach to strengthen flexibility and
27 implement climate change adaptation actions, given deep uncertainty about the future (Ranger et al., 2013;
28 Klijn et al., 2015; Bloemen et al., 2019; Hall et al., 2019; Pot et al., 2019), more traditional water
29 management approaches still dominate across Europe (OECD, 2013; OECD, 2015; Wiering et al., 2017).
30 Current measures focus on structural flood protection and water resources supply and play an important role
31 to preserve present land use and development patterns. The long-term effectiveness of such measures is
32 increasingly challenged by their reinforcing path-dependency (e.g. flood defence and water supply attract
33 developments which require further protection and supply); this path-dependency limits the solution space
34 and may hamper implementation of transformative measures such as land use change to accommodate the
35 water system (medium confidence) (CCP2, Di Baldassarre et al., 2015; Kreibich et al., 2015; Alfieri et al.,
36 2016; Gralepois et al., 2016; Welch et al., 2017; Di Baldassarre et al., 2018; Haer et al., 2020).

37

38 Water laws, policies and guidance documents increasingly mainstream climate impacts and adaptation
39 options (Runhaar et al., 2018; Mehryar and Surminski, 2021), though not everywhere. Differences are
40 apparent for example in coastal adaptation where most but not all countries are planning for SLR (Figure
41 13.5) (McEvoy et al., 2021). Although the planning horizon of 2100 and 1m SLR are most common
42 (adjusted for local conditions), there are significant differences between countries (e.g. the high-end SLR
43 value in 2100 ranges from 0.3 to 3m), which may lead to unequal impacts, over time (McEvoy et al., 2021).

44

45

Do Not Cite, Quote or Distribute  13-16       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.5: Sea level rise (SLR) vulnerability and national planning in Europe. (a) Map of countries in Europe
 3 summarizing the amount of SLR each country is planning for, at different time horizons (blue bars) and present
 4 population (2020) at risk of a 100-year coastal flood event (orange bars) (Haasnoot et al., 2021b). The amounts of SLR
 5 and time horizons reflect national guidance or planning; local or project-based levels may differ (McEvoy et al., 2021).
 6 (b) Projected population at risk to a 1 in 10-year coastal flood event under RCP2.6-SSP1 and RCP8.5-SSP5 assuming
 7 present protection and population, as well as population change according to respectively, SSP1 and SSP5, based on
 8 Merkens (2016); and (c) projected population at risk to 1 in 100-year coastal flood event under RCP2.6-SSP1 and
 9 RCP8.5-SSP5, assuming present protection and population, as well as population change according to respectively,
10 SSP1 and SSP5, based on Merkens (2016) (Haasnoot et al., 2021b).

11

12

13 13.2.2.1 Flood Risk Management

14

15 Across Europe a range of measures have been implemented to address flood risk (Figure 13.6), with
16 protection as a most used strategy (high confidence). Early warning and flood protection have been
17 successful in reducing vulnerability to coastal and riverine flooding (Jongman et al., 2015; Kreibich et al.,
18 2015; Bouwer and Jonkman, 2018). Consequently, fatalities due to river flooding have decreased in Europe,
19 despite similar numbers of people exposed (1990-2010 compared to 1980-1989) (Jongman et al., 2015;
20 Paprotny et al., 2018a).

21

22 Coastal flood risk management

Do Not Cite, Quote or Distribute  13-17       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Further protection against coastal flooding is considered economically beneficial for densely populated areas
 3 (Lincke and Hinkel, 2018; Tiggeloven et al., 2020). At least 83% of flood damages due to coastal flooding
 4 could be avoided by elevating dykes along ~23-32% of Europe's coastline by 2100 (RCP4.5-SSP1, RCP8.5-
 5 SSP5) (Vousdoukas et al., 2020). Limitations of building flood defences include cost-benefit considerations
 6 in rural areas, available land, and social acceptability in densely populated areas (Haasnoot et al., 2018;
 7 Hinkel et al., 2018; Meyerhoff et al., 2021).

 8

 9 Nature-based (NbS, e.g. wetlands) and sediment-based (e.g. sand nourishment) solutions are increasingly
10 considered for environmental, economic and/or societal reasons (Cross-Chapter Box NATURAL in Chapter
11 2, Stive et al., 2013; Pranzini et al., 2015; Pinto et al., 2020; de Schipper et al., 2021). Coastal wetlands can
12 be effective to reduce wave height and form habitats, but their feasibility and effectiveness is limited for
13 densely populated areas with competing land use, runoff of pollution, sediment starved deltas like the Rhine
14 delta (Edmonds et al., 2020) and rapid SLR (Kirwan et al., 2016; Oppenheimer et al., 2019; Haasnoot et al.,
15 2020b). While losses of wetlands could be minor if warming stays below 1.7癈 GWL, at high warming or
16 SLR above 0.5m large scale losses of these habitats impact their ecological importance, ecosystem function
17 (Section 13.4, Key Risk 1, 13.10) and their ability to protect coastlines (Roebeling et al., 2013; van der Spek,
18 2018; Wang et al., 2018; Xi et al., 2021). A combination with structural defences could reduce risk in
19 urbanized coastal regions (high confidence). Accommodation through elevated or floating houses have been
20 implemented and proposed locally within cities as part of a hybrid strategy together with protection and as a
21 way of innovative urban development (13.6.2, CCP2, Penning-Rowsell, 2020; Storbj鰎k and Hjerpe, 2021).

22

23 Avoidance through restricting new developments in flood prone areas is applied along the coast of WCE and
24 SEU (Harman et al., 2015; Lincke et al., 2020) and is considered a low-cost alternative to coastal defence at
25 lower SLR. In SEU, an Integrated Coastal Zone Management (ICZM) protocol has been developed which
26 requires a setback zone of 100m from the coast in unprotected areas. Setback zones are projected to reduce
27 impacts considerably in urbanized regions (Lincke et al., 2020). Planned relocation is increasingly
28 considered as a realistic adaptation option in case of extreme SLR (Haasnoot et al., 2021a; Lincke and
29 Hinkel, 2021; Mach and Siders, 2021), e.g UK Shoreline Management Plans (Nicholls et al., 2013; Buser,
30 2020). Retreat is rarely applied in Europe (medium confidence), though it can have larger benefit-cost
31 outcomes than protection, particularly in less populated parts of Europe (Lincke and Hinkel, 2021). Along
32 parts of the coast in the UK (e.g., The Wash), Germany (e.g., Langeoog Island), and the Netherlands (e.g.,
33 Westerschelde) retreat has been applied to restore salt marshes and to aid coastal defence (Haasnoot et al.,
34 2019; Kiesel et al., 2020). (Lincke and Hinkel, 2021)

35

36 Riverine and pluvial flood risk management

37

38 Structural flood protection (e.g. levees) is considered economically beneficial in densely populated areas
39 (Alfieri et al., 2016; Dottori et al., 2020) and could reduce flood damage of ~45% is estimated under 1.5癈
40 GWL and ~70% under 3癈 GWL (Dottori et al., 2020).

41

42 Providing more room for water through NbS is increasingly considered (Kreibich et al., 2015) as they can
43 reduce risk effectively at lower costs, except in places with limited space or in areas with large protection.
44 Such measures include (forest) restoration for upstream retention, restoration of river channels, and widening
45 riverbeds for natural flood retention (Kreibich et al., 2015; Barth and D鰈l, 2016; Wyga et al., 2018).
46 Natural retention areas are estimated to be the most effective option to reduce riverine flood risk across
47 Europe in the 21st century, followed by protection (low evidence) (Dottori et al., 2020).

48

49 Wet and dry proofing of buildings can be applied at household level. While measures taken at household
50 levels can reduce the risk of flooding, there is often insufficient investment (Bamberg et al., 2017; Aerts et
51 al., 2018) (medium confidence). Reasons include low awareness or under-estimation of the risk (Kellens et
52 al., 2013), low perceived efficacy of adaptation measures (van Valkengoed and Steg, 2019), and lack of
53 financial support (Kreibich, 2011). In the long-term, risk reduction measures by governments are projected
54 to outweigh floodproofing at household level, in particular in WCE, while for near-term household
55 adaptation or regionally in SEU this could reduce risk more effectively (Haer et al., 2019). Relocation of
56 households has occurred in response to river flood events, e.g. the 2013 flood events along the Danube river

Do Not Cite, Quote or Distribute  13-18       Total pages: 143
    FINAL DRAFT                                                                          Chapter 13                                    IPCC WGII Sixth Assessment Report

1 in Austria, with financial compensation playing a crucial role (Mayr et al., 2020; Thaler and Fuchs, 2020;
2 Thaler, 2021).

3

4 Urban drainage infrastructure is designed based on historical rainfall intensities, and thus may not have
5 sufficient capacity for increased future intensities (Dale et al., 2018). Adaptation options to pluvial flooding
6 include large retention ponds , local green spaces and green roofs within cities (Z鰈ch et al., 2017; Maragno
7 et al., 2018; Babovic and Mijic, 2019; Ribas et al., 2020).

8

           Effectiveness & feasibility of adaptation options for water-related climate impacts & risk in Europe

    Impact Type         Adaptation Option                                 Effectiveness Feasibility                                                                       Confidence
                                                                                                                                                                          Evidence
                                                                               Economic              Technological Institutional  Socio-cultural Ecological  Geophysical              Agreement
                                                                                                                                                                   H 1           H           H
    Flooding - Coast/River Flood defenses (Protect)                       H    M                          H                H      NL   L                            NL           H           M

                        Flood preparedness & early warning                L 2  H                          H                H      NL   H

                        plans (Protect/Accommodate)

                        Planned relocation (Retreat)                      H    H                          H                H      L    H                     H            M           H

                        No-build zone, restrict new                       M    H                          H                M      NL   NL                    NL           M           H

                        developments (Avoidance)

                        Flood insurance (Supporting)                      L    M                          H                H      H    M                     NL           M           H

    Flooding - Coast    Ecosystem based (e.g. wetlands, oyster            M    H                          H   NL                  NL   H                     M 1          M           H

                        rSeeedfism) e(Pnrtobteacste)d (e.g. nourishment)  M    M                          M                H      NL   M                     M 4          H           H

                        (Protect)

                        Wet & dry proofing (Accommodate)                  L    M                          NL               H      H    NL                    NL           L           H

    Flooding - River    Ecosystem based (e.g. floodplain                  H    M                          M   NL                  NL   H                     M            M           H

                        restoration, widening riverbed) (Protect)

                        Retention & diversion (Accommodate)               M    H                          H   NL                  NL   NL                    NL           L           H

                        Wet & dry proofing (Accommodate)                  M    M                          M                H      H    NL                    NL           H           M

    Flooding - Pluvial  Retention: green roofs (Accommodate)              L    NL                         NL  NL                  NL   H                     M            M           H

                        Retention: parks (Accommodate)                    H    NL                         M   NL                  NL   H                     M            M           H

                        Update drainage systems & pumps                   NL   L                          NL  NL                  NL   NL                    M            M           H

                        (Accommodate)

    Water Scarcity      Supply: Storage (reservoirs)                      M    L                          M   M 5                 M    L                     M            H           M

                        Supply: Water diversion & transfer                L 6  L                          M                H      M    M                     NL           M           H

                        Supply: Desalinization                            H 7  L                          H                H      H    L                     H            M           H

                        Supply: Water reuse                               M 7  M                          M                M      L    M                     NL           H           H

                        Demand: Water saving & efficiency                 L    M                          M                H      M    M                     M            H           M

                        Demand: Regulate distribution                     M    H                          M                M      NL   NL                    NL           M           H

                        Demand: Economic instruments                      M    H                          M                H      M    NL                    NL           H           M

                        Demand: Land management & cover                   M    H                          M                H      NL   M                     M            M           L

                        change

                        Monitoring & operational management,              L    H                          M                H      NL   H                     NL           M           M

                        drought early warning systems

    Legend                                                                1 Physically hampered in highly urbanized regions

     High = H                                                             2 Low on preventing damage, medium on preventing fatalities

     Medium = M                                                           3 Availability of sand can hamper feasibility in SEU

     Low = L                                                              4 Low in SEU, high in NLU, WCE

    No/Limited Evidence = NL                                              5 In SEU, no evidence for other parts of Europe

9                                                                         6 Medium in SEU & high in WCE/NLU

10 Figure 13.6: Effectiveness and feasibility of water-related adaptation options to achieve objectives under increasing

11 climate hazards (SM13.1, SM13.9).

12

13

14 Early warning systems, insurance and behaviour change can complement protect and accommodate
15 measures to limit residual risk (high confidence). Early warning systems have high monetary benefits
16 (Pappenberger et al., 2015). Behavioural adaptation to flooding relies on recognition of the threat and
17 capacity to respond, both of which are often lacking (Section 13.11.2.2, Bamberg et al., 2017; Haer et al.,

18 2019). Flood risk insurance and compensation systems vary across European countries, ranging from post-
19 disaster payments by governments, compulsory flood insurance, to public-private partnerships where the
20 state acts as reinsurer (Keskitalo et al., 2014; Surminski et al., 2015; Hanger et al., 2018). Risk-based
21 insurance premiums can induce risk averting behaviour but may become unaffordable to poor households
22 and some households in high risk zones (Hudson, 2018; Surminski, 2018). Increasing future flood risks due

23 to both climatic and socioeconomic change could overburden government budgets (medium confidence)
24 (Section 13.11.2, Paudel et al., 2015; Mysiak and Perez-Blanco, 2016; Schinko et al., 2017; Mochizuki et al.,
25 2018), result in unavailable or unaffordable insurance for private customers (13.8.3, Hudson et al., 2016;
26 Surminski, 2018), and underfunding and insufficient solvency of insurance companies (Section 13.6.2.5,
27 Lamond and Penning-Rowsell, 2014). Local knowledge about disastrous flood events in the past can be lost
28 across generations, reverting behaviours to avoid settlement in risky areas (Fanta et al., 2019).

29

30 Limits to adaptation to extremely high sea level rise scenarios have been identified for coastal defences, such
31 as the Venice MoSE barrier (box 13.1), Thames Barrier in the UK (Ranger et al., 2013) and the Maeslant

32 Barrier in the Netherlands (Kwadijk et al., 2010; Haasnoot et al., 2020b). However, the scale and pace of

    Do Not Cite, Quote or Distribute                                                                 13-19                                                                Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 adaptation required to face high-end SLR scenarios along all coasts of Europe is poorly studied. Given the
 2 lead and long lifetime of large critical infrastructure, there is a growing need to look beyond 2100 to support
 3 the design of new infrastructure (Cross-Chapter Box SLR in Chapter 3).

 4

 5 13.2.2.2 Water Resources Management

 6

 7 Planning adaptation to water scarcity has centred on increasing availability and supply of fresh water through
 8 water storage, diversification of sources and water diversion and transfer (high confidence). Reservoirs are
 9 costly, have negative environmental impacts, and will not be sufficient under higher warming levels in every
10 place (Papadaskalopoulou et al., 2015a; Di Baldassarre et al., 2018; Garnier and Holman, 2019). Waste
11 water reuse is considered a low-cost and effective measure where wastewater is available (Lavrnic et al.,
12 2017; De Roo et al., 2020), but public acceptance for domestic reuse is presently limited (high confidence)
13 (Papadaskalopoulou et al., 2015b; Morote et al., 2019). Increasing desalination capacity is used particularly
14 in SEU but has high energy demands and produces brine waste (Garnier and Holman, 2019; Jones et al.,
15 2019; Morote et al., 2019).

16

17 Adaptation measures on the demand side include monitoring (e.g., water meters, early warning systems of
18 drought) and regulating demand, e.g. water restrictions, water pricing, water saving and efficiency measures
19 and land management and cover change (Papadaskalopoulou et al., 2015b; Varela-Ortega et al., 2016;
20 Manouseli et al., 2018; Garnier and Holman, 2019). Prolonged water restrictions and prioritising sectoral
21 supply could result in economic losses e.g. for irrigated agriculture (Section 13.5.2, Wimmer et al., 2014;
22 Salmoral et al., 2019). Economic instruments, such as water pricing, can be effective when combined with
23 incentives for water saving and efficiency (Kayaga and Smout, 2014; Esteve et al., 2018; Crespo et al.,
24 2019). Water saving and efficiency measures, such as leakage repair, education and improved irrigation,
25 could limit conflicts across sectors, but necessitate technological advances and changes of practice together
26 with a willingness to cooperate (Garnier and Holman, 2019; Papadimitriou et al., 2019; Teot髇io et al.,
27 2020). Increased irrigation efficiency has reduced water scarcity, particularly in SEU (13.5, De Roo et al.,
28 2020), and occur at farm level in WCE and NEU (Papadaskalopoulou et al., 2015b; van Duinen et al., 2015;
29 Rey et al., 2017), but come with increasing path-dependency on supply and trade-offs which may not be
30 sustainable on the long-term (high confidence) (Di Baldassarre et al., 2018).

31

32 The assessment of the effectiveness and feasibility of adaptation options shows that a portfolio of supply and
33 demand measures is needed to reduce water scarcity (Key Risk 3, Section 13.10.3), although locally
34 demand-side measures could be sufficient (Kingsborough et al., 2016). Under high warming levels,
35 adaptation to drought and low flows by water saving and efficiency measures may not be sufficient to
36 counteract reduced availability (medium agreement, low evidence) (Collet et al., 2015; De Roo et al., 2020).
37 Successful adaptation in the water sector depends on integrating water considerations into sectoral policies
38 (Collet et al., 2015; Papadaskalopoulou et al., 2016). Inclusive and participatory approaches where (local)
39 stakeholders are actively involved in the initiation and execution of water management can enhance problem
40 ownership, quality and democratic legitimacy of processes and decisions, enhance support and accelerate
41 decisions (Edelenbos et al., 2017; Begg, 2018).

42

43 13.2.3 Knowledge Gaps

44

45 An assessment of the full solution space of adaptation options and pathways under low to high GWLs
46 including the long-term is lacking. A quantification of the effectiveness of measures in reducing risk is
47 limited in the scientific literature. Available assessments consider adaptation by incremental measures.
48 Transformative options, such as land use changes, planned relocation from exposed areas or restricting future
49 development, are rarely considered. While high-end scenarios, describing low confidence processes and
50 scenarios beyond 2100 are considered to be useful for risk-averse decision making, in particular coastal
51 adaptation (Hinkel et al., 2019; Haasnoot et al., 2020b), they are rarely considered in practice.

52

53

54 13.3 Terrestrial and Freshwater Ecosystems and their Services

55

56 13.3.1 Observed Impacts and Projected Risks

57

Do Not Cite, Quote or Distribute  13-20       Total pages: 143
FINAL DRAFT                       Chapter 13                          IPCC WGII Sixth Assessment Report

1 13.3.1.1 Observed Impacts on Terrestrial and Freshwater Ecosystems

2
3

 4

 5 Figure 13.7: K鰌pen-Geiger climate classification and biodiversity hotspots in Europe. Boundaries of the (a) Northern
 6 (NEU), (b) Western-Central (WCE), (c) Southern (SEU), and (d) Eastern (EEU) European regions for 1985-2014 (left)
 7 and 2076-2100 (right, A1FI scenario, ~4癈 GWL), based on Rubel and Kottek (2010).

 8

 9

10 European land and freshwater ecosystems (Figure 13.7) are already strongly impacted by a range of
11 anthropogenic drivers (very high confidence), particularly habitats at the southern and northern margins,
12 along the coasts, up mountains and in freshwater systems (CCP1). Interacting with climate change are non-
13 climatic hazards, such as habitat loss and fragmentation, over-exploitation, water abstraction, nutrient
14 enrichment, and pollution, all of which reduce resilience of biotas and ecosystems (very high confidence).
15 Peatlands in NEU and EEU and other historically important cultural landscapes in Europe are overexploited
16 for forestry, agriculture, and peat mining (Page and Baird, 2016; Tanneberger et al., 2017; Ojanen and
17 Minkkinen, 2020). Inland wetland RAMSAR convention sites in Europe, which constitute 47% of the global
18 sites, have lost area in WCE and gained in SEU from 1980 to 2014 (Xi et al., 2021). Forests in WCE were
19 impacted by the extreme heat and drought event of 2018, with effects lasting into 2019 (Schuldt et al., 2020)
20 and losses in conifer timber sales in Europe (Hl醩ny et al., 2021).

21

22 Extirpation, e.g. local losses of species, have been observed in response to climate change in Europe
23 (medium confidence) (Wiens, 2016; EEA, 2017a; Soroye et al., 2020). Strong climate-induced declines have
24 been detected in thermosensitive taxa (Hellmann et al., 2016), including many freshwater groups, insects
25 (Habel et al., 2019; Harris et al., 2019; Seibold et al., 2019; Soroye et al., 2020), amphibians, reptiles
26 (Falaschi et al., 2019), birds (Lehikoinen et al., 2019) and fishes (Myers et al., 2017a; Jari et al., 2019 ). The
27 loss of native species, especially specialised taxa, is changing biodiversity; however overall biodiversity
28 could remain stable because losses may be compensated by range shifts of native and the establishment of
29 non-native species (Dornelas et al., 2014; McGill et al., 2015; Hillebrand et al., 2018; Outhwaite et al.,
30 2020).

31

32

Do Not Cite, Quote or Distribute  13-21                               Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.8: Summary of major impacts on and risks for terrestrial and freshwater ecosystems in Europe for 1.5篊 and
 3 3篊 GWL (Table SM13.2).

 4

 5

 6 Range shifts are leading to northward and upward expansions of warm-adapted taxa (very high confidence)
 7 (Figure 13.8 and Chapter 2). These shifts altered species living in the boreal and alpine tundra (Elmhagen et
 8 al., 2015; Post et al., 2019; Mekonnen et al., 2021) and are greening the high Arctic tundra with shrubs and
 9 trees (Myers-Smith et al., 2020). Plants display more stable distributions at low than at higher mountain
10 altitudes (Rumpf et al., 2018 ). Microclimatic variability in some locations can buffer warming impacts
11 (medium confidence) (Suggitt et al., 2018; Zellweger et al., 2020; Carnicer et al., 2021). Northward shifts of
12 tree species distributions is documented in North Western Europe (Bryn and Potthoff, 2018; Mamet et al.,
13 2019) but not consistently detected (Cudl韓 et al., 2017; Vil�-Cabrera et al., 2019).

14

15 The timing of many processes, including spring leaf unfolding and autumn senescence and flight dates
16 changed in response to changes in seasonal temperatures, water and light availability (very high confidence)
17 (Chapter 2, Szab� et al., 2016; Asse et al., 2018; Peaucelle et al., 2019; Menzel et al., 2020; Rosbakh et al.,
18 2021), resulting e.g. in earlier arrival dates for many birds and butterflies (Karlsson, 2014; Bobretsov et al.,
19 2019; Lehikoinen et al., 2019). Greatest growing season lengthening in plants has been detected in WCE,
20 NEU and EEU, but shortening in parts of SEU driven by later senescence (Garonna et al., 2014), increasing
21 population growth for butterflies and moths (Macgregor et al., 2019) and birds (Halupka and Halupka,
22 2017), and residence time for migrant birds (Newson et al., 2016).

23

24 13.3.1.2 Projected Risks for Terrestrial and Freshwater Ecosystems

25

Do Not Cite, Quote or Distribute  13-22       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

1 Risks for terrestrial ecosystems will increase with warming (very high confidence) with high impacts at
2 > 2.4篊 GWL and very high impacts > 3.5篊 GWL (medium confidence) (13.10.3.1). Land use changes will
3 increase extirpation and extinction risk (Vermaat et al., 2017) (very high confidence). In NEU, biodiversity
4 vulnerability is projected to be lower as new climate and habitat space is becoming available (Warren et al.,
5 2018; Harrison et al., 2019). Warming < 1.5篊 GWL would limit risks to biodiversity, while 4篊 GWL and
6 intensive land use may lead to a loss of suitable climate and habitat space for most species (low confidence)
7 (Warren et al., 2018; Harrison et al., 2019).

8

9

10

11 Figure 13.9: Species projected to remain within their suitable climate conditions at increasing levels of climate change.
12 Colour shading represent proportion of species projected to remain within their suitable climates averaged over 21
13 CMIP5 climate models (Warren et al., 2018). Areas shaded in green retain a large number of species with suitable
14 climate conditions, while those in purple represent areas where climates become unsuitable for more than 80% of the
15 without dispersal (Table SM13.3).

16

17

18 Disruption of habitat connectivity reduces resilience and is projected to impact 30% of lake and river
19 catchments in Europe by 2030, through drought and reduced river flows (Markovic et al., 2017) (medium
20 evidence). Average wetland area is not projected to change at 1.7癈 GWL across Europe, while for > 4癈
21 GWL expanding sites in NEU are not sufficient to balance losses in SEU and WCE (high confidence) (Xi et
22 al., 2021). At 3癈 GWL the alpine tundra habitat and its associated species are projected to be lost in the
23 Pyrenees and shrink dramatically in NEU, WCE and EEU (Anisimov et al., 2017; Barredo et al., 2020).

24

25 Population range shifts (Figure 13.7, 13.10) are projected to continue (Figure 13.8) (medium confidence at
26 1.5� GWL, high confidence at 3.0癈 GWL). The largest losses of suitable climatic conditions are projected
27 for plants and insects, with different taxon-specific regions of highest risk, while proportions of species
28 projected to loose suitable climates are lower for other groups (medium confidence) (Figure Box 13.1.1,
29 Table SM13.3, Warren et al., 2018). > 1.5癈 GWL will lead to a progressive subtropicalisation in SEU,
30 expanding into WCE at > 3癈 GWL, a northward shifting of the temperate domain into NEU (Feyen et al.,
31 2020) (medium confidence), and an expansion of desert biomes in EEU (Sergienko and Konstantinov, 2016).
32 Changes in distribution are projected for major tree species in all European regions at 1.7癈 GWL (Dyderski
33 et al., 2018; Leskinen et al., 2020), with economic implications for managed forests (13.5.1.4). The longer

Do Not Cite, Quote or Distribute  13-23       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 growth season in NEU and WCE will support the establishment of invasive species (CCP1). < 1.5篊 GWL
 2 would limit expansion and novel appearances of pests while > 3.4篊 GWL will make large parts of SEU and
 3 WCE suitable for pest, e.g. wood beetles (Urvois et al., 2021), and increase economic losses due to lower
 4 harvest quality of timber (Toth et al., 2020).

 5

 6 Risks emerging from climate change for phenology are uncertain, given asynchrony between species, taxa
 7 and trophic responses (Thackeray et al., 2016; Posledovich et al., 2018; Keogan et al., 2021) and the
 8 complexity of phenological events and their cues (Delgado et al., 2020; Ettinger et al., 2020) (medium
 9 confidence). Spring events may continue to occur earlier (Ga鼁鑢e et al., 2016), but reduced chilling may
10 decrease this temporal shift (Wang et al., 2020). Projections for autumn are mixed, with continuing delays
11 (Prislan et al., 2019) or earlier onset of leaf senescence (Wu et al., 2018), but reduced chilling may also
12 decrease these developments (Wang et al., 2020). Advancement, combined with longer autumn growth, may
13 extend the growing season of trees by two days per decade in SEU (Prislan et al., 2019). Warming to > 3癈
14 GWL will impact forest planning in NEU (Caffarra et al., 2014).

15

16 13.3.1.3 Observed Impacts and Projected Risks of Wildfires

17

18 Fires affect over 400,000 ha every year in the European Union (San-Miguel-Ayanz et al., 2019), with 85% of
19 the area located in SEU (Khabarov et al., 2016; de Rigo et al., 2017; Costa et al., 2020), where `fire weather'
20 conditions (determined by temperature, precipitation, wind speed and relative humidity) are most
21 pronounced (Figure 13.10). Fire hazard conditions, including heat waves (Boer et al., 2017), have increased
22 throughout Europe from 1980 to 2019 (Figure 13.10), with substantive increases in SEU and WCE (high
23 confidence) (Urbieta et al., 2019; Di Giuseppe et al., 2020; Fargeon et al., 2020). Extreme wildfires have
24 been observed in recent years, including 2017 in Portugal, 2018 in Sweden (Krikken et al., 2021) and 2021
25 in south-eastern Europe. In SEU, WCE and NEU human activities caused more than 90-95% of the fires,
26 while natural ignition accounts for a substantial portion of burnt area in EEU (Wu et al., 2015; Filipchuk et
27 al., 2018).

28

29 Except for Portugal, burnt area in SEU has shown a slightly decreasing trend since 1980, with high inter-
30 annual variability (CCP 4, Turco et al., 2016; de Rigo et al., 2017). In SEU, burned terrestrial biomass
31 declined from 2003 to 2019 (Turco et al., 2016), despite increasing fire risks. This trend is parallel to
32 increasing fire management measures implemented (Fernandez-Anez et al., 2021). The slight increase in
33 burned biomass in WCE and NEU is associated with more hazardous landscape configurations and warming
34 in recent decades (Turco et al., 2016; Urbieta et al., 2019).

35

36 Projections of wildfire risks are uncertain due to multiple factors, including compound events, fire-
37 vegetation interaction and social factors (Thompson and Calkin, 2011; San-Miguel-Ayanz et al., 2019).
38 Wildfire risks can increase across all regions of Europe at 1.5癈 and 3癈 GWL (medium to high confidence)
39 (Figure 13.8). In SEU, the frequency of heat-induced fire-weather is projected to increase by 14% at 2.5癈
40 GWL and rising to 30% at 4.4癈 GWL (Turco et al., 2018; Costa et al., 2020; Ruffault et al., 2020). In the
41 European Arctic, the extent and duration of extreme fire seasons will increase because of increasing extreme
42 fire weather, increased lightning activity, and drier vegetation and ground fuel conditions due to prolonged
43 droughts (McCarty et al., 2021). Projections suggest that new fire-prone regions in Europe could emerge,
44 particularly in WCE and NEU where wildfires have been uncommon and fire management capacity is slowly
45 increasing (Wu et al., 2015; Forzieri et al., 2021).

46

47

Do Not Cite, Quote or Distribute  13-24       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.10: Geographical variability and dynamic changes in fire danger in Europe over the last decades. Significant
 3 increases in fire hazard at the multidecadal scale and unprecedented years of elevated fire hazard have occurred over the
 4 last decade in Southern and Western and Central Europe (SEU, WCE). The environmental conditions required for fires
 5 to spread and intensify were evaluated using fire hazard estimates (`Fire Weather Index' (FWI) based on meteorological
 6 variables such as temperature, precipitation, wind speed and relative humidity). FWI trends were calculated with the
 7 ECMWF ERA-5 FWI reanalysis dataset (Copernicus, 2019; Copernicus, 2020a; Copernicus, 2020b).

 8

 9

10 13.3.1.4 Observed Impacts and Projected Risks on Ecosystem Functions and Regulating Services

11

12 European temperate and boreal forests, wetlands and peatlands hold important carbon stocks (Bukvareva and
13 Zamolodchikov, 2016; Yousefpour et al., 2018). Effects of warming and increasing droughts on soil
14 moisture, respiration and carbon sequestration have been detected across European regions (high confidence)
15 (Tab. 13.3, Sangin閟 de C醨cer et al., 2018; Carnicer et al., 2019; Green et al., 2019; Schuldt et al., 2020).
16 Forest expansion in boreal regions results in net warming (Bright et al., 2017), possibly influencing cloud
17 formation and rainfall patterns (medium confidence) (Teuling et al., 2017). These changes are affecting
18 climate, pollination and soil protection services (Table 3.3, Verhagen et al., 2018). If not managed through
19 increased reforestation/revegetation or peatland restoration, future climate change impacts will progressively
20 limit the climate regulation capacity of European terrestrial ecosystems (medium confidence) (Figure 13.8),
21 especially in SEU (Pe駏elas et al., 2018; Xu et al., 2019). Predominantly positive CO2 fertilization effects at
22 current warming will change into increasingly negative effects of warming and drought on forests at higher
23 temperatures (medium confidence) (Pe駏elas et al., 2017; Green et al., 2019; Ito et al., 2020; Wang 2020; Yu
24 et al., 2021). In NEU and EEU, peatlands are projected to shrink with 1.7癈 GWL, and become carbon
25 sources at 3癈 GWL (Qiu et al., 2020), peat bogs to lose 50% carbon at 2癈 GWL, and blanket peatland to
26 shrink or regionally disappear (Gallego-Sala et al., 2010; Ferretto et al., 2019)

27

28 Declines in pollinator ranges in response to climate are occurring for many groups in Europe (high
29 confidence) (Figure Box 13.1.1, Table 13.3, Kerr et al., 2015; Soroye et al., 2020; Zattara and Aizen, 2020),
30 with observed shifts to higher elevations of southern and lower elevation in northern species (Kerr et al.,
31 2015) resulting in higher pollinator richness in NEU (Franz閚 and 謈kinger, 2012). Lags in responses to
32 climate change suggest current impacts on pollination have not been fully realized (IPBES, 2018).

Do Not Cite, Quote or Distribute  13-25       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 Pollinators are also declining due to lack of suitable habitat, pollution, pesticides, pathogens and competing
 2 invasive alien species (Settele et al., 2016; Steele et al., 2019).

 3

 4 Projected climate impacts on pollinators show mixed responses across Europe, but are greater under 3癈
 5 GWL (medium confidence) (Rasmont et al., 2015). Increasing homogenisation of populations may increase
 6 vulnerability to extreme events (Vasiliev and Greenwood, 2021). Geographic changes to the climatic niche
 7 of pollinators are similar to insects, with mixed trends, depending on group and location (Figure 13.9,
 8 Kaloveloni et al., 2015; Rasmont et al., 2015; Radenkovi et al., 2017). In NEU, species richness may
 9 increase for some groups (Rasmont et al., 2015), with unclear trends for bumblebees (Fourcade et al., 2019;
10 Soroye et al., 2020). Future land use will have important effects on pollinator distribution (Marshall, 2018)
11 as habitat fragmentation in densely populated Europe decreases opportunities for range shifts and micro-
12 climatic buffering (Vasiliev and Greenwood, 2021).

13

14 Soil erosion varies across Europe, with higher rates in parts of SEU and WCE, but lower in NEU (high
15 confidence) (Table 13.3, Petz et al., 2016; Polce et al., 2016; Borrelli et al., 2020), related to vegetation type
16 and amount of cover, slope and soil type (Panagos et al., 2015a). Short-term, land use change and
17 management may impact soil erosion more than climate (Verhagen et al., 2018). Where conservation
18 agriculture is practised or vegetation cover increasing, erosion is slightly decreasing (Panagos et al., 2015b;
19 Guerra et al., 2016). Reduced soil loss due to reduced spring snow melt has been observed in EEU (Golosov
20 et al., 2018) , while fire exacerbates soil loss, especially in SEU (Borrelli et al., 2016; Borrelli et al., 2017).

21

22 Projected increase in rainfall could increase soil erosion, while warming enhances vegetation cover, leading
23 to overall mixed responses (medium confidence) (Berberoglu et al., 2020; Ciampalini et al., 2020). In
24 Europe, rainfall erosivity could increase by >81% (Panagos et al., 2017) at 2癈 GWL, especially in NEU
25 (Borrelli et al., 2020) where risks can be limited by soil erosion control (Polce et al., 2016). Decreased
26 rainfall projected for parts of SEU could reduce erosion, although increases in rainfall intensity could offset
27 this (Serpa et al., 2015). Soil losses from fire will increase in SEU in response to 2癈 GWL (Pastor et al.,
28 2019), especially if combined with extreme rainfall (Mor醤-Ord篑ez et al., 2020). In northern regions,
29 reduced soil losses are projected during spring snowmelt (Svetlitchnyi, 2020).

30

31 13.3.2 Solution Space and Adaptation Options

32

33 Autonomous species adaptation, via range shifts towards higher latitudes and altitudes and changes in
34 phenology, but extirpation have been documented in all European regions (Figure 13.8) (very high
35 confidence). Lowering vulnerability by reducing other anthropogenic impacts (Gillingham et al., 2015), such
36 as land use change, habitat fragmentation (Eigenbrod et al., 2015; Oliver et al., 2017; Wessely et al., 2017),
37 pollution, and deforestation (Chapter 2), enhances adaptation capacity and biodiversity conservation (high
38 confidence) (Ockendon et al., 2018). Protected areas, such as the EU Natura 2000 network, have contributed
39 to biodiversity protection (medium confidence) (Ga鼁鑢e et al., 2016; Sanderson et al., 2016; Santini et al.,
40 2016; Hermoso et al., 2018) but 60% of terrestrial species in these sites could lose suitable climate niches at
41 4癈 GWL (Figure Box 13.1.1, EEA, 2017a).

42

43 Most protected areas are static and thus do not take species migration into consideration (high confidence)
44 (Gillingham et al., 2015; Heikkinen et al., 2020b). More dynamic areas of protection, such as networks of
45 protected areas with corridors, buffer zones and zoning, can facilitate population shifts (Barredo et al., 2016;
46 Nila et al., 2019; Crick et al., 2020; Keeley et al., 2021) and thereby reduce but not eliminate vulnerability
47 (Wessely et al., 2017; Pav髇-Jord醤 et al., 2020).

48

49 Rehabilitation and restoration of land (Prober et al., 2019), particularly abandoned agricultural areas in SEU
50 and NEU (Terres et al., 2015), are long-term strategies to improve regulating services and enhance
51 biodiversity conservation (Morecroft et al., 2019; Campos et al., 2021). Their success will depend on
52 consideration of the future climate niche when restoring peatlands (Bellis et al., 2021) or long-lived species
53 with limited mobility (Hazarika et al., 2021) (high confidence). Combination of supporting the resilience of
54 species, increasing functional diversity of habitats, and assisted migration of species at the limit of their
55 adaptive capacity (Park and Talbot, 2018) are needed to protect and restore ecosystems, e.g. forests (Boiffin
56 et al., 2017; Messier et al., 2019). Successful interventions consider habitat and the ecological and evolution

Do Not Cite, Quote or Distribute  13-26       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 interactions of species (Seho et al., 2019; Diallo et al., 2021) combined with monitoring to assess their
 2 effectiveness (Casazza et al., 2021).

 3

 4 Fire management plans and programs are in place in most of SEU, and increasingly developed in the parts of
 5 Europe where wildfires are less common (Fernandez-Anez et al., 2021). Capacity to implement and maintain
 6 these options remains limited, however (medium confidence). The dominant fire management paradigm of
 7 fire suppression in some regions of SEU has been questioned, as it contributes to fuel accumulation.
 8 Approaches are advocated which combining fire-risk mitigation, prevention, and preparation (Moreira et al.,
 9 2020), recovery through post-fire management (Lucas-Borja et al., 2021), diverse fuels treatment (Mirra et
10 al., 2017), including prescribed burning (Fernandes et al., 2013).

11

12 Ecosystem-based adaptations (EbA) and nature-based solutions (NbS) that restore or recreate ecosystems,
13 build resilience and produce synergies with adaptation and mitigation in other sectors are increasingly used
14 in Europe (high confidence)(Cross-Chapter Box NATURAL in Chapter 2, Berry et al., 2015; Chausson et al.,
15 2020). Planting trees or recreating wetlands can function as part of natural flood management (Dadson et al.,
16 2017; Cooper et al., 2021), whilst urban green infrastructure can reduce flooding (13.2.2) and heat stress and
17 provide recreation opportunities and health benefits (13.6.2.3; Box 13.3) (Kabisch et al., 2016; Choi et al.,
18 2021).
19 Appropriately implemented ecosystem-based mitigation, such as reforestation with climate-resilient native
20 species (13.3.1.4), peatland and wetland restoration and agroecology (13.5.2), can enhance carbon
21 sequestration or storage (medium confidence) (Seddon et al., 2020). Saltmarsh protection or recreation can
22 increase carbon storage capacity, enhance coastal flood protection and provide cultural services (Beaumont
23 et al., 2014; Bindoff et al., 2019). Trade-offs between ecosystem protection, their services and human
24 adaptation and mitigation needs can generate challenges, such as loss of habitats, increased emissions from
25 restored wetlands (G黱ther et al., 2020) and conflicts between carbon capture services, and provisioning of
26 bioenergy, food, timber and water (Lee et al., 2019; Krause et al., 2020) (medium confidence).

27

28 The solution space for responding to climate-change risks for terrestrial ecosystem has increased in parts of
29 Europe (medium confidence). For example, EbA and NbS figure prominently in the EU Adaptation Strategy
30 (2021a) and climate change adaptation is mainstreamed in the EU Biodiversity Strategy for 2030 (European
31 Comission, 2020), the EU Forest Strategy for 2030 (European Comission, 2021b), the EU Green
32 Infrastructure Strategy (European Comission, 2013a), as well as several national and regional policies. Yet,
33 in the northern parts of EEU and NEU (e.g. Greenland, Iceland, NW Russian Arctic), areas which are often
34 sites of pronounced biodiversity shifts and changes, solutions are lacking or slow in emergence, due to
35 remoteness, lack of resources and sparse populations (Canosa et al., 2020). In the EU, innovative financing
36 schemes such as the Natural Capital Financing Facility are being explored by the European Investment Bank
37 and the European Commission which supports projects delivering on biodiversity and climate adaptation
38 through tailored loans and investments. Multiple EU-level service platforms have been promoted to track
39 climate change impacts on land ecosystems and adaptation (e.g. Climate-Adapt, Copernicus Land and Fire
40 Monitoring Service, Forest Information System of Europe) (13.11.1).

41

42 Despite an expanding solution space, widespread implementation and monitoring of natural and planned
43 adaptation across Europe is currently limited, due to high management costs, undervaluation of nature, and
44 conservation laws and regulations that do not consider species shifts under future socioeconomic and
45 climatic changes (high confidence) (Kabisch et al., 2016; Prober et al., 2019; Fernandez-Anez et al., 2021).
46 Climate risks are not perceived as urgent due to a continuing perception of high adaptive capacity of
47 ecosystems (Uggla and Lidskog, 2016; Esteve et al., 2018; Vulturius et al., 2018). Limited financial
48 resources prevent widespread implementation of large-scale and connected conservation areas (high
49 confidence) (Hermoso et al., 2017; Lee et al., 2019; Krause et al., 2020). Particularly in WCE, competition
50 for land use with other functions, including mitigation options, is a critical barrier to implementation of
51 adaptation. Risks to terrestrial and freshwater ecosystems are rarely integrated into regional and local land
52 use planning, land development plans, and agro-system management (medium confidence) (Nila et al., 2019;
53 Heikkinen et al., 2020a) .

54

55 13.3.3 Knowledge gaps

56

Do Not Cite, Quote or Distribute  13-27       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 Despite growing evidence of climate change impacts and risk, including attributed changes to terrestrial
 2 ecosystems (13.10.1), this information is geographically not equally distributed, leaving clear gaps for some
 3 processes or regions (high confidence). For processes such as wildfire, the Fire Weather Index (13.3.1.3)
 4 suggests increasing risks for fires in Europe but robust projections on incidents and magnitudes of wildfire
 5 and their impacts on ecosystems and other sectors is currently limited, particularly for NEU, EEU and WCE
 6 (high confidence).

 7

 8 Many studies consider only individual climate drivers, though new research shows strong interactions
 9 between hazards such as warming and drought (13.3.1), as well as non-climatic drivers (Chapter 2). This
10 creates uncertainty about the emergence of extinctions and the magnitudes of impacts for European
11 ecosystems and the services they provide (high confidence), such as pollination on food production. RCP-
12 SSP combinations to assess risks are only just emerging (Harrison et al., 2019).

13

14 Assessments of the long-term effectiveness of adaptation actions is missing, due to the time lag in
15 determining effectiveness of an action and attributing risk reduction (Morecroft et al., 2019). For example,
16 many landscape restoration actions are discussed but it is unclear which would bring highest benefits and
17 which species should be used for the restoration (Ockendon et al., 2018). Further, adaptation actions will
18 depend on local implementation and benefit from being assessed using cultural and Indigenous knowledge
19 where applicable, but this is hardly studied (medium confidence).

20

21

22 13.4 Ocean and Coastal Ecosystems and their Services

23

24 13.4.1 Observed Impacts and Projected Risks

25

26 13.4.1.1 Observed Impacts

27

28 Warming continues to be the key climate hazard for European seas (Figure 13.1). Interacting with other
29 climatic and non-climatic drivers, it has detectable and attributable impacts at a wide range of biological and
30 ecological organisational levels (Figure 13.11).

31
32

Do Not Cite, Quote or Distribute  13-28       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.11: Major impacts and risks for marine and coastal ecosystems in Europe for observed and projected 1.5 癈
 3 and 3.0 癈 GWL (Table SM13.4).

 4

 5

 6 Particularly habitat loss in shallow coastal waters and at the coasts themselves, and northward distribution
 7 shifts of populations and communities are evident across all European marine subregions (Figure 13.11: high
 8 confidence; Chapter 3). Marine heatwaves have had severe ecological impacts in SEUS (high confidence)
 9 (CCP 4), threatening sessile benthic biotas and coastal habitats (Munari, 2011; Kersting et al., 2013; Rivetti
10 et al., 2014; Garrabou et al., 2019). Range contractions, extirpations (medium confidence) (Smale, 2020) and
11 species redistributions have been observed (high confidence) in TEUS (Cottier-Cook et al., 2017) and SEUS
12 (Castellanos-Galindo et al., 2020). Habitat losses, range shifts, species invasions and species thermal
13 preferences altered community compositions (Vasilakopoulos et al., 2017), resulting in the
14 `subtropicalisation' of TEUS and `tropicalisation' of SEUS (Chapter 3; CCP 4) and temperature-dependent
15 timing of abundance and reproduction cycles (Hjerne et al., 2019; Polte et al., 2021; Uriarte et al., 2021).

16

17 Reductions in growth and reproductive success of calcifying species are not yet unambiguously detected and
18 attributed in European seas (Figure 13.11) (medium confidence), as many show resilience (Kroeker et al.,
19 2010; Wall et al., 2015). However, fish population sizes are shrinking (Queir髎 et al., 2018; Ikpewe et al.,
20 2021), and growth, reproduction and recruitment are negatively impacted (Lindegren et al., 2018; Goldberg
21 et al., 2019; Hidalgo et al., 2019; Vieira et al., 2019; Denechaud et al., 2020; Maynou et al., 2020; Polte et
22 al., 2021), though positive effects also occur (Sguotti et al., 2019; Tanner et al., 2019). Biodiversity changes
23 depend on region, habitat, and taxon (Figure 13.11) (medium confidence) overall resulting in the
24 redistribution of biodiversity in Europe (Garc韆 Molinos et al., 2016), and biodiversity declines in some
25 subregions (high confidence) (IPBES, 2018).

26

27 Biological and ecological impacts have cascading effects for marine ecosystem functioning (Chivers et al.,
28 2017; Baird et al., 2019) and biogeochemical cycling (Huete-Stauffer et al., 2011; Munari, 2011; Kersting et

Do Not Cite, Quote or Distribute  13-29       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 al., 2013; Rivetti et al., 2014; Garrabou et al., 2019). In TEUS, increased water-column stratification
 2 (Section 13.1) and decreasing eutrophication, result in reduced primary production (high confidence)
 3 (Figure 13.11, Capuzzo et al., 2018) and productivity at higher trophic levels (Free et al., 2019) (high
 4 confidence), while in NEUS sea-ice decline resulted in primary production increase by 40-60% (Figure
 5 13.11) (high confidence) (Arrigo and van Dijken, 2015; Borsheim, 2017; Lewis et al., 2020). Climate-related
 6 deoxygenation impacts are small in most European waters (Figure 13.11) (medium confidence), expect for
 7 semi-enclosed seas such as the Baltic and Black Seas (Frolov et al., 2014; Jacob et al., 2014; Reusch et al.,
 8 2018). Here warming and eutrophication altered ecosystem functioning (high confidence), reduced potential
 9 fish yield, increased harmful algal blooms (Alekseev et al., 2014; Carstensen et al., 2014; Berdalet et al.,
10 2017; Daskalov et al., 2017; Riebesell et al., 2018; Stanev et al., 2018), and the risks of Vibrio pathogens and
11 vibriosis (Section 13.7.1, Baker-Austin et al., 2017; Semenza et al., 2017). Across all European seas there is
12 only low confidence of a consistent change in provisioning ecosystem services (e.g., fishing yields, Section
13 13.5), because of interregional variability, but high confidence in the decrease in regulating services and
14 coastal protection because of cascading effects of ecosystem impacts (Figure 13.11).

15

16 13.4.1.2 Projected Risks

17

18 Risks to marine and coastal European ecosystems are very likely to intensify (Figure 13.11) in response to
19 projected further warming. Since the capacity of natural systems for autonomous adaptation is limited
20 (Thomsen et al., 2017; Miller et al., 2018; Bindoff et al., 2019) (medium confidence), pronounced changes in
21 community composition and biodiversity patterns are projected by 2100 for TEUS and the eastern
22 Mediterranean Sea (SEUS) for> 3癈 GWL (Garc韆 Molinos et al., 2016), challenging conservation efforts
23 (Corrales et al., 2018; Cramer et al., 2018; Kim et al., 2019). At 1.5癈 GWL, particularly in winter,
24 Mediterranean coastal fish communities are projected to lose ~10% of species, increasing to ~60% at 4癈
25 GWL (Dahlke et al., 2020), exacerbating regime shifts linked to overexploitation (Clark et al., 2020)
26 (medium confidence). Warming at this level will threaten many species currently living in Marine Protected
27 Areas (MPA) in TEUS and NEUS (Bruno et al., 2018). Increasing marine heatwaves (MWH), particularly in
28 SEUS at 4癈 GWL (Darmaraki et al., 2019a), elevate risks for species (Galli et al., 2017), coastal
29 biodiversity, and ecosystem functions, goods and services (Smale et al., 2019). However, MWH-related risk
30 levels differ among biotas (Pansch et al., 2018) and across European seas (Smale et al., 2015).

31

32 Marine primary production is projected to further decrease by 2100 in most European seas between 0.3% at
33 1.5癈 GWL to 2.7% at 4癈 GWL (Figure 13.11) (high confidence), mainly caused by stratification-driven
34 reductions in nutrient availability, impacting food webs (Doney et al., 2012; Laufkoetter et al., 2015;
35 Wakelin et al., 2015; Salihoglu et al., 2017; Holt et al., 2018; Bryndum-Buchholz et al., 2019; Carozza et al.,
36 2019; Kwiatkowski et al., 2019). In the Barents Sea, however, largely stable primary production is projected
37 under all scenarios in response to sea-ice decline (Slagstad et al., 2011) and in the eastern Mediterranean due
38 to reduced stratification (Macias et al., 2015; Moullec et al., 2019). These changes in productivity are
39 projected to increase of fish and macroinvertebrate biomass between 5 and 22% (Moullec et al., 2019).
40 Decreasing net primary production will impact higher trophic levels (Section 13.5.1), e.g., in TEUS (Holt et
41 al., 2016; Holt et al., 2018). Marine animal biomass is projected to likely decline in most European waters,
42 with decreases < 10% under all scenarios until the 2030s but losses growing to 25% at 2癈 GWL and 50% at
43 4癈 GWL in coastal waters of the NE Atlantic (Lotze et al., 2019; Bryndum-Buchholz et al., 2020).

44

45 Ocean acidification and its biological and ecological risks are projected to rise in European waters by
46 impeding growth and reproductive success of vulnerable calcifying organisms (Figure 13.11) (medium
47 confidence). Coralline algae are projected to reduce skeletal performance at 3癈 GWL, with negative
48 consequences for habitat formation (Ragazzola et al., 2016) (medium confidence). Regionally (Brodie et al.,
49 2014), differences in species-specific vulnerability will result in community shifts from calcifying
50 macroalgae (Ragazzola et al., 2013) (medium confidence) to non-calcifying macroalgae (Gordillo et al.,
51 2016) (high confidence). Experimental studies demonstrated high resilience of some important habitat
52 formers, such as the deep-water coral Lophelia pertusa (Wall et al., 2015; Morato et al., 2020), and habitat
53 engineers, such as Mediterranean limpets (Langer et al., 2014), . facilitated by energy reallocation. However,
54 if not supported by sufficient food availability (Thomsen et al., 2013; Clements and Darrow, 2018), such
55 energy reallocation will negatively impact growth or reproduction (medium confidence) (B黶cher et al.,
56 2017). (Thomsen et al., 2013)This suggests that acidification risks will be amplified by increased

Do Not Cite, Quote or Distribute  13-30       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 stratification and reduced primary production (medium confidence). The emergence of harmful algal blooms
 2 and pathogens at higher GWLs is unclear across all European seas (Figure 13.11) (low confidence).

 3

 4 Risks to marine biotas and ecosystems in European seas are projected to impact important ecosystem
 5 services (Figure 13.11). Elevated CO2 levels predicted at 4癈 GWL will affect the C/N ratio of organic-
 6 matter export and, hence, the efficiency of the biological pump (low confidence), depending on the shifts in
 7 plankton composition and, hence, food-web structure (Taucher et al., 2020). Atlantic herring (Clupea
 8 harengus) will benefit with enhanced larval growth and survival from indirect food-web effects (Sswat et al.,
 9 2018a), whereas Atlantic cod (Gadus morhua) will face overall negative impacts (medium confidence)
10 (Section 13.5, Stiasny et al., 2018; Stiasny et al., 2019). Anoxic dead zones in the Black (Altieri and Gedan,
11 2015) and the Baltic (Jokinen et al., 2018; Reusch et al., 2018) Seas are projected to increase, e.g., by 5% in
12 the Baltic Sea at 4癈 GWL (Saraiva et al., 2019). Europe's coastal vegetated 'blue-carbon' ecosystems
13 (subtidal seagrass meadows and intertidal salt marshes) are highly vulnerable (Spencer et al., 2016; Schuerch
14 et al., 2018; Spivak et al., 2019), particularly in microtidal areas such as the Baltic and Mediterranean coast.
15 Losses are projected for Posidonia oceanica seagrass habitats in the Mediterranean by of up to 75% at 2.5癈
16 GWL (low confidence) (Chapter 3). The Wadden Sea, the world's largest system of intertidal flats, is
17 projected to reduce in surface area and height, as the sediment transport capacity limits the possibility of
18 growth with rapidly rising sea levels (Wang et al., 2018; Jiang et al., 2020). For the Dutch Wadden Sea, the
19 critical rate of 6 to 10 mm yr-1, at which intertidal flats will start to `drown', will be reached by 2030 at 1.5癈
20 GWL (medium confidence), or even earlier through subsidence due to human activities (van der Spek, 2018).
21 European coastal zones provided a total of 494 billion of ecosystem services in 2018, and 4.2 to 5.1% of
22 this value will be lost due to coastal erosion by 2100 at 2.5癈 and 4.6癈 GWL, respectively (medium
23 confidence) (Paprotny et al., 2021).

24

25 13.4.2 Solution Space and Adaptation Options

26

27 Human adaptation options for marine systems encompass socio-institutional adaptation, technology, and
28 measures supporting autonomous adaptation (Chapter 3). Integrated Coastal Zone Management (ICZM) and
29 Marine Spatial Planning (MSP) are frameworks for addressing climate-change adaptation needs, as well as
30 operationalizing and enforcing marine conservation. However, ICZM and MSP do commonly not explicitly
31 take climate-change adaptation into consideration (Elliott et al., 2015), Transboundary ICZM and/or MSP
32 (Gormley et al., 2015) will become even more important with the projected acceleration of range extensions
33 and ecological regime shifts due to climate change (IPCC, 2019).

34

35 Many climate-change adaptation governance and implementation measures are embedded in international
36 strategies, such as HELCOM (Baltic Marine Environment Protection Commission (HELCOM) (Backer et
37 al., 2010), OSPAR (Convention for the Protection of the Marine Environment of the North-East Atlantic)
38 (OSPAR, 2009), and the Marine Strategy Framework Directive (MFSD) and European Water Framework
39 Directive (EWFD) of the European Union. In the Russian Arctic, mainly the Barents Sea, conservation
40 priority areas (CPA) have been identified as Ecologically and Biologically Significant Areas (EBSA)
41 (Solovyev et al., 2017). However, plans are generally at a relatively early stage (Miller et al., 2018), and
42 assessments of the effectiveness of these policy frameworks to accelerate climate-change adaptation are
43 ongoing (Haasnoot et al., 2020a).

44

45 `Green' adaptations, either `Ecosystem based Adaptations' or `Nature based Solutions, are part of adaptive
46 management strategies (European Comission, 2011) that facilitate coastal flood protection (Section 13.2.2;
47 Chapter 3; CCC SLR) and generate benefits beyond habitat creation (medium confidence), e.g., from avoided
48 expenditures for flood defence infrastructure and avoided loss of the built assets (Gedan et al., 2010). Marine
49 Protected Areas (MPAs) have been identified as adaptation options for natural areas, including permitted and
50 non-permitted uses (Chapter 3, Selig et al., 2014; Hopkins et al., 2016a; Roberts et al., 2017). The extent of
51 MPAs has been increasing in Europe, albeit with strong regional variations (Figure 13.12). MPAs provide
52 protection from local stressors, such as commercial exploitation, and enhance the resilience of marine and
53 coastal ecosystems and thus lessen the impacts of climate change (medium confidence) (Narayan et al., 2016;
54 Roberts et al., 2017). However, climate change risk reduction is only a limited MPA objective (Hopkins et
55 al., 2016b; Rilov et al., 2019). The implementation of the legal frameworks, such as the EC Habitats
56 Directive and EC Birds Directive, allows for enabling adaptation (Verschuuren, 2015) as does the
57 incorporation of climate considerations in management of Natura 2000 sites (European Comission, 2013b).

Do Not Cite, Quote or Distribute  13-31       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 There is evidence that better international cooperation is required to increase effectiveness of the MSFD
 2 (Cavallo et al., 2019), and the Good Environmental Status is currently not effectively monitored (Machado et
 3 al., 2019).

 4

 5 The greatest benefits are obtained from large, long established, no-take MPAs (Edgar et al., 2014). Yet most
 6 MPAs in Europe are partially protected or multi-use areas, and existing no-take areas tend to be very small
 7 (< 50 km2). No take areas are accounting in total for less than 0.4% of the area European waters (Figure
 8 13.12) and are often nested within multi-use MPAs. In some partially protected MPAs, local stressors, such
 9 as fishing, are higher than adjacent unprotected areas (medium confidence) (Zupan et al., 2018a; Mazaris et
10 al., 2019). Despite evidence for climate mitigation benefits of no-take zones (Roberts et al., 2017), the
11 efficacy of partial protected MPAs is debated and dependent on local management (Zupan et al., 2018b).
12 MPAs of all types require effective management to contribute to mitigating climate change impacts,
13 including effective monitoring and enforcement (Watson et al., 2014), yet the management effectiveness of
14 European MPAs has repeatedly been called into question (Batista and Cabral, 2016; Amengual and Alvarez-
15 Berastegui, 2018; Fraschetti et al., 2018; Rilov et al., 2019). Many MPAs lack management plans, and
16 insufficient resources are frequently an issue (羖varez-Fern醤dez et al., 2017; Sch閞� et al., 2020). Thus,
17 whilst substantial in potential, the current capacity of the European MPA network to reduce climate change
18 impacts is limited (Jones et al., 2016; Claudet et al., 2020).

19

20

21

22 Figure 13.12: Marine Protected Areas (MPA) in European seas. Proportions of designated and proposed MPAs in the
23 total areas of northern (NEUS), temperate (TEUS) and southern (SEUS) European seas, as well as the shares of no-take,
24 partial, unimplemented and unknown protection levels of designated MPAs (Marine Conservation Institute, 2021).
25 Moreover, the average increase of surface-sea temperatures (SST) at 4.0癈 GWL by 2100 in NEUS, TEUS and SEUS
26 is indicated.

27

28

29 Conservation approaches (MPAs, climate refugia), habitat restoration efforts (Bekkby et al., 2020), and
30 further ecosystem-based management policies do support alleviation of or adaptation to climate-change

Do Not Cite, Quote or Distribute  13-32       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 impacts (medium confidence) but are themselves impacted by climate change (Chapter 3). Moreover, the
 2 interaction of adaptation and mitigation measures poses risks to marine systems. Many coastal regions of the
 3 North Sea, especially in the south, are particularly susceptible to rising sea levels because of the strong tidal
 4 regime and the effects of storm surges (Figure 13.3). Hard measures to protect human infrastructure against
 5 sea level rise (Section 13.2) will lead to loss of coastal habitats, with negative impacts on marine biodiversity
 6 (Cross-Chapter Box SLR in Chapter 3, Airoldi and Beck, 2007; Cooper et al., 2016). While rising sea levels
 7 will also directly threaten intertidal and beach ecosystems, coastal wetlands will benefit (medium
 8 confidence), in case lateral accommodation space and the opportunity for systems to migrate landwards and
 9 upwards is provided, enhancing their ability to capture and store carbon (Rogers et al., 2019) (WGIII AR6
10 Chapter 4). In general, European coastal blue-carbon ecosystems, e.g. seagrass meadows, kelp forests, tidal
11 marshes (Bekkby et al., 2020) are potentially effective as carbon sinks in climate mitigation, akin to
12 reforestration efforts on land (section 13.3). However, their expansion has the potential to interfere with
13 other ecosystem services (Cadier et al., 2020) and biodiversity conservation (Howard et al., 2017; Chausson
14 et al., 2020). The `Blue Growth' strategy of the European Commission with the aim to increase offshore
15 activities (European Comission, 2012) will increase the pressures on the marine environments (medium
16 confidence). Large-scale offshore wind-park infrastructure is currently developed in European seas, mostly
17 in the North Sea (WindEuropeBusinessIntelligence, 2019), as a major component of climate-change
18 mitigation efforts (WGIII AR6 Chapter 6). The introduction of novel hard-substrate intertidal habitats has
19 and will have profound ecological ramifications for marine systems, including hydrodynamic changes,
20 stepping-stones for non-native species, noise and vibration, and changes of the food web (Lindeboom et al.,
21 2011; De Mesel et al., 2015; Gill et al., 2018; Dannheim et al., 2019) (high confidence).

22

23 13.4.3 Knowledge Gaps

24

25 Major knowledge gaps are uncertainties and shortcomings in our understanding of combined, cascading and
26 interacting impacts of climatic and non-climatic pressures on European marine and coastal socio-ecological
27 systems (Korpinen et al., 2021). Further observational, experimental, and modelling work will enhance the
28 insight into multiple drivers, processes and their interactions, strengthen the confidence of risk projections
29 and provide a foundation for future adaptation actions.

30

31 There is limited knowledge about the connectivity among populations, species, and ecosystems which would
32 provide new recruits, enable gene flow in Marine Protected Areas (MPA) networks (Dubois et al., 2016;
33 Sahyoun et al., 2016), and facilitate assisted migration. MPAs cover a wide range of protection status with
34 limited evidence which level of protection and connectivity is needed to achieve adaptations goals in
35 response to future warming.

36

37 Although European seas and coasts are comparatively well-studied on a global scale, the spatial and
38 temporal resolution and coverage of open-access data is still limited in many regions, particularly in eastern
39 Europe. The detection and attribution of ongoing or emerging environmental and biological changes is
40 therefore limited. Some efforts are in place, such as the six 'Sea-basin Checkpoints' (North Sea,
41 Mediterranean Sea, Arctic, Atlantic, Baltic, Black Sea) that were established since 2013 under The European
42 Marine Observation and Data Network, but high quality observations of key ocean characteristics at the level
43 of regional sea-basins are still too scarce to support decision-making for marine adaptation (M韌uez et al.,
44 2019).

45

46

47 13.5 Food, Fibre, and Other Ecosystem Products

48

49 13.5.1 Observed Impacts and Projected Risks

50

51 13.5.1.1 Crop Production

52

53 Agriculture is the primary user of land in Europe. In 2013, Europe provided 28% of cereals, 59% of sugar
54 beet and 60% of wine produced globally, as well as being part of a globalized food system with a third of the
55 commodities produced and consumed in Europe traded internationally (FAOSTAT, 2019).

56

Do Not Cite, Quote or Distribute  13-33       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 Observed climate change has led to a northward movement of agro-climatic zones in Europe and earlier
 2 onset of the growing season (Ceglar et al., 2019) (high confidence). Warming and precipitation changes since
 3 1990 explain continent-wide reductions in yield of wheat and barley and increases in maize and sugar beet
 4 (high confidence) (Fontana et al., 2015; Moore and Lobell, 2015; Ray et al., 2015; Ceglar et al., 2017). Heat
 5 stress has increased in southern Europe in spring, in summer throughout central and southern Europe, and
 6 recently expanded into the southern boreal zone (Fontana et al., 2015; Ceglar et al., 2019). Drought,
 7 excessive rain, and the compound hazards of drought and heat (13.2.1, 13.3.1, 13.10.2) increased costs and
 8 cause economic losses in forest productivity (Schuldt et al., 2020) and annual and permanent crops and
 9 livestock farming (Stahl et al., 2016), including losses in wheat production in the EU (van der Velde et al.,
10 2018) and EEU (Ivanov et al., 2016; Loboda et al., 2017) (high confidence), with the severity of impacts
11 from extreme heat and drought tripling over last 50 years (Br醩 et al., 2021). Meteorological extremes due to
12 compound effects of cold winters, excessive autumn and spring precipitation, and summer drought caused
13 production losses (up to 30% relative to trend expectations) in 2012, 2016, 2018 (Ben-Ari et al., 2018; van
14 der Velde et al., 2018; Zscheischler et al., 2018; Toreti et al., 2019b) that were exceptional compared to
15 recent decades (Webber et al., 2020). Regionally, warming caused increases in yields of field grown fruiting
16 vegetables, decreases in root vegetables, tomatoes and cucumbers (Potopov� et al., 2017) and earlier
17 flowering of olive trees (Garcia-Mozo et al., 2015) (high confidence). Delayed harvest due to both wet
18 conditions and earlier harvests in central Europe in response to warming impacted wine quality (Cook and
19 Wolkovich, 2016; van Leeuwen and Darriet, 2016; Di Lena et al., 2019).

20

21 Evidence for growing regional differences of projected climate risks is increasing since AR5 (high
22 confidence). While there is high agreement of the direction of change, the absolute yield losses are uncertain
23 due to differences in model parameterization and whether adaptation options are represented (high
24 confidence) (Donatelli et al., 2015; Moore and Lobell, 2015; Knox et al., 2016; Webber et al., 2018). At
25 1.5癈 GWL, compound events which led to recent large wheat losses are projected to become 12% more
26 frequent (Ben-Ari et al., 2018). Growing regions will shift northward or expand for melons (Bisbis et al.,
27 2019), tomatoes and grapevines reaching NEU and EEU in 2050 under 1.5癈 GWL (Hannah et al., 2013;
28 Litskas et al., 2019) (high confidence), while warming would increase yields of onions, Chinese cabbage and
29 French beans (Bisbis et al., 2019) (medium confidence). In response to 2癈 GWL, agro-climatic zones in
30 Europe are expected to move northward 25-135 km/decade, fastest in EEU (Ceglar et al., 2019). Negative
31 impacts of warming and drought are counterbalanced by CO2 fertilization for crops such as winter wheat
32 (medium confidence, medium agreement), resulting in some regional yield increases with climate change
33 (Zhao et al., 2017; Webber et al., 2018).

34

35 Reductions in agricultural yields will be higher in the south at 4癈 GWL, with lower losses or gains in the
36 north (Figure 13.5, Trnka et al., 2014; Webber et al., 2016; Szewczyk et al., 2018) (high confidence). Largest
37 impacts of warming are projected for maize in SEU (Deryng et al., 2014; Knox et al., 2016) (high
38 confidence) with yield losses across Europe of 10-25% at 1.5-2癈 GWL and 50-100% at 4癈 GWL (Deryng
39 et al., 2014; Webber et al., 2018; Feyen et al., 2020).

40

41 Use of longer season varieties can compensate for heat stress on maize in WCE and lead to yield increases
42 for Northern Europe, but not SEU for 4癈 GWL (Siebert et al., 2017; Ceglar et al., 2019) (medium
43 confidence). Irrigation can reduces projected heat and drought stress, e.g., for wheat and maize (Ruiz-Ramos
44 et al., 2018; Feyen et al., 2020), but use is limited by water availability (13.2.1; KR3). The advantages of a
45 longer growing season in NEU and EEU are outbalanced by the increased risk of early spring and summer
46 heat waves (Ceglar et al., 2019).

47

48 Warming causes range expansion and alters host pathogen association of pests, diseases and weeds affecting
49 health for European crops (Caffarra et al., 2012; Pushnya and Shirinyan, 2015; Latchininsky, 2017) (high
50 confidence) with high risk for contamination of cereals (Moretti et al., 2019). Regionally predicted reduction
51 in rainfall (13.1) can lead to carryover of herbicides (Karkanis et al., 2018).

52

53 Net yield losses will reduce economic output in the EU from agriculture, reaching a reduction of 7% for the
54 EU and UK combined, and 10% in SEU at 4癈 GWL (Naumann et al., 2021). Farmland values are projected
55 to decrease by 5-9% per degree of warming in SEU (Van Passel et al., 2017). Increased heat and drought
56 stress and reduced irrigation water availability will decreased profitability and cause abandonment of
57 farmland in SEU (Holman et al., 2017) (limited evidence, low confidence).

Do Not Cite, Quote or Distribute  13-34       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 13.5.1.2 Livestock Production

 3

 4 Heat and humidity affect livestock directly exposed in open barns and outdoors, such as dairy cows and
 5 goats (Gauly et al., 2013; Bernabucci et al., 2014; Silanikove and Koluman, 2015), and cold adapted
 6 husbandry (Box 13.2, Section 13.8.3) (high confidence). Heat impacts animal health (Sanker et al., 2013;
 7 Lambertz et al., 2014), nutrition, behaviour and welfare (Heinicke et al., 2019), performance and product
 8 quality (Gauly and Ammer, 2020). Climate change also impacts grassland production, fodder composition
 9 and quality, particularly in SEU (Dumont et al., 2015) and EEU (Bezuglova et al., 2020), as well as altering
10 the prevalence, distribution and load of pathogens and their vectors (2.4.2.7.3) (Morgan et al., 2013; Charlier
11 et al., 2016) (high confidence). Projected impacts on poultry and pigs are low due to temperature control in
12 large parts of Europe, but greater in SEU where open systems prevail (Chapter 5).

13

14 Warming increases the pasture growing season and farming period in NEU and at higher altitudes (Fuhrer et
15 al., 2014), while longer drought periods and thunderstorms can influence abandonment of remote Alpine
16 pastures, reducing cultural and landscape ecosystem services and losing traditional farming practices
17 (Herzog and Seidl, 2018) (high confidence) (Section 13.8.3). At 2-4癈 GWL grassland biomass production
18 for forage-fed animals will increase in NEU and the northern Alps, while forage production will decrease in
19 SEU and the southern Alps due to heat and water scarcity (Gauly et al., 2013; J鋑er et al., 2020), causing
20 regional reductions of cow milk production in WCE and SEU (Silanikove and Koluman, 2015) (high
21 confidence).

22

23 13.5.1.3 Aquatic Food Production

24

25 Seafood production in Europe provides jobs for >250,000 people, predominantly in SEU (Carvalho et al.,
26 2017). Marine fisheries contribute 80% to European aquatic food production, while marine aquaculture
27 provides 18% and freshwater production 3% (Blanchet et al., 2019). The Russian Federation provides 1/4 of
28 seafood production in Europe (FAOSTAT, 2019).

29

30 Climate change has impacted European marine food production (high confidence). However, extraction is
31 still the major impact on commercially important fish stocks in Europe (Mullon et al., 2016), with 69% of
32 stocks overfished and 51% outside safe biological limits (Froese et al., 2018). The North Sea, the Iberian
33 coastal Sea and Celtic Sea-Biscay Shelf are globally among the areas most negatively affected by warming
34 with losses of 15-35% in maximum sustainable yields (MSY) during the last decades (Free et al., 2019).
35 Warming caused ongoing northward movement and range expansion of Northeast Atlantic fish stocks (13.4,
36 Baudron et al., 2020). In the North Sea, cuttlefish (van der Kooij et al., 2016; Oesterwind et al., 2020) and
37 tuna (Bennema, 2018; Faillettaz et al., 2019) became new target species (medium confidence). In SEU,
38 warm-water species increasingly dominate fisheries landings (Fortibuoni et al., 2015; Teixeira et al., 2016;
39 Vasilakopoulos et al., 2017).

40

41 European countries are assessed to be globally among the least vulnerable to the impacts of climate change
42 on fisheries-related food security risks (high confidence) due to low levels of exposure to climate hazards,
43 low dependency of economies on fisheries and a high adaptive capacity (Barange et al., 2014; Ding et al.,
44 2017). European freshwater production is suggested to be less vulnerable than marine sectors and marine
45 production vulnerability increases with latitude (Blanchet et al., 2019). In the aquaculture sector Norway is
46 highly vulnerable due to high sensitivity of salmon farming to warming and high per-capita production
47 (Handisyde et al., 2017). In the fisheries sector, vulnerability for fishing communities is highest in SEU and
48 UK (Figure 13.9A , Handisyde et al., 2017; Payne et al., 2021), while for aquaculture sectors it is highest in
49 SEU and some NEU and WCE countries (Figure 13.9B, 2020).

50

51

Do Not Cite, Quote or Distribute  13-35       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.13: Future vulnerability and risks for aquatic food production. (a) Vulnerability for fisheries in 105 coastal
 3 regions across 26 countries based on biological traits and physiological metrics of 556 resource populations (Payne et
 4 al., 2021); (b) Vulnerability of major aquaculture species in European countries on physiological attributes, farming
 5 methods and economic output (Peck et al., 2020); (c-d) Differences (%) between projected changes for 1.5癈 and 4癈
 6 global warming (Peck et al., 2020), with (c) changes in abundance of major fish species by region, and (d) changes in
 7 productivity of major aquaculture species by country.

 8

 9

10 Future vulnerabilities, risks and opportunities are projected to strongly vary regionally and between major
11 fisheries and aquaculture species (Peck et al., 2020) (Figure 13.13 c,d). Assuming MSY-management,
12 projections suggest reduced abundance of most commercial fish stocks in European waters of 35% (up to
13 90% for individual stocks) between 1.5癈 and 4.0癈 GWL (Peck et al., 2020; Payne et al., 2021) (medium
14 confidence) (Figure 13.13). In response to 4癈 GWL, higher trophic level biomass is projected to increase in
15 the SEUS mainly due to increases of small pelagic and thermophilic, often exotic species (Moullec et al.,
16 2019).

17

18 Ocean acidification (Section 13.4, Chapter 4) will develop into a major risk for marine food production in
19 Europe under 4癈 GWL (high confidence), affecting recruitment of important European fish stocks such as
20 those of cod in the Western Baltic and Barents Sea by 8 and 24%, respectively (Sswat et al., 2018b; Stiasny
21 et al., 2018; Voss et al., 2019). Acidification is also projected to negatively affect marine shellfish production
22 and aquaculture in Europe with 4癈 GWL (medium confidence) (Fernandes et al., 2017; Narita and Rehdanz,
23 2017; Mangi et al., 2018).

24

25 13.5.1.4 Forestry and Forest Products

26

27 Climate change is altering the structure and function of European forests via changes in temperature,
28 precipitation and atmospheric CO2 as well as through interaction with pests and fire (13.3.1) (Moreno et al.,
29 2018; Morin et al., 2018; Senf et al., 2018; Orlova-Bienkowskaja et al., 2020) (high confidence). Species-
30 specific responses of trees to drier summers (Vitali et al., 2018) shape regional variability in European forest
31 productivity in response to water and nutrient availability, heat wave and evaporative demand (Reyer et al.,
32 2014; Kellom鋕i et al., 2018). While warming and extended growing seasons have positive impacts on forest

Do Not Cite, Quote or Distribute  13-36       Total pages: 143
    FINAL DRAFT                                                                    Chapter 13                       IPCC WGII Sixth Assessment Report

 1 growth in cold areas in WCE and NEU (Pretzsch et al., 2014; Matskovsky et al., 2020), EEU (Tei et al.,
 2 2017) and higher altitude (Sedm醟ov� et al., 2019), drought stress across Europe has been increasing
 3 (Primicia et al., 2015; Marqu閟 et al., 2018; Ruiz-P閞ez and Vico, 2020) (high confidence). Combined with
 4 land-use, climate change has increased large scale forest mortality since the 1980s (Senf et al., 2018).
 5 Extreme events such as the 2018 drought in WCE caused widespread leaf shedding and mortality of trees
 6 (Buras et al., 2020) with carryovers into 2019 (Schuldt et al., 2020) and bark beetle outbreaks (Netherer et
 7 al., 2019) resulting in felling and cuttings of more than 1 Million ha of spruce forest and disrupting timber
 8 markets (Mauser, 2021).

 9

10 In response to 3癈 GWL, forest productivity is projected to increase in NEU and altitudes, show mix trends
11 in WCE and decreases in SEU (Reyer et al., 2014) (medium confidence). This trend is driven by increases in
12 productivity of pine and spruce and decreases of beech and oak and excludes disturbances and management
13 options (Reyer et al., 2014). Water stress exacerbates the incidence from and effects of fire and other natural
14 disturbances (13.3.1), resulting in forest productivity declines or cancelling out productivity gains from CO2
15 (Seidl et al., 2014; Reyer et al., 2017) (high confidence). In response to 1.7 癈 GLW, managed forest and
16 unmanaged woodland areas are projected to decrease only minimally, while at >2.5癈 GLW declines
17 increasing for managed forest and unmanaged woodland area increases (Harrison et al., 2019). Reducing
18 warming from 4癈 GLW to below 1.7 癈 GLW would reduce the Europe wide impacts on managed forest
19 by 34% (Harrison et al., 2019).

20

21 13.5.2 Solution Space and Adaptation Options

22

23 The solution space for climate change adaption for food and timber includes production related options
24 (13.5.2.1 - 13.5.2.3) and market-based changes to consumer demand and trade (13.5.2.4). The assessment of
25 effectiveness and feasibility of options in the food system is summarised in Figure 13.14.

26

27
           Effectiveness & feasibility of adaptation options for food system to climate impacts & risk in Europe

    Impact Type  Adaptation Option                      Effectiveness Feasibility                                                                       Confidence

                                                            Economic               Technological Institutional  Socio-cultural Ecological  Geophysical  Evidence Agreement
                                                                                                                                                  L
    Heat stress  Irrigation                             M   M                      H           M                L   L                             H     M           M
                                                                                                                                                  H
                 Change of sowing/harvest date          M   H                      H           NL               M   M                             L     H           M
                                                                                                                                                  H
                 Change of cultivars                    L   M                      M           NL               M   M                             H     M           M
                                                                                                                                                 M
    Drought      Irrigation                             H   H                      M           M                H   L                             L     H           H
                                                                                                                                                 M
                 Change of sowing/harvest date          M   H                      H           NL               M   M                                   M           M

                 Change of cultivars                    L   M                      M           NL               M   M                                   H           M

                 Soil management                        M   M                      M           H                M   H                                   L           M

    Flooding     Change of sowing/harvest date          L   L                      M           NL               H   M                                   L           M

    Compound &   Plant & livestock breeding,            M   M                      L           L                M   M                                   M           M

    extreme weather including GMO

                 Mixed use - agroecology &              H   M                      M           L                L   H                      M            M           M

                 agroforestry

                 Agricultural policy changes            M   M                      M           M                M   M                      H            L           H

                 Training & information                 L   M                      NL          M                M   M                      H            L           M

                 Crop selection changes                 M   H                      H           NL               L   L                      L            L           L

                 Land cover change, incl. agricultural  L   M                      M           L                L   L                      L            L           L

                 land abandonment

    Disease pathogen Plant & livestock breeding,        NL  NL                     L           L                L   NL                     NL           L           NL

    & vectors    including GMO

                 Management, including high             NL  NL                     NL          NL               NL  NL                     NL           L           NL

                 frequency rotations

    Combined impacts International trade changes        M   M                      NL          L                M   L                      M            L           M

    on productivity Consumer shifts in consumption      NL  M                      NL          NL               L   NL                     NL           L           M

    Legend

     High = H

     Medium = M

     Low = L

28  No/Limited Evidence = NL

29 Figure 13.14: Effectiveness and feasibility of the main adaptation options for food systems in Europe. (SM13.9 and

30 Table SM13.5).

31

32

33 13.5.2.1 Crops and Livestock

34

    Do Not Cite, Quote or Distribute                                               13-37                                                                Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 Farm management adaptations options to climate change include changing sowing and harvest dates,
 2 changes in cultivars, irrigation and selecting alternative crops (Figure 13.14, 13.15) (Donatelli et al., 2015).
 3 Irrigation is effective at reducing yield loss from heat stress and drought, e.g., for wheat and maize (Figure
 4 13.14, 13.15), but increases demand for water withdrawals (Siebert et al., 2017; Ruiz-Ramos et al., 2018;
 5 Feyen et al., 2020). Where sufficient water and infrastructure is available, irrigation of wheat reverses yield
 6 losses across Europe at 2 GWL to become gains, while yield losses in maize in SEU are reduced from up
 7 to 80% to 11% (Feyen et al., 2020). Extensive droughts during the last two decades caused many irrigated
 8 systems in southern Europe to cease production (Stahl et al., 2016) indicating limited adaptive capacity to
 9 heat and drought (medium confidence). Water management for food production on land is becoming
10 increasingly complex due to the need to satisfy other social and environmental water demands (KR3, 13.10)
11 and is limited by costs and institutional coordination (Iglesias and Garrote, 2015). Agricultural water
12 management adaptation practices include irrigation, reallocating of water to other crops, improving use
13 efficiency, and soil water conservation practices (Iglesias and Garrote, 2015). In-season forecasts of climate
14 impacts on yield have successfully been used in the 2018 drought for European wheat (van der Velde et al.,
15 2018).

16

17

18

19 Figure 13.15: Projected yield changes with climate change for 1.5 癈 GWL (RCP 2.6), 1.7 癈 (RCP4.5) and 2癈 GWL

20 (RCP8.5) and altered crop management, and associated water demand, showing: (a) relative yield changes under

21 climate change and elevated CO2 for current production systems, i.e. rainfed and irrigated simulations weighted by

Do Not Cite, Quote or Distribute  13-38       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 current share of rainfed and irrigated areas; (b) yield increase if current predominantly rainfed areas are full irrigated;
 2 (c) additional yield increases for irrigated production systems if new varieties used to avoid losses associated with faster
 3 development and earlier maturity under climate change; and, (d) water demand for irrigated systems with current
 4 varieties in currently rainfed areas (Webber et al., 2018). Relative yield changes to a period centred on 2055 to a
 5 baseline period centred on 1995. Boxplots are Europe aggregate results considering current production areas (a) or
 6 current rainfed areas (b, c) with boxplots showing uncertainty across crop models and GCMs. The maps are for the crop
 7 model median for RCP 4.5 (1.7癈 GWL) with GFDL-CM3.

 8

 9

10 Changes to cultivars and sowing dates can reduce yield losses (Figure 13.15), but are insufficient to fully
11 ameliorate losses projected >3 GWL, with an increase of risk from north to south and for crops growing
12 later in the season such as maize and wheat (high confidence) (Ruiz-Ramos et al., 2018; Feyen et al., 2020).
13 Adaptations for early maturing reduce yield loss by moving the cycle towards a cooler part of year, and also
14 constrains the increases in irrigation water demands, but reduce the period for photosynthesis and grain
15 filling (high confidence) (Ruiz-Ramos et al., 2018; Holzk鋗per, 2020). Crop breeding for drought and heat
16 tolerance can improve sustainability of agricultural production under future climate (Costa et al., 2019),
17 particularly in SEU where drought tolerant varieties provide 30% higher yields than drought-sensitive
18 varieties at 3癈 GWL (Senapati et al., 2019). Soil management practices, such as crop residue retention or
19 improved crop rotations, generally undertaken as a mitigation option to increase soil carbon sequestration,
20 are not commonly evaluated for adaptation in European agriculture (Hamidov et al., 2018).

21

22 Adaptation practices for livestock systems on European farms commonly focus on controlling cooling, shade
23 provision and management of feeding times (Gauly et al., 2013). These options are used in indoors reared
24 species (Gauly et al., 2013), but limited in mountain pastures (Del間lise et al., 2019) (high confidence).
25 Response options to insufficient amount and quality of fodder include changing feeding strategies (Kaufman
26 et al., 2017; Ammer et al., 2018), feed additives (Ghizzi et al., 2018), relocating livestock linked to improved
27 pasture management, organic farming (Rojas-Downing et al., 2017; EEA, 2019c), importing fodder and
28 reducing stock (Toreti et al., 2019b). Dairy systems that maximize the use of grazed pasture are considered
29 more environmentally sustainable, but are not fully supported by policy and markets (Hennessy et al., 2020)
30 (medium confidence). Genetic adaptation of crops, pasture and animals could be a long-term adaptation
31 strategy (Anzures-Olvera et al., 2019; Del間lise et al., 2019). Control strategies for pathogens and vectors
32 include indoor or outdoor rearing and applying new diagnostic tools or drugs (Bett et al., 2017; Vercruysse et
33 al., 2018), and regulations to ensure safe trade and reduce risks of introducing or spreading pests (European
34 Comission, 2016 ).

35

36 Agro-ecological systems provide adaptation options that rely on ecological process (e.g., soil organic matter
37 recycling and functional diversification) to lower inputs without impacting productivity (Cross-Chapter Box
38 NATURAL in Chapter 2 , Aguilera et al., 2020). High frequency rotational grazing and mixed livestock
39 systems are agro-ecological strategies to control pathogens (Aguilera et al., 2020). Agroforestry, integrating
40 trees with crops (silvoarable), livestock (silvopasture), or both (agrosilvopasture) can enhance resilience to
41 climate change (Chapter 5), but implementation in Europe needs improved training programs and policy
42 support (Hern醤dez-Morcillo et al., 2018) (high confidence).

43

44 Technological innovations including "smart farming" and knowledge training can strengthen farmers'
45 responses to climate impacts (Del間lise et al., 2019; Kernecker et al., 2019), although strong belief in
46 "technosalvation" by farmers (Ricart et al., 2019) can reduce the solution space and timing of adaptation
47 options. Agricultural policy, market prices, new technology and socio-economic factors play a more
48 important role in short-term farm-level investment decisions than climate change impacts (Juhola et al.,
49 2016; Hamidov et al., 2018) (high confidence).

50

51 Effective policy guidance is needed to increase the climate-resilience of agriculture (Spinoni et al., 2018;
52 Toreti et al., 2019b). Financial measures include simplifying procedures for obtaining subsidies and
53 insurance premiums and interest rates that incentivise adoption of climate friendly agricultural methods
54 (Garrote et al., 2015; Iglesias and Garrote, 2015; Zakharov and Sharipova, 2017; Hamidov et al., 2018;
55 Wir閔n, 2018). The EU's Common Agricultural Policy has increasingly focused on environmental outcomes
56 (AllianceEnvironnement, 2018) but does not sufficiently provide for adaptation measures (Leventon et al.,
57 2017; Pe'er et al., 2020). Limits to European farm-level adaptation include lack of resources for investment,

Do Not Cite, Quote or Distribute  13-39       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 political urgency to adapt, institutional capacity, access to adaptation knowledge and information from other
 2 countries (EEA, 2019c).

 3

 4 13.5.2.2 Aquatic Food

 5

 6 Climate-resilient fish production in Europe is the goal of the European Union's Common Fisheries Policy
 7 (CFP) rebuilding fish stocks to maximum sustainable yield (MSY) levels, but success has been variable
 8 (Froese et al., 2018; Stecf, 2019). Adaptation is largely ignored in related EU policy frameworks such as the
 9 CFP, the Marine Strategy Framework Directive, and the "Strategic guidelines for the sustainable
10 development of EU aquaculture." (Pham et al., 2021). A major governance challenge for adaptation will be
11 the redistribution of the fixed allocation scheme for total allowable catches (Harte et al., 2019; Baudron et
12 al., 2020). Inflexible and non-adaptive allocation schemes can result in conflicts among European countries
13 (medium confidence), as demonstrated by the case of the North East Atlantic mackerel (Spijkers and
14 Boonstra, 2017).

15

16 The development of adaptation strategies for seafood production since the Paris Agreement is insufficient in
17 Europe (high confidence) (Kalikoski et al., 2018; Pham et al., 2021). Concrete plans for adaptation planning
18 towards climate-ready fisheries and aquaculture are lacking in all parts of Europe (European Comission,
19 2018), especially accounting for the expected reduced landings of traditional target species and in
20 preparation for a new portfolio of resource species (Blanchet et al., 2019).

21

22 Recent scientific progress towards adaptation in European fisheries and aquaculture include conceptual
23 guidance and demonstration cases on climate adaptation planning (Pham et al., 2021) and climate
24 vulnerability assessments (Blanchet et al., 2019; Peck et al., 2020; Payne et al., 2021). Socio-political
25 scenarios for European aquatic resources have been developed and have the potential to inform adaptation
26 planning by European fisheries and aquaculture sectors (Kreiss et al., 2020; Hamon et al., 2021; Pinnegar et
27 al., 2021).

28

29 13.5.2.3 Forests

30

31 Forest management has been adopted as a frequent strategy to cope with drought, reduce fire risk, and
32 maintain biodiverse landscapes and rural jobs (Hl醩ny et al., 2014; Fern醤dez-Manjarr閟 et al., 2018).
33 Successful adaptation strategies include altering the tree species composition to enhance the resilience of
34 European forests (Schelhaas et al., 2015; Zubizarreta-Gerendiain et al., 2017; Pukkala, 2018) (high
35 confidence). Greater diversity of tree species reduces vulnerability to pests and pathogens (Felton et al.,
36 2016) and increases resistance to natural disturbances (Jactel et al., 2017; Pukkala, 2018; Pardos et al., 2021)
37 (high confidence). Depending on forest successional history (Sheil and Bongers, 2020), tree composition
38 change can increase carbon sequestration (Liang et al., 2016), biodiversity and water quality (Felton et al.,
39 2016) (high confidence). Conservation areas can also help climate change adaptation by keeping the forest
40 cover intact, creating favourable microclimates and protecting biodiversity (Jantke et al., 2016) (low
41 confidence).

42

43 Reforestation reduces warming rates (Zellweger et al., 2020) and extremely warm days (Sonntag et al., 2016)
44 inside forests reducing natural disturbances and fires (high confidence). Active management approaches can
45 limit the impact of fires (13.3.1) on forest productivity, including fuel reduction management, prescribed
46 burning, changing from conifers to deciduous, less flammable species, and recreating mixed forests (Feyen
47 et al., 2020) and agroforestry (Damianidis et al., 2020).

48

49 13.5.2.4 Demand and Trade

50

51 An increasing globalized food system makes European nations sensitive to supply chain disturbances in
52 other parts of the world, but also provides capacity to adapt to production shifts within Europe through
53 changes in international trade (Section 13.9.1) (Alexander et al., 2018; Challinor et al., 2018; Ercin et al.,
54 2021). Consumer demand for food and timber products can adapt to productivity changes and be mediated
55 by price (e.g., in response to production changes or policies on food related taxation), reflect changes in
56 preferences (e.g., towards plant-based foods motivated by environmental, ethical or health concerns), or
57 reductions in food waste (Alexander et al., 2019; Willett et al., 2019) (high confidence). Although mitigation

Do Not Cite, Quote or Distribute  13-40       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 potentials of dietary changes have received increasing attention, evidence is lacking on potential for
 2 adaptation through changes in European food consumption and trade, despite these socio-economic factors
 3 being a strong driver for change (Harrison et al., 2019, Kebede, 2021, Integrated assessment of the food-
 4 water-land-ecosystems nexus in Europe: Implications for sustainability) (medium confidence). Calls are
 5 increasing across Europe for sustainable and resilient agri-food systems acknowledging interdependencies
 6 between producers and consumers to deliver healthy, safe and nutritional foods and services (Venghaus and
 7 Hake, 2018) (Section 13.7).

 8

 9 13.5.3 Knowledge Gaps

10

11 Aggregated projections of impacts, especially of combined hazards, are still rare despite many physiological
12 papers on species specific response to warming in all food sectors (high confidence). This is specifically true
13 for scenarios that consider land use change and population growth, though Agri SSPs are currently being
14 developed (Mitter et al., 2019). Effectiveness of adaptation options is predominantly qualitatively mentioned
15 but not assessed and effectiveness of combinations of measures is rarely assessed (Ewert et al., 2015;
16 Holman et al., 2018; M黮ler et al., 2020) (high confidence). Effective adaptation planning would be
17 supported by better modelling and scenario development including improved coupled nature-human
18 interactions, e.g., with more realistic representation of behaviours beyond economic rationality and `bottom-
19 up' autonomous farmer adaptations, as well as greater stakeholder involvement.

20

21 Coverage of impacts and adaptation options in Europe are biased towards the EU28 and have gaps within the
22 eastern part of WCE and EEU, despite dramatic changes in land use over the recent decades in Russia and
23 Ukraine (high confidence) which have the potential to increase production and export of agricultural
24 products, especially wheat, meat and milk (Swinnen et al., 2017).

25

26 A bias towards modelling of cereals, specifically wheat and maize, results in gaps in knowledge for fruit and
27 vegetables, especially for temperate regions in Europe (Bisbis et al., 2019). The assessment of irrigation
28 needs and the impact of CO2 and O3 tend to focus on individual species and processes hindering upscaling to
29 multiple stressors and mixed production (Challinor et al., 2016; Webber et al., 2016) (high confidence).

30

31 There is a lack of actionable adaptation strategies for European fisheries and aquaculture. Knowledge gaps
32 include adaptive capacities of local fishing communities to a new mix of target species and the consumer
33 acceptance of the product. Increased knowledge on the effects on freshwater fisheries and their resources is
34 also needed.

35

36

37 13.6 Cities, Settlements and Key Infrastructure

38

39 Urban areas in Europe offer home to 547 million inhabitants, corresponding to 74% of the total European
40 population (UN/DESA, 2018). In the EU-28, 39% of the total population lives in metropolitan regions (i.e.,
41 areas with at least one million inhabitants) where 47% of the total GDP is generated (Eurostat, 2016). Apart
42 from urban settlements, this section also covers energy and transport systems, as well as tourism, industrial
43 and business sectors which are key for livelihood, economic prosperity and well-being of residents.

44

45 13.6.1 Observed Impacts and Projected Risks

46

47 13.6.1.1 Energy Systems

48

49 The energy sector in Europe already faces impacts from climate extremes (high confidence). Significant
50 reductions and interruptions of power supply have been observed during exceptionally dry and/or hot years
51 of the recent 20-year period, e.g. in France, Germany, Switzerland and UK during the extremely hot summer
52 of 2018 which led to water cooling constraints on power plants (van Vliet et al., 2016b; Abi-Samra, 2017;
53 Vogel et al., 2019). Heating degree days decreased and cooling degree days increased during 1951-2014,
54 with clearer trends after 1980 (De Rosa et al., 2015; Spinoni et al., 2015; EEA, 2017b). Projected climate
55 risks for energy supply are summarized in Figure 13.16.

56

57

Do Not Cite, Quote or Distribute  13-41       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.16: Projected climate change risks for energy supply in Europe for major sources and under 1.5 癈, 2 癈 and
 3 >3 癈 GWL (Tables SM13.5.6-13).

 4

 5

 6 New studies reinforce the findings of AR5 on risks for thermoelectric power and regional differences
 7 between NEU and SEU regarding risks for hydropower (Figure 13.16). In NEU and EEU, extremely high
 8 water inflows to dams are projected to increase flooding risks for plant and nearby settlements (Chernet
 9 Haregewoin et al., 2014; Porfiriev et al., 2017), while increasing temperatures could reduce the efficiency of
10 steam and gas turbines (Porfiriev et al., 2017; Cronin et al., 2018; Klimenko et al., 2018a). Water scarcity
11 may limit onshore carbon capture and storage in some regions (Byers et al., 2016; Murrant et al., 2017; EEA,
12 2019a).

13

14 Reduced surface wind speeds during 1979-2016 (Frolov et al., 2014; Perevedentsev and Aukhadeev, 2014;
15 Tian et al., 2019) support projected trends in decreasing onshore wind energy potential. Seasonal changes
16 may result in reductions in many areas in summer (by 8-30% in Southern Europe) and increases in most of
17 NEU during winter. Increasing probabilities and persistence of high winds over the Aegean and Baltic Seas
18 (Weber et al., 2018a) could create new opportunities for offshore wind. The future configuration of the wind
19 fleet will affect the spatial and temporal variability of wind power production (Tobin et al., 2016). Total
20 backup energy needs in Europe could increase by 4-7% by 2100 (Wohland et al., 2017) with potentially
21 larger seasonal changes (Weber et al., 2018b).

22

23 There is low evidence and limited agreement on projections of solar power potential due to differences in the
24 integration of aerosols and the estimated cloud cover between climate models (Bartok et al., 2017; Bo� et al.,
25 2020; Guti閞rez et al., 2020). Studies on climate risks for bioenergy are also limited.

26

27 Energy demand is projected to display regional differences in response to global warming beyond 2oC GWL,
28 with a the significant southwest-to-northeast decrease of heating degree days by 2100 (particularly in
29 northern Scandinavia and Russia), and a smaller north-to-south increase of cooling degree days (Porfiriev et

Do Not Cite, Quote or Distribute  13-42       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 al., 2017; Spinoni et al., 2018; Coppola et al., 2021). Under present population numbers, total energy demand
 2 would decrease in almost all Europe, whereas it could increase in some countries (e.g. UK, Spain, Norway)
 3 when considering Eurostat's population projections (Klimenko et al., 2018b; Spinoni et al., 2018).There is
 4 medium confidence that peak load will increase in SEU and decrease in NEU (Damm et al., 2017; Wenz et
 5 al., 2017; Bird et al., 2019). Beyond 2oC GWL, a shift of peak load from winter to summer in many countries
 6 is possible (Wenz et al., 2017). Together with water-cooling constraints for thermal power, this change in
 7 load may challenge the stability of electricity networks during heatwaves (EEA, 2019a). Technological
 8 factors, increased electricity use and adaptation influence significantly the temperature sensitivity of
 9 electricity demand and consequently risks (Damm et al., 2017; Wenz et al., 2017; Cassarino et al., 2018;
10 Figueiredo et al., 2020). Potential power curtailments or outages during climatic extremes may increase
11 electricity prices (Pechan and Eisenack, 2014; Steinh鋟ser and Eisenack, 2020).

12

13 13.6.1.2 Transport

14

15 Heatwaves in 2015 and 2018 in parts of WCE and NEU caused road melting, railway asset failures, and
16 speed restrictions to reduce the likelihood of track buckling (Ferranti et al., 2018; Vogel et al., 2019). Recent
17 studies on projected risks focus mainly on infrastructure and much less on transport flows and disruptions.

18

19 Sea level rise (section 13.2) may disrupt port operations and surrounding areas, mainly in parts of NEU and
20 WCE (Christodoulou et al., 2018), while changes of waves agitation could increase non-operability hours of
21 some Mediterranean ports beyond 2癈 GWL (Sierra et al., 2016; Camus et al., 2019; Izaguirre et al., 2021).
22 Low water level days at some critical locations for inland navigation at the Rhine river are projected to
23 increase beyond 2oC GWL, while decreases at the Danube river are possible (van Slobbe et al., 2016;
24 Christodoulou et al., 2020).

25

26 Risks of rutting and blow-ups of roads (particularly in low altitudes) due to high summer temperatures are
27 expected to increase in WCE and EEU at 3oC GWL (medium confidence) (Frolov et al., 2014; Matulla et al.,
28 2018; Yakubovich and Yakubovich, 2018). In EEU and northern Scandinavia, the higher number of
29 freezing-thawing cycles of construction materials will increase risks for roads (Frolov et al., 2014;
30 Yakubovich and Yakubovich, 2018; Nilsen et al., 2021) while warming beyond 2oC GWL could
31 significantly reduce road maintenance costs in NEU (Lorentzen, 2020), but limiting off-road overland
32 transport in north-western Russia (G鋎eke et al., 2021). Beyond 3oC GWL, more frequent hourly
33 precipitation extremes are projected over WCE and NEU in summer (e.g. 2-fold and 10-fold increase for
34 events exceeding the present-day 99.99th percentile in Germany and UK) but more widely across Europe in
35 autumn and winter (increase higher than 10-fold for 99.99th percentile events in southern Europe in autumn
36 (Chan et al., 2020), potentially severely damaging roads as happened in Mandra, Greece in 2017 (Diakakis et
37 al., 2020). Landslide risks in WCE and SEU could increase beyond a 2oC GWL, threatening road networks
38 (Schlogl and Matulla, 2018; Rianna et al., 2020).

39

40 Current flood risk for railways could double or triple at 1.5-3癈 GWL, particularly in WCE, increasing
41 public expenditure for rail transport in Europe by 1.22 billion annually under 3癈 GWL and no adaptation
42 (Bubeck et al., 2019). Thermal discomfort in urban underground railways is expected to increase, even at a
43 high level of saloon cooling (Jenkins et al., 2014a).

44

45 The number of airports vulnerable to inundation from sea level rise and storm surges may double between
46 2030 and 2080 without adaptation, especially close to the North Sea and Mediterranean coasts
47 (Christodoulou and Demirel, 2018). Rising temperatures reducing lift generation could impose weight
48 restrictions for large aircraft at 2oC GWL and beyond in airports of France, UK and Spain (Coffel et al.,
49 2017). There is a lack of studies quantifying the effect of future extreme events on flight arrivals at and
50 departures from European airports.

51

52 13.6.1.3 Business and Industry

53

54 European industrial and service sectors contribute 85% to Gross Value Added in EU-28 (Eurostat, 2020);
55 while their direct exposure and vulnerability is smaller compared to sectors directly reliant on weather, they
56 are directly and indirectly affected by heat, flooding, water scarcity and drought (Weinhofer and Busch,
57 2013; Gasbarro and Pinkse, 2016; Meinel and Schule, 2018; Schiemann and Sakhel, 2018; TEG, 2019). Heat

Do Not Cite, Quote or Distribute  13-43       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 reduces the productivity of labour particularly in construction, agriculture and manufacturing (Garc韆-Le髇
 2 et al., 2021; Schleypen et al., 2021) (section 13.7.1). Direct losses from floods in Europe are highest for
 3 manufacturing, utilities, transportation; indirect losses arise e.g. for manufacturing, construction, and
 4 banking and insurance (Koks et al., 2019a; Sieg et al., 2019; Mendoza-Tinoco et al., 2020). Drought and
 5 water scarcity directly affect European industries in the sectors of pulp and paper, chemical and plastic
 6 manufacturing and food and beverages (Gasbarro et al., 2019; Teot髇io et al., 2020); additionally, drought
 7 may indirectly affect sectors relying on shipping, hydropower, or public water supply (Naumann et al.,
 8 2021). The European financial and insurance sector is affected by climate change impacts via their customers
 9 and financial markets (Bank of England, 2015; Georgopoulou et al., 2015; Battiston et al., 2017; TCFD,
10 2017; Bank of England, 2019; de Bruin et al., 2020; Monasterolo, 2020).

11

12 The vulnerability to climate hazards varies by European region, type of risk, sector and business
13 characteristics (Gasbarro et al., 2016; Forzieri et al., 2018; ECB, 2021a; Kouloukoui et al., 2021). Current
14 damages are mainly related to river floods and storms, but heat and drought will become major drivers in the
15 future (medium confidence); until 2050, the probability of default of firms located in particularly exposed
16 locations may increase to up to four times of that of an average firm in all sectors (ECB, 2021a).

17

18 Many European sectors are exposed to multiple and cross-cutting risks (Gasbarro et al., 2019; Schleypen et
19 al., 2021). Indirect effects via supply chains, transport and electricity networks can be as high as or
20 substantially higher than direct effects (medium confidence) (Koks et al., 2019a; Koks et al., 2019b; Knittel
21 et al., 2020).

22

23 13.6.1.4 Tourism

24

25 Snow cover duration and snow depth in the Alps decreased since the 1960s (Klein et al., 2016; Sch鰊er et
26 al., 2019; Matiu et al., 2021). Despite snowmaking, the number of skiers to French resorts at low elevations
27 during the extraordinary warm/dry winters of 2006/2007 and 2010/2011 was 12-26% lower (Falk and Vanat,
28 2016).

29

30 Due to reduced snow availability and hotter summers, damages are projected for the European tourism
31 industry, with larger losses in SEU (high confidence) and some smaller gains in the rest of Europe (medium
32 confidence) (Ciscar et al., 2014; Roson and Sartori, 2016; Dellink et al., 2019).

33

34 At 2癈 GWL, the operation of low altitude resorts without snowmaking will likely be discontinued, while
35 beyond 3癈 GWL snowmaking will be a necessary but not always sufficient for most resorts in many
36 European mountains and parts of NEU (Pons et al., 2015; Joly and Ungureanu, 2018; Scott et al., 2019;
37 Spandre et al., 2019). Expanding snowmaking is capital-intensive and will strongly increase water and
38 energy consumption, particularly at 3癈 GWL and beyond (Spandre et al., 2019; Morin et al., 2021),
39 adversely affecting the financial stability of small resorts (Pons et al., 2015; Falk and Vanat, 2016; Spandre
40 et al., 2016; Joly and Ungureanu, 2018; Moreno-Gen� et al., 2018; Steiger and Scott, 2020). Permafrost
41 degradation due to rising temperatures is expected to create stability risks for ropeway transport
42 infrastructure at high-altitude Alpine areas (Duvillard et al., 2019).

43

44 Climatic conditions from May to October at 1.5-2癈 GWL are projected to become more favourable for
45 summer tourism in NEU and parts of WCE and EEU, while there is medium confidence on opposite trends
46 for SEU from June to August (Grillakis et al., 2016; Scott et al., 2016; Jacob et al., 2018; Koutroulis et al.,
47 2018). The amenity of European beaches may decrease as a result of sea level rise amplifying coastal erosion
48 and inundation risks, although less in NEU (Ebert et al., 2016; Toimil et al., 2018; Lopez-Doriga et al., 2019)
49 (Section 13.2 and Ranasinghe et al., 2021 Section 12.4.5).

50

51 13.6.1.5 Built Environment, Settlements and Communities

52

53 Expected shift of European residents to large cities and coastal areas will increase assets at risk (Section
54 13.2). The share of urban population in Europe is projected to increase from 74% in 2015 to 84% in 2050,
55 corresponding to 77 million new urban residents (UN/DESA, 2018), with most of this increase in SEU and
56 WCE (particularly in Turkey and France). In the EU-28, urban residents in 2100 may increase by about 30
57 million under SSP1 and SSP5, and decrease by 90-110 million under SSP3 and SSP4 (Terama et al., 2019).

Do Not Cite, Quote or Distribute  13-44       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 About 32% of 571 European cities in the GISCO Urban Audit 2014 dataset show a medium to high or
 3 relatively high vulnerability against heatwaves, drought and floods (Tapia et al., 2017). Under current
 4 vulnerabilities, future climate hazards will augment climate risks for several cities, particularly beyond 3癈
 5 GWL (Figure 13.17). In many NEU cities, a high increase in pluvial flooding risk by the end of the century
 6 is possible, while in WCE cities may face a high increase in pluvial flooding risks, moderate to very high
 7 increase in extreme heat risk, and to some extent moderate to high increase in drought risk. Many SEU cities
 8 could face a high to very high increase in risks from extreme heat and meteorological drought.

 9

10

11

12 Figure 13.17: Projected changes in pluvial flooding, extreme heat and meteorological drought risks for the 65 largest
13 cities in EU-28 plus Norway and Switzerland for 2.5癈 GWL and 4.4癈 GWL compared with the baseline (1995-2014)
14 (Tapia et al., 2017). Exposure is expressed in terms of current population. Values of climatic impact-drivers are derived
15 from the Euro-CORDEX regional climate model ensemble.

16

17

18 13.6.1.5.1. Risks from coastal, river and pluvial flooding
19 New studies increase confidence in AR5 statements that flood damages will increase in coastal areas due to
20 sea level rise and changing social and economic conditions (section 13.2.1.1). Except for areas affected by
21 land uplift, it is projected that further adaptation will be required to maintain risks at the present level for
22 most coastal cities and settlements (Haasnoot et al., 2013; Ranger et al., 2013; Malinin et al., 2018; Hinkel et
23 al., 2019; Umgiesser, 2020).

24

Do Not Cite, Quote or Distribute  13-45       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 In many cities, the sewer system is older than 40 years, potentially reducing their capacity to deal with more
 2 intense pluvial flooding (EEA, 2020c). Apart from climate change, urbanization is an important driver for
 3 increases in flooding risks as it results in growth of impervious surfaces. Flash floods are particularly
 4 challenging, causing the overburdening of drainage systems (Dale et al., 2018), urban transport disruptions,
 5 and health and pollution impacts due to untreated sewage discharges (Kourtis and Tsihrintzis, 2021).

 6

 7 More than 25% of the population in nearly 13% of the EU-cities lives within potential river floodplains. In
 8 many of these, e.g. 50% of UK cities, a significant increase of the 10-year high river flow is possible beyond
 9 2癈 GWL under a high-impact scenario (i.e. 90th percentile of projections) (Guerreiro et al., 2018; EEA,
10 2020c).

11

12 13.6.1.5.2 Risks from heatwaves, cold waves and drought
13 Heatwave days and number of long heatwaves increased in most capitals from 1998-2015 compared to 1980-
14 1997 (Morabito et al., 2017; Seneviratne et al., 2021). In the summer of 2018, many cities suffered from
15 heatwaves attributed to climate change (Vogel et al., 2019; Undorf et al., 2020). As a result, indoor
16 overheating and reduced outdoor thermal comfort, often coupled with urban heat island (UHI) effect, have
17 already impacted European cities (Di Napoli et al., 2018; EEA, 2020c) (see also Section 13.7.1).

18

19 Heatwaves are likely to become a major threat not only for SEU but also for WCE and EEU cities (Russo et
20 al., 2015; Guerreiro et al., 2018; Lorencova et al., 2018; Smid et al., 2019).At 2oC GWL and SSP3, half of1.1
21 the European population will be under very high risk of heat stress in summer (Rohat et al., 2019). The UHI
22 effect will further increase urban temperatures (Estrada et al., 2017). In many cities, hospitals and social
23 housing tend to be located within the intense UHI, thus increasing exposure to vulnerable groups (EEA,
24 2020c). There is high confidence that overheating during summer in buildings with insufficient ventilation
25 and/or solar protection will increase strongly , with thermal comfort hours potentially decreasing by 74% in
26 locations of southern Europe at 3oC GWL (Jenkins et al., 2014a; Hamdy et al., 2017; Heracleous and
27 Michael, 2018; Dino and Meral Akg黮, 2019; Shen et al., 2020). Highly insulated buildings, following
28 present building standards, will be vulnerable to overheating, particularly under high GWL levels, unless
29 adequate adaptation measures are applied (Williams et al., 2013; Virk et al., 2014; Mulville and
30 Stravoravdis, 2016; Fosas et al., 2018; Ibrahim and Pelsmakers, 2018; Salem et al., 2019; Tian et al., 2020).
31 Cities in NEU and WCE are more vulnerable due to limited solar shading and fewer air conditioning
32 installations (Ward et al., 2016; Thomson et al., 2019). Cooling energy demand in SEU buildings has been
33 projected to increase by 81-104% by 2035 and by 91-244% after 2065 compared to 1961-1990 depending on
34 GWL (Cellura et al., 2018). Increases of 31-73% by 2050 and by 165-323% by 2100 compared to 1996-2005
35 were estimated for buildings in NEU (Dodoo and Gustavsson, 2016) with risks modified by adaptation
36 (Vigui� et al., 2020) (see section 13.6.2). Cold waves beyond 3oC GWL will not represent an effective threat
37 for European cities at the end of the century, and only a marginal hazard under 2oC GWL (Smid et al., 2019).

38

39 At 2oC GWL and beyond, cities in SEU and large parts of WCE would exceed the historical maximum 12-
40 month Drought Severity Index of the past 50-years (on drought risks see Section 13.2) and 30% will have at
41 least 30% probability of exceeding this maximum every month (Guerreiro et al., 2018). This could adversely
42 affect the operation of municipal water services (Kingsborough et al., 2016). For example, under 2oC GWL,
43 the reservoir storage volume is predicted to decrease for all of England and Wales catchments, resulting in a
44 probability of years with water use restrictions doubling by 2050 and quadrupling by 2100 compared to
45 1975-2004 (Dobson et al., 2020). The combination of high temperatures, drought, and extreme winds,
46 potentially coupled with insufficient preparedness and adaptation, may amplify the damage of wildfires in
47 peri-urban environments (13.3.1.3). High fuel load combined with proximity of the built environment to
48 wildland highly increases fire risks (EEA, 2020c).

49

50 Extreme heat and drought causes shrinking and swelling of clays, threatening the stability of small houses in
51 peri-urban environments (Pritchard et al., 2015), with damage costs of  0.9-1 billion during the 2003
52 heatwave (Corti et al., 2011). In WCE and SEU, mean annual damage costs could increase by 50% for 2癈
53 GWL, and by a factor of 2 for 3癈 GWL (Naumann et al., 2021).

54

55 13.6.1.5.3 Risks from thaw of permafrost and mudflows
56 Increasing temperatures in NEU and the Alps lead to accelerated degradation of permafrost, negatively
57 affecting the stability of infrastructures (Stoffel et al., 2014; Beniston et al., 2018; Duvillard et al., 2019). In

Do Not Cite, Quote or Distribute  13-46       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 the Caucasus, glacial mudflows due to permafrost degradation and modern tectonic processes pose a
 2 significant danger to the infrastructure (Vaskov, 2016). In the last 30 years, the permafrost temperature in the
 3 European part of the Russian Arctic has increased by 0.5-2.0癈, resulting to damages of buildings, roads
 4 and pipelines and to significant expenditure for stabilizing soils (Porfiriev et al., 2017; Konnova and Lvova,
 5 2019). Beyond 3oC GWL, the bearing capacity for infrastructure in the permafrost region of the European
 6 Russia could decrease by 32-75% by mid-century and by 95% by 2100, potentially affecting settlements in
 7 northern EEU (Shiklomanov et al., 2017; Streletskiy et al., 2019). Increasing number of cycles of freezing
 8 and thawing, observed in EEU, led to accelerated aging of building envelopes (Frolov et al., 2014)
 9 (13.8.1.4). Permafrost degradation due to higher temperatures could increase the potential of debris flow
10 detachment in Alpine locations (13.6.1.4, Damm and Felderer, 2013).

11

12 Increased precipitation falling on local topography can increase landslide and mudflow risks, as seen in
13 settlements at the Caucasus mountainous region (Marchenko et al., 2017; Efremov and Shulyakov, 2018;
14 Kerimov et al., 2020). At the Umbria region in Italy, landslide events could increase by 16-53% under 2oC
15 GWL and by 24-107% beyond 3oC GWL, mostly during winter (Ciabatta et al., 2016). Risks from shallow
16 landslides are expected to increase in the Alps and Carpathians if no adequate risk mitigation measures are in
17 place (CCP5.3.2, Gariano and Guzzetti, 2016).

18

19 13.6.2 Solution Space and Adaptation Options

20

21 Monetary assessments of future damages from climate extremes on critical infrastructures show a sevenfold
22 escalating figures by 2080s (Figure 13.18) compared to the baseline (Forzieri et al., 2018), highlighting the
23 need for adaptation.

24

25

26

27 Figure 13.18: Climate risks to critical infrastructures, aggregated at European (EU+) level under the SRES A1B
28 scenario (Forzieri et al., 2018). Baseline: 1981-2010. 2020s: 2011�2040. 2050s: 2041�2070. 2080s: 2071�2100.

29

30

31 13.6.2.1 Current Status of Adaptation

32

33 There is new evidence on increasing adaptation planning in cities, settlements, and key infrastructure, but
34 less on implemented adaptation (Table 13.1; Box 13.3; Figure 13.36), adaptation by private actors and by
35 cities against sea level rise (Chapter 16; CCP2).

36

37

38 Table 13.1: Present status of planned and implemented adaptation in cities, energy sector, tourism sector, transport and
39 industry in Europe

Do Not Cite, Quote or Distribute  13-47       Total pages: 143
FINAL DRAFT                                Chapter 13                                   IPCC WGII Sixth Assessment Report

              General commitments / Adaptation Plans                                    Implemented adaptation actions

Cities         Increasing number of cities acknowledging the critical role of            Large cities are in the process of implementation (e.g., Helsinki,

                 adaptation in building resilience to climate change.                       Copenhagen, Rotterdam, Barcelona, Madrid, London, Moscow).

               Of 9609 European municipalities in the Covenant of Mayors for             Many cities have implemented measures potentially supporting

                 Climate & Energy (CoM), 2221 reported on adaptation through the            adaptation but not labelled as such.
                 CoM platform, 429 provided some information on adaptation goals,
                 risk and vulnerability assessments/action plans, and 230-290 reported   Current climate policies implemented at city-scale are primarily
                 adaptation goals and funds for adaptation. Extreme heat, droughts
                 and forest fires are the most often reported hazards.                      addressing mitigation and, to a lesser extent adaptation.

               Most urban adaptation plans include ecosystem-based measures (but         Increasing use of Nature-based Solutions (NbS) and ecosystem-based

                 often with insufficient baseline information and lack of convincing        adaptation (EbA) to address urban heating and the discontinuity of the
                 implementation actions).                                                   urban water cycle due to surface sealing and limited infiltration.

               Adaptation to risks from climate extremes (mostly flooding) is often      Strategic, tactical, and emergency measures applied for drought

                 addressed through municipal emergency plans.                               management (e.g., London, Istanbul)

Energy         29 countries (in place in 14 and in progress in 15).                      11 countries (actions implemented in 5 and in progress in 6)
               Few countries have considered specific adaptation actions (mostly
                                                                                         Measures undertaken by some distribution system operators (DSOs) and
                 preparatory) in their national or energy-specific risk assessments.
                                                                                            energy companies, focusing on adaptation of transmission lines, water
                                                                                            cooling, dams for avoiding flooding during intense precipitation events,
                                                                                            actions to avoid flooding and to secure fuel supply.

Tourism        Consideration of tourism in national adaptation strategies is limited,    18-67% of ski slopes (67% in Austria, 39% in Switzerland, 18% in

                 and national tourism strategies rarely mention adaptation.                 Bavaria-Germany, 20% in French Alps, and 45% in Spain apply
                                                                                            snowmaking.
               Tourism operators do not consider longer term adaptation strategies
                                                                                         Resorts offering nocturnal skiing (e.g., Spain) and other snow-based
                 to be relevant.
                                                                                            activities. Transformation to all-year mountain resorts (e.g., 70% of
               Legally binding consideration of climate change when constructing            Spanish ski resorts).

                 new tourism units (e.g., the 2016 French Mountain Act).                 Some diversification of tourism products offered in Mediterranean

                                                                                            coastal destinations.

                                                                                         Implementation of water saving measures, primarily for cost reduction.

Transport      Only 10 countries have started coordination activities or identified      Only in 5 countries.
                                                                                         Majority of actions are preparatory.
                 adaptation measures.                                                    Actions mostly focus on infrastructure and much less on services,

               An integrated, trans-modal approach to adaptation is lacking.                although the latter have started gaining ground (e.g., operational forecasts
               Few countries are mainstreaming of adaptation within transport               for water levels in rivers).

                 planning and decision-making (e.g., the `Low-water Rhine' action        Transport modes often compete for public funds, and political priorities
                 plan and in Germany).
                                                                                            for specific modes often influence adaptation

               Some action is undertaken in the public and private sector, e.g.,         Some public and private actors are moving faster: new railway drainage

                 revised manuals/guidelines/ protocols to consider climate change           standards (Network Rail/ UK), prediction of adverse weather events
                 impacts and extreme events (e.g., Deutsche Bahn, Norwegian Public          (Spanish rail service operator), measures against coastal flooding
                 Roads Administration).                                                     (Copenhagen Metro), measures for sea level rise (Rotterdam port and
                                                                                            France).

Industry and   Recommendation of the High Level Expert Group on Sustainable              50 of European large listed companies publicly disclosed their climate
   business
                 Finance ((HLEG) that the European Commission endorses and                  risks in 2020; yet only a small share provided specifics on sectoral risks,
                 implements the guidelines provided by the Task force on Climate-           how risks differ over time and by different climate scenarios
                 Related Financial Disclosure in 2019.
                                                                                         Large national/multinational companies, and companies regulated by

                                                                                            mitigation policy are first movers in corporate adaptation, while small
                                                                                            and medium-sized enterprises often lack knowledge and resources to
                                                                                            address risks and adaptation options.

                                                                                         Development of different tools (stress testing, scenario analysis, value at

                                                                                            risk) by climate service providers, insurance companies and central banks

              Well-established adaptation  Advancing adaptation                         Low adaptation

 1

 2

 3 Although urban adaptation is underway, many small, economically weak (i.e. with low GDP/capita) or cities
 4 facing high climate change risks lack adaptation planning (Reckien et al., 2015; EEA, 2016). While almost
 5 all large municipalities in NEU and WCE report implemented actions at least in one sector, this is not the
 6 case for 39% of municipalities in SEU (Aguiar et al., 2018). In the UK, the legal requirement to develop
 7 urban adaptation plans has been a significant driver for their widespread adoption (Reckien et al., 2015). The
 8 availability of and access to funding for adaptation is also crucial for plan development (section 13.11.1).
 9 Network membership (e.g., ICLEI, C40, Covenant of Mayors for Climate & Energy) is an important driver
10 for city planning and transfer of best practices (Heikkinen et al., 2020a). Stakeholder engagement is key for
11 successful adaptation (Bertoldi et al., 2020)( see Chapter 17).

12

13 Only 29% of local adaptation plans are mainstreamed in cities, which could reduce the effectiveness of
14 implementing adaptation (Reckien et al., 2019) (13.11.1.2). Although large municipalities usually fund the

Do Not Cite, Quote or Distribute                      13-48                                             Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

1 implementation of their adaptation plans, smaller and less populated municipalities (particularly in SEU and
2 EEU) often depend on intergovernmental, international and national funding.

3

4 13.6.2.2 Adaptation Options as a Function of Impacts

5

6 Examples of adaptation options in Europe are presented in Figure 13.19.

7

8

 9

10 Figure 13.19: Adaptation options in cities, settlements, and key infrastructure in Europe (Table SM13.7).

11

12

Do Not Cite, Quote or Distribute  13-49       Total pages: 143
    FINAL DRAFT                                                        Chapter 13                        IPCC WGII Sixth Assessment Report

 1 Nature-based solutions (NbS) and ecosystem-based adaptation (EbA) such as green spaces, ponds, wetlands
 2 and green roofs for urban stormwater management and vegetation for heat mitigation represent an emerging
 3 adaptation option in cities. Combined with traditional water infrastructure, they can contribute to managing
 4 urban flood events (Kourtis and Tsihrintzis, 2021), playing a role in mitigating flood peaks (Pour et al.,
 5 2020) and protecting critical urban infrastructure (Ossa-Moreno et al., 2017). For example, in the
 6 Augustenborg district of Malm�, Sweden, using nature to manage stormwater runoff resulted in capturing an
 7 estimated 90% of runoff from impervious surfaces and reduced the total annual runoff volume from the
 8 district by about 20% compared to the conventional system (EEA, 2020c). Urban greening is associated with
 9 lower ambient air temperature and relatively higher thermal comfort during warm periods (Bowler et al.,
10 2010; Oliveira et al., 2011; Cohen et al., 2012; Cameron et al., 2014). The scale and relative degree of
11 management or integration of approaches drawing on nature with `engineered' solutions affect their
12 vulnerability to climate change. Small-scale urban NbS are relatively less vulnerable due to increased
13 capacity for intervention, while the relatively greater contact between stakeholders and urban NbS
14 (compared with larger-scale, rural approaches) provides greater opportunity for human intervention to ensure
15 the survival of urban vegetation during droughts or heatwaves.

16

17 When selecting and combining adaptation options, challenges remain on how to address uncertainties of
18 climate projections and climatic extremes (Fowler et al., 2021) and to translate scientific input into practical
19 guidance for adaptation (Dale, 2021) (Section 13.11.1.3).

20

21 An assessment of the feasibility and effectiveness of main adaptation options, based on literature, is
22 presented in Figure 13.20; for adaptation to flood risk, see Figure 13.6.

23

24

            Effectiveness & feasibility of main adaptation options to climate impacts & risk for cities, settlements & key infrastructure in Europe

    Impact types          Adaptation options                   Effectiveness Feasibility                                                                      Confidence

                                                                   Economic               Technological Institutional Socio-cultural Ecological  Geophysical  Evidence Agreement
                                                                                                                                                        M
    Reduction of thermal  Interventions in the building shell  M   M                      L      M   L   NL                                             M     M           M
    comfort due to        Ventilation (natural/mechanical,
    increasing            incl. night)                         M   M                      H      M   L   NL                                                   H           M
    temperatures &        Air conditioning
    extreme heat          Shading                              H   L                      NL     NL  L   NL                                      M            L           M

                                                               M   L                      H      L   M   NL                                      L            M           M

                          Green roofs, green walls             L   M                      M      L   M   M                                       M            M           M

                          Urban green spaces                   L   L                      M      L   M   M                                       M            L           M

                          Use of 'cool' paints & coatings      L   H                      M      L   M   NL                                      H            M           M

                          Escape to nearby non-urban           NL  NL                     NL     NL  M   NL                                      NL           L           H

                          destinations

    Loss of critical services Improvements in cooling systems  M   L                      M      M   NL  NL                                      M            H           M

    due to heatwaves & Shifting production to less water-      M   M                      M      L   NL  NL                                      NL           L           H

    drought               intensive plants

                          Regulatory measures                  L   M                      NL     M   NL  NL                                      NL           L           M

                          Management measures                  M   M                      M      M   NL  L                                       M            M           M

                          Use of heat-resilient materials      L   H                      L      M   M   NL                                      NL           L           H

                          Replace vulnerable infrastructure    L   M                      NL     NL  NL  NL                                      NL           L           H
                          with resilient one

    Legend

     High = H

     Medium = M

     Low = L

25  No/Limited Evidence = NL

26 Figure 13.20: Effectiveness and feasibility of main adaptation options for cities, settlements and key infrastructure in

27 Europe (SM13.9 and Table SM13.8).

28

29

30 There are gaps of knowledge on the social, environmental and geophysical dimension of feasibility for many
31 options and a holistic assessment of different options is largely lacking. This latter could reveal unintended
32 impacts from and synergies or trade-offs between options, as in water and wastewater services (Dobson and
33 Mijic, 2020).

34

35 13.6.2.3 Adaptation Limits, Residual Risks, Incremental and Transformative Adaptation

36

37 Adaptation in cities, settlements and key infrastructure in Europe faces technical, environmental, economic
38 and social limits (Figure 13.21).

39

    Do Not Cite, Quote or Distribute                                                      13-50                                                               Total pages: 143
      FINAL DRAFT                 Chapter 13  IPCC WGII Sixth Assessment Report

1

 2

 3 Figure 13.21: Indicative adaptation limits in cities, settlements and key infrastructure in Europe (Table SM13.16).

 4

 5

 6 Adaptation options for many sectors will not be sufficient to remove residual risks, e.g. regarding
 7 overheating in buildings under high GWL, (Tillson et al., 2013; Virk et al., 2014; Dodoo and Gustavsson,
 8 2016; Mulville and Stravoravdis, 2016; Hamdy et al., 2017; Heracleous and Michael, 2018; Dino and Meral
 9 Akg黮, 2019); snowmaking beyond 3癈 GWL (Scott et al., 2019; Steiger et al., 2020; Steiger and Scott,
10 2020); hydropower (Gaudard et al., 2013; Ranzani et al., 2018); electricity transmission and demand
11 (Bollinger and Dijkema, 2016; EEA, 2019a; Palkowski et al., 2019); urban subways (Jenkins et al., 2014a);
12 and flood mitigation in cities (Skougaard Kaspersen et al., 2017; Umgiesser, 2020). Some adaptation actions
13 in a sector may also have side-effects on others, increasing their vulnerability (Pranzini et al., 2015) (sections
14 13.2.2 and 13.2.3).

15

16 Examples of transformative adaptation in urban areas are observed (e.g., the Benthemplein water square, the
17 Floating Pavilion in Rotterdam and the Hafencity flood proofing in Hamburg), but they often remain policy
18 experiments and prove challenging to upscale (Jacob, 2015; Restemeyer et al., 2015; Restemeyer et al.,
19 2018; Holscher et al., 2019). The active involvement of local stakeholders, public administration and
20 political leaders are drivers for community transformation, whereas lack of local resources and/or capacities
21 are frequently reported barriers to change (F黱fgeld et al., 2019; Thaler et al., 2019).

22

23 13.6.2.4 Governance and Insurance

24

25 Urban adaptation plans can enhance resilience and their development is mandatory in the UK, France, and
26 Denmark (Reckien et al., 2019). There is medium confidence that the development of urban adaptation
27 planning is much more influenced by a city's population size, present adaptive capacity and GDP per capita
28 than by anticipated climate risks (Reckien et al., 2018). A high organizational capacity in a municipality may
29 not be a necessary condition for forward-looking investment decisions on urban water infrastructure,
30 although enablers differ for small versus medium-to-large municipalities (Pot et al., 2019). There is large in-
31 country variation in policy mixes utilized by local governments for supporting adaptation (Lesnikowski et
32 al., 2019). In early adapter cities (e.g., Rotterdam), adaptation is institutionally embedded in climate,
33 resilience and sustainability-related actions and collaboration between city departments, government levels,
34 businesses, and rest stakeholders (Holscher et al., 2019). In most other cities however, adaptation planners

Do Not Cite, Quote or Distribute  13-51       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 rarely consider collaborations with citizens and there are difficulties in departmental coordination and
 2 upscaling from pilot projects (Brink and Wamsler, 2018).

 3

 4 The level and type of collaboration between the public and private sector in managing climate risks varies
 5 across Europe (Wiering et al., 2017; Alkhani, 2020). For example, in flood management (also section 13.2),
 6 the private sector involvement in Rotterdam is much more pronounced and there are joint public璸rivate
 7 responsibilities throughout most of the policy process due to the large share of private ownership of land and
 8 real estate (Mees et al., 2014).

 9

10 In large infrastructure networks, the lack of a leading and powerful institutional body, with sufficient
11 research resources targeted to climate change risk assessment, may limit adaptive capacity, as for example in
12 railways (Rotter et al., 2016).

13

14 The European insurance industry has developed tailored products for specific climate risks threatening cities,
15 settlements and key infrastructure, such as risk-based flood insurance for homeowners and companies
16 (section 13.2.3). The European insurance industry is developing new services (such as risk analysis and
17 catastrophe modelling embedding climate change, early warning and post-event recovery recommendations)
18 and it has recently started to play a role as communicator of future risks and as institutional investor with the
19 aim of risk reduction (Jones and Phillips, 2016; Marchal et al., 2019).

20

21 13.6.2.5 Links between Adaptation and Mitigation

22

23 Evidence from transport in Europe shows that adaptation actions do not consider enough long-term transition
24 paths embedded in mitigation, while mitigation strategies are often not assessed under future climate
25 scenarios (Aparicio, 2017). Without rapid decarbonization of electricity supply, greenhouse gas emissions
26 will increase due to the increased use of air-conditioning installations in cities. This trade-off can be reduced
27 to some extent through use of more efficient cooling technologies (IEA, 2018) and complementary
28 adaptation measures such as large-scale urban greening, building policies and behavioural changes in air
29 conditioning use (Vigui� et al., 2020; Sharifi, 2021; Vigui� et al., 2021). Greenhouse gas emissions from
30 transport may increase due to the temporary relocation of city residents to cooler locations during heatwaves
31 (Juschten et al., 2019), and from increased energy use for snowmaking in European ski resorts (Scott et al.,
32 2019).

33

34 13.6.3 Knowledge Gaps

35

36 A key knowledge gap is the lack of a quantitative European-wide integrated assessment of future climate
37 change risks on water and energy, including different socio-economic futures. Models capable of
38 representing integrated policies for energy and water are lacking (Khan et al., 2016) including quantitative
39 modelling of impacts on energy transmission and coastal energy infrastructure (Cronin et al., 2018). These
40 lacks are especially pertinent when combined with the small number of studies considering SSP population
41 projections, and adaptation tipping points. The limited social vulnerability assessments, mapping, and
42 validation (Rufat et al., 2019) contribute further to these knowledge gaps.

43

44 While compound, concurrent, and consecutive climate extremes become more frequent, there is limited
45 knowledge on sectoral risks or on cascading risks for through transport, telecommunications, water, and
46 banking and finance. While heat is well studies, studies on risks for cities and key infrastructure from
47 hailstorms and lightning are missing.

48

49 Empirical data on the damage of transport infrastructure (e.g., railways) covering different European
50 countries is not systematically collected and indirect economic effects of interruptions of transport networks
51 are not well studied (Bubeck et al., 2019). These deficits result in uncertainties associated with impacts of
52 climate change on transport flows and indirect impacts (delays, economic losses).

53

54 There is limited knowledge on interactions created by synchronous adaptation in ski tourism supply and
55 demand, and models are not yet including individual snowmaking capacity and a higher time resolution
56 (Steiger et al., 2019). Furthermore, there is no European-wide assessment of coastal flooding risks on
57 tourism.

Do Not Cite, Quote or Distribute  13-52       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Many studies lack consideration of market characteristics (e.g., competitors) in their risk assessment which
 3 would be improved by location- and sector-specific knowledge on climate risks for firm assets, operations,
 4 business, industry, finance and insurance needed to inform adaptation actions (de Bruin et al., 2020; Feridun
 5 and G黱g鰎, 2020; Monasterolo, 2020).

 6

 7

 8 13.7 Health, Wellbeing and the Changing Structure of Communities

 9

10 13.7.1 Observed Impacts and Projected Risks

11

12 13.7.1.1 Mortality due to Heat and Other Extreme Events

13

14 Attribution studies show that human-induced climate change is increasing the frequency and intensity of heat
15 waves and has already impacted human health in Europe (Section 13.10.1, Vicedo-Cabrera et al., 2021); for
16 example the 2010 heatwave in EEU resulted in 55,000 heat-related deaths (Barriopedro et al., 2011; Russo et
17 al., 2015); the 2018 heatwave in NEU (Ebi et al., 2021) and the 2019 heatwave in WCE and NEU both had
18 significant health impacts (Cross-Chapter Box DISASTER in Chapter 4, Vautard et al., 2020; Watts et al.,
19 2021). Elderly, children, (pregnant) women, socially isolated people and those with low physical fitness are
20 particularly exposed and vulnerable to heat-related risks, as are those people suffering from pre-existing
21 medical conditions, including cardiovascular disease, kidney disorders, diabetes and respiratory diseases
22 (de'Donato et al., 2015; Sheridan and Allen, 2018; Naik et al., 2021). An aging population in Europe is
23 increasing the pool of vulnerable individuals, resulting in higher risk of heat-related mortality (Montero et
24 al., 2012; Carmona et al., 2016b; WHO, 2018b; Watts et al., 2021).

25

26 1.5癈 GWL warming could result in 30,000 annual deaths due to extreme heat, with up to three-fold the
27 number under 3癈 GWL (high confidence) (Rold醤 et al., 2015; Forzieri et al., 2017; Kendrovski et al.,
28 2017; Naumann et al., 2020). The risk of heat stress, including mortality and discomfort, is dependent on
29 socioeconomic development (Rohat et al., 2019; Ebi et al., 2021) (Figure 13.22). Heat stress risks will be
30 lower under SSP1 than SSP3 or SSP4 scenarios (high confidence) (Hunt et al., 2017; Rohat et al., 2019;
31 Wang et al., 2020; Ebi et al., 2021). The incidence of heat-related mortality and morbidity will be highest in
32 SEU, where their magnitude is also expected to increase more rapidly (Forzieri et al., 2017; Gasparrini et al.,
33 2017; Guo et al., 2018; D韆z et al., 2019; Vicedo-Cabrera et al., 2021). WCE, NEU, SEU will experience
34 accelerating negative consequences beyond 1.5癈 GWL, particularly under SSP3 and SSP4 due to higher
35 vulnerability compared with SSP1 (Figure 13.22, Rohat et al., 2019). The number of heat-related respiratory
36 hospital admissions is projected to increase from 11,000 (1981-2010) to 26,000 annually (2021-2050),
37 particularly in SEU mainly due to a relative increase in the number of extremely hot days (舠tr鰉 et al.,
38 2013). Cold spells are projected to decrease across Europe, particularly in southern Europe, but do not
39 compensate for the additional heat related deaths projected (Lhotka and Kysely, 2015; Carmona et al.,
40 2016a; Martinez et al., 2018).

41

42 74% of Europeans live in urban areas (section 13.6), where the effect of heat waves on human health is
43 exacerbated by microclimates due to buildings and infrastructure, heat island effects, and air pollution
44 (WHO, 2018a; Smid et al., 2019). In large European cities, stabilizing climate warming at 1.5癈 GWL
45 would decrease premature deaths by 15�22% in summer compared with stabilization at 2癈 GWL (Mitchell
46 et al., 2018) (high confidence).

47

48 Although there is very high confidence that risk consequences will inevitably be more pervasive and
49 widespread in a warmer Europe, evidence of higher heat tolerance is also emerging across most European
50 regions (Todd and Valleron, 2015; 舠tr鰉 Daniel et al., 2016 Sweden, 1901�2009; Follos et al., 2020).
51 Future projections of mortality rates in Europe under the assumption of complete acclimatization suggest
52 constant or even decreasing rates of mortality in spite of global warming (舠tr鰉 et al., 2017; Guo et al.,
53 2018; D韆z et al., 2019). However, there are large uncertainties in the ability to adapt to future heat extremes
54 which might fall outside of historical ranges (Vanos et al., 2020).

55

56

Do Not Cite, Quote or Distribute  13-53       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.22: Scenario matrix for multi-model median heat stress risks for the baseline 1986-2005, and different SSP-
 3 RCP combinations for the period 2040-2060. SSPs are extended SSPs for Europe (EU28+). Heat stress risk is
 4 calculated by geometric aggregation of the hazard (heat wave days), population vulnerability and exposure. Risk values
 5 are normalised using a z-score rescaling with a factor 10-shift. Details of the methodology are provided in Rohat et al.
 6 (2019).

 7

 8

 9 Other extreme events already result in major health risks across Europe. Between 2000 and 2014, for
10 example, floods in Russia killed approximately 420 people, mainly older women (Belyakova et al., 2018).
11 Fatalities associated with coastal and riverine flooding (13.2.2), wildfires (13.3.4), and windstorms could rise
12 substantially by 2100 (Forzieri et al., 2017; Feyen et al., 2020). Lifetime exposure to extreme weather events
13 for children born in 2020 will be about 50% larger at 3.5癈 compared to 1.5癈 GWL (Thiery et al., 2021).

14

15 13.7.1.2 Air Quality

16

17 Air pollution is already one of the biggest public health concerns in Europe; in 2016, roughly 412,000 people
18 died prematurely due to long-term exposure to ambient PM2.5, 71,000 due to NO2, and more than 15,000
19 premature mortalities occurred due to near surface ozone (EEA, 2019b; Lelieveld et al., 2019). Impacts of air
20 pollution determined by air quality policies, changes to temperature, humidity, and precipitation (Naik et al.,
21 2021). Climate change could increase air pollution health effects, with the size of the effect differing across
22 European regions and pollutants (Jacob and Winner, 2009; Orru et al., 2017; Tarin-Carrasco et al., 2021)
23 (medium confidence). Increases in temperature and changes in precipitation will impact future air quality due
24 to increased risk of wildfires and related air pollution episodes. Data on the health impacts of wildfires in

Do Not Cite, Quote or Distribute  13-54       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 Europe is currently limited (Section 13.3.1.4), but examples such as the 2017 fires suggest that more than
 2 100 people died prematurely in Portugal alone as a result of poor air-quality (Oliveira et al., 2020).

 3

 4 At 2.5癈 GWL, mortalities due to exposure to PM2.5 are projected to increase by up to 73% in Europe
 5 (medium confidence) (Silva et al., 2017; Lelieveld et al., 2019; Tarin-Carrasco et al., 2021). At 2癈 GWL,
 6 annual premature mortalities due to exposure to near-surface ozone are projected to increase up to 11% in
 7 Western Central and Southern Europe and to decrease up to 9% in Northern Europe (under RCP4.5) (Orru et
 8 al., 2019) (medium confidence). Projected increase in wildfires and reduced air quality is expected to
 9 increase respiratory morbidity and mortality, especially in SEU (Slezakova et al., 2013; de Rigo et al., 2017).
10 Constant or lower emissions combined with stricter regulations and new policy initiatives, might improve air
11 quality in coming decades (medium agreement, low evidence). Aging population in Europe augments the
12 future air quality mortality burden by 3-13% in 2050 (Geels et al., 2015; Orru et al., 2019). Beside ambient
13 air quality, projected increases in flood risk and heavy rainfall could decrease indoor air quality (13.6.1.5.2)
14 due to dampness and mould, leading to increased negative health impacts, including allergies, asthma and
15 rhinitis (EASAC, 2019; EEA, 2019b).

16

17 13.7.1.3 Climate Sensitive Infectious Diseases

18

19 Figure 13.23 summarizes the observed and projected changes in climatic suitability and assesses the risk for
20 selected climate-sensitive infectious diseases in Europe.

21

22 Among the tick-borne diseases, Lyme disease is the most prevalent disease in Europe. There has been a
23 temperature-dependent range expansion of ticks that is projected to expand further north in Sweden, Norway
24 and the Russian Arctic. (Jaenson et al., 2012; Jore et al., 2014; Tokarevich et al., 2017; Waits et al., 2018)
25 and to higher elevations in Austria and the Czech Republic (Daniel et al., 2003; Heinz et al., 2015) (medium
26 confidence). A potential habitat expansion of these ticks of 3.8% across Europe, relative to 1990-2010, is
27 projected for 2癈 GWL (Porretta et al., 2013; Boeckmann and Joyner, 2014). In contrast, there are projected
28 habitat contractions for these ticks in SEU due to unfavourable climatic conditions (Semenza and Suk,
29 2018).

30

31 The Asian tiger mosquito (Aedes albopictus) is present in many European countries and can transmit
32 Dengue, Chikungunya and Zika (Liu-Helmersson et al., 2016; Tjaden et al., 2017; Messina et al., 2019).
33 There is a moderate climatic suitability projected for chikungunya transmission, notably across France,
34 Spain, and Germany but also contractions, particularly in Italy. Europe experienced an exceptionally early
35 and intense transmission season of West Nile virus in 2018, with elevated spring temperature abnormalities
36 (Haussig et al., 2018; Marini et al., 2020). Projections for Europe show West Nile virus risk to expand; by
37 2025 the risk is projected to increase in SEU and south and eastern parts of WCE (medium confidence)
38 (Semenza et al., 2016). Although climatic suitability for malaria transmission in Europe is increasing and
39 will lead to a northward spread of the occurrences of Anopheles vectors, the risk from malaria to human
40 health in Europe remains low due to economic and social development and access to health care (medium
41 confidence) (Sudre et al., 2013; Hertig, 2019).

42

43 Water-borne diseases are also associated with changes in climate such as heavy precipitation events
44 (Semenza, 2020). Warming has been linked with elevated incidence of campylobacteriosis outbreaks in
45 different European countries (Yun et al., 2016; Lake et al., 2019). Marine bacteria such as vibrio thrive under
46 elevated sea surface temperature and low salinity such as the Baltic Sea. Under further warming, the number
47 of months with risk of Vibrio transmission increases and the seasonal transmission window expands, thereby
48 increasing the risk to human health in the future (high confidence) (Baker-Austin et al., 2017; Semenza et al.,
49 2017).

50

51

Do Not Cite, Quote or Distribute  13-55       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.23: Assessment of climate sensitive infectious diseases. The assessment considers the main drivers of hazard
 3 (climatic impact drivers, pathogens and vectors), vulnerability (lack of safeguards and a predisposition to these hazards)
 4 and exposure (humans to be affected by these pathogens and vectors), the direction of change of climatic suitability (i.e.
 5 temperature, precipitation, relative humidity, extreme weather events) of observed changes and at 1.5癈 and 3癈 GWL,
 6 and assesses the overall infectious disease risks across Europe (Chapter 7.3 and 7.4, Lindgren et al., 2012; Semenza and
 7 Paz, 2021). The assessment does not consider incidence of disease infections through autochthonous transmission.
 8 (Table SM13.18).

 9

10

11 13.7.1.4. Allergies and Pollen

12

13 The main drivers of allergies are predominantly non-climatic (e.g. increased urbanization, adoption of
14 westernized lifestyles, social and genetic factor) but climate change strongly contributes to the spread of
15 some allergenic plants, thus exacerbating existing and causing new allergies to humans across Europe (high
16 confidence) (D'Amato et al., 2016; EASAC, 2019). The prevalence of hay fever (allergic rhinitis), for
17 example, is between 4% and 30% among European adults (Pawankar et al., 2013). The invasive common
18 ragweed is a key species already causing major allergy in late summers (including hay fever and asthma),
19 particularly in Hungary, Romania and parts of Russia (Ambelas Skj鴗h et al., 2019). Across Europe,
20 sensitization to ragweed is expected to increase from 33 million people in 1986�2005 to 77 million people at
21 2癈 GWL (Lake et al., 2017).

22

23 Warming will result in an earlier start of the pollen season and extending it, but this differs across regions,
24 species, traits and flowering periods (Ziello et al., 2012; Bock et al., 2014; EASAC, 2019; Revich et al.,
25 2019). For instance in different parts of WCE and NEU, the start of the birch season flowering has been
26 shifted and extended up to two weeks earlier during recent decades (Biedermann et al., 2019). Airborne
27 pollen concentrations are projected to increase across Europe (Ziello et al., 2012). In south-eastern Europe,
28 where pollen already has a substantive impact, the pollen count could increase more than 3 to 3.5 times at

Do Not Cite, Quote or Distribute  13-56       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 2.5癈 GWL and can become a more widespread health problem across Europe, particularly where it is
 2 currently uncommon (medium agreement, low evidence) (Lake et al., 2017).

 3

 4 13.7.1.5. Labour Productivity and Occupational Health

 5

 6 Extreme heat and cold waves have been linked to an increased risk of occupational injuries (Martinez-
 7 Solanas et al., 2018) and changes in labour productivity (Orlov et al., 2019; Garc韆-Le髇 et al., 2021) while
 8 evidence on the consequences of other extreme events is lacking. The sectors with a high percentage of high
 9 intensity outdoor work in Europe, mainly agriculture and construction, have the highest risk of increased
10 injury and labour productivity losses, but also manufacturing and service sectors can be affected when air
11 conditioning is not available (Gosling et al., 2018; Szewczyk et al., 2018; Dellink et al., 2019; Orlov et al.,
12 2019) (13.6.1.3). The heatwaves of August 2003, July 2010 and July 2015 concentrated in SEU and led to
13 reductions in monthly worker productivity of on average 3-3.5% in SEU, ranging up to 8-9% in Cyprus
14 (2003, 2010) and in Italy (2015) (Orlov et al., 2019); in contrast, the heatwave of 2018 centred on NEU, but
15 also led to pronounced productivity reductions in WCE and SEU (Garc韆-Le髇 et al., 2021); each of these
16 major European heatwaves led to considerable economic losses in agriculture and construction (high
17 confidence) and reduced GDP in Europe (except EEU) by 0.3-0.5% (Garc韆-Le髇 et al., 2021). At 2.5癈
18 GWL and beyond, GDP losses are projected to increase fivefold compared to 1981-2010, ranging from 2-
19 3.5% in SEU, to 0.5-1.5% in WCE, and below 0.5% in NEU and EEU (Roson and Sartori, 2016; Takakura et
20 al., 2017; Szewczyk et al., 2018; Dellink et al., 2019; Garc韆-Le髇 et al., 2021) (13.10.3).

21

22 13.7.1.6. Food Quality and Nutrition

23

24 There is growing evidence that climate change will negatively affect food quality (diversity of food, nutrient
25 density, and food safety) and food access, although the risks for European citizens are significantly lower
26 compared to other regions (Fanzo et al., 2018; IFPRI, 2018). Projected changes in crop and livestock
27 production (13.5.1), particularly reduced access to fruits and vegetables and foods with lower nutritional
28 quality, will impact already vulnerable groups (Swinburn et al., 2019). The effects of climate change on food
29 quality and access varies by income, livelihood, and nutrient requirements, with low income and more
30 vulnerable societal groups in Europe most affected (IFPRI, 2018). Spikes in food prices due to changing
31 growing conditions in Europe (13.5.1), increased competition for land (e.g., land-based climate change
32 mitigation), and feedbacks from international markets, are expected to decrease access to affordable and
33 nutritious food for European citizens (13.9.1) (EASAC, 2019; Loopstra, 2020). Reduced access to healthy
34 and varied food could contribute to overweight and obesity which is a growing health concern across
35 European countries (Springmann et al., 2016). Increased rates of obesity and diabetes further exacerbate
36 risks from heat-related events (EASAC, 2019).

37

38 13.7.1.7. Mental Health and Wellbeing

39

40 Extreme weather events can trigger post-traumatic stress disorder (PTSD), anxiety and depression; this is
41 well-documented for flooding in Europe (high confidence), but less for other extreme weather events. For
42 example, in the UK, flooded residents suffered stress and identity loss from the flood event itself, but also
43 from subsequent disputes with insurance and construction companies (Carroll et al., 2009; Greene et al.,
44 2015). Residents displaced from their homes for at least one year due to 2013-2014 floods in England were
45 significantly more likely to experience PTSD, depression and anxiety, with stronger effects in the absence of
46 advance warning (Munro et al., 2017; Waite et al., 2017). There is emerging evidence across Europe that
47 young people may be experiencing anxiety about climate change, though it is unclear how widespread or
48 severe this is (Hickman, 2019). In northern Italy, the number of daily emergency psychiatric visits and mean
49 daily air temperature has been linked (Cervellin et al., 2014).

50

51 13.7.2 Solution Space and Adaptation Options

52

53 Adaptation to health impacts has generally received less attention compared to other climate impacts across
54 Europe (EASAC, 2019). Progress on health adaptation can be observed. Between 2012 and 2017, at least 20
55 European countries instituted new governance mechanisms, such as interdepartmental coordinating bodies
56 for health adaptation and adopted health adaptation plans (Kendrovski and Schmoll, 2019). Progress on city

Do Not Cite, Quote or Distribute  13-57       Total pages: 143
    FINAL DRAFT                                                              Chapter 13                       IPCC WGII Sixth Assessment Report

 1 level health adaptation is generally limited (Araos et al., 2015) , with most activities occurring in SEU (Paz
 2 et al., 2016) (high agreement, medium evidence).

 3

 4 Figure 13.24 presents the assessment of the feasibility and effectiveness of key heat-related health adaptation
 5 actions. It shows that substantial social-cultural and institutional barriers complicate widespread
 6 implementation of measures; studies on the implementation of new blue-green spaces in existing urban
 7 structures in, for example Sweden (Wihlborg et al., 2019), UK (Carter et al., 2018), the Netherlands (Aalbers
 8 et al., 2019), point to important feasibility challenges (e.g., access to financial resources, societal opposition,
 9 competition for space) (high confidence). Lower perception of health risks has been observed amongst
10 vulnerable groups which, in conjunction with perceived high costs of protective measures, act as barriers to
11 implementing health adaptation plans (van Loenhout et al., 2016; Macintyre et al., 2018; Martinez et al.,
12 2019). Key barriers to mental health adaptation actions include lack of funding, coordination, surveillance,
13 and training (e.g., psychological first aid) (Hayes and Poland, 2018). Existing health measures, such as
14 monitoring and early warning systems, play an important role in detecting and communicating emerging
15 climate risks and weather extremes (Confalonieri et al., 2015; Casanueva et al., 2019; Linares et al., 2020)}
16 (high confidence). Stricter enforcement of existing health regulation and policy can have a positive effect in
17 reducing risks (Berry et al., 2018).

18

19
           Effectiveness & feasibility of main adaptation options to reduce heat related impacts & risks to human health in Europe

    Impact Type       Adaptation Option           Effectiveness Feasibility                                                                       Confidence

    Mortality,                                       Economic                Technological Institutional  Socio-cultural Ecological  Geophysical  Evidence Agreement
    morbidity,                                                                                                                             NL
    exposure, stress  Behavior change measures    M  NL                      NL          H                H   NL                           NL     M           L
    from heat                                                                                                                               H
                      Natural cooling             L  NL                      M           NL               M   NL                            M     H           L
                                                                                                                                            M
                      Building interventions      M  M                       M           NL               M   NL                           NL     H           L
                                                                                                                                            H
                      Green infrastructures       M  M                       M           M                M   H                                   M           H

                      Heat proof land management  H  NL                      M           M                H   H                                   M           M

                      Heat health action plans    H  NL                      M           H                H   NL                                  H           H

                      Bundle of options           H  H                       NL          NL               NL  NL                                  M           L

    Legend

     High = H

     Medium = M

     Low = L

20  No/Limited Evidence = NL

21 Figure 13.24: Effectiveness and feasibility of the main adaptation options to reduce heat related impacts and risks to

22 human health in Europe. (SM13.9 and Table SM13.19).

23

24

25 The effectiveness of most options in reducing climate induced health risks is determined by many co-
26 founding factors, including the extent of the risk, existing socio-political structure and culture, and other

27 adaptation options in place (high agreement, medium evidence). Successful examples include the

28 implementation of heat wave plans (Schifano et al., 2012; van Loenhout and Guha-Sapir, 2016; De'Donato et
29 al., 2018), improvements in health services, and infrastructure of homes (Vandentorren et al., 2006) (Section
30 13.10.2.1). A study of nine European cities, for example, showed lower numbers of heat-related deaths in
31 Southern European cities, and attributed this to the implementation of heat prevention plans, a greater level
32 of individual and household adaptation, and growing awareness of citizens about exposure to heat
33 (de'Donato et al., 2015). Long-term national prevention programs in Northern Europe have been shown to

34 reduce temperature-related suicide (Helama et al., 2013). Physical fitness of individuals may increase

35 resilience to extreme heat (Schuster et al., 2017). Combining multiple types of adaptation options into a
36 consistent policy portfolio may have an amplifying effect in reducing risks, particularly at higher GWL

37 (Lesnikowski et al., 2019) (medium confidence) (Chapter 7).

38

39 Health adaptation actions have demonstrable synergies and trade-offs (Cross-Chapter Box HEALTH in

40 Chapter 7). For example, increasing green-blue spaces in Europe's densely populated areas can be effective

41 in improving micro-climates, reducing the impact of heat waves, improving air quality, and improving
42 mental health by increasing access to fresh air and green (restorative) environment (Gascon et al., 2015;
43 Kondo et al., 2018; Kumar et al., 2019). Health adaptations can also have negative trade-offs, be inconsistent
44 with mitigation ambitions, and could lead to maladaptation. Green-blue spaces, for example, may create new

45 nesting grounds for carriers of vector-borne diseases, increase pollen and allergies (Kabisch et al., 2016),

    Do Not Cite, Quote or Distribute                                         13-58                                                                Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 enlarge freshwater use for irrigation (Reyes-Paecke et al., 2019), and could raise climate equity and justice
 2 issues such as green gentrification (Yazar et al., 2019). Similarly, air conditioning and cooling devices are
 3 considered highly effective but have low economic and social feasibility and negative trade-offs due to
 4 increasing energy consumption, raising energy costs which is particularly challenging for the poor (Section
 5 13.8.1.1), enhancing the heat island effect, and increasing noise pollution (Fernandez Milan and Creutzig,
 6 2015; Hunt et al., 2017; Macintyre et al., 2018).

 7

 8 The solution space for implementing health adaptation options is slowly expanding in Europe. Health
 9 adaptation can build on, and integrate into, established health system infrastructures but these differ
10 significantly across Europe, as do existing capacities to deal with climate-related extreme events (Austin et
11 al., 2016; Austin et al., 2018; Orru et al., 2018; Watts et al., 2018; Austin et al., 2019; Martinez et al., 2019).
12 Despite some progress, limited mainstreaming of climate change is observed, particularly due to low societal
13 pressure to change, confidence in existing health systems, and lack of awareness of links between human
14 health and climate change (medium confidence) (Austin et al., 2016; WHO, 2018b; Watts et al., 2021).
15 Coordination of health adaptation actions across scales and between public sectors is needed to ensure timely
16 and effective responses for a diversity of health impacts (high confidence) (Austin et al., 2018; Ebi et al.,
17 2018). Key enabling conditions to extend the solution space include increasing the role for national and
18 regional governments in facilitating knowledge sharing across scales, allocating dedicated financial
19 resources, and creating dedicated knowledge and policy programs on climate and health (Wolf et al., 2014;
20 Akin et al., 2015; Curtis et al., 2017). Investing in public healthcare systems more broadly increases their
21 capacity to respond to climate-related extreme events and will ensure wider societal benefits as the COVID-
22 19 pandemic has demonstrated (Cross-Chapter Box COVID in Chapter 7).

23

24 Despite a range of options available, there are limits to how much adaptation can take place and residual
25 risks remain. These are predominantly discussed in the context of excess mortality and morbidity to heat
26 extremes (Hanna and Tait, 2015; Martinez et al., 2019). Future heat waves are expected to stretch existing
27 adaptation interventions well beyond levels observed in response to the observed events of 2003 and 2010
28 (Hanna and Tait, 2015), see Section 13.10.2.1.

29

30 13.7.3 Knowledge Gaps

31

32 Literature on the link between public health, climate impacts, vulnerability and adaptation is skewed across
33 Europe, with most studies focusing on region-specific impacts (e.g., flood injuries in WCE, heatwaves in
34 SEU). In general, attributing health impacts to climate change remains challenging, particularly for mental
35 health and wellbeing, (mal)nutrition and food quality, and climate sensitive infectious diseases, where other
36 socio-economic determinants play an important role. The connection between climate change and health
37 risks under different socio-economic development pathways is hardly studied comprehensively for Europe,
38 with some exceptions for extreme events. However, these interactions seem to play an important role in
39 better understanding projected risks and inform choices on adaptation planning.

40

41 Some climate-related health issues are emerging but evidence is too limited for a robust assessment, for
42 example the links between climate change and violence in Europe (Fountoulakis et al., 2016; Mares and
43 Moffett, 2016; Sanz-Barbero et al., 2018; Koubi, 2019).

44

45 The solution space for public health adaptation in Europe, and the effectiveness of levers for interventions,
46 are hardly assessed. Although health adaptations are documented, these are particularly around mortality and
47 injuries due to extreme events (predominantly floods (13.2.1) and heat waves (13.7.1.1). There are very few
48 studies assessing the barriers and enablers of health adaptations, nor systematic assessment of the
49 effectiveness of (portfolio of) options. Limited insights into what works and where hamper upscaling these
50 insights across Europe and constrains the ability to evaluate whether investments in health adaptation have
51 actually reduced risks.

52

53

54 13.8 Vulnerable Livelihoods and Social Inequality

55

Do Not Cite, Quote or Distribute  13-59       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 This section addresses social consequences of climate change for Europe, by looking into consequences for
 2 poor households and minority groups; migration and displacement of people; livelihoods particularly
 3 vulnerable to climate change (indigenous and traditional communities); and cultural heritage.

 4

 5 13.8.1 Observed Impacts and Projected Risks

 6

 7 13.8.1.1 Poverty and Social Inequality

 8

 9 While climate change is not the main driver of social inequality in Europe, poor households and
10 marginalized groups in Europe are affected more strongly than other social groups by flooding, heat and
11 drought and risks to spreading diseases (medium confidence).

12

13 Urban poor and ethnic minorities often settle in more vulnerable settlement zones, and are therefore
14 impacted more by flooding (medium confidence) (Medd et al., 2015; Zupari-Ilji, 2017; Efendi, 2018;
15 Fielding, 2018; Winsemius et al., 2018; Puak, 2019; Inuit Circumpolar Council, 2020). Yet, in some
16 western European residential waterside developments this pattern is reversed by flooding impacting high
17 income residents more strongly (Walker and Burningham, 2011).

18

19 The health of the poor is disproportionately affected e.g. during heat waves in the Mediterranean (Jouzel and
20 Michelot, 2016). Women, those with disabilities and the elderly are disproportionately affected by heat
21 (Section 13.7.1). Floods in the Western Balkans in 2014 resulted in heavy metal pollution of water and land
22 threatening the health condition of the poorer rural population (Filijovi and orevi, 2014). Access to
23 water and sanitation is less available to poorer households and marginalized groups in Europe (Ezbakhe et
24 al., 2019; Anthonj et al., 2020) which could be intensified by increasing water scarcity in parts of Europe
25 under future climate change (Section 13.10.3).

26

27 Food self-provisioning is a widespread practice in many parts of Europe (Aleynikov et al., 2014; Corcoran,
28 2014; Church et al., 2015; Mustonen and Huusari, 2020), reaching over half of German rural areas (V醰ra et
29 al., 2018). While it strengthens resilience for disadvantaged households (Church et al., 2015; Boost and
30 Meier, 2017; Promberger, 2017; V醰ra et al., 2018; Anci et al., 2019; Pungas, 2019) and renews their local
31 knowledge, it can become at risk in regions with projected crop yield reductions (high confidence)
32 (Hallegatte et al., 2016; Quiroga and Su醨ez, 2016; Myers et al., 2017b; Inuit Circumpolar Council, 2020),
33 and after extreme weather events (Filijovi and orevi, 2014).

34

35 Energy poor households often live in thermally inefficient homes and cannot afford air conditioning to adapt
36 to overheating in summer (Sanchez-Guevara et al., 2019; Thomson et al., 2019). While energy poverty is
37 much more prevalent in southern and eastern Europe (Bouzarovski and Petrova, 2015; Pye et al., 2015;
38 Atsalis et al., 2016; Monge-Barrio and S醤chez-Ostiz Guti閞rez, 2018), climate change will also exacerbate
39 energy poverty in European regions where heating was so far the major share of energy costs (medium
40 confidence) (Sanchez-Guevara et al., 2019; Randazzo et al., 2020).

41

42 13.8.1.2 Migration and Displacement of People

43

44 Most migration and displacement due to climate change is taking place within national borders and single
45 regions (Cross-Chapter Box MIGRATE in Chapter 7). There is low confidence in climate change
46 contributing to migration from outside Europe into Europe (Gemenne, 2011; Topilin, 2016; Gemenne and
47 Blocher, 2017; Selby et al., 2017). Some economic models project that asylum applications to the EU might
48 increase by a third at 2.5癈 GWL and more than double beyond 4癈 GWL by end of the century (Missirian
49 and Schlenker, 2017), but empirical evidence shows that applications might decrease due to growing
50 economic and legal barriers in the capacity of populations to emigrate from Africa or other regions (Kelley et
51 al., 2015; Zickgraf, 2018; Borderon et al., 2019).

52

53 Migration of people within Europe is predominantly triggered by economic disparities among European
54 countries (Fischer and Pfaffermayr, 2018). There is limited and inconclusive evidence for climate-driven
55 impacts on these movements (Hoffmann et al., 2020). Small scale climate-induced displacement within
56 Europe occurs in the aftermath of flood and drought disasters and over short distances (Cattaneo et al.,
57 2019). The unequal distribution of future climate risks (13.1) and adaptive capacity across European regions

Do Not Cite, Quote or Distribute  13-60       Total pages: 143
    FINAL DRAFT                          Chapter 13                                IPCC WGII Sixth Assessment Report

 1 may increase pressure for internal migration (Williges et al., 2017; Forzieri et al., 2018). For instance,
 2 projected sea level rise (13.2.1, Cross-Chapter Box SLR in Chapter 3) may result in planned relocation of
 3 coastal settlements and inland migration in the UK, the Netherlands and the northern Mediterranean
 4 (Mulligan et al., 2014; Antonioli et al., 2017). The number of people living in areas at risk in Europe is
 5 projected to increase with future SSPs increasing exposure (Merkens et al., 2016; Byers et al., 2018;
 6 Harrison et al., 2019).

 7

 8 13.8.1.3 Loss and Damage to Vulnerable Livelihoods in Europe

 9

10 A number of livelihoods maintaining unique cultures in Europe is particularly vulnerable to climate change
11 (Table 13.2): indigenous communities in the European polar region because of their dependence on
12 cryosphere ecosystems (high confidence) (CCP Polar, Hayashi, 2017; Huntington et al., 2017; Hock et al.,
13 2019; Meredith et al., 2019; Inuit Circumpolar Council, 2020; Douville et al., 2021; Fox-Kemper et al.,
14 2021) and communities dependent on small-scale fisheries, traditional farming and unique cultural
15 landscapes (medium confidence) (Kovats et al., 2014; Ruiz-D韆z et al., 2020).

16

17 For S醡i reindeer herding impacts cascade due to a lack of access to key ecosystems, lakes and rivers
18 thereby threatening traditional livelihoods, food security, cultural heritage (e.g. burial grounds, seasonal
19 dwellings and routes), mental health (Box 13.2 and Figure 13.13, Feodoroff, 2021), and growing costs for
20 example as a result of the need for artificial feeding of reindeer.

21

22

23 Table 13.2: Examples of losses and damages to vulnerable livelihoods in Europe, differentiating for different categories
24 of non-economic loss and damage. (Table SM13.20.).

    Human life                                          Communal and production sites and intrinsic value

    Sense of place                                      Agency and identity

    Cultural artefacts                                  Psychological and emotional distress

    Biodiversity and ecosystems

25

                                         Change in exposure and                    Observed impact / projected

    Climate hazard                       vulnerability                             risk

    Loss of livelihood, culture, health and wellbeing of the S醡i and the Nenets.

    Decrease and alterations in snow     Land-use change (e.g. expansion of        Loss of livelihood (e.g. reindeer
    and ice sheet, unstable winter       renewable energy) resulting in            herding), loss of food security (cold
    weather, especially in the form of   pasture loss and disconnection of         dependent species), culture, health
    rain-on-snow events; increased       ecosystems.                               (impact on safety; psychological
    precipitation and thawing                                                      impacts from stress to reindeer and
    permafrost, in tundra; unstable                                                Indigenous way of life), and
    loss/flux of marine ice cover.                                                 cultural and linguistic wellbeing;
                                                                                   release of anthrax from permafrost
                                                                                   soils in the Nenets area.

    Loss of key species in high-Arctic freshwater habitats, proliferation of introduced species and disruption of local
    food systems in Greenland, Finland, Sweden, NW Russia and Scotland.

    Warmer water temperatures in         Introduced Pacific Pink Salmon has        Shifts in freshwater aquatic habitats
    high-Arctic freshwater habitats      expanded in range since 1970s,            and loss of endemic cold-dependent
    (13.3.1) increase productivity in    affecting endemic species through         fish, such as Arctic Char and Arctic
    oligotrophic systems and eventually  competition and reducing their            Salmon, cause disruptions to local

    Do Not Cite, Quote or Distribute                    13-61                                 Total pages: 143
FINAL DRAFT                                         Chapter 13               IPCC WGII Sixth Assessment Report

lead to loss of oxygen in water;      abundance. Increased nutrient          food supply, and local extinctions

warming temperatures and changes loading of rivers and rapid                 threatens livelihood safety and

to ice cover and cryosphere lead to expansion of algae increase the          cultural well-being.

access issues to freshwater           risks for cold-dependent fish.

fisheries.

Warmer winters lead to loss of income from ice fishing and cultural heritage in Finland.

The start of ice cover on lakes, e.g. The quality of the water in the lakes Lack of winter ice combined with

lake Puruvesi (Finland), has          used for fishing depend on ice         delayed freeze up and earlier ice

changed from November to              cover during most of the year, and breakup reduce fish harvest for

February; ice breakup occurs much the season of open water is now            important species by up to 50% and

earlier in the year.                  much longer, increasing nutrient impacts local safety, ecosystems,

                                      flow and loss of water quality in      oral history maintenance and the

                                      these lakes.                           local economy.

Changes to marine food web results in loss of Indigenous knowledge and food insecurity in Greenland.

Warmer ocean waters moving            Traditional practices and              Loss of Indigenous knowledge of
further north (so-called              knowledge based on sea ice uses        how to deal with and use sea ice
"atlantification" of Greenland        and hunting are being lost; species    regarding species and navigation,
waters); higher temperatures          are being replaced with southern       loss of access to seals and walruses,
removing sea ice                      fish.                                  food insecurity.

Reduced yields on managed alpine grasslands decreases the self-sufficiency of pastoral livestock farming in the
Austrian, French and Swiss Alps

Increase in heat, precipitation       Land-use change resulting in           Abandonment of summer pastures
variability and agricultural as well  natural reforestation of abandoned     and farms, with negative
as hydrological drought; less snow    pastoral land; shifts in alpine plant  consequences for farming income,
on the ground, increase in glacier    communities; more intensive            tourism, cultural and aesthetic
melt, landslides susceptibility and   cultivation of grasslands; change in   values.
erosion                               agricultural markets and support
                                      policy.

Reduced yields on semi-natural grasslands, compromising livestock feeding in winter, and ultimately decreasing
viability of pastoralism in the Spanish Pyrenees

Higher temperatures and more          Demographic change, change in          Decreasing viability of pastoralism,
variable precipitation, less snow,    policy and market conditions,          concentration of pastoral
change in seasonality and drought     simplification of pastoral practices   production on most profitable
                                      and agroecosystems, land               locations for intensive rearing of
                                      abandonment or afforestation of        livestock with abandonment of the
                                      marginal pastoral lands and            rest of the land; pastoral land
                                      intensification of more favourable     encroachment both by shrubs and
                                      lands in the lowlands, troublesome     other activities; grassland
                                      coexistence with tourism and nature    degradation; biodiversity loss
                                      conservation initiatives

Retreating glaciers and changes in the landscape lead to loss of identity, culture and self-reliance in the Italian
Alps (Alto Adige)

Do Not Cite, Quote or Distribute                    13-62                                          Total pages: 143
   FINAL DRAFT                               Chapter 13                           IPCC WGII Sixth Assessment Report

   Glacier volume loss from               Vulnerability is mainly driven by       Loss of sense of community
   increasing temperatures                reliance on tourism                     through shared memories, and
                                                                                  history. Sadness caused by the loss
                                                                                  of what feels like "home". Loss of
                                                                                  well-being due to uncertainty and
                                                                                  fear of the future.

   Drought results in a reduction of provisioning (water) and regulating services (protection against floods) in
   Western and Eastern Alps, Iberian Mountains, Dinaric Mountains

   Increase in drought, particularly      Forest management strategies,           Critical importance of alpine

   under high-end GWL                     including that of natural forests, can natural forests and meadows for

                                          enhance or reduce vulnerability.        regulating services; Negative

                                                                                  impacts of climate change are

                                                                                  found mainly at low elevations and

                                                                                  for specific species (Norway

                                                                                  spruce); decrease in soil moisture

                                                                                  due to abandonment of pastoralism

                                                                                  result in reduced water provision

                                                                                  for downstream water users

   Increase of sea temperature leads to shifts in distribution of cold water species, reducing productivity at lower

   latitudes. Artisanal fisheries in Southern European coastal areas (Mediterranean) that rely on local, nearshore

   stocks can have difficulties to adapt

   Increase in sea temperature            Substitution of artisanal fisheries by  Due to their low investment
                                          industrial fisheries; less support by   capacity and boat size, fishers are
                                          governments, shift in employment        limited in their movement to other
                                          (e.g. tourism) which do not match       fishing places when local fish
                                          the skill sets, education or desires    stocks decline. Increasing sea
                                          of small-scale fishers; national        temperatures are increasing the
                                          quotas system leads to prices to        threat of invasive species in coastal
                                          high buy or lease quotas and            ecosystems.
                                          immense amount of bureaucracy
                                          and regulations

1

2

3 [START BOX 13.2 HERE]

4

5 Box 13.2: S醡i Reindeer Herding in Sweden

 6

 7 Reindeer (Rangifer tarandus) are keystone species in northern landscapes (Vors and Boyce, 2009). Reindeer
 8 herding is a traditional, semi-nomadic livelihood of the S醡i. Reindeer migrate between seasonal pastures
 9 that cover 55% of Sweden and are simultaneously used for multiple other purposes (Sandstr鰉 et al., 2016).
10 Reindeer herding is recognized as an indigenous right, protected by the UN Declaration on the Rights of
11 Indigenous Peoples, several UN conventions and through Swedish national legislation.

12

13 Temperatures in Arctic and sub-Arctic regions have increased on average by 2癈 over the last 30 years (very
14 high confidence) (Ranasinghe et al., 2021). Future warming is expected to further increase winter
15 precipitation (high confidence) (Ranasinghe et al., 2021) and rain-on-snow events, creating a hard ice crust
16 on the snow after refreezing (Bokhorst et al., 2016; Rasmus et al., 2018).

17

18 The documented and projected impacts on reindeer are complex and varied. Warming and CO2 increase
19 result in higher plant productivity (Section 13.3), changes in plant community composition, and higher
20 parasite harassment; unstable ice conditions affect migration; extreme weather conditions during critical
21 winter months, more frequent forest fires and changes in plant community composition reduce pasture
22 quality (medium confidence) (Mallory and Boyce, 2018) (Figure Box 13.2.1). High snow depth and rain-on-

   Do Not Cite, Quote or Distribute          13-63                                Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

1 snow events impede reindeer access to ground lichen in winter and delay spring green-up during critical
2 calving period; both cause malnutrition and negative impacts on reindeer health, mortality and reproductive
3 success (medium confidence) (Hansen et al., 2014; Forbes et al., 2016; Mallory and Boyce, 2018). Lower
4 slaughter weights and increased mortality reduce the income of herders (high confidence) (Tyler et al., 2007;
5 Helle and Kojola, 2008).

6

7

 8

 9 Figure Box 13.2.1: Cumulative impacts of climate and land use change on reindeer herding as a traditional, semi-
10 nomadic S醡i livelihood (Table SM13.21).

11

12

13 Reindeer herders already autonomously adapts to changing conditions through flexible use of pastures and
14 supplementary feeding (high confidence), reducing and thereby hiding some of the negative impacts of
15 climate change (Uboni et al., 2016). However, adaptive herding practices have themselves added significant
16 burden through increased workload, costs and stress (high confidence) (Furberg et al., 2011; L鰂, 2013;

Do Not Cite, Quote or Distribute  13-64       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 Rosqvist et al., 2021). Supplementary feeding increases risk for infectious diseases and implies culturally
 2 undesirable herding practices (low confidence) (Lawrence and Kl鴆ker Larsen, 2019; Tryland et al., 2019).

 3

 4 Rapid land use change reduces the ability to adapt (high confidence) (Tyler, 2010; L鰂, 2013). National and
 5 EU policies expand land uses for mining, wind energy and bioeconomy in the area, causing loss,
 6 fragmentation and degradation of pastures, and increasing human disturbance to animals (medium
 7 confidence) (Kivinen et al., 2012; Skarin and 舎man, 2014; Kivinen, 2015; Skarin et al., 2015; Sandstr鰉 et
 8 al., 2016; Beland Lindahl et al., 2017; 謘terlin and Raitio, 2020). The cumulative impacts of these land-uses
 9 on pastures are not adequately assessed or recognized in land use planning (Kl鴆ker Larsen et al., 2017;
10 Kl鴆ker Larsen et al., 2018 ). Herding communities face strong barriers to protecting their rights and halting
11 further degradation of pastures (medium confidence) (Allard, 2018; Kl鴆ker Larsen and Raitio, 2019; Raitio
12 et al., 2020). Attempts by herding communities to stop mining projects have led to conflicts with other
13 actors, including racist hate speech (Persson et al., 2017; Beland Lindahl et al., 2018). Combined with land
14 use conflicts, climate impacts cause reduced psycho-social health and increase suicidal thoughts among
15 herders (low confidence) (Kaiser et al., 2010; Furberg et al., 2011).

16

17 Reindeer herding is significantly affected by climate change directly and indirectly (Figure Box 13.2.1)
18 (Pape and L鰂fler, 2012; Andersson et al., 2015). The cumulative effects of land use and climate change
19 have already increased vulnerability and reduced the adaptive capacity of reindeer herding to the extent that
20 its long-term sustainability is threatened (medium confidence) (L鰂, 2013; Horstkotte et al., 2014; Kl鴆ker
21 Larsen et al., 2017).

22

23 Maintaining and improving the solution space to adapt reindeer herding is crucial for reducing existing
24 impacts and projected risks of climate and land use change (Andersson et al., 2015; Turunen et al., 2016;
25 AMAP, 2017; Hausner et al., 2020). Lack of control over land use is the biggest and most urgent threat to the
26 adaptive capacity of reindeer herding and the right of S醡i to their culture (high confidence) (Pape and
27 L鰂fler, 2012; Andersson et al., 2015; Kl鴆ker Larsen and Raitio, 2019).

28

29 [END BOX 13.2 HERE]

30

31

32 13.8.1.4 Cultural and Natural Heritage

33

34 Climate change poses a serious threat to preservation of cultural heritage in Europe, both tangible and
35 intangible (high confidence) (Haugen and Mattsson, 2011; Daire et al., 2012; Dupont and Van Eetvelde,
36 2013; Macalister, 2015; Phillips, 2015; Fatori and Seekamp, 2017; Graham et al., 2017; Carroll and
37 Aarrevaara, 2018; Sesana et al., 2018; Iosub et al., 2019; Daly et al., 2020). At higher GWL, building
38 exteriors and valuable indoor collections become at risk (Leissner et al., 2015). Coastal heritage such along
39 the North Sea and Mediterranean are under water-related threats (Reimann et al., 2018b; Walsh, 2018;
40 Harkin et al., 2020) (Box 13.1 Venice; WGII AR6 CCP4).

41

42 Disappearing cultural heritage can reduce incomes due to loss of tourism (Hall et al., 2016), as exemplified
43 by glacier retreat e.g. in the Swiss Alps and Greenland (Bjorst and Ren, 2015; Bosson et al., 2019)
44 (CCP5.3.2.4). Glacier retreat can create a sense of discomfort, loss of sense of place, displacement and
45 anxiety in people (Section 13.7) (Albrecht et al., 2007; Brugger et al., 2013; Allison, 2015; Jurt et al., 2015).
46 Intangible cultural heritage, such as place names, and lost traditional practices can also be affected
47 (Mustonen, 2018; Dastgerdi et al., 2019).

48

49 13.8.2 Solution Space and Adaptation Options

50

51 As climate change is interacting with many other drivers of poverty, improving the social position of the
52 currently poor may increase their climate resilience (low confidence) (Hallegatte and Rozenberg, 2017;
53 Fronzek et al., 2019). Some adaptation actions have the potential to alleviate poverty (Section 13.11.3), but
54 adaptation can also increase social inequalities, e.g. when practices of disaster recovery focus on high
55 visibility areas and not on low-income neighbourhoods or marginalized spaces (D'Alisa and Kallis, 2016).
56 Risk communication and management reliant on new information technologies can exclude elderly and
57 populations with lower educational attainment (Kesetovi et al., 2017).

Do Not Cite, Quote or Distribute  13-65       Total pages: 143
    FINAL DRAFT                       Chapter 13     IPCC WGII Sixth Assessment Report

1

2 Unlike migration within the European Union, migration from outside Europe to Europe is heavily
3 constrained by restrictive migration and asylum policies (Fielding, 2011; Mulligan et al., 2014), eventually

4 leaving people to stay in more exposed and risk-prone regions (Benveniste et al., 2020). To reduce

5 vulnerability in these regions, Europe can contribute to adaptation and development in regions outside
6 Europe (section 13.9.4).

7

8 Indigenous and local knowledge, embedded e.g. in fishermen, farmers and navigators, can be a vehicle for

 9 detecting, monitoring and observing impacts (Arctic Council, 2013; Brattland and Mustonen, 2018; Madine
10 et al., 2018; Meredith et al., 2019) (section 13.11.1.3). Regarding risks to northern traditional livelihoods and
11 indigenous communities, small-scale adaptation is taking place, for example by ecological restoration of
12 habitats (section 13.3) (Mustonen and Kontkanen, 2019). However, limited access to resources outside the
13 jurisdictions of the communities limits the scope of community-based adaptation (Arctic Council, 2013;
14 Mustonen et al., 2018; Meredith et al., 2019).

15

16 European cultural heritage in general and world heritage sites specifically lack adaptation strategies to
17 preserve key cultural assets (Haugen and Mattsson, 2011; Howard, 2013; Heathcote et al., 2017; Reimann et

18 al., 2018b; Harkin et al., 2020). Key reasons are the underdeveloped adaptation actions available, resources
19 for implementing them, and absence of overarching policy guidance (Phillips, 2015; Fernandes et al., 2017;
20 Sesana et al., 2018; Daly et al., 2020) (Sesana et al., 2018; Fatori and Biesbroek, 2020; Sesana et al., 2020).

21

22 13.8.3 Knowledge Gaps

23

24 There is limited understanding of how different social groups are affected by the four European key risks

25 under future climate change (13.11.2), and by adaptation to them. Similarly, the interaction of multiple risks

26 across sectors and how this interaction results in displacement, migration, or immobility of people both
27 within and from outside Europe is insufficiently understood. For indigenous and traditional livelihoods in
28 Europe, the understanding of how risks will change at different warming levels is very limited, due to
29 complex interactions with socio-economic and political change. For European cultural heritage, there is also
30 a lack of tailored knowledge and understanding of the impacts and how to translate these into adaptation
31 measures.

32

33

34 13.9 Interregional Impacts, Risks and Adaptation

35

36 This section addresses interregional risks between Europe and other parts of the world. Global risk pathways
37 affecting sectors and supply chains relevant for European economies and societies involve (1) ecosystems,

38 (2) people (e.g., through migration), (3) financial flows, and (4) trade, and these pathways ultimately impact

39 security, health, wellbeing and food supply (Yokohata et al., 2019) (Cross-Chapter Box INTEREG in
40 Chapter 16).

41

42 13.9.1 Consequences of Climate-change Driven Impacts, Risks and Adaptation Emerging in Other Parts

43  of the World for Europe

44

45 Recent literature (Wenz and Levermann, 2016; Hedlund et al., 2018; Benzie et al., 2019) strengthens the

46 confidence in the AR5 statement that "with increasing globalization, the impacts of climate change outside
47 the European region are likely to have implications for countries within the region" (Kovats et al., 2014). The
48 exposure of European countries to trans-European climate impact and risk pathways varies depending on
49 their territorial settings, national policies and position in the global supply chain (high confidence) (Berry et
50 al., 2015; Hedlund et al., 2018; Benzie et al., 2019). There is limited evidence that Europe is more exposed to

51 interregional risks than North America, and less than Africa and Asia (Hedlund et al., 2018). The social and
52 governance context in Europe make the region less vulnerable to conflicts driven by climate change than
53 other regions, at least up to 2癈 GWL (Buhaug et al., 2014; Mach et al., 2019; Ide et al., 2020).

54

55 Climate risks in other parts of the world can be transmitted to European economies via trade networks

56 (Figure 13.25). European agricultural imports exert a high water footprint in originating countries already
57 today (Dolganova et al., 2019; Ercin et al., 2019), and some crop imports such as tropical fruits are highly

    Do Not Cite, Quote or Distribute  13-66          Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 vulnerable to future climate change (Br醩 et al., 2019). Simultaneous breadbasket failures, and trade
 2 restrictions increase risks to food supply (medium confidence) (Fellmann et al., 2014; d'Amour et al., 2016;
 3 Gaupp et al., 2017; Gaupp et al., 2020). There is high confidence that the European economy could be
 4 negatively affected by supply chain disruptions due to flooding destroying facilities, heatwaves and malaria
 5 reducing productivity in labour intensive industries and regions (13.7.1), and sea level rise affecting ports
 6 and cities along coastlines (13.6.1.2) (Nicholls and Kebede, 2012; Challinor, 2016; Wenz and Levermann,
 7 2016; Hedlund et al., 2018; Koks, 2018; Szewczyk et al., 2018; Willner et al., 2018; Knittel et al., 2020;
 8 Kulmer et al., 2020; Carter et al., 2021).

 9

10

11

12 Figure 13.25: Trans-European climate risks in trade: virtual water flows embodied in agricultural imports to Europe in
13 2018 and the vulnerability to climate change of the most important crops in the originating countries (Dolganova et al.,
14 2019; Ercin et al., 2019).

15

16

17 13.9.2 Interregional Consequences of Climate Risks and Adaptation Emerging from Europe

18

19 New literature since AR5 suggests that climate risks in Europe can propagate worldwide in response to 3癈
20 GWL (medium confidence). Key concerns include climate impacts on European agriculture threatening
21 global food security (Berry et al., 2017; van der Velde et al., 2018) (Section 13.5.1) and the European
22 demand limiting the adaptation potential for ecosystems in South-America, Africa and Asia (IPBES, 2018;
23 Pendrill et al., 2019; Fuchs et al., 2020). Emerging literature suggests that coastal and riverine flood risks in
24 Europe could be amplified through the global financial system, and generate a systemic financial crisis
25 (Mandel et al., 2021) (Figure 13.26). For 3癈 GWL and without adaptation, northern Atlantic flight routes
26 and European ports are projected to be increasingly disrupted by changing winds, waves, and sea level rise
27 (Section 13.6.1.2) (Williams and Joshi, 2013; Irvine et al., 2016; Williams, 2016; Becker et al., 2018; Camus
28 et al., 2019; Verschuur et al., 2020).

29

30

Do Not Cite, Quote or Distribute  13-67       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.26: The transmission of coastal and riverine flood risks via finance flows from Europe to the Rest of the
 3 World. (Mandel et al., 2021).

 4

 5

 6 13.9.3 European Territories Outside Europe

 7

 8 European territories outside Europe are critically exposed to climate risks such as increased forest fires
 9 (Russian Siberia) (Sitnov et al., 2017) (Chapter 10), climate change-induced biodiversity losses and sea level
10 rise (UK, Spanish, Portuguese, French and Dutch overseas regions and territories) (Ferdinand, 2018; Sieber
11 et al., 2018) (Chapters 12 and 15). Climate risks emerging from these territories include smoke and dust from
12 Siberian forest fires (Sitnov et al., 2017), and, depending on European health-risk mitigation measures,
13 dengue and other mosquito-transmitted diseases (13.7) (Schaffner and Mathis, 2014). Some marine protected
14 areas (MPA, 13.4.3) in European overseas territories are increasingly affected by changes originating in far-
15 field upstream areas. These changes ultimately undermine their ability to curb biodiversity losses and
16 provide ecosystem services (Schaffner and Mathis, 2014; Robinson et al., 2017). Adaptation options and
17 regulations developed within Europe apply in these territories, despite low confidence that they meet local
18 and regional adaptation challenges and address the aspiration for social justice, promotion of local solutions
19 and consideration of traditional knowledge (Ferdinand, 2018; Terorotua et al., 2020).

20

21 13.9.4 Solution Space and Adaptation Options

22

23 European countries can address interregional risks at the place of origin or destination, e.g., by developing
24 local adaptation capacity in trading partner countries and in European territories outside Europe(Petit and
25 Prudent, 2008; Benzie et al., 2019; Adams et al., 2020; Terorotua et al., 2020), by providing international
26 adaptation finance (Dzebo and Stripple, 2015; BMUB, 2017), by developing insurance mechanisms suitable
27 for adaptation, or European climate services to support global adaptation (Linnerooth-Bayer and Mechler,
28 2015; Brasseur and Gallardo, 2016; Street, 2016; Cavelier et al., 2017) (Cross-Chapter Box INTEREG in
29 Chapter 16). Along the supply chain, risks can be reduced by trade diversification and alternative sourcing
30 (Benzie and Persson, 2019; Adams et al., 2020). Within Europe, risks can be reduced by integrating
31 interregional climate risks into national adaptation strategies and plans and mainstreaming into EU policies

Do Not Cite, Quote or Distribute  13-68       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 (e.g., Common Agricultural Policy, trade agreements) (Benzie et al., 2019; Benzie and Persson, 2019;
 2 Groundstroem and Juhola, 2019; Adams et al., 2020). There is high confidence that the exposure of
 3 European countries to interregional risks can be reduced by international governance (Dzebo and Stripple,
 4 2015; Cramer et al., 2018; Persson and Dzebo, 2019) (Cross-Chapter Paper 4), e.g., fulfilling the targets of
 5 environmental agreements such as the Convention for Biological Diversity (IPBES, 2018). There is
 6 emerging evidence that supporting adaptation outside Europe may generate economic co-benefits for Europe
 7 (Rom醤 et al., 2018).

 8

 9

10 13.10 Detection and Attribution, Key Risks and Adaptation Pathways

11

12 13.10.1 Detection and Attribution of Impacts

13

14 Since AR5, scientific documentation of observed changes attributed to global warming have proliferated
15 (high confidence). These include ecosystem changes detected in previous assessments, such as earlier annual
16 greening and onset of faunal reproduction processes, and relocation of species towards higher latitudes and
17 altitudes (high confidence), and impacts of heat on human health, and productivity (high confidence) (Figure
18 13.27 and Table SM13.22) (Vicedo-Cabrera et al., 2021). Formal attribution of impacts of compound events
19 to anthropogenic climate change is just emerging for example in the recent crop failures due to heat and
20 drought (Toreti et al., 2019a). Also, there is high agreement and medium evidence that particular events
21 attributed to climate change have induced cascading impacts and other impact interactions (Smale et al.,
22 2019; Vogel et al., 2019). In the recent decades (2000-2015), economic losses intensified in southern Europe
23 (high confidence) and were detected for parts of WCE and NEU (medium confidence). The methodology for
24 detection and attribution is presented in Chapter 16.2.

25

26

27

28 Figure 13.27: Detected changes and attribution (D&A) of climate-related impacts on land (top) and in the ocean
29 (bottom). Assessment based on peer reviewed literature in this chapter that reported observed evidence with at least
30 90% significance and usually with 95% significance or more (Table SM13.22).

31

32

Do Not Cite, Quote or Distribute  13-69       Total pages: 143
    FINAL DRAFT                           Chapter 13  IPCC WGII Sixth Assessment Report

 1 13.10.2 Key Risks Assessment for Europe

 2

 3 Key risks (KRs) are defined as a subset of climate risks which can potentially become or are already severe
 4 now (Section 16.5). The selection process included a review of KRs already identified in AR5 Chapter 23
 5 and a review of the large number of new evidence on projected risks presented in Sections 13.2-13.9. Key
 6 risks are reinforced by evidence from the detection and attribution assessment (Section 13.10.1) and new
 7 evidence from WGI AR6 Chapters 11 and 12 on regional climatic impact drivers and extremes (Ranasinghe
 8 et al., 2021; Seneviratne et al., 2021). Several expert opinion workshops of lead and contributing authors led
 9 to further refinements, adjustment and consensus building around the key risks' characteristics, which
10 ultimately guided the construction of the burning embers (Figures 13.28-13.32) (SM13.10). There is high
11 confidence that under low or medium adaptation, high to very high risks are projected at 3癎WL (Figure
12 13.28 and Sections 13.10.2.1-13.10.2.4). Most risks are assessed as moderate up to 1.5癎WL (Figure 13.28).

13

14 This section also includes an assessment of the solution space using illustrative adaptation pathways which
15 show alternative sequences of options to reduce risks as climate changes (SM13.10). Low effectiveness
16 measures are followed by measures of a higher effectiveness, while accounting for path-dependency of
17 decisions (Toreti et al., 2019b; Haasnoot et al., 2020a). The process to derive the pathways draws on
18 evidence from the feasibility and effectiveness assessments (Sections 13.2, 13.5, 13.6, 13.7).

19

20

21

22 Figure 13.28: Burning ember diagrams for low to medium adaptation. More details on each burning ember are

23 provided in Sections 13.10.2.1-13.10.2.4 and SM13.10. Some burning embers are shown again in Figure 13.29-13.34

24 alongside burning embers with high adaptation.

25

26

27 13.10.2.1 KR1: Risks of human mortality and heat stress and of ecosystems disruptions due to heat extremes

28  and increase in average temperatures

29

30 This key risk cut across humans and ecosystems and severe consequences are mainly driven by an increasing
31 frequency, intensity and duration of heat extremes and increasing average temperatures (high confidence)

32 (Urban, 2015; Forzieri et al., 2017; Feyen et al., 2020; Naumann et al., 2020; Ranasinghe et al., 2021). The
33 risk of human heat stress and mortality is largely influenced by underlying socio-economic pathways, with

34 consequences being more severe under SSP3, SSP4 and SSP5 scenarios than SSP1 (very high confidence)
35 (Figure 13.22, Section 13.6.1.5.2, Section 13.7.1.1, Hunt et al., 2017; Kendrovski et al., 2017; Rohat et al.,
36 2019; Casanueva et al., 2020). SSPs impact natural systems as well but are not yet well studied. The impact

37 of warming in marine systems are often synergistic with SLR in coastal systems and ocean acidification
38 driven by the rise in CO2, while habitat fragmentation and land use have important synergies in terrestrial
39 systems (Sections 13.3.1.2 and 13.4.1.2, high confidence). More intense heatwaves on land and in the ocean,

40 particularly in Mediterranean Europe (Section 13.4, CCP 4, Darmaraki et al., 2019b; Fox-Kemper et al.,
41 2021), are expected to cause mass mortalities of vulnerable species and species extinction, altering the
42 provision of important ecosystem goods and services (Marb� and Duarte, 2010).

43

44 The burning embers on risks for humans (Figure 13.29a) differentiate between present and medium

45 adaptation conditions, drawing on SSP2 and SSP4 (and to a lesser extent SSP3), and high adaptation

    Do Not Cite, Quote or Distribute      13-70       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 conditions, drawing on SSP1 and papers using various temperature adjustment methods (Table SM13.25).
 2 There is high confidence that the risk is already moderate now because it has been detected and attributed
 3 with high confidence (13.10.1). The transition from moderate to high risk for human health is assessed to
 4 happen after 1.5癈 GWL in a scenario with present to medium adaptation and implies 2- 3 fold increase
 5 (compared to moderate risk levels) in magnitude of consequences such as mortality, morbidity, heat stress
 6 and thermal discomfort (Rohat et al., 2019; Casanueva et al., 2020; Naumann et al., 2020). At this level, the
 7 risk will also become more persistent across the continent due to increase in heat events exceeding critical
 8 thresholds for health (Ranasinghe et al., 2021) (high confidence on the direction of change and temperature
 9 transition, but medium confidence on the magnitude).

10

11 The burning embers on risk for terrestrial and marine ecosystems and some of their services is shown in
12 Figure 13.28, second and third ember from left (Tables SM13.26 SM13.27). The transition to moderate risk
13 is currently happening as warming already results in changes in timing of development, species migration
14 northwards and upwards, desynchronization of species interactions, especially at the range limits, with
15 cascading and cumulative impacts through ecosystems and food webs (Sections 13.3 and 13.4; Figure 13.8
16 and 13.12) (high confidence). While some terrestrial ecosystems are already impacted today such as Alpine,
17 cryosphere and peatlands, the impacts are not widespread and severe yet across a wide range of terrestrial
18 systems. Around 2癈 GWL, losses accelerate in marine ecosystem and appear across systems, including
19 habitat losses especially in coastal wetlands (Roebeling et al., 2013; Clark et al., 2020), biodiversity and
20 biomass losses (Bryndum-Buchholz et al., 2019; Lotze et al., 2019) and ecosystem services such as fishing
21 (Raybaud et al., 2017) (high confidence on the direction of change, but medium confidence on the local and
22 regional magnitude). The transition is happening at slightly higher warming in terrestrial systems due to
23 higher number of thermal refugia in terrestrial systems causing relocation but not already severe impacts
24 (medium confidence) (Chapter 2).

25

26 There is medium confidence that high adaptation or conditions posing low challenges for adaptation (e.g.
27 SSP1) in the context of human health can delay the transition from moderate to high risk (舠tr鰉 et al.,
28 2017; Ebi et al., 2021). The illustrative adaptation pathways in Figure 13.29b,c show the sequencing of
29 options to a high adaptation future for NEU and SEU. Whether or not adaptation measures are effective to
30 reduce risk severity for people's health depends on local context (high confidence) (Figure 13.29, Section
31 13.6.2 and 13.7.2). Some adaptation options are found to be highly effective across Europe irrespective of
32 warming levels, including air conditioning and urban planning (high confidence) (Sections 13.6.2 and 13.7.2,
33 Jenkins et al., 2014b; Donner et al., 2015; Dodoo and Gustavsson, 2016; 舠tr鰉 et al., 2017; Dino and
34 Meral Akg黮, 2019; Venter et al., 2020), although air conditioning increasingly faces some feasibility
35 constraints (Figure 13.20). Building interventions alone have low to medium effectiveness independent of
36 the region. Many behavioural changes such as personal and home heat protection have already been
37 implemented in SEU (Section 13.7.2, Martinez et al., 2019). To reach high adaptation, a combination of low,
38 medium, and high effectiveness measures in different sectors and sub-regions is needed, many of which
39 entail systems' transformations (Chapter 16) (e.g., heat proof land management) and remain effective at
40 higher warming levels (medium confidence) (D韆z et al., 2019). These transformations have long lead times,
41 therefore requiring timely start of implementation including regions that are not yet experiencing high heat
42 stress (e.g. NEU) (high agreement, medium evidence).

43

44 Autonomous adaptation of species via migration in response to climate change is well documented in
45 contemporary, historical and geological records (Chapter 2, Cross-Chapter Box PALEO in Chapter 1).
46 However, the projected rate of climate change can exceed migration potential, leading to evolutionary
47 adaptation or increased extinction risk (Chapters 2 and 3; Sections 13.3 and 13.4). A reduction of non-
48 climatic stressors, such as nutrient loads, resource extraction, habitat fragmentation or pesticides on land, are
49 considered important adaptation options to increase the resilience to climate-change impacts (Sections 13.3
50 and 13.4, Ram韗ez et al., 2018) (high confidence). A major governance tool to reduce climatic and non-
51 climatic impacts is the establishment of networks of protected areas (Sections 13.3.2 and 13.4.2) especially
52 when aggregated, zoned or linked with corridors for migration (high confidence), as well as a cost-effective
53 adaptation strategy with multiple additional co-benefits (Berry et al., 2015; Roberts et al., 2017).
54 Reforestation, rewilding and habitat restoration are long term strategies for reducing risk for biodiversity loss
55 supported by assisted migration and evolution (Section 13.3.2, 13.4) though current laws and regulations do
56 not include species migration (high confidence) (Prober et al., 2019; Fernandez-Anez et al., 2021).

57

Do Not Cite, Quote or Distribute  13-71       Total pages: 143
    FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

1 Very high risks are expected beyond 3癈 GWL due to the magnitude and increased likelihood of serious
2 consequences, as well as to the limited ability of humans and ecosystems to cope with these impacts. There
3 is high confidence that even under high adaptation scenarios for human systems or autonomous adaptation of
4 natural systems, the risk will still be high at 3癈 GWL and beyond (Section 13.7.2, Hanna and Tait, 2015;
5 Spencer et al., 2016) with medium confidence on the temperature range of the transition. Projected sea level
6 rise will strongly impact coastal ecosystems (high confidence), minimising their contribution to shoreline
7 protection (Section 13.10.2.4).

8

9

10

11 Figure 13.29: (a) Burning ember diagrams for the risk to human health from heat. The low/medium adaptation scenario
12 corresponds to present, SSP2 and SSP4 socio-economic conditions. The high adaptation includes SSP1 and adaptation
13 needed to maintain current risk levels. (b-c) Illustrative adaptation pathways for NEU (top) and SEU (bottom) and key
14 messages based on the feasibility and effectiveness assessment in Figures 13.20 and 13.24. Grey shading means long
15 lead time and dotted lines signal reduced effectiveness. The circles imply transfer to another measure and the bars that
16 the measure have reached a tipping point. (Tables SM13.24, SM13.25)

17

18

19 13.10.2.2 KR2: Risk of losses in crop production, due to compound heat and dry conditions, and extreme

20  weather

21

22 KR2 encompasses agriculture productivity (Figure 13.30a). It is mainly driven by the increase in the

23 likelihood of compound heat and dry conditions and extremes and their impact on crops. There is high

24 confidence that climate change will increase the likelihood of concurrent extremely dry (Table SM13.28)

25 and hot warm seasons with higher risks for WCE, EEU (particularly north-western Russia) and SEU leading
26 to enhanced risk of crop failure and decrease in pasture quality (Section 13.5.1) (Zscheischler and
27 Seneviratne, 2017; Sedlmeier et al., 2018; Seneviratne et al., 2021). The risk is already moderately severe

28 due to multiple crop failures in the last decade in WCE and Russia (Section 13.5.1, Hao et al., 2018;
29 Pfleiderer et al., 2019; Vogel et al., 2019). Under high-end scenarios, heat and drought extremes are
30 projected to become more frequent and widespread as early as mid-century (Toreti et al., 2019a). For present
31 to moderate adaptation and at least up to 2.5癎WL, negative consequences are mostly in SEU (Bird et al.,
32 2016; EEA, 2019c; Moretti et al., 2019; Feyen et al., 2020). The transition from moderate to high risk is

33 projected to happen around 2.7癈 GWL when hazards and risk will become more persistent and widespread
34 in other regions (Section 13.1, Deryng et al., 2014; Donatelli et al., 2015; Webber et al., 2018; Ceglar et al.,
35 2019; Ranasinghe et al., 2021; Seneviratne et al., 2021). This temperature increase will trigger shift of
36 agricultural zones, onset of early heat stress, losses in maize yield of up to 28% across EU-28, and regional

37 disparity in losses and gains in wheat, which are not able to offset losses across the continent (Deryng et al.,

    Do Not Cite, Quote or Distribute  13-72       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 2014; Szewczyk et al., 2018; Ceglar et al., 2019). There will be also broader adverse impacts such as
 2 reduction of grassland biomass production for fodder, increases in weeds and reduction in pollination
 3 (Castellanos-Frias et al., 2016; Nielsen et al., 2017; Br醩 et al., 2019) (medium confidence). Combined with
 4 socio-economic development, increased heat and drought stress and reduced irrigation water availability in
 5 SEU are projected to lead to abandonment of farmland (Holman et al., 2017). Around 4癈 GWL, the risk is
 6 very high due to persistent heat and dry conditions (Ben-Ari et al., 2018) and the emergence of losses also in
 7 NEU which would be much higher without the assumed CO2 fertilisation (Deryng et al., 2014; Szewczyk et
 8 al., 2018; Harrison et al., 2019).

 9

10 Farmers have historically adapted to environmental changes and such autonomous adaptation will continue.
11 Higher CO2 levels have a fertilisation effect on plants that is considered to decrease production risks (Deryng
12 et al., 2014). Adaptation solutions to heat and drought risks include changes in sowing and harvest dates,
13 increased irrigation, changes in crop varieties, the use of cover crops, and mixed agricultural practices
14 (Section 13.5.2; Figure 13.14 and Figure 13.30b). Under high adaptation, the use of irrigation can
15 substantially reduce risk by both reducing canopy temperature and drought impacts (high confidence)
16 (Section 13.5.2, Webber et al., 2018). Some reductions of maize yields in SEU are still possible, but are
17 balanced by gains in other crops and regions (Deryng et al., 2014; Donatelli et al., 2015; Webber et al., 2018;
18 Feyen et al., 2020). At 3癈 GWL and beyond, the adaptive capacity is reduced (Ruiz-Ramos et al., 2018).
19 Crop production is a major consumer of water in agriculture (Gerveni et al., 2020), yet a potentially scarcer
20 supply of water in some regions must be distributed across many needs (KR3, Section 13.10.2.3), limiting
21 availability to agriculture which is currently the main user of water in many regions of Europe (Section
22 13.5.1) (high confidence). Where the ability to irrigate is limited by water availability, other adaptation
23 options are insufficient to mitigate crop losses in some sub-regions, particularly at 3 GWL and above, with
24 an increase of risk from north to south and higher risk for late-season crops such as maize (high confidence).
25 Under these conditions, land abandonment is projected (Holman et al., 2017) (low confidence).

26

27

28

29 Figure 13.30: (a) Burning ember diagrams for losses in crop production with present or medium adaptation condition
30 and with high adaptation. Panel (b) Illustrative adaptation pathways and key messages based on the feasibility and
31 effectiveness assessment in Figure 13.14. Grey shading means long lead time and dotted lines signal reduced
32 effectiveness. The circles imply transfer to another measure and the bars that the measure have reached a tipping point
33 (Table SM13.28).

34

35

36 13.10.2.3 KR3: Risk of water scarcity to multiple interconnected sectors

37

38 Risks related to water scarcity across multiple sectors can become severe in WCE and to a much larger
39 extent in SEU based on projections of drought damage, population and sectors exposed and increases in
40 water exploitation (Figure 13.31a; Table SM13.29). In EEU, uncertainty in hydrological drought projections
41 and risk consequences is higher (Greve et al., 2018; Ranasinghe et al., 2021; Seneviratne et al., 2021) and the
42 available number of publications is lower, not allowing a conclusion on how risk levels change with GWL.
43 Yet, there is emerging evidence that drought related risks increase with warming beyond 3癈 GWL also in

Do Not Cite, Quote or Distribute  13-73       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 EEU (Seneviratne, 2021 #12291for hydrological drought and 4癈 GWL, Kattsov and Porfiriev, 2020).
 2 Evidence from the D&A assessment suggests that the risk is already moderate in SEU (e.g. 48 million people
 3 exposed to moderate water scarcity between 1981-2010) (Section 13.10.1, Figure 13.31a) (high confidence).

 4

 5 Risks have a high potential to lead to cascading impacts well beyond the water sector since water scarcity
 6 affects a number of highly interconnected sectors in Europe, from agriculture and livestock farming to
 7 energy (hydropower and cooling of thermal power plants) and industry (e.g., shipping) (Blauhut et al., 2015;
 8 Stahl et al., 2016; Bisselink et al., 2020; Cammalleri et al., 2020). Extensive water extraction will augment
 9 pressures on water reserves, impacting the ecological status of rivers and ecosystems dependent on them
10 (Grizzetti et al., 2017). Socioeconomic conditions contributing to severe consequences are when more
11 residents settle in drought-prone regions, or when the share of agriculture in GDP declines (high confidence).
12 For Europe, risks of water scarcity will be higher under SSP5 and SSP3 than under SSP1 (medium
13 confidence) (Byers et al., 2018; Arnell et al., 2019; Harrison et al., 2019). Transition to high risks is
14 projected to occur below 2癈 GWL in SEU and associated with more persistent droughts (Section 13.1.3)
15 and at 2癈 GWL with 54% increase of the population facing at least moderate levels of water shortage
16 (Byers et al., 2018). This transition will happen at higher warming in WCE since risks are projected to
17 increase less rapidly (transition between 2-3癈 GWL) (medium confidence) (Section 13.2.1.2, Byers et al.,
18 2018). At 3癈 GWL and beyond, water scarcity will become much more widespread and severe in already
19 water scarce areas in SEU (high confidence) and will expand to currently non-water scarce regions in WCE
20 (medium confidence) (Section 13.2.1.2, Bisselink et al., 2018; Naumann et al., 2018; Harrison et al., 2019;
21 Koutroulis et al., 2019; Cammalleri et al., 2020; Spinoni et al., 2020). Decrease in hydropower potential in
22 SEU and WCE are expected beyond 3癎WL (Figure 13.16).

23

24 To reduce risk severity, adaptation measures both at supply and demand level have been suggested (Section
25 13.2.2, Figure 13.6, Figure 13.31b, Garnier and Holman, 2019; Hagenlocher et al., 2019). Several measures
26 are already in place, showing high technical and institutional feasibility (Section 13.2.2.2 and 13.5.2.1). The
27 effectiveness of options varies regionally (in particular between northern and southern regions). For
28 example, in SEU many water reservoirs are already in place. Irrigation is used to support agriculture where
29 rainfed supplies are not sufficient (13.5.2). Their future extension depends on available precipitation. Also,
30 wastewater reuse can only be effective if sufficient wastewater is available. Improvements in water
31 efficiency and behavioural changes are very effective in SEU (>25% of damages avoided) (Section 13.2.2.2).
32 Investments in large water infrastructures and advanced technologies (incl. storage), water transfer, water
33 recycling and reuse and desalination allow to buy time and therefore to cope with additional warming
34 (Papadaskalopoulou et al., 2016; Greve et al., 2018). Beyond 2.5癈 GWL, transformational adaptation is
35 needed to lower risk levels, such as planned relocation of industry, abandonment of farmland or the
36 development of alternative livelihoods (Holman et al., 2017). In WCE the solution space to water scarcity is
37 expanding with considerable potential for investments in large water infrastructure and advanced
38 technologies (incl. storage), for reducing risks above 3癈 GWL (Greve et al., 2018). Under medium
39 warming a larger portfolio of measures might be needed in SEU in particular, although it may not be able to
40 completely avoid water shortages at high warming.

41

42

Do Not Cite, Quote or Distribute  13-74       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.31: (a) Burning ember diagrams for the risk of water scarcity with no/low adaptation and with high
 3 adaptation for SEU and WCE. (b) Illustrative adaptation pathways and key messages (see Figure 13.6). Grey shading
 4 means long lead time and dotted lines signal reduced effectiveness. The circles imply transfer to another measure and
 5 the bars that the measure have reached a tipping point (Table SM13.29).

 6

 7

 8 13.10.2.4 KR4: Risks to people, economies and infrastructures due to coastal and inland flooding

 9

10 Damages and losses from coastal and river floods are projected to increase substantially in Europe over the
11 21st century (high confidence) (Section 13.2.1, SM13.10). Coastal areas have already started to be affected
12 by sea level rise (Box13.1; Section 13.10.1) and human exposure to coastal hazards is projected to increase
13 in the next decades (high confidence), but less under SSP1 (20%) than SSP5 (50%) by the end of the century
14 (medium confidence) (Merkens et al., 2016; Reimann et al., 2018a). Under low adaptation (i.e. coastal
15 defences are maintained but not further strengthened), severe consequences include increase in expected
16 annual damage by a factor of at least 20 for 1.5-2.1癈 GWL (i.e.. high risks) and by 2-3 orders of magnitude
17 between 2 and 3癈 GWL in EU-28 (i.e. very high risk) (medium confidence) (Figure 13.28, 13.34c; Section
18 13.2.1.1); (Vousdoukas et al., 2018b; Haasnoot et al., 2021b). Under high adaptation (i.e. lowlands are
19 protected where it is economically efficient), expected annual damages still increase by a factor of 5 above
20 2癈 GWL (Section13.2, Vousdoukas et al., 2020). Sea-levels are committed to rise for (Fox-Kemper et al.,
21 2021), submerging at least 10% of the territory in 12 countries in Europe after millennia if GWL exceed 1.5-
22 2.5癈 (Clark et al., 2016), and this represents a major threat for the European and Mediterranean cultural
23 heritage (Figure 13.28, Cross-Chapter Box SLR in Chapter 3, CCP4, Marzeion and Levermann, 2014;
24 Reimann et al., 2018b).

25

26 Pluvial and riverine flood events in Europe have been attributed to climate change, but the associated
27 damages and losses also depend on land use planning and flood risk management practices (medium
28 confidence) (Section 13.10.1, Ranasinghe et al., 2021). Exposure to urban flooding will increase with
29 urbanization (Jongman et al., 2012; Jones and O'Neill, 2016; Dottori et al., 2018; Paprotny et al., 2018b).
30 Flooding is projected to rise with temperature in Europe, with e.g. a doubling of damage costs and people
31 affected from river flood for low adaptation above 3癈 GWL (Alfieri et al., 2018). Inland flooding represents
32 a key risk for Europe due to the extent of settlements exposed, the frequency of the hazards, the risks to
33 human lives associated with flash floods and the limited adaptation potential to pluvial flooding (e.g.
34 difficulty to upgrade urban drainage systems (Dale et al., 2018; Dale, 2021). Hence, risks can become very
35 high from 3癈 GWL (Figure 13.32a).

36

37 A range of adaptation options to coastal flooding exists and adaptation is possible in many European regions
38 if started on time (Section 13.2 and Figure 13.32d). Continuing a protection pathway is cost-effective in
39 urbanized regions for this century (Vousdoukas et al., 2020), but there is high agreement that it comes with

Do Not Cite, Quote or Distribute  13-75       Total pages: 143
    FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 residual risk if coastal defences fail during a storm. This residual risk can be reduced through early warning
 2 and evacuations, insurance and accommodate measures (Section 13.2.2). Soft limits to protection have been
 3 identified under high GWL, in particular due to the rate of change and delayed impacts of long-term SLR
 4 (medium confidence) (Hinkel et al., 2018; Haasnoot et al., 2020a). Ecosystem based solutions such as
 5 wetlands can reduce waves' propagation, provide co-benefits for the environment and climate mitigation,
 6 and reduce costs for flood defences (medium confidence) (Section 13.2.2.1). At higher GWL, ecosystems are
 7 projected to experience reduced effectiveness due to temperature increase and increased rate of SLR
 8 combined with lack of sediment and human pressures (Cross-Chapter Box SLR in Chapter 3). Retention and
 9 diversion can be effective for compound flooding or for estuaries with a limited storm surge duration, but
10 there is lack of knowledge on their effectiveness (Sections 13.2.2).

11

12 In the case of river flooding, adaptation has the potential to contain damage and losses up to 3癈 GWL
13 (Figure 13.32b, Jongman et al., 2014; Alfieri et al., 2016), provided they are implemented on time and that
14 the technical, social and financial barriers are addressed (Sections 13.2.2 and 13.6.2). Residual risks can be
15 reduced through early warning and evacuations, insurance and accommodate measures (see Section 13.2.2,
16 Kreibich et al., 2015). Accommodation strategies such as retention and ecosystem-based solutions require
17 space, which is not always available in cities. Both protection and flood retention are effective in reducing
18 inland flooding risk across Europe, but with regional variation in the benefit-to-cost ratio (Alfieri et al.,
19 2016; Dottori et al., 2020) (medium confidence). Furthermore, upgrading drainage systems to accommodate
20 increase in pluvial flooding is costly, technically complex and requires time (Dale et al., 2018; Dale, 2021).

21

22 Avoiding developments in risk-prone areas can reduce both coastal and inland flooding risks and can be
23 followed by planned relocation, particularly in less-populated areas. To align relocation with social goals and
24 to achieve positive outcomes long lead times are needed (Haasnoot et al., 2021a).

25

26

27

    Do Not Cite, Quote or Distribute  13-76       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.32: (a) Burning ember diagram for the risks from riverine and pluvial flooding with and without adaptation.
 3 (b) Illustrative adaptation pathways to riverine flooding risks. (c) Burning ember diagram for the risks from coastal
 4 flooding with and without adaptation. (d) Illustrative adaptation pathways to coastal flooding risks. Grey shading means
 5 long lead time and dotted lines signal reduced effectiveness. The circles imply transfer to another measure and the bars
 6 that the measure have reached a tipping point.(Tables SM13.30, SM13.31)

 7

 8

 9 13.10.3 Consequences of Multiple Climate Risks for Europe

10

11 European regions are affected by multiple key risks simultaneously; while there is a wide range in
12 quantifications, there is high agreement that consequences for socio-economic and natural systems can be
13 substantial, with more severe consequences in the south than in the north (very high confidence); there is
14 some indication also for a west to east gradient, with higher uncertainty in eastern part of WCE and EEU
15 which makes adaptation more challenging (medium confidence). Furthermore, the food-water-energy-land
16 use nexus plays an important role in amplifying overall risk levels in Europe (medium confidence) (Forzieri
17 et al., 2016; Harrison et al., 2016; Byers et al., 2018; Arnell et al., 2019; Harrison et al., 2019; Kebede et al.,
18 2021). Southern Europe, European cities and coastal areas are projected to become hotspots of multiple risks
19 (Cramer et al., 2018; Forzieri et al., 2018; Guerreiro et al., 2018) (high confidence). The number of people
20 exposed to multiple KRs in Europe are projected to at least double at 3癈 GWL compared to 1.5癈 GWL
21 (Forzieri et al., 2017; Byers et al., 2018; Arnell et al., 2019), but risk levels are already higher at 1.5癈 GWL
22 than today for a number of KRs (Figure 13.28) (medium confidence).

23

24 Economic losses and damages for European economies from multiple KRs are projected increase (high
25 confidence) (Figure 13.34, Szewczyk et al., 2018; Feyen et al., 2020; Kalkuhl and Wenz, 2020), potentially
26 quadruple at 3癈 GWL compared to 1.5癈 GWL (Feyen et al., 2020). Existing estimates of projected
27 economic costs for Europe, based on integrated assessment or computable general equilibrium models, are
28 however likely to be underestimations of the true costs because of incomplete coverage of biophysical
29 impacts, in particular low-probability high impact events, and disruptive risk propagation channels (Lamperti
30 et al., 2018; Stoerk et al., 2018; Schewe et al., 2019; Piontek et al., 2021). The main driver for this increase
31 in economic losses and damages is mortality due to heat stress (medium confidence), followed by reduced
32 labour productivity, coastal and inland flooding, water scarcity and drought (medium confidence) (Figure
33 13.33; Section 13.6.1.3). While losses are highest in SEU for both 1.5癈 and 3癈 GWL and increase by a
34 factor of more than three between these GWLs, the projected economic damages and losses also increase
35 significantly in WCE (by a factor of 4 from 1.5癈 to 3癈 GWL; 40% of total losses in EU-28 at 3癈 GWL)
36 and in NEU (almost 10% of total losses at 3癈 GWL) (Szewczyk et al., 2018; Szewczyk et al., 2020).
37 Adaptation is projected to reduce macroeconomic costs, but residual costs remain, particularly for warming
38 above 3癈 GWL (medium confidence) (De Cian et al., 2016; Bosello et al., 2018; Parrado et al., 2020).

39

40

Do Not Cite, Quote or Distribute  13-77       Total pages: 143
      FINAL DRAFT                 Chapter 13  IPCC WGII Sixth Assessment Report

1

 2

 3 Figure 13.33: Economic losses/damages and gains due to projected climate risks, for 1.5癈 and 3癈 GWL relative to
 4 no additional warming; macroeconomic effects measured in GDP or welfare. Effect for EEU report for Russia as a
 5 whole country, deviating from the definition of EEU in this chapter. Effects may deviate from sectoral assessments in
 6 Sections 13.2-13.7 due to different degree of coverage of risk channels (Table SM13.23).

 7

 8

 9 13.10.4 Knowledge Gaps

10

11 Information on risk levels and development are available for 1.7癈, 2.5癈 and > 4癈 GWL, making the
12 determination of transitions for the burning embers challenging and impairing a comprehensive assessment
13 across key risks. Further efforts to extend the SSP narratives to Europe can contribute to a more
14 disaggregated understanding of risk severity for different vulnerability and exposure conditions, but the
15 evidence to date remains limited to few sectors (CCP4, Kok et al., 2019; Pedde et al., 2019; Rohat et al.,
16 2019). There is only very limited evidence on the extent and timing of residual risks under different GWL,
17 even with high adaptation.

18

19 There is medium confidence on the effectiveness of adaptation beyond 3癈 GWL particularly where risks are
20 high to very high (Figures 13.28-13.32). There is limited evidence on the effectiveness of specific adaptation
21 options at different levels of warming that also include consideration of lead and lifetimes. An integrated
22 assessment, which projects the impacts on crop production by examining the potential availability of water
23 for agricultural purposes together with other adaptation measures, is missing.

24

25 Transboundary risks, interactions between commodity and financial markets, market imperfections, non-
26 linear socio-economic responses, and loss of ecosystem services, may amplify losses for European
27 economies. Available models may underestimate the full costs of climate change as they generally neglect
28 systemic risks, tipping points, indirect and intangible losses and limits to adaptation (Dafermos et al., 2018;
29 Lamperti et al., 2018; van Ginkel et al., 2020; Dasgupta et al., 2021; Ercin et al., 2021; Piontek et al., 2021).
30 With increasing global warming, compound, low likelihood, or unprecedented extremes such as the
31 European dry and hot summer 2018 or the extreme rainfall following storm Desmond in the UK in 2015,
32 become more frequent (AR6 WGI Cross-Chapter Box 11.2). These events can have catastrophic

Do Not Cite, Quote or Distribute  13-78       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 consequences for Europe, but the extent of economic and non-economic damages and losses remain largely
 2 uncertain.

 3

 4

 5 13.11 Societal Adaptation to Climate Change across Regions, Sectors and Scales

 6

 7 Building on our sectoral analysis in previous sections, this section looks across European sectors, vulnerable
 8 groups, and regions to assess how climate change impacts are being responded to in general by state (Section
 9 13.11.1) and non-state (Section 13.11.2) actors, and their synergies and dependencies. Sections 13.11.3
10 assess if and how system transformations have emerged and implications for the SDGs and climate resilient
11 development pathways (CRDPs).

12

13 13.11.1 Policy Responses, Options and Pathways

14

15 13.11.1.1 Progress on Adaptation Planning and Implementation

16

17 The solution space for climate change adaptation has expanded across European regions since AR5 (high
18 confidence). European countries are increasingly planning to adapt to observed impacts and projected
19 climate risks across scales of government (Lesnikowski et al., 2016; Russel et al., 2020) (high confidence).
20 Whereas in 2009, only nine EU countries had developed a National Adaptation Strategy (NAS) (Biesbroek et
21 al., 2010; EEA, 2014), by mid-2020 all EU Member States and several other European countries had adopted
22 at least a NAS and/or revised and updated prior strategies (Figure 13.34, bottom, Klostermann et al., 2018;
23 EEA, 2020a). Progress is also observed at the level of the European Union with the adoption of the new EU
24 strategy on adaptation to climate change in 2021 (European Comission, 2021a) and regionally, particularly in
25 federalist and decentralised states (Steurer and Clar, 2018; EEA, 2020c; Pietrapertosa et al., 2021), and
26 locally, with an increasing number of European cities planning for climate risks (Aguiar et al., 2018; Reckien
27 et al., 2018; Grafakos et al., 2020) (Section 13.6.2.1;Box 13.3;Chapter 6) (high confidence). There is
28 evidence of action across sectors and scales, even in European countries where national adaptation
29 frameworks are absent (Figure 13.34) (De Gregorio Hurtado et al., 2015; Pietrapertosa et al., 2018; Reckien
30 et al., 2018) (medium confidence). However, the implementation gap identified in AR5 (Chambwera et al.,
31 2014), i.e. the gap between defined goals and ambitions and actual implemented actions on the ground,
32 persists in Europe (Aguiar et al., 2018; Russel et al., 2020; UNEP, 2021).

33

34 The drivers of adaptation progress in Europe differ across sectors and regions. Common drivers include:
35 experienced climatic events, improved climatic information, societal pressures to act, projected economic
36 and societal costs of climate change, participation in (city) networks, societal and political leadership, and
37 changes in national and European policies and legislation (medium evidence, high agreement) (EEA, 2014;
38 Massey et al., 2014; Reckien et al., 2018). The availability of human, knowledge, and financial resources
39 appears important for proactive adaptation(Termeer et al., 2012; Sanderson et al., 2018)(Termeer et al.,
40 2012; Sanderson et al., 2018)(Termeer et al., 2012; Sanderson et al., 2018)(Termeer et al., 2012; Sanderson
41 et al., 2018), while adaptation is also strongly dependent on economic and social development (Sanderson et
42 al., 2018) (high confidence). How adaptation is governed differs substantially across Europe (Clar, 2019;
43 Lesnikowski et al., 2021). Political commitment, persistence and consistent action across scales of
44 government is critical to move beyond planning for adaptation (Step A to C in Figure 13.34) and to ensure
45 adequacy of implementation (Step D and E in Figure 13.34) (Howlett and Kemmerling, 2017; Lesnikowski
46 et al., 2021; Patterson, 2021).

47

48 The scope of climate risks included in European adaptation policies and plans (Step B in Figure 13.34) is
49 generally broad (EEA, 2018a). Systemic and cascading risks (Section 13.10) are often recognized, but most
50 conventional risk assessment methods that inform adaptation planning are ill-equipped to deal with these
51 effects (Adger et al., 2018). For example, transboundary risks emerging in regions outside of Europe are
52 considered only by a few countries such as the UK and Germany (Section 13.9.3). European climate change
53 adaptation strategies and national policies are generally weak on gender, LGBTQI+, and other social equity
54 issues (Cross-Chapter Box GENDER in Chapter 18, Boeckmann and Zeeb, 2014; Allwood, 2020).

55

56

Do Not Cite, Quote or Distribute  13-79       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.34: Progress of National Adaptation in Europe in 2018 and Status of National Adaptation Plans and
 3 Strategies in 2020. Data on the progress of national adaptation are from the self-reported status of EU members, as
 4 documented in the Adaptation Scoreboard for Country fiches (SWD(2018)460). The status of national adaptation plans
 5 and strategies data are from EEA Report 6/2020 (2020a), the ClimateADAPT portal (EEA, 2021a) and from the
 6 Grantham Institute database Climate Change Laws of the World (Grantham Research Institute, 2021).

 7

 8

 9 Many near-term investment decisions have long-term consequences, and planning and implementation (Step
10 C and D in Figure 13.34) can take up to decades, particularly for critical infrastructure planning in Europe
11 (Zandvoort et al., 2017; Pot et al., 2018). Consequently, there are calls to expand planning horizons, to
12 consider long-term uncertainties to prevent lock-in decision dependencies, to seize opportunities and
13 synergies from other investments (e.g., socio-economic developments and systems transitions) and to
14 broaden the range of considered possible impacts (e.g. Frantzeskaki et al., 2019; Marchau, 2019;

Do Not Cite, Quote or Distribute  13-80       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 Oppenheimer et al., 2019; Haasnoot et al., 2020b ). Yet, high GWL scenarios beyond 2100 are often not
 2 considered in climate change adaptation planning due to a lack of perceived usability, missing socio-
 3 economic information, constraining institutional settings, and conflicting decision-making timeframes
 4 (medium confidence) (Lourenco et al., 2019; Taylor et al., 2020). High GWL scenarios are often seen as
 5 having a low probability of occurrence, resulting in inaction or incremental rather than transformative
 6 adaptation responses to projected climate risks (Dunn et al., 2017). Extending planning horizons to beyond
 7 2100 increases deep uncertainties for decision-makers as a result of unclear future socio-economic and
 8 climatic changes. For adaptation to sea level rise along Europe's coast, for example, there are already
 9 considerable uncertainties during this century (Fox-Kemper et al., 2021).

10

11 Adaptive planning and decision making are still limited across Europe (high confidence). Prominent
12 examples of adaptive plans include the flood defence systems for the City of London (Ranger et al., 2013;
13 Kingsborough et al., 2016; Hall et al., 2019) and the Netherlands (Van Alphen, 2016; Bloemen et al., 2019)
14 Adaptation pathways have been also developed for planning of urban water supply (Kingsborough et al.,
15 2016; Erfani et al., 2018), urban drainage (Babovic and Mijic, 2019) and wastewater systems (Sadr et al.,
16 2020) (Cross-Chapter Box DEEP in Chapter 17). Flexible strategies are increasingly considered by European
17 countries (e.g. Stive et al., 2013; Kreibich et al., 2015; Bubeck et al., 2017; Haasnoot et al., 2019), but
18 require appropriate design to be effective (Metzger et al., 2021).

19

20 Monitoring and evaluation of adaptation action is only done in some European countries (Figure 13.34, part
21 E), but is important for adjusting planning, if needed (Hermans et al., 2017; Haasnoot et al., 2018), and
22 enhancing transparency and accountability of progress (Mees and Driessen, 2019). In the Netherlands, a
23 comprehensive monitoring system has been put in place, including signals for adaptation that support
24 decisions on when to implement adaptation options or to adjust plans (Hermans et al., 2017; Haasnoot et al.,
25 2018; Bloemen et al., 2019).

26

27 13.11.1.2 Mainstreaming and Coordination

28

29 Coordinated responses are necessary to prevent inefficient and costly action (Biesbroek, 2021), balance
30 under- and over-reaction to climate risks (Peters et al., 2017; Biesbroek and Candel, 2019), avoid
31 redistributing vulnerability and maladaptive actions (Atteridge and Remling, 2018; Albizua et al., 2019;
32 Neset et al., 2019), and ensure timely implementation (Benson and Lorenzoni, 2017) (high confidence).
33 Since AR5, progress has been made to increase coordinated adaptation actions, but so far this is limited to a
34 few sectors (mostly water management and agriculture) and European countries and regions (mostly SEU,
35 WCE depending on impact) (Lesnikowski et al., 2016; Biesbroek and Delaney, 2020; Booth et al., 2020)
36 (high confidence) (Section 13.11.2). Despite evidence of emerging bottom-up (e.g., citizens and business
37 initiatives) and top-down initiatives (e.g., governmental plans and instruments to ensure action), there are
38 considerable barriers to mainstreaming adaptation (Runhaar et al., 2018) (high confidence).

39

40 While mainstreaming of adaptation into other policy domains has been advocated as an enabler for
41 adaptation, it may have resulted incremental rather than transformational adaptation, and may not be
42 sufficient to close the adaptation gap (Andersson and Keskitalo, 2018; Remling, 2018; Scoville-Simonds et
43 al., 2020).

44

45 13.11.1.3 Climate Services and Local Knowledge

46

47 Climate services to support adaptation decision-making of governments and businesses across Europe have
48 rapidly increased since AR5, partly as a result of national and EU investments such as the Copernicus C3S
49 service (high confidence) (Street, 2016; Soares and Buontempo, 2019). These services are increasingly used
50 in NEU, SEU and WCE for example in energy and risk prevention in coastal and riverine cities, stimulating
51 regulations and bottom-up initiatives (Cavelier et al., 2017; Le Cozannet et al., 2017; Reckien et al., 2018;
52 Howard et al., 2020). However, climate service efficacy is rarely systematically evaluated (Cortekar et al.,
53 2020). Barriers to use include: lack of perceived usefulness of climate information to organisations and
54 expertise to use the information, outdated statistics, mismatch between needs and type of information made
55 available, insufficient effective engagement between providers and recipients of climate information and
56 lack of business models to sustain climate services over time (high evidence, medium agreement) (Cavelier
57 et al., 2017; R鋝鋘en et al., 2017; Bruno Soares et al., 2018; Christel et al., 2018; Oberlack and Eisenack,

Do Not Cite, Quote or Distribute  13-81       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 2018; Hewitt et al., 2020). Adaptation decision support platforms also face challenges regarding updating,
 2 training and engagement with users (EEA, 2015; Palutikof et al., 2019).

 3

 4 In addition to scientific knowledge, traditional and local knowledges can enable adaptation action
 5 (Huntington et al., 2017) as is the case with indigenous-led ecosystem restoration in the European Arctic
 6 (Brattland and Mustonen, 2018). There is a need to draw on surviving Indigenous knowledge systems in
 7 Europe (Greenland, Nenets, Khanty, S醡i, Veps, Ingrian) as unique, endemic ways of knowing the world
 8 that can position present and historical change in context and offer unique reflections of change in the future
 9 (Ogar et al., 2020; Mustonen et al., 2021).

10

11 13.11.1.4 Financing Adaptation and Financial Stability

12

13 Dedicated financial resources for the implementation of NAS and plans are a key enabling factor for
14 successful adaptation (high confidence) (Russel et al., 2020) (Chapter 17). Yet, only 14 EU countries have
15 announced such budget allocations in their plans and strategies; and even if budget numbers are available,
16 they are difficult to compare (EEA, 2020b). Current adaptation spending varies greatly across and within
17 European countries, partly reflecting (sub-)national adaptation priorities or financing sources targeting
18 investment projects (L髉ez-D髍iga et al., 2020; Russel et al., 2020) and competing statutory priorities (Porter
19 et al., 2015). European government budgets are also burdened by climate change damages today, particularly
20 after huge flooding events, and austerity following financial crises, limiting anticipatory action (Penning-
21 Rowsell and Priest, 2015; Miskic et al., 2017; Schinko et al., 2017; Slav韐ov� et al., 2020). National
22 adaptation funding in EU member states is complemented by EU funding (e.g. European Structural and
23 Investment Funds (ESIF), European Regional Development Funds, and LIFE program). While the EU
24 spending target on climate action increased from 20% in 2016-2020 to 25% in 2021-2026, most spending is
25 going into mitigation, not adaptation (Berkhout et al., 2015; Hanger et al., 2015; EEA, 2020b).

26

27 With higher warming levels, financing needs are likely to increase (high confidence) (Mochizuki et al., 2018;
28 Bachner et al., 2019; Parrado et al., 2020) governments can address this higher need by cutting other
29 expenditures, increasing taxes, or by increasing the fiscal deficit (Miskic et al., 2017; Mochizuki et al., 2018;
30 Bachner et al., 2019). Yet, the requirement for fiscal consolidation that will be needed after the COVID-19
31 pandemic (Cross-Chapter Box COVID in Chapter 7) may also lead to a cessation of adaptation spending, as
32 evidenced by the expenditure drop in coastal protection in Spain after the financial crisis 2008 (L髉ez-
33 D髍iga et al., 2020). Governments can shift the financial burden to beneficiaries of adaptation, as e.g.,
34 suggested for coastal protection and riverine flooding (Jongman et al., 2014; Penning-Rowsell and Priest,
35 2015; Bisaro and Hinkel, 2018). There is also an increase in financial mechanisms to accelerate private
36 adaptation actions, including adaptation loans, subsidies, direct investments, and novel public-private
37 arrangements. For example, the European Investment Bank created a finance facility to support European
38 regions through loans to implement adaptation projects (EEA, 2020b).

39

40 Since AR5, new evidence has emerged that climate change may deteriorate financial stability both at the
41 global and European scale (Campiglio et al., 2018; Dafermos et al., 2018; Lamperti et al., 2019; ECB,
42 2021a). The European Central Bank, the European Systemic Risk Board, and several national central banks
43 in NEU and WCE have started to systematically assess the consequences of climate risks for financial
44 stability and plan to integrate climate stress testing into their supervisory tools (Batten et al., 2016; ECB,
45 2021a; ECB, 2021b).

46

47 13.11.2 Societal Responses, Options and Pathways

48

49 13.11.2.1 Private-sector

50

51 Within the private sector, there tends to be a preference for `soft' (e.g., knowledge generation) than `hard'
52 (e.g., infrastructure) adaptation measures (Goldstein et al., 2019), in contrast to government-led responses
53 typically favouring hard measures (Pranzini et al., 2015). However, there also remains diversity across
54 sectors and organisations in the degree and type of adaptation response (Traw鰃er, 2014; Dannevig and
55 Hovelsrud, 2016; Ray et al., 2017; Ricart et al., 2018). Whereas some sectors such as flood management,
56 banking and insurance, and energy (Bank of England, 2015; Gasbarro and Pinkse, 2016; Bank of England,
57 2019; Wouter Botzen et al., 2019) have generally made moderate progress on adaptation planning across

Do Not Cite, Quote or Distribute  13-82       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 Europe, there are key vulnerable economic sectors that are in earlier stages, including aviation (Burbidge,
 2 2015), ports and shipping (Becker et al., 2018; Ng et al., 2018), and ICT (EEA, 2018b) (high confidence).
 3 There is also some evidence of `short-sighted' adaptation or maladaptation; for example, in winter tourism
 4 there is a preference for technical and reactive solutions (e.g., artificial snow) that will not be sufficient under
 5 high levels of warming (Section 13.6.1.4).

 6

 7 Where adaptation is considered by companies, it is typically triggered either by the experience of extreme
 8 weather events that led to business disruptions (McKnight and Linnenluecke, 2019) or is included into
 9 corporate risk management in response to regulatory, shareholder or customer pressure (Averchenkova et al.,
10 2016; Gasbarro et al., 2017). For instance, following the implementation of the recommendations of the Task
11 Force on Climate-Related Financial Disclosure (TFCD) by the European Commission in 2019, 50 publicly
12 listed companies revealed their exposure to their physical climate risks in 2020 (CDSB, 2020). But even if
13 companies experience extreme weather events or stakeholder pressure, they may not adapt because they
14 underestimate their vulnerability (Pinkse and Gasbarro, 2019) (Table 13.1). For example, key barriers to
15 adaptation among Greek firms include both external (e.g., lack of support/guidance) and internal factors
16 (e.g., few resources, managerial perceptions (Halkos et al., 2018). Lack of knowledge, feeling climate
17 change is not a salient risk, and lack of social learning or collaboration, appear to be key barriers to private-
18 sector adaptation (Dinca et al., 2014; Andr� et al., 2017; Romagosa and Pons, 2017; Esteve et al., 2018; Lu韘
19 et al., 2018; Ng et al., 2018) (Section 13.16.2.2). There remains little research on private-sector awareness of
20 or responses to cascading or compound risks associated with climate change (Miller and Pescaroli, 2018;
21 Pescaroli, 2018).

22

23 13.11.2.2 Communities, Households and Citizens

24

25 Planned behavioural adaptation remains limited amongst European households (high confidence), with few
26 examples that can be considered transformative (e.g., structural, long-term, collective (Wilson et al., 2020))
27 (medium confidence). One Swedish survey of householders at risk of extreme weather events (e.g., floods,
28 storms) found evidence of some organisational measures (e.g., bringing possessions inside prior to a storm,
29 preparing for power cuts with candles, etc.) but very few households took any other (technical, social,
30 nature-based, or economic) measures (Brink and Wamsler, 2019). Similarly, few at risk of flooding are
31 taking action (Stojanov et al., 2015) (Sections 13.2.1, 13.6.1); for example, little public take-up of available
32 municipal support for individual adaptation in Germany (Wamsler, 2016). Water efficiency measures in
33 anticipation of, or response to, drought are also limited (Bryan et al., 2019), although water reuse in
34 Mediterranean and some other EU (e.g., UK, Netherlands) countries is increasing (Aparicio, 2017) (Section
35 13.2). Amongst the adaptation responses recorded, few are perceived as opportunities (Taylor et al., 2014;
36 Simonet and Fatori, 2016). There is currently little European research on public responses to risks other
37 than flooding, heat stress and drought, such as vector-borne disease, and to multiple and cascading risks (van
38 Valkengoed and Steg, 2019) (Section 13.7).

39

40 Perceived personal responsibility for tackling climate change remains low across the EU (Figure 13.35) and
41 partly explains why household adaptation remains limited (high confidence) (Taylor et al., 2014; van
42 Valkengoed and Steg, 2019), despite risk perception apparently growing (Figure Box 13.2.1, Capstick et al.,
43 2015; Poppel et al., 2015; BEIS, 2019). Householders' risk perception and concern about climate change
44 fluctuates in response to media coverage and significant weather or socio-political events (high confidence)
45 (Capstick et al., 2015). On average across Europe, and particularly in relation to gradual change, non-experts
46 continue to under-estimate climate change risks compared to experts (medium confidence) (Taylor et al.,
47 2014), have low awareness of adaptation options, and confuse adaptation and mitigation (Harcourt, 2019),
48 suggesting a need for improved climate literacy amongst the public. Indeed, fostering learning and coping
49 capacity supports robust adaptation pathways (J鋑er et al., 2015).

50

51 There is strong public support for adaptation policy (e.g., building flood defences), particularly within the
52 UK, France, Norway and Germany (Doran et al., 2018). Although, in some cases such public adaptation can
53 undermine motivation for householders to take adaptation measures (Section 13.2), public adaptation can
54 also increase householder motivations, with perceived efficacy of action a strong predictor of adaptation
55 (high confidence) (Moser, 2014; van Valkengoed and Steg, 2019). However, there are also structural and
56 economic barriers to household adaptation due to lack of policy incentives or regulations. For example,
57 water-saving devices in homes could halve consumption, but lack of economic benefits to householders are

Do Not Cite, Quote or Distribute  13-83       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

1 barriers to adoption; while lack of standards as well as societal hesitation may explain low levels of water
2 reuse in Europe (EEA, 2017c) (Section 13.2). Conversely, water meters and higher tariffs have been found to
3 reduce water consumption only in combination with other measures (EEA, 2017c; Bryan et al., 2019).

4

5

 6

 7 Figure 13.35: Trends in perceived climate change risks and responsibility for tackling climate change across EU-28;
 8 data collected from around 1,000 respondents per country for each year surveyed (European Comission, 2017)

 9

10

11 As well as temporal trends in climate change risk perception, the literature since AR5 continues to show
12 much heterogeneity (both within and between nations) amongst householders in respect of risk perception
13 (high confidence). Higher climate change risk perceptions observed in Spain, Portugal, Iceland, and
14 Germany (Figure 13.2); at individual level, women, younger age groups, more educated, left-leaning, and
15 those with more `self-transcendent' values perceive more negative impacts from climate change, although
16 the strength of these relationships varies across European nations (Clayton et al., 2015; Doran et al., 2018;
17 Poortinga et al., 2019; Duijndam and van Beukering, 2021). Stronger evidence exists since AR5 that
18 experience of extreme weather events can shape climate change risk perceptions, if these events are
19 attributed to climate change or evoke negative emotions (high confidence) (Clayton et al., 2015; Demski et
20 al., 2017; Ogunbode et al., 2019). Proximity to climate hazards does not predict adaptation responses in a
21 straightforward way: in Portugal, those living by the coast were more likely to attribute local natural hazards
22 to climate change and to take some adaptive measures (Lu韘 et al., 2017); while waterside residents in flood-
23 prone regions of France and Austria were more resistant to relocation, due to higher place attachment (Adger
24 et al., 2013; Rey-Valette et al., 2019; van Valkengoed and Steg, 2019; Seebauer and Winkler, 2020).
25 Migration from threatened regions is discussed in Section 13.8.1.3.

26

27 13.11.3 Adaptation, Transformation and Sustainable Development Goals

28

29 The implementation of far-reaching and rapid systemic changes, including both adaptation and mitigation
30 options (de Coninck et al., 2018) remains less researched in societal systems than natural ones (Salomaa,
31 2020) that enhance multilevel governance and institutional capabilities, and enables lifestyle/behavioural
32 change and technology innovation. Adaptation responses across European regions and sectors, are more
33 often incremental than transformative (medium confidence), with possible exceptions including water-related
34 examples in for example the Netherlands (Section 13.2.2) and some cities (Box 13.3). Transformative
35 options may be better able to exploit new opportunities and co-benefits (EEA, 2019a) (Box 13.3) (Cross-
36 Chapter Box HEALTH in Chapter 7). Transitions towards more adaptive and climate resilient systems are

Do Not Cite, Quote or Distribute  13-84       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 often the result of responses to crises which create windows of opportunity for systemic changes
 2 (Johannessen et al., 2019) (cf. Chapter 18). This includes extreme weather events, financial crises (e.g.,
 3 Malm�; (Anderson, 2014; Isaksson and Heikkinen, 2018)) and the COVID-19 pandemic (e.g., Milan), which
 4 have disrupted the status quo and accelerated innovation and implementation (e.g., Milan; Box 13.3, Cross-
 5 Chapter Box COVID in Chapter 7).

 6

 7

 8 [START BOX 13.3 HERE]

 9

10 Box 13.3: Climate Resilient Development Pathways in European Cities

11

12 Climate resilient development (CRD) in European cities offers synergies and co-benefits from integrating
13 adaptation and mitigation with environmental, social and economic sustainability (Geneletti and Zardo,
14 2016; Grafakos et al., 2020). Climate networks (e.g., Covenant of Mayors), funding (e.g., Climate-KIC),
15 research programs (e.g., Horizon Europe), European and national legislation, international treaties, and the
16 identification of co-benefits, contribute to the prioritisation of climate action in European cities (Heidrich et
17 al., 2016; Reckien et al., 2018; CDP, 2020). Still, mitigation and adaptation remain largely siloed and
18 sectoral (Heidrich et al., 2016; Reckien et al., 2018; Grafakos et al., 2020). An assessment of the integration
19 of mitigation and adaptation in urban climate change action plans in Europe found only 147 cases in a
20 representative sample of 885 cities (Reckien et al., 2018).

21

22 In European cities, CRD is most evident in the areas of green infrastructure, energy efficient buildings and
23 construction, and active and low-carbon transport (Pasimeni et al., 2019; Grafakos et al., 2020). Nature-
24 based solutions (NbS), such as urban greening, often integrate adaptation and mitigation in sustainable urban
25 developments and are associated with increasing natural and social capital in urban communities, improving
26 health and wellbeing, and raising property prices (Geneletti and Zardo, 2016; Pasimeni et al., 2019; Grafakos
27 et al., 2020). Barriers to CRD in European cities include limitations in: funding, local capacity, guidance
28 documents and quantified information on costs, co-benefits and trade-offs (Grafakos et al., 2020). Pilot
29 projects are used to initiate CRD transitions (Nagorny-Koring and Nochta, 2018). Malm� (Sweden) and
30 Milan (Italy) are two examples to illustrate the strategies and challenges of two European cities attempting to
31 implement CRDP.

32

33 Malm� (population 0.3M): Since the 1990s, Malm� has been transitioning toward an environmentally,
34 economically and socially sustainable city, investing in eco-districts (redeveloped areas that integrate and
35 showcase the city's sustainability strategies) and adopting ambitious adaptation and mitigation targets. The
36 city has focused on energy efficient buildings and construction, collective and low carbon transportation, and
37 green spaces and infrastructure (Anderson, 2014; Malmo Stad, 2018). Malm� has developed creative
38 implementation mechanisms, including a "climate contract" between the city, the energy distributor and the
39 water and waste utility to co-develop the climate-smart district, Hyllie (Isaksson and Heikkinen, 2018;
40 Kanters and Wall, 2018; Parks, 2019). Flagship eco-districts play a central role in the city's transition, in the
41 wider adoption of CRD and in securing implementation partners (Isaksson and Heikkinen, 2018; Stripple
42 and Bulkeley, 2019). The city has also leveraged its status as a CRD leader to attract investment. The private
43 sector views CRD as profitable, due to the high demand and competitive value of these developments
44 (Holgersen and Malm, 2015). Malm� adopted the SDGs as local goals and the city's Comprehensive Plan is
45 evaluated on these, e.g., considering gender in the use, access and safety of public spaces, and emphasizing
46 development that facilitates climate resilient lifestyles (Malmo Stad, 2018). Malm� also engages
47 stakeholders via dialogue with residents, collaboration with universities and partnerships with industry and
48 service providers (Kanters and Wall, 2018; Parks, 2019). Despite measurable and monitored targets, and
49 supportive institutional arrangements, sustainability outcomes for the flagship districts have been tempered
50 by developers' market-oriented demands (Holgersen and Malm, 2015; Isaksson and Heikkinen, 2018) and
51 there is limited low-income housing in climate-resilient districts (Anderson, 2014; Holgersen and Malm,
52 2015).

53

54 Milan (population 1.4M): Milan is taking a CRD approach to new developments (Comune di Milano,
55 2019). From 2020, new buildings must be carbon neutral and reconstructions must reduce the existing land
56 footprint by at least 10%. The Climate and Air Plan (CAP) and the city's Master Plan (Comune di Milano,
57 2019) focus on low-carbon, inclusive and equitable development. The CAP is directed at municipal and

Do Not Cite, Quote or Distribute  13-85       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 private assets, and individual to city-scale actions. In 2020, Milan released a revised Adaptation Plan and the
 2 Open Streets project to ensure synergies between the COVID-19 response and longer-term CRD. Examples
 3 include strengthening neighbourhood-scale disaster response and reallocating street space for walking and
 4 cycling (Comune di Milano, 2020). Milan emphasizes institutionalization of CRD via a dedicated resilience
 5 department, and through active participation in climate networks and projects that support learning and
 6 exchange. Climate network commitments are cited in the city's Master Plan and CAP guidelines as driving
 7 more ambitious deadlines and emissions targets (Comune di Milano, 2019). Implementation of Milan's plans
 8 remains a challenge, despite dedicated resources and commitment.

 9

10 [END BOX 13.3 HERE]

11

12

13 Considerable barriers exist that prevent system transitions from taking place in Europe, including
14 institutional and behavioural lock-ins such as administrative routines, certain types of legislation, and
15 dominant paradigms of problem solving (Johannessen et al., 2019; Roberts and Geels, 2019) (high
16 confidence). For example, near-term and sectoral decision-making constrains transformative options for
17 water-related risks (Section 13.2). Breaking through these lock-ins requires substantive (political) will,
18 (un)learning of practices, resources, and evidence of what works. Trade-offs exist between the depth, scope,
19 and pace of change in transforming from one system to another, suggesting that designing system
20 transformations is a careful balancing act (Termeer et al., 2017). Aspiring in-depth and comprehensive
21 transformational changes might create a consensus frame to work towards; but it might not offer concrete
22 perspectives to act on the ground. Taking small steps and quick-wins offer an alternative pathway (Termeer
23 and Dewulf, 2018).

24

25 Adaptation responses can also be understood in terms of their trade-offs and synergies with SDGs
26 (Papadimitriou et al., 2019; Bogdanovich and Lipka, 2020). In terms of synergies, analysis of the Russian
27 NAP found that successful completion of the NAP's first phase could lead to significant progress towards 15
28 of the 17 goals (Bogdanovich and Lipka, 2020). European water adaptation (e.g., flood protection) can
29 similarly support freshwater provision; and water-secured environments support socio-economic growth
30 (Sadoff et al., 2015) since people and assets tend to accumulate in areas protected from flooding and
31 supplied with water, reducing the incentive for autonomous adaptation (de Moel et al., 2011; Hartmann and
32 Spit, 2016; Di Baldassarre et al., 2018). In health, behavioural measures to reduce mental health impacts
33 (e.g., gardening, active travel) can have broader health benefits (SDG 3) as well as help reduce emissions
34 (Section 13.7; SDGs 7 and 13). Conversely, growing use of air conditioning for humans and livestock
35 represents a potential trade-off between adaptation and mitigation (Sections 13.5,13.6, 13.7, 13.10). As noted
36 in Section 13.8, addressing poverty (SDG 1) - including energy poverty (SDG 7) and hunger (SDG 2) - and
37 inequalities (SDG 10) - including gender inequality (SDG 5) - improves resilience to climate impacts for
38 those groups that are disproportionately affected (women, low-income and marginalised groups). Also, more
39 inclusive and fair decision-making can enhance resilience (SDG 16; Section 13.4.4); although adaptation
40 measures may also lead to resource conflicts (SDG 16; Section 13.7). Climate adaptation, particularly
41 nature-based solutions, also supports ecosystem health (SDGs 14 and 15) (Dzebo et al., 2019). Economic
42 trade-offs appear to be more common across adaptation strategies, for example reduced employment arising
43 from land use change measures (Papadimitriou et al., 2019). There are also trade-offs between large-scale
44 mitigation measures (e.g., wind farms) and adaptation options that rely on ecosystem services (e.g., water
45 regulation; Section 13.3-13.4); and conversely, some adaptation options (e.g., air conditioning) may
46 negatively impact mitigation. Figure 13.36 summarises the synergies between adaptation and SDGs as
47 identified by 167 European cities in 2019; particularly prominent are reported biodiversity and health
48 benefits most often arising from societal (e.g., informational) and structural (e.g., technological/engineering)
49 measures. Beyond the urban context, biodiversity co-benefits from agro-ecology are also recognised (Section
50 13.5). Sustainable behaviour change measures have been found to be particularly likely to lead to synergies
51 with SDGs (Papadimitriou et al., 2019).

52

53

Do Not Cite, Quote or Distribute  13-86       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 Figure 13.36: Co-benefits for SDGs from adaptation actions. Shows how European cities have assessed the
 3 sustainability co-benefits of taking adaptation actions. Data was extracted from the Carbon Disclosure Project (CDP)
 4 database using the 2019 dataset; of the 861 European cities submitting data, 167 provided data on their adaptation
 5 actions and so this data is shown here (CDP, 2019). CDP categories of climate hazards were recategorized into WG1
 6 Climate Impact Drivers (e.g., cold spell, heavy precipitation); CDP adaptation actions were reclassified into AR5
 7 adaptation options (`social', `structural' and `institutional'; `other' includes actions falling outside these AR5
 8 categories); and CDP co-benefits were recategorized as SDGs. Panel 2 shows that all SDGs except one (SDG 17) were
 9 identified as a co-benefit of adaptation, although more environmental co-benefits were identified than social or
10 economic ones. Panel 2 shows that societal actions were most common, followed by structural, then institutional.
11 Informational measures were particularly common. Panel 3 shows how many actions were taken by different European
12 cities.

13

14

15 [START FAQ13.1 HERE]

16

17 FAQ 13.1: How can climate change affect social inequality in Europe?

18

19 The poor and those practising traditional livelihoods are particularly exposed and vulnerable to climate
20 change. They rely more often on food self-provisioning and settle in flood-prone areas. They also often lack
21 the financial resources or the rights to successfully adapt to climate-driven changes. Good practice
22 examples demonstrate that adaptation can reduce inequalities.

23

24 Social inequalities in Europe arise from disparities in income, gender, ethnicity, age, as well as other social
25 categorisations. In the European Union, about a fifth of the population (109 million people) at present, live
26 under conditions of poverty or social exclusion. Moreover, poverty is unequally distributed across Europe,

Do Not Cite, Quote or Distribute  13-87       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 with higher poverty levels in the eastern parts of Europe. The oldest and youngest in society are often most
 2 vulnerable.

 3

 4 The poor and those practising traditional livelihoods are particularly vulnerable and exposed to climate risks.
 5 Many depend on food self-provisioning from lakes, the sea, and the land. With higher temperatures, the
 6 availability of these sources of food is likely to be reduced, particularly in southern Europe. Poorer
 7 households often settle in flood-prone areas and are therefore more exposed to flooding. Traditional
 8 pastoralist and fishing practices are also negatively affected by climate change across Europe. Semi-
 9 migratory reindeer herding, a way of life among Indigenous and traditional communities (Komi, Saami,
10 Nenets) in the European Arctic, is threatened by reduced ice and snow cover. Almost 15% of the EU
11 population (in some countries more than 25%) already cannot meet their healthcare needs for financial
12 reasons, while they are at risk of health impacts from warming.

13

14 In addition to being more exposed to climate risks, socially vulnerable groups are also less able to adapt to
15 these risks, because of financial and institutional barriers. More than 20% of people in the South and East of
16 Europe live in dwellings that cannot be cooled to comfortable levels during summer. These people are
17 particularly vulnerable to risks from increasing heatwave days in European cities e.g. when they already face
18 energy poverty. They may also lack the means to protect against flooding or heat, e.g., when they do not own
19 the property. Risk-based insurance premiums, which are intended to help people reduce climate risks, are
20 potentially unaffordable for poor households. The ability to adapt is also often limited for Indigenous People,
21 as they often lack the rights and governance of resources, particularly when in competition with economic
22 interests such as resource mining, oil and gas, forestry, and expansion of bioenergy.

23

24 Adaptation actions by governments can both increase and decrease social inequality. The installation of new
25 or the restoration of existing green spaces may increase land prices and rents due to a higher attractiveness of
26 these areas, leading to potential displacement of population groups who cannot afford higher prices. On the
27 other hand, rewilding and restoration of ecosystems can improve the access of less privileged people to
28 ecosystem services and goods, such as the availability of freshwater. At city level, there are examples of
29 good practice in climate-resilient development that consider social equity which integrate a gender-inclusive
30 perspective in its sustainable urban planning, including designing public spaces and transit to ensure women,
31 disabled people and other groups can access and feel safe using these public amenities (see Box 13.3).

32

33 [END FAQ13.1 HERE]

34

35

36 [START FAQ13.2 HERE]

37

38 FAQ 13.2: What are the limits of adaptation for ecosystems in Europe?

39

40 Land, freshwater and ocean organisms and ecosystems across Europe are facing increasing pressures from
41 human activities. Climate change is rapidly becoming an additional and, in the future, a primary threat.
42 Ongoing and projected future changes are too severe and happen too fast for many organisms and
43 ecosystems to adapt. More expensive and better implemented environmental conservation and adaptation
44 measures can slow down, halt, and potentially reverse biodiversity and ecosystem declines, but only at low
45 or intermediate warming.

46

47 Ecosystem degradation and biodiversity loss have been evident across Europe since 1950, mainly due to land
48 use and overfishing. However, climate-change is becoming a key threat. The unprecedented pace of
49 environmental change has already surpassed the natural adaptive capability of many species, communities,
50 and ecosystems in Europe. For instance, the space available for some land ecosystems has shrunk, especially
51 in Europe's polar and mountain areas, due to warming and thawing of permafrost. Across Europe, heatwaves
52 and droughts and their impacts such as wildfires add further acute pressures, as seen in the 2018 heatwave,
53 which impacted forest ecosystems and their services. In the Mediterranean Sea, plants and animals cannot
54 shift northward and are negatively affected by marine heatwaves. Food-web dynamics of European
55 ecosystems are disrupted as climate change alters the timing of biological processes, such as spawning and
56 migration of species, and ecosystem composition. Moreover, warming fosters the immigration of invasive
57 species that compete with璦nd can even out-compete璽he native flora and fauna.

Do Not Cite, Quote or Distribute  13-88       Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1

 2 In a future with further and even stronger warming, climate change and its many impacts will become
 3 increasingly more important threats. Several species and ecosystems are projected to be already at high risk
 4 at 2癈 global warming, including fishes and lake and river ecosystems. At 3癈 global warming, many
 5 European ecosystems, such as coastal wetlands, peatlands, and forests, are projected to be at much higher
 6 risk of being severely disrupted than in a 2癈 warmer world. For example, Mediterranean seagrass meadows
 7 will very likely become extinct due to more frequent, longer, and more severe marine heat waves by 2050.
 8 Several wetland and forest plants and animals will be at high risk to be replaced by invasive species that are
 9 better adapted to increasingly dry conditions, especially in boreal and Arctic ecosystems.

10

11 Current protection and adaptation measures, such as the Natura 2000 network of protected areas, have some
12 positive effects for European ecosystems. However, these policies are not sufficient to effectively curb
13 overall ecosystem decline, especially for the projected higher risks above 2癈 global warming. Nature-based
14 solutions, such as the restoration of wetlands, peatlands, and forests, can serve both ecosystem protection and
15 climate-change mitigation through strengthening carbon sequestration. Some climate change mitigation
16 measures such as reforestation and restoration of coastal ecosystems can strengthen conservation measures.
17 These approaches are projected to reduce risks for European ecosystems and biodiversity, especially when
18 internationally coordinated.

19

20 Not all climate-change adaptation options are beneficial to ecosystems. When planning and implementing
21 adaptation options and Nature-based Solutions, trade-offs and unintended side effects should be considered.
22 On one hand, engineering coastal protection measures (seawalls, breakwaters, and similar infrastructure) in
23 response to sea level rise reduce the space available for coastal ecosystems. One the other hand, Nature-
24 based Solutions can also have unintended side effects, such as increased methane release from larger wetland
25 areas and large-scale tree planting changing the albedo of the surface.

26

27 [END FAQ13.2 HERE]

28

29

30 [START FAQ13.3 HERE]

31

32 FAQ 13.3: How can people adapt at individual and community level to heat waves in Europe?

33

34 Heatwaves will become more frequent, more intense and will last longer. A range of adaptation measures
35 are available for communities and individuals before, during and after a heat wave strikes. Implementing
36 adaptation measures are important to reduce the risks of future heat waves.

37

38 Heat waves will affect people in different ways, risks are higher for the elderly, pregnant women, small
39 children, people with pre-existing health conditions and low-income groups. By 2050, about half of the
40 European population may be exposed to high or very high risk of heat stress during summer, particularly in
41 Southern Europe and increasingly in Eastern and Western and Central Europe. The severity of heat-related
42 risks will be highest in large cities, due to the urban heat island effect.

43

44 In southern Europe people are already aware of the risks of heat extremes. Consequently, governments and
45 citizens have implemented a range of adaptation responses to reduce the impacts of heat waves. However,
46 there are limits to how much adaptation can be implemented. At 3癈 global warming there will be
47 substantial risks to human lives and productivity, which cannot be avoided. In the parts of Europe where heat
48 waves are a relatively new phenomena, such as many parts of North and Western and Central Europe, public
49 awareness of heat extremes is increasing and institutional capacity to respond is growing.

50

51 Preparing for heat waves is an important first step. Implementing and sustaining effective measures, such as
52 national or regional early warning and information systems, heat wave plans and guidelines, and raising
53 public awareness through campaigns are successful responses. Evidence suggests that such measures have
54 contributed to reduced mortality rates in Southern and Western and Central Europe. At city level, preparing
55 for heat waves can sometimes require urban redesign. For example, green-blue spaces, such as recreational
56 parks and ponds in cities, have been shown to reduce the average temperature in cities dramatically and to
57 provide co-benefits, such as improved air quality and recreational space. The use of cool materials in asphalt,

Do Not Cite, Quote or Distribute  13-89       Total pages: 143
    FINAL DRAFT                       Chapter 13      IPCC WGII Sixth Assessment Report

1 increasing reflectivity, green roofs, and building construction measures are being considered in urban
2 planning for reducing heat risks. Citizens can prepare themselves by using natural ventilation, using
3 approaches to stay cool in heat waves, green roofs and green facades in their buildings.

4

5 During heat waves, public information that is targeted at people and social care providers is critical,
6 particularly for the most vulnerable citizens. Governments and NGOs play an important role in informing

7 people about how to prepare and what to do to avoid health impacts and reduce mortality. Coordination
8 between vital emergency and health services is critical. Individuals can take several actions to effectively

 9 protect themselves from heat: 1) decreasing exposure to high temperatures (e.g., avoid outdoor during hottest
10 times of the day, access cool areas, wear protective and appropriate clothing); 2) keep hydrated (e.g., drink
11 enough proper fluids, avoid alcohol, etc.), 3) be sensitive to the symptoms of heat illness (dizziness, heavy

12 sweating, fatigue, cool and moist skin with goosebumps when in heat, etc.).

13

14 Once the heat wave has ended, evaluation of what worked well and how improvements can be made is key to

15 prepare for the next heat wave. Governments can, for example, evaluate whether the early warning systems
16 provided timely and useful information, whether coordination went smoothly, or assess estimated number of
17 lives saved as results of the measures implemented. Sharing these lessons learned is critical to allow other

18 cities and regions to plan for heat extremes. After the heatwave, citizens can reflect if their responses were
19 sufficient, whether investments are needed to be better prepared, and draw key lessons about what (not) to do
20 when the next heat wave strikes.

21

22 [END FAQ13.3 HERE]

23

24

25 [START FAQ13.4 HERE]

26

27 FAQ 13.4: What opportunities does climate change generate for human and natural systems in

28  Europe?

29

30 Not all climate change impacts across Europe pose challenges and threats to natural communities and

31 human society. In some regions, and for some sectors, opportunities will emerge . Although these

32 opportunities do not outweigh the negative impacts of climate change, considering these in adaptation
33 planning and implementation is important to benefit from them. Nevertheless, Europe will face difficult

34 decisions balancing the trade-offs between the adaptation needs of different sectors, regions and adaptation

35 and mitigation actions.

36

37 Opportunities of climate change can be 1) positive effects of warming for specific sectors and regions, such

38 as agriculture in northern Europe and 2) co-benefits of transformation of cities or transport measures that

39 reduce the speed and impact of climate-change while improving air quality, mental health and wellbeing.
40 Windows of action for transformation opportunities for large-scale transitions and transformation of our
41 society may be accelerated through new policy initiatives in response to the COVID crisis, such as the

42 European New Green Deal and Building Back Better.

43

44 As warming and droughts impact southern Europe most strongly, direct opportunities from climate change

45 are primarily in northern regions, thereby increasing existing inequalities across Europe. Across Europe,

46 positive effects of climate change are fewer than negative impacts and are typically limited to some aspects
47 of agriculture, forestry, tourism, and energy sectors. In the food sector, opportunities emerge by the

48 northward movement of food production zones, increases in plant growth due to CO2 fertilisation, and
49 reduction of heating costs for livestock during cold winters. In the energy sector, positive effects include
50 increased wind energy in the southwestern Mediterranean, and reduced energy demand for heating across

51 Europe. While climatic conditions for tourist activities are projected to decrease for winter tourism (e.g.
52 lacking sufficient snow) and summer tourism in some parts of Europe (e.g. too hot), conditions may improve

53 during spring and autumn in many European locations. Fewer cold waves will reduce risks on transport
54 infrastructure, such as cracking of road surface, in parts of Northern and Eastern Europe particularly by the
55 end of the century.

56

    Do Not Cite, Quote or Distribute  13-90                                     Total pages: 143
FINAL DRAFT                       Chapter 13  IPCC WGII Sixth Assessment Report

 1 Indirect opportunities emerge from the co-benefits of implementing adaptation actions. Some of these co-
 2 benefits are wide-spread but need careful consideration in order to be utilized. For example, Nature-based
 3 Solutions to adaptation can make cities and settlements more liveable, increase the resilience of agriculture,
 4 and protect biodiversity. Ecosystem-based adaptation can attract tourists and create recreational space. There
 5 are opportunities to mainstream adaptation into other developments and transitions, including the energy or
 6 agricultural transitions and COVID19 recovery plans. Transformative solutions to achieve sustainability may
 7 be accelerated through larger changes of, for example behaviour, energy, food or transport, to better exploit
 8 new opportunities and co-benefits. Implementation of adaptation actions can also help to make progress
 9 towards achieving the Sustainable Development Goals.

10

11 Inclusive, equitable and just adaptation is critical for climate resilient development considering SDGs,
12 gender, as well as Indigenous knowledge and local knowledge and practices. Implementation requires
13 political commitment, persistence and consistent action across scales of government. Upfront mobilization of
14 political, human and financial capital in implementation of adaptation actions is key, even when the benefits
15 are not immediately visible.

16

Do Not Cite, Quote or Distribute  13-91       Total pages: 143