diff --git "a/ipcc_statements_dataset.tsv" "b/ipcc_statements_dataset.tsv" new file mode 100644--- /dev/null +++ "b/ipcc_statements_dataset.tsv" @@ -0,0 +1,8095 @@ +statement_idx report page_num sent_num statement confidence score split +0 AR6_WGI 20 22 Since 2011 (measurements reported in AR5), concentrations have continued to increase in the atmosphere, reaching annual averages of 410 parts per million (ppm) for carbon dioxide (CO 2), 1866 parts per billion (ppb) for methane (CH 4), and 332 ppb for nitrous oxide (N 2O) in 2019.6 Land and ocean have taken up a near-constant proportion (globally about 56% per year) of CO 2 emissions from human activities over the past six decades, with regional differences high 2 train +1 AR6_WGI 21 8 Mid-latitude storm tracks have likely shifted poleward in both hemispheres since the 1980s, with marked seasonality in trends medium 1 train +2 AR6_WGI 21 18 The average rate of sea level rise was 1.3 [0.6 to 2.1] mm yr–1 between 1901 and 1971, increasing to 1.9 [0.8 to 2.9] mm yr–1 between 1971 and 2006, and further increasing to 3.7 [3.2 to 4.2] mm yr–1 between 2006 and 2018 high 2 train +3 AR6_WGI 24 2 Since 1750, increases in CO2 (47%) and CH4 (156%) concentrations far exceed – and increases in N2O (23%) are similar to – the natural multi-millennial changes between glacial and interglacial periods over at least the past 800,000 years very high 3 test +4 AR6_WGI 24 4 Temperatures during the most recent decade (2011–2020) exceed those of the most recent multi-century warm period, around 6500 years ago13 [0.2°C to 1°C relative to 1850–1900] medium 1 train +5 AR6_WGI 24 5 Prior to that, the next most recent warm period was about 125,000 years ago, when the multi-century temperature [0.5°C to 1.5°C relative to 1850–1900] overlaps the observations of the most recent decade medium 1 train +6 AR6_WGI 24 7 Late summer Arctic sea ice area was smaller than at any time in at least the past 1000 years medium 1 train +7 AR6_WGI 24 8 The global nature of glacier retreat since the 1950s, with almost all of the world’s glaciers retreating synchronously, is unprecedented in at least the last 2000 years medium 1 train +8 AR6_WGI 24 10 The global ocean has warmed faster over the past century than since the end of the last deglacial transition (around 11,000 years ago) medium 1 train +9 AR6_WGI 24 11 A long-term increase in surface open ocean pH occurred over the past 50 million years high 2 train +10 AR6_WGI 24 12 However, surface open ocean pH as low as recent decades is unusual in the last 2 million years medium 1 train +11 AR6_WGI 24 17 Marine heatwaves have approximately doubled in frequency since the 1980s high 2 train +12 AR6_WGI 24 19 Human-induced climate change has contributed to increases in agricultural and ecological droughts15 in some regions due to increased land evapotranspiration16 medium 1 train +13 AR6_WGI 25 1 Over South Asia, East Asia and West Africa, increases in monsoon precipitation due to warming from GHG emissions were counteracted by decreases in monsoon precipitation due to cooling from human-caused aerosol emissions over the 20th century high 2 train +14 AR6_WGI 25 2 Increases in West African monsoon precipitation since the 1980s are partly due to the growing influence of GHGs and reductions in the cooling effect of human-caused aerosol emissions over Europe and North America medium 1 train +15 AR6_WGI 25 5 Event attribution studies and physical understanding indicate that human-induced climate change increases heavy precipitation associated with tropical cyclones high 2 train +16 AR6_WGI 25 7 This includes increases in the frequency of concurrent heatwaves and droughts on the global scale (high confidence), fire weather in some regions of all inhabited continents (medium confidence), and compound flooding in some locations medium 1 train +17 AR6_WGI 27 21 The remainder is due to improved scientific understanding and changes in the assessment of aerosol forcing, which include decreases in concentration and improvement in its calculation high 2 train +18 AR6_WGI 27 23 The observed average rate of heating of the climate system increased from 0.50 [0.32 to 0.69] W m–2 for the period 1971–200619 to 0.79 [0.52 to 1.06] W m–2 for the period 2006–201820 high 2 train +19 AR6_WGI 27 24 Ocean warming accounted for 91% of the heating in the climate system, with land warming, ice loss and atmospheric warming accounting for about 5%, 3% and 1%, respectively high 2 train +20 AR6_WGI 27 28 Together, ice-sheet and glacier mass loss were the dominant contributors to global mean sea level rise during 2006–2018 high 2 train +21 AR6_WGI 27 30 Based on multiple lines of evidence,21 the very likely range of equilibrium climate sensitivity is between 2°C (high confidence) and 5°C medium 1 train +22 AR6_WGI 27 31 The AR6 assessed best estimate is 3°C with a likely range of 2.5°C to 4°C high 2 train +23 AR6_WGI 31 7 For example, every additional 0.5°C of global warming causes clearly discernible increases in the intensity and frequency of hot extremes, including heatwaves (very likely), and heavy precipitation (high confidence), as well as agricultural and ecological droughts30 in some regions high 2 train +24 AR6_WGI 31 8 Discernible changes in intensity and frequency of meteorological droughts, with more regions showing increases than decreases, are seen in some regions for every additional 0.5°C of global warming medium 1 train +25 AR6_WGI 31 9 Increases in frequency and intensity of hydrological droughts become larger with increasing global warming in some regions medium 1 train +26 AR6_WGI 31 11 Projected percentage changes in frequency are larger for rarer events high 2 train +27 AR6_WGI 31 13 The Arctic is projected to experience the highest increase in the temperature of the coldest days, at about three times the rate of global warming high 2 train +28 AR6_WGI 31 14 With additional global warming, the frequency of marine heatwaves will continue to increase (high confidence), particularly in the tropical ocean and the Arctic medium 1 train +29 AR6_WGI 32 1 At the global scale, extreme daily precipitation events are projected to intensify by about 7% for each 1°C of global warming high 2 train +30 AR6_WGI 32 2 The proportion of intense tropical cyclones (Category 4–5) and peak wind speeds of the most intense tropical cyclones are projected to increase at the global scale with increasing global warming high 2 train +31 AR6_WGI 35 20 The portion of the global land experiencing detectable increases or decreases in seasonal mean precipitation is projected to increase medium 1 train +32 AR6_WGI 35 25 The monsoon season is projected to have a delayed onset over North and South America and West Africa (high confidence) and a delayed retreat over West Africa medium 1 train +33 AR6_WGI 36 1 This is projected to result in a higher proportion of emitted CO 2 remaining in the atmosphere high 2 train +34 AR6_WGI 36 3 Under the very low and low GHG emissions scenarios (SSP1-1.9, SSP1-2.6), where CO 2 concentrations peak and decline during the 21st century, the land and ocean begin to take up less carbon in response to declining atmospheric CO 2 concentrations (high confidence) and turn into a weak net source by 2100 under SSP1-1.9 medium 1 train +35 AR6_WGI 36 4 It is very unlikely that the combined global land and ocean sink will turn into a source by 2100 under scenarios without net negative emissions (SSP2-4.5, SSP3-7.0, SSP5-8.5).32 {4.3, 5.4, 5.5, 5.6, Box TS.5, TS.3.3} B.4.3 The magnitude of feedbacks between climate change and the carbon cycle becomes larger but also more uncertain in high CO 2 emissions scenarios very high 3 train +36 AR6_WGI 36 5 However, climate model projections show that the uncertainties in atmospheric CO 2 concentrations by 2100 are dominated by the differences between emissions scenarios high 2 train +37 AR6_WGI 36 6 Additional ecosystem responses to warming not yet fully included in climate models, such as CO 2 and CH 4 fluxes from wetlands, permafrost thaw and wildfires, would further increase concentrations of these gases in the atmosphere high 2 train +38 AR6_WGI 37 7 Based on multiple lines of evidence, upper ocean stratification (virtually certain), ocean acidification (virtually certain) and ocean deoxygenation high 2 train +39 AR6_WGI 37 8 Changes are irreversible on centennial to millennial time scales in global ocean temperature (very high confidence), deep-ocean acidification (very high confidence) and deoxygenation medium 1 train +40 AR6_WGI 37 10 Loss of permafrost carbon following permafrost thaw is irreversible at centennial time scales high 2 train +41 AR6_WGI 37 14 Relative to 1995–2014, the likely global mean sea level rise by 2100 is 0.28–0.55 m under the very low GHG emissions scenario (SSP1-1.9); 0.32–0.62 m under the low GHG emissions scenario (SSP1-2.6); 0.44–0.76 m under the intermediate GHG emissions scenario (SSP2-4.5); and 0.63–1.01 m under the very high GHG emissions scenario (SSP5-8.5); and by 2150 is 0.37–0.86 m under the very low scenario (SSP1-1.9); 0.46–0.99 m under the low scenario (SSP1-2.6); 0.66–1.33 m under the intermediate scenario (SSP2-4.5); and 0.98–1.88 m under the very high scenario (SSP5-8.5) (medium confidence).35 Global mean sea level rise above the likely range – approaching 2 m by 2100 and 5 m by 2150 under a very high GHG emissions scenario (SSP5-8.5) low 0 train +42 AR6_WGI 37 16 Over the next 2000 years, global mean sea level will rise by about 2 to 3 m if warming is limited to 1.5°C, 2 to 6 m if limited to 2°C and 19 to 22 m with 5°C of warming, and it will continue to rise over subsequent millennia low 0 test +43 AR6_WGI 37 17 Projections of multi-millennial global mean sea level rise are consistent with reconstructed levels during past warm climate periods: likely 5–10 m higher than today around 125,000 years ago, when global temperatures were very likely 0.5°C–1.5°C higher than 1850–1900; and very likely 5–25 m higher roughly 3 million years ago, when global temperatures were 2.5°C–4°C higher medium 1 train +44 AR6_WGI 39 32 For example, internal decadal variability and variations in solar and volcanic drivers partially masked human-caused surface global warming during 1998–2012, with pronounced regional and seasonal signatures high 2 train +45 AR6_WGI 39 33 Nonetheless, the heating of the climate system continued during this period, as reflected in both the continued warming of the global ocean (very high confidence) and in the continued rise of hot extremes over land medium 1 train +46 AR6_WGI 40 1 At global and regional scales, near-term changes in monsoons will be dominated by the effects of internal variability medium 1 train +47 AR6_WGI 40 2 In addition to the influence of internal variability, near-term projected changes in precipitation at global and regional scales are uncertain because of model uncertainty and uncertainty in forcings from natural and anthropogenic aerosols medium 1 train +48 AR6_WGI 40 8 Further decreases are projected in permafrost; snow, glaciers and ice sheets; and lake and Arctic sea ice (medium to high confidence).40 These changes would be larger at 2°C global warming or above than at 1.5°C high 2 train +49 AR6_WGI 40 9 For example, extreme heat thresholds relevant to agriculture and health are projected to be exceeded more frequently at higher global warming levels high 2 train +50 AR6_WGI 40 11 Also, more frequent and/or severe agricultural and ecological droughts are projected in a few regions in all inhabited continents except Asia compared to 1850–1900 (medium confidence); increases in meteorological droughts are also projected in a few regions medium 1 train +51 AR6_WGI 40 12 A small number of regions are projected to experience increases or decreases in mean precipitation medium 1 train +52 AR6_WGI 40 14 Heavy precipitation and associated flooding events are projected to become more intense and frequent in the Pacific Islands and across many regions of North America and Europe (medium to high confidence).40 These changes are also seen in some regions in Australasia and Central and South America medium 1 train +53 AR6_WGI 40 16 A small number of regions in Africa, Australasia, Europe and North America are also projected to be affected by increases in hydrological droughts, and several regions are projected to be affected by increases or decreases in meteorological droughts, with more regions displaying an increase medium 1 train +54 AR6_WGI 40 17 Mean precipitation is projected to increase in all polar, northern European and northern North American regions, most Asian regions and two regions of South America high 2 train +55 AR6_WGI 41 1 Region-specific changes include intensification of tropical cyclones and/or extratropical storms medium 1 train +56 AR6_WGI 41 3 Approximately two-thirds of the global coastline has a projected regional relative sea level rise within ±20% of the global mean increase medium 1 train +57 AR6_WGI 41 4 Due to relative sea level rise, extreme sea level events that occurred once per century in the recent past are projected to occur at least annually at more than half of all tide gauge locations by 2100 high 2 train +58 AR6_WGI 41 5 Relative sea level rise contributes to increases in the frequency and severity of coastal flooding in low-lying areas and to coastal erosion along most sandy coasts high 2 train +59 AR6_WGI 41 7 Urbanization also increases mean and heavy precipitation over and/or downwind of cities (medium confidence) and resulting runoff intensity high 2 train +60 AR6_WGI 41 8 In coastal cities, the combination of more frequent extreme sea level events (due to sea level rise and storm surge) and extreme rainfall/ riverflow events will make flooding more probable high 2 train +61 AR6_WGI 41 11 Concurrent extremes at multiple locations, including in crop-producing areas, become more frequent at 2°C and above compared to 1.5°C global warming high 2 train +62 AR6_WGI 43 4 The probability of low-likelihood, high-impact outcomes increases with higher global warming levels high 2 train +63 AR6_WGI 43 5 Abrupt responses and tipping points of the climate system, such as strongly increased Antarctic ice-sheet melt and forest dieback, cannot be ruled out high 2 train +64 AR6_WGI 45 13 When adjusted for emissions since previous reports, estimates of remaining carbon budgets are therefore of similar magnitude compared to SR1.5 but larger compared to AR5 due to methodological improvements.44 {5.5, Box 5.2, TS.3.3} (Table SPM.2) D.1.4 Anthropogenic CO 2 removal (CDR) has the potential to remove CO 2 from the atmosphere and durably store it in reservoirs high 2 train +65 AR6_WGI 45 15 CDR methods can have potentially wide-ranging effects on biogeochemical cycles and climate, which can either weaken or strengthen the potential of these methods to remove CO 2 and reduce warming, and can also influence water availability and quality, food production and biodiversity45 high 2 train +66 AR6_WGI 46 1 CDR would lower atmospheric CO 2 by an amount approximately equal to the increase from an anthropogenic emission of the same magnitude high 2 train +67 AR6_WGI 46 2 The atmospheric CO 2 decrease from anthropogenic CO 2 removals could be up to 10% less than the atmospheric CO 2 increase from an equal amount of CO 2 emissions, depending on the total amount of CDR medium 1 train +68 AR6_WGI 46 4 For instance, it would take several centuries to millennia for global mean sea level to reverse course even under large net negative CO 2 emissions high 2 train +69 AR6_WGI 46 6 In the long term, this net warming is lower in scenarios assuming air pollution controls combined with strong and sustained CH 4 emissions reductions high 2 train +70 AR6_WGI 46 8 Because of the short lifetime of both CH 4 and aerosols, these climate effects partially counterbalance each other, and reductions in CH 4 emissions also contribute to improved air quality by reducing global surface ozone high 2 train +71 AR6_WGI 46 12 Emissions pathways that reach and sustain net zero GHG emissions defined by the 100-year global warming potential are projected to result in a decline in surface temperature after an earlier peak high 2 train +72 AR6_WGI 46 14 Under these contrasting scenarios, discernible differences in trends of global surface temperature would begin to emerge from natural variability within around 20 years, and over longer time periods for many other climatic impact-drivers high 2 train +73 AR6_WGI 46 16 Global and regional climate responses to this temporary forcing are, however, undetectable above natural variability high 2 train +74 AR6_WGI 46 17 Atmospheric CO 2 concentrations continued to rise in 2020, with no detectable decrease in the observed CO 2 growth rate medium 1 train +75 AR6_WGI 47 1 Scenarios with targeted reductions of air pollutant emissions lead to more rapid improvements in air quality within years compared to reductions in GHG emissions only, but from 2040, further improvements are projected in scenarios that combine efforts to reduce air pollutants as well as GHG emissions, with the magnitude of the benefit varying between regions high 2 train +76 AR6_WGI 47 5 The response of many other climate variables would emerge from natural variability at different times later in the 21st century high 2 train +77 AR6_WGI 47 7 By the end of the century, scenarios with very low and low GHG emissions would strongly limit the change of several CIDs, such as the increases in the frequency of extreme sea level events, heavy precipitation and pluvial flooding, and exceedance of dangerous heat thresholds, while limiting the number of regions where such exceedances occur, relative to higher GHG emissions scenarios high 2 test +78 AR6_WGI 47 8 Changes would also be smaller in very low compared to low GHG emissions scenarios, as well as for intermediate (SSP2-4.5) compared to high or very high GHG emissions scenarios high 2 train +79 AR6_WGI 57 25 Since 2012, strong warming has been observed, with the past five years (2016–2020) being the hottest five- year period in the instrumental record since at least 1850 high 2 train +80 AR6_WGI 62 1 Changes in temperature by latitude in response to multiple forcings show that polar amplification (stronger warming at high latitudes than the global average) is a prominent feature of the climate system across multiple climate states, and the ability of models to simulate this polar amplification in past warm climates has improved since AR5 high 2 train +81 AR6_WGI 62 2 Over the past millennium, and especially since about 1300 CE, simulated global surface temperature anomalies are well within the uncertainty of reconstructions medium 1 test +82 AR6_WGI 62 7 Model uncertainties in (a) and (b) are 5–95% ranges of multi-model ensemble means; reconstructed uncertainties are 5–95% ranges medium 1 train +83 AR6_WGI 62 15 Climate models are able to reproduce decadal drought variability on large regional scales, including the severity, persistence and spatial extent of past megadroughts known from proxy records medium 1 train +84 AR6_WGI 63 1 Although past and future global warming differ in their forcings, evidence from paleoclimate records and modelling show that ice-sheet mass and global mean sea level (GMSL) responded dynamically over multiple millennia high 2 train +85 AR6_WGI 63 3 For example, under a past global warming levels of around [2.5°C to 4°C] relative to 1850–1900, like during the mid-Pliocene Warm Period, sea level was [5 to 25 m] higher than 1900 (medium confidence); under past global warming levels of [10°C to 18°C], like during the Early Eocene, the planet was essentially ice free high 2 train +86 AR6_WGI 63 8 The paleo context supports the assessment that ongoing increase in ocean heat content (OHC) represents a long-term commitment, essentially irreversible on human time scales high 2 test +87 AR6_WGI 63 16 At the same time, the storage of remineralized carbon in the ocean interior increased by as much as 750–950 PgC, sufficient to balance the removal of carbon from the atmosphere (200 PgC) and terrestrial biosphere reservoirs combined high 2 train +88 AR6_WGI 64 16 Increasing horizontal resolution in global climate models improves the representation of small-scale features and the statistics of daily precipitation high 2 train +89 AR6_WGI 64 17 Earth system models, which include additional biogeochemical feedbacks, often perform as well as their lower-complexity global climate model counterparts, which do not account for these additional feedbacks medium 1 train +90 AR6_WGI 64 26 Earth system models, characterized by additional biogeochemical feedbacks, often perform at least as well as related, more constrained, lower-complexity models lacking these feedbacks medium 1 train +91 AR6_WGI 64 29 The CMIP6 models also reproduce surface temperature variations over the past millennium, including the cooling that follows periods of intense volcanism medium 1 train +92 AR6_WGI 65 4 Because a negative cloud optical depth feedback in response to surface warming results from ‘brightening’ of clouds via active phase change from ice to liquid cloud particles (increasing their shortwave cloud radiative effect), the extratropical cloud shortwave feedback in CMIP6 models tends to be less negative, leading to a better agreement with observational estimates medium 1 train +93 AR6_WGI 65 7 Increasing horizontal resolution in global climate models improves the representation of small-scale features and the statistics of daily precipitation high 2 train +94 AR6_WGI 65 11 The overall performance of CMIP6 models in simulating the intensity and frequency of extreme precipitation is similar to that of CMIP5 models high 2 train +95 AR6_WGI 65 21 Models reproduce observed spatial features and variance of the Southern Annular Mode (SAM) and Northern Annular Mode (NAM) very well high 2 train +96 AR6_WGI 65 22 The summertime SAM trend is well captured, with CMIP6 models outperforming CMIP5 models medium 1 train +97 AR6_WGI 65 25 Model performance is limited in reproducing sea surface temperature anomalies for decadal modes of variability, despite improvements from CMIP5 to CMIP6 medium 1 train +98 AR6_WGI 65 30 These higher sensitivity values can, in some models, be traced to changes in extratropical cloud feedbacks medium 1 test +99 AR6_WGI 66 12 The signal of temperature change has emerged more clearly in tropical regions, where year-to-year variations tend to be small over land, than in regions with greater warming but larger year- to-year variations high 2 train +100 AR6_WGI 66 16 Global ocean heat content continued to increase throughout this period, indicating continuous warming of the entire climate system very high 3 train +101 AR6_WGI 66 17 Hot extremes also continued to increase during this period over land high 2 train +102 AR6_WGI 66 18 Even in a continually warming climate, periods of reduced and increased trends in global surface temperature at decadal time scales will continue to occur in the 21st century very high 3 train +103 AR6_WGI 67 1 Simulations and understanding of modes of climate variability, including teleconnections, have improved since AR5 medium 1 train +104 AR6_WGI 67 3 While anthropogenic forcing has contributed to multi-decadal mean precipitation changes in several regions, internal variability can delay emergence of the anthropogenic signal in long-term precipitation changes in many land regions high 2 train +105 AR6_WGI 67 6 Several impact-relevant changes have not yet emerged from natural variability but will emerge sooner or later in this century depending on the emissions scenario high 2 train +106 AR6_WGI 67 7 Ocean acidification and deoxygenation have already emerged over most of the global open ocean, as has a reduction in Arctic sea ice high 2 train +107 AR6_WGI 68 4 New techniques and analyses drawing on several lines of evidence have provided greater confidence in attributing changes in regional weather and climate extremes to human influence high 2 train +108 AR6_WGI 68 11 For example, the difference in observed warming trends between cities and their surroundings can partly be attributed to urbanization very high 3 train +109 AR6_WGI 68 13 Multiple attribution approaches support the contribution of human influence to several regional multi-decadal mean precipitation changes high 2 train +110 AR6_WGI 68 14 Understanding about past and future changes in weather and climate extremes has increased due to better observation-based datasets, physical understanding of processes, an increasing proportion of scientific literature combining different lines of evidence, and improved accessibility to different types of climate models high 2 train +111 AR6_WGI 71 8 In SSP1-2.6 and SSP2-4.5, changes in ERF also explain about half of the changes in the range of warming medium 1 train +112 AR6_WGI 71 9 For SSP5-8.5, higher climate sensitivity is the primary reason behind the upper end of the CMIP6- projected warming being higher than for RCP8.5 in CMIP5 medium 1 train +113 AR6_WGI 71 14 The differences in the few ESMs for which both emissions and concentration-driven runs were available for the same scenario are small and do not affect the assessment of global surface temperature projections discussed in Cross-Section Box TS.1 and Section TS.2 high 2 train +114 AR6_WGI 73 8 Multiple lines of evidence can be used to construct climate information on a global to regional scale and can be further distilled in a co-production process to meet user needs high 2 train +115 AR6_WGI 73 12 Many global and regional climatic impact-drivers have a direct relation to global warming levels high 2 train +116 AR6_WGI 75 4 Climate change has already altered CID profiles and resulted in shifting magnitude, frequency, duration, seasonality and spatial extent of associated indices high 2 train +117 AR6_WGI 75 6 These include heat, cold, wet and dry hazards, both mean and extremes; cryospheric hazards (snow cover, ice extent, permafrost) and oceanic hazards (marine heatwaves) high 2 train +118 AR6_WGI 75 12 Changes in GMST and GSAT over time differ by at most 10% in either direction high 2 train +119 AR6_WGI 75 16 Temperatures as high as during the most recent decade (2011–2020) exceed the warmest centennial-scale range reconstructed for the present interglacial, around 6500 years ago [0.2°C to 1°C] medium 1 train +120 AR6_WGI 75 17 The next most recent warm period was about 125,000 years ago during the last interglacial when the multi-centennial temperature range [0.5°C to 1.5°C] encompasses the 2011–2020 values medium 1 train +121 AR6_WGI 75 22 Global surface temperature in any individual year could exceed 1.5°C relative to 1850–1900 by 2030 with a likelihood between 40% and 60% across the scenarios considered here medium 1 train +122 AR6_WGI 75 24 Periods of reduced and increased global surface temperature trends at decadal time scales will continue to occur in the 21st century very high 3 train +123 AR6_WGI 76 5 It is likely that there was a net anthropogenic forcing of 0.0–0.3 Wm–2 in 1850–1900 relative to 1750 medium 1 train +124 AR6_WGI 76 7 Beginning approximately 6500 years ago, global surface temperature generally decreased, culminating in the coldest multi-century interval of the post-glacial period (since roughly 7000 years ago), which occurred between around 1450 and 1850 high 2 train +125 AR6_WGI 76 8 Over the last 50 years, global surface temperature has increased at an observed rate unprecedented in at least the last two thousand years high 2 train +126 AR6_WGI 76 9 Temperatures as high as during the most recent decade (2011–2020) exceed the warmest centennial-scale range reconstructed for the present interglacial, around 6500 years ago [0.2°C to 1°C] medium 1 train +127 AR6_WGI 76 10 The next most recent warm period was about 125,000 years ago during the Last Interglacial when the multi-centennial temperature range [0.5°C to 1.5°C] encompasses the 2011–2020 values medium 1 train +128 AR6_WGI 76 11 During the mid-Pliocene Warm Period, around 3.3–3.0 million years ago, global surface temperature was 2.5°C to 4°C warmer medium 1 train +129 AR6_WGI 76 20 Furthermore, the heating of the climate system continued during this period, as reflected in the continued warming of the global ocean (very high confidence) and in the continued rise of hot extremes over land medium 1 train +130 AR6_WGI 76 21 Since 2012, global surface temperature has risen strongly, with the past five years (2016–2020) being the hottest five-year period between 1850 and 2020 high 2 train +131 AR6_WGI 78 3 Vertical bars are 5–95th percentile ranges of estimated global surface temperature for the Last Interglacial and mid-Holocene medium 1 train +132 AR6_WGI 78 22 The other half arises because for central estimates of climate sensitivity, most scenarios show stronger warming over the near term than was estimated as ‘current’ in SR1.5 medium 1 train +133 AR6_WGI 78 25 If climate sensitivity lies near the lower end of the assessed very likely range, crossing the 1.5°C warming level is avoided in scenarios SSP1-1.9 and SSP1-2.6 medium 1 train +134 AR6_WGI 78 26 Global surface temperature in any individual year, in contrast to the 20-year average, could by 2030 exceed 1.5°C relative to 1850–1900 with a likelihood between 40% and 60%, across the scenarios considered here medium 1 train +135 AR6_WGI 78 34 The uncertainty ranges for the period 2081–2100 continue to be dominated by the uncertainty in equilibrium climate sensitivity and transient climate response very high 3 train +136 AR6_WGI 79 14 Continued GHG emissions greatly increase the likelihood of potentially irreversible changes in the global climate system (Box TS.9), in particular with respect to the contribution of ice sheets to global sea level change high 2 train +137 AR6_WGI 83 11 Global mean concentrations of anthropogenic aerosols peaked in the late 20th century and have slowly declined since in northern mid-latitudes, although they continue to increase in South Asia and East Africa high 2 train +138 AR6_WGI 83 12 The total anthropogenic effective radiative forcing (ERF) in 2019, relative to 1750, was 2.72 [1.96 to 3.48] W m–2 medium 1 train +139 AR6_WGI 83 14 The average magnitude and variability of volcanic aerosols since 1900 has not been unusual compared to at least the past 2500 years medium 1 train +140 AR6_WGI 85 2 The centennial rate of change of CO 2 since 1850 has no precedent in at least the past 800,000 years (Figure TS.9), and the fastest rates of change over the last 56 million years were at least a factor of four lower low 0 train +141 AR6_WGI 85 6 The increase since 1750 of 1137 ± 10 ppb (157.8%) far exceeds the range over multiple glacial–interglacial transitions of the past 800,000 years high 2 train +142 AR6_WGI 85 12 N 2O concentration trends since 1980 are largely driven by a 30% increase in emissions from the expansion and intensification of global agriculture high 2 train +143 AR6_WGI 85 16 Abundances of HFCs, which are replacements for CFCs and HCFCs, are increasing high 2 train +144 AR6_WGI 85 19 Ice cores show increases in aerosols across the Northern Hemisphere mid-latitudes since 1700 and reductions since the late 20th century high 2 train +145 AR6_WGI 85 20 Aerosol optical depth (AOD), derived from satellite- and ground-based radiometers, has decreased since 2000 over the mid-latitude continents of both hemispheres, but increased over South Asia and East Africa high 2 train +146 AR6_WGI 85 22 Global carbonaceous aerosol budgets and trends remain poorly characterized due to limited observations, but black carbon (BC), a warming aerosol component, is declining in several regions of the Northern Hemisphere low 0 train +147 AR6_WGI 85 23 Total aerosol ERF in 2019, relative to 1750, is −1.1 [−1.7 to −0.4] W m−2 medium 1 train +148 AR6_WGI 85 25 Since the mid-20th century, tropospheric ozone surface concentrations have increased by 30–70% across the Northern Hemisphere medium 1 train +149 AR6_WGI 85 26 Future changes in surface ozone concentrations will be primarily driven by changes in precursor emissions rather than climate change high 2 train +150 AR6_WGI 85 27 Stratospheric ozone has declined between 60°S–60°N by 2.2% from 1964–1980 to 2014–2017 high 2 train +151 AR6_WGI 85 31 Model estimates suggest no significant change in oxidizing capacity from 1850 to 1980 low 0 test +152 AR6_WGI 86 6 The proportion of tropical cyclones that are intense is expected to increase (high confidence), but the total global number of tropical cyclones is expected to decrease or remain unchanged medium 1 train +153 AR6_WGI 87 2 In the tropics, since at least 2001 (when new techniques permit more robust quantification), the upper troposphere has warmed faster than the near-surface medium 1 train +154 AR6_WGI 87 6 This has been accompanied by a strengthening of the Hadley Circulation in the Northern Hemisphere medium 1 train +155 AR6_WGI 87 7 It is likely that human influence has contributed to the poleward expansion of the zonal mean Hadley cell in the Southern Hemisphere since the 1980s, which is projected to further expand with global warming high 2 train +156 AR6_WGI 87 26 The frequency of intense extratropical cyclones is projected to decrease medium 1 train +157 AR6_WGI 87 27 Projected changes in the intensity depend on the resolution of climate models medium 1 test +158 AR6_WGI 88 6 With increasing global warming, some very rare extremes and some compound events (multivariate or concurrent extremes) with low likelihood in past and current climate will become more frequent, and there is a higher chance that events unprecedented in the observational record occur high 2 train +159 AR6_WGI 88 25 Continued Amazon deforestation, combined with a warming climate, raises the probability that this ecosystem will cross a tipping point into a dry state during the 21st century low 0 train +160 AR6_WGI 89 5 Compound events and concurrent extremes contribute to increasing probability of low-likelihood, high-impact outcomes and will become more frequent with increasing global warming high 2 train +161 AR6_WGI 90 1 Over the past four to six decades, it is virtually certain that the global ocean has warmed, with human influence extremely likely the main driver since the 1970s, making climate change irreversible over centuries to millennia medium 1 train +162 AR6_WGI 90 4 A long-term increase in surface open ocean pH occurred over the past 50 million years, and surface ocean pH as low as recent times is uncommon in the last 2 million years medium 1 test +163 AR6_WGI 90 8 Stratification (virtually certain), acidification (virtually certain), deoxygenation (high confidence) and marine heatwave frequency high 2 train +164 AR6_WGI 90 11 The ocean is currently warming faster than at any other time since at least the last deglacial transition (medium confidence), with warming extending to depths well below 2000 m very high 3 train +165 AR6_WGI 90 14 Ocean warming is irreversible over centuries to millennia, but the magnitude of warming is scenario-dependent from about the mid-21st century medium 1 test +166 AR6_WGI 90 15 The warming will not be globally uniform, with heat primarily stored in Southern Ocean water-masses and weaker warming in the subpolar North Atlantic high 2 train +167 AR6_WGI 90 18 Marine heatwaves have become more frequent over the 20th century (high confidence), approximately doubling in frequency (high confidence) and becoming more intense and longer since the 1980s medium 1 train +168 AR6_WGI 90 24 It is extremely likely that human influence has contributed to this salinity change and that the large-scale pattern will grow in amplitude over the 21st century medium 1 train +169 AR6_WGI 90 27 Direct observational records since the mid-2000s are too short to determine the relative contributions of internal variability, natural forcing and anthropogenic forcing to AMOC change high 2 train +170 AR6_WGI 90 31 Western boundary currents and subtropical gyres have shifted poleward since 1993 medium 1 train +171 AR6_WGI 90 32 Subtropical gyres, the East Australian Current Extension, the Agulhas Current, and the Brazil Current are projected to intensify in the 21st century in response to changes in wind stress, while the Gulf Stream and the Indonesian Throughflow are projected to weaken medium 1 train +172 AR6_WGI 90 33 All of the four main eastern boundary upwelling systems are projected to weaken at low latitudes and intensify at high latitudes in the 21st century high 2 train +173 AR6_WGI 92 1 Ocean acidification and associated reductions in the saturation state of calcium carbonate – a constituent of skeletons or shells of a variety of marine organisms – is expected to increase in the 21st century under all emissions scenarios high 2 train +174 AR6_WGI 92 2 A long-term increase in surface open ocean pH occurred over the past 50 million years (high confidence), and surface ocean pH as low as recent times is uncommon in the last 2 million years medium 1 train +175 AR6_WGI 92 4 Over the past 2–3 decades, a pH decline in the ocean interior has been observed in all ocean basins high 2 train +176 AR6_WGI 92 6 Deoxygenation is projected to continue to increase with ocean warming high 2 train +177 AR6_WGI 92 9 The range of a smaller subset of organisms has shifted equatorward and to shallower depths high 2 train +178 AR6_WGI 92 10 Phenological metrics associated with the life cycles of many organisms have also changed over the last two decades or longer high 2 train +179 AR6_WGI 92 16 The Arctic Ocean is projected to become practically sea ice-free in late summer under high CO 2 emissions scenarios by the end of the 21st century high 2 train +180 AR6_WGI 92 18 Glaciers will continue to lose mass at least for several decades even if global temperature is stabilized very high 3 train +181 AR6_WGI 92 21 Since the late 1970s, Arctic sea ice area and thickness have decreased in both summer and winter, with sea ice becoming younger, thinner and more dynamic very high 3 train +182 AR6_WGI 92 22 It is very likely that anthropogenic forcing, mainly due to greenhouse gas increases, was the main driver of this loss, although new evidence suggests that anthropogenic aerosol forcing has offset part of the greenhouse gas-induced losses since the 1950s medium 1 train +183 AR6_WGI 92 24 This practically sea ice-free state will become the norm for late summer by the end of the 21st century in high CO 2 emissions scenarios high 2 train +184 AR6_WGI 92 25 Arctic summer sea ice varies approximately linearly with global surface temperature, implying that there is no tipping point and observed/ projected losses are potentially reversible high 2 train +185 AR6_WGI 92 29 For each additional 1°C of warming (up to 4°C above the 1850–1900 level), the global volume of perennially frozen ground to 3 m below the surface is projected to decrease by about 25% relative to the present volume medium 1 train +186 AR6_WGI 92 30 However, these decreases may be underestimated due to an incomplete representation of relevant physical processes in ESMs low 0 train +187 AR6_WGI 93 3 Under RCP2.6 and RCP8.5, respectively, glaciers are projected to lose 18% ± 13% and 36% ± 20% of their current mass over the 21st century medium 1 train +188 AR6_WGI 93 6 It is virtually certain that the Greenland Ice Sheet has lost mass since the 1990s, with human influence a contributing factor medium 1 train +189 AR6_WGI 93 10 Projections of future Greenland ice-mass loss (Box TS.4, Table 1; Figure TS.11e) are dominated by increased surface melt under all emissions scenarios high 2 train +190 AR6_WGI 93 13 The total Antarctic ice mass losses were dominated by the West Antarctic Ice Sheet, with combined West Antarctic and Peninsula annual loss rates increasing since about 2000 very high 3 test +191 AR6_WGI 93 16 Mass losses from West Antarctic outlet glaciers, mainly induced by ice shelf basal melt (high confidence), outpace mass gain from increased snow accumulation on the continent very high 3 train +192 AR6_WGI 93 22 Human activities were very likely the main driver of observed GMSL rise since 1971, and new observational evidence leads to an assessed sea level rise over the period 1901 to 2018 that is consistent with the sum of individual components contributing to sea level rise, including expansion due to ocean warming and melting of glaciers and ice sheets high 2 train +193 AR6_WGI 93 24 Sea level responds to greenhouse gas (GHG) emissions more slowly than global surface temperature, leading to weaker scenario dependence over the 21st century than for global surface temperature high 2 train +194 AR6_WGI 93 25 This slow response also leads to long-term committed sea level rise, associated with ongoing ocean heat uptake and the slow adjustment of the ice sheets, that will continue over the centuries and millennia following cessation of emissions high 2 train +195 AR6_WGI 93 26 By 2100, GMSL is projected to rise by 0.28– 0.55 m (likely range) under SSP1-1.9 and 0.63–1.01 m (likely range) under SSP5-8.5 relative to the 1995–2014 average medium 1 train +196 AR6_WGI 93 33 New analyses and paleo-evidence since AR5 show this rate is very likely faster than during any century over at least the last three millennia high 2 train +197 AR6_WGI 93 34 Since AR5, there is strengthened evidence for an increase in the rate of GMSL rise since the mid-20th century, with an average rate of 2.3 [1.6–3.1] mm yr–1 over the period 1971–2018 increasing to 3.7 [3.2–4.2] mm yr–1 for the period 2006–2018 high 2 test +198 AR6_WGI 95 11 By 2300, GMSL will rise 0.3–3.1 m under low CO 2 emissions (SSP1-2.6) low 0 train +199 AR6_WGI 95 12 Under high CO 2 emissions (SSP5-8.5), projected GMSL rise is between 1.7 and 6.8 m by 2300 in the absence of MICI and by up to 16 m considering MICI low 0 train +200 AR6_WGI 95 19 However, over the 21st century, the majority of coastal locations have a median projected regional sea level rise within ±20% of the projected GMSL change medium 1 train +201 AR6_WGI 95 23 Projections show that while land and ocean sinks absorb more CO 2 under high emissions scenarios than low emissions scenarios, the fraction of emissions removed from the atmosphere by natural sinks decreases with higher concentrations high 2 train +202 AR6_WGI 96 6 The combustion of fossil fuels and land-use change for the period 1750–2019 resulted in the release of 700 ± 75 PgC (likely range, 1 PgC = 1015 g of carbon) to the atmosphere, of which about 41% ± 11% remains in the atmosphere today high 2 train +203 AR6_WGI 96 9 During the last decade (2010–2019), average annual anthropogenic CO 2 emissions reached the highest levels in human history at 10.9 ± 0.9 PgC yr–1 high 2 train +204 AR6_WGI 96 10 Of these emissions, 46% accumulated in the atmosphere (5.1 ± 0.02 PgC yr–1), 23% (2.5 ± 0.6 PgC yr–1) was taken up by the ocean and 31% (3.4 ± 0.9 PgC yr–1) was removed by terrestrial ecosystems high 2 train +205 AR6_WGI 96 12 This coherence between emissions and the growth in ocean and land sinks has resulted in the airborne fraction of anthropogenic CO 2 remaining at 44 ± 10% over the past 60 years high 2 train +206 AR6_WGI 96 13 Interannual and decadal variability of the ocean and land sinks indicate that they are sensitive to changes in the growth rate of emissions as well as climate variability and are therefore also sensitive to climate change high 2 train +207 AR6_WGI 96 15 Since the 1980s, carbon fertilization from rising atmospheric CO 2 has increased the strength of the net land CO 2 sink medium 1 train +208 AR6_WGI 96 21 Carbon dioxide emissions-driven simulations account for uncertainty in these feedbacks, but do not significantly change the projected global surface temperature changes high 2 train +209 AR6_WGI 96 22 Although land and ocean sinks absorb more CO 2 under high emissions than low emissions scenarios, the fraction of emissions removed from the atmosphere decreases high 2 train +210 AR6_WGI 96 23 This means that the more CO 2 that is emitted, the less efficient the ocean and land sinks become high 2 train +211 AR6_WGI 96 26 Under SSP3-7.0 and SSP5-8.5, the initial growth of both sinks in response to increasing atmospheric concentrations of CO 2 is subsequently limited by emerging carbon–climate feedbacks high 2 train +212 AR6_WGI 96 30 Under SSP1-1.9, models project that combined land and ocean sinks will turn into a weak source by 2100 medium 1 train +213 AR6_WGI 98 18 The frequency and intensity of heavy precipitation events have increased over a majority of those land regions with good observational coverage high 2 train +214 AR6_WGI 98 19 Over the past half century, key aspects of the biosphere have changed in ways that are consistent with large- scale warming: climate zones have shifted poleward, and the growing season length in the Northern Hemisphere extratropics has increased high 2 train +215 AR6_WGI 98 20 The amplitude of the seasonal cycle of atmospheric CO 2 poleward of 45°N has increased since the 1960s (very high confidence), with increasing productivity of the land biosphere due to the increasing atmospheric CO 2 concentration as the main driver medium 1 train +216 AR6_WGI 98 21 Global-scale vegetation greenness has increased since the 1980s high 2 train +217 AR6_WGI 98 24 Warming of the land surface during the period 1971–2018 contributed about 5% of the increase in the global energy inventory (Section TS.3.1), nearly twice the estimate in AR5 high 2 train +218 AR6_WGI 98 26 The warming pattern will likely vary seasonally, with northern high latitudes warming more during winter than summer medium 1 train +219 AR6_WGI 98 33 Human-induced climate change has contributed to increases in agricultural and ecological droughts in some regions due to increases in evapotranspiration medium 1 train +220 AR6_WGI 100 3 Earlier onset of snowmelt has contributed to seasonally dependent changes in streamflow high 2 train +221 AR6_WGI 100 9 The projected increase in heavy precipitation extremes translates to an increase in the frequency and magnitude of pluvial floods high 2 train +222 AR6_WGI 100 11 Concurrent heatwaves and droughts have become more frequent over the last century, and this trend will continue with higher global warming high 2 train +223 AR6_WGI 100 12 The probability of compound flooding (storm surge, extreme rainfall and/or river flow) has increased in some locations and will continue to increase due to both sea level rise and increases in heavy precipitation, including changes in precipitation intensity associated with tropical cyclones high 2 train +224 AR6_WGI 100 14 At the same time an increase in the amplitude of the seasonal cycle of atmospheric CO 2 poleward of 45°N since the early 1960s (high confidence) and a global-scale increase in vegetation greenness of the terrestrial surface since the early 1980s high 2 train +225 AR6_WGI 100 17 Reactive nitrogen, ozone and aerosols affect terrestrial vegetation and carbon cycle through deposition and effects on large-scale radiation high 2 train +226 AR6_WGI 100 20 The SRCCL concluded that continued warming will exacerbate desertification processes (medium confidence) and that ecosystems will become increasingly exposed to climates beyond those that they are currently adapted to high 2 train +227 AR6_WGI 100 23 There is low confidence in the magnitude of these changes, but the probability of crossing uncertain regional thresholds (e.g., fires, forest dieback) increases with further warming high 2 train +228 AR6_WGI 100 24 The response of biogeochemical cycles to the anthropogenic perturbation can be abrupt at regional scales, and irreversible on decadal to century time scales high 2 train +229 AR6_WGI 101 1 Global land precipitation has likely increased since 1950, with a faster increase since the 1980s medium 1 train +230 AR6_WGI 101 3 Annual global land precipitation will increase over the 21st century as global surface temperature increases high 2 train +231 AR6_WGI 101 4 Human influence has been detected in amplified surface salinity and precipitation minus evaporation (P–E) patterns over the ocean high 2 train +232 AR6_WGI 101 5 The severity of very wet and very dry events increase in a warming climate high 2 train +233 AR6_WGI 101 6 Water cycle variability and related extremes are projected to increase faster than mean changes in most regions of the world and under all emissions scenarios high 2 test +234 AR6_WGI 101 7 Over the 21st century, the total land area subject to drought will increase and droughts will become more frequent and severe high 2 train +235 AR6_WGI 101 8 Near-term projected changes in precipitation are uncertain mainly because of internal variability, model uncertainty and uncertainty in forcings from natural and anthropogenic aerosols medium 1 train +236 AR6_WGI 101 9 Over the 21st century and beyond, abrupt human-caused changes to the water cycle cannot be excluded medium 1 train +237 AR6_WGI 101 11 Global land precipitation has likely increased since 1950, with a faster increase since the 1980s medium 1 test +238 AR6_WGI 101 13 The overall effect of anthropogenic aerosols is to reduce global precipitation through surface radiative cooling effects high 2 train +239 AR6_WGI 101 14 Over much of the 20th century, opposing effects of GHGs and aerosols on precipitation have been observed for some regional monsoons high 2 train +240 AR6_WGI 101 16 Inter-model differences and internal variability contribute to a substantial range in projections of large-scale and regional water cycle changes high 2 train +241 AR6_WGI 101 17 The occurrence of volcanic eruptions can alter the water cycle for several years high 2 train +242 AR6_WGI 101 20 Near-surface specific humidity has increased over the ocean (likely) and land (very likely) since at least the 1970s, with a detectable human influence medium 1 train +243 AR6_WGI 101 21 Human influence has been detected in amplified surface salinity and precipitation minus evaporation (P–E) patterns over the ocean high 2 train +244 AR6_WGI 101 24 In response to cryosphere changes (Section TS.2.5), there have been changes in streamflow seasonality, including an earlier occurrence of peak streamflow in high-latitude and mountain catchments high 2 train +245 AR6_WGI 101 25 Projected runoff (Box TS.6, Figure 1c) is typically decreased by contributions from small glaciers because of glacier mass loss, while runoff from larger glaciers will generally increase with increasing global warming levels until their mass becomes depleted high 2 train +246 AR6_WGI 102 1 Greater warming over land than over the ocean alters atmospheric circulation patterns and reduces continental near-surface relative humidity, which contributes to regional drying high 2 train +247 AR6_WGI 102 3 Projected increases in evapotranspiration due to growing atmospheric water demand will decrease soil moisture over the Mediterranean region, south-western North America, South Africa, South-Western South America and south-western Australia high 2 train +248 AR6_WGI 102 4 Some tropical regions are also projected to experience enhanced aridity, including the Amazon basin and Central America high 2 train +249 AR6_WGI 102 5 The total land area subject to increasing drought frequency and severity will expand (high confidence), and in the Mediterranean, South-Western South America, and Western North America, future aridification will far exceed the magnitude of change seen in the last millennium high 2 train +250 AR6_WGI 102 7 Large-scale deforestation likely decreases evapotranspiration and precipitation and increases runoff over the deforested regions relative to the regional effects of climate change medium 1 train +251 AR6_WGI 102 8 Urbanization increases local precipitation (medium confidence) and runoff intensity high 2 train +252 AR6_WGI 102 9 Increased precipitation intensities have enhanced groundwater recharge, most notably in tropical regions medium 1 train +253 AR6_WGI 102 12 A warmer climate increases moisture transport into weather systems, which intensifies wet seasons and events high 2 train +254 AR6_WGI 102 13 The magnitudes of projected precipitation increases and related extreme events depend on model resolution and the representation of convective processes high 2 train +255 AR6_WGI 102 14 Increases in near-surface atmospheric moisture capacity of about 7% per 1ºC of warming lead to a similar response in the intensification of heavy precipitation from sub-daily up to seasonal time scales, increasing the severity of flood hazards high 2 train +256 AR6_WGI 102 15 The average and maximum rain-rates associated with tropical and extratropical cyclones, atmospheric rivers and severe convective storms will therefore also increase with future warming high 2 train +257 AR6_WGI 102 17 In the tropics year-round and in the summer season elsewhere, interannual variability of precipitation and runoff over land is projected to increase at a faster rate than changes in seasonal mean precipitation (Figure TS.12e,f) medium 1 train +258 AR6_WGI 102 18 Sub-seasonal precipitation variability is also projected to increase, with fewer rainy days but increased daily mean precipitation intensity over many land regions high 2 train +259 AR6_WGI 107 5 Improved quantifications of ERF, the climate system radiative response, and the observed energy increase in the Earth system for the period 1971–2018 demonstrate improved closure of the global energy budget (i.e., the extent to which the sum of the integrated forcing and the integrated radiative response equals the energy gain of the Earth system) compared to AR5 high 2 train +260 AR6_WGI 108 6 Changes in sulphur dioxide (SO 2) emissions make the dominant contribution to the ERF from aerosol– cloud interactions high 2 train +261 AR6_WGI 108 7 Over the 1750–2019 period, the contributions from the emitted compounds to global surface temperature changes broadly match their contributions to the ERF high 2 train +262 AR6_WGI 109 1 The ERF due to aerosol– cloud interactions (ERFaci) contributes most to the magnitude of the total aerosol ERF (high confidence) and is assessed to be –1.0 [–1.7 to –0.3] W m–2 (medium confidence), with the remainder due to aerosol–radiation interactions (ERFari), assessed to be –0.3 [–0.6 to 0.0] W m–2 medium 1 train +263 AR6_WGI 109 13 Feedback processes are expected to become more positive overall (more amplifying of global surface temperature changes) on multi-decadal time scales as the spatial pattern of surface warming evolves and global surface temperature increases, leading to an ECS that is higher than was inferred in AR5 based on warming over the instrumental record high 2 train +264 AR6_WGI 109 15 Based on process understanding, climate modelling, and paleoclimate reconstructions of past warm periods, it is expected that future warming will become enhanced over the eastern Pacific Ocean (medium confidence) and Southern Ocean high 2 train +265 AR6_WGI 109 21 There is a high level of agreement among the different lines of evidence (Figure TS.16c) high 2 train +266 AR6_WGI 111 7 Because the total biogeophysical and non-CO 2 biogeochemical feedback is assessed to have a central value that is near zero low 0 train +267 AR6_WGI 111 15 The CMIP5 and CMIP6 ranges of cloud feedback are similar to this assessed range, with CMIP6 having a slightly more positive median cloud feedback high 2 train +268 AR6_WGI 113 7 This near-linear relationship further implies that mitigation requirements for limiting warming to specific levels can be quantified in terms of a carbon budget high 2 train +269 AR6_WGI 113 9 Several factors, including estimates of historical warming, future emissions from thawing permafrost, variations in projected non-CO 2 warming, and the global surface temperature change after cessation of CO 2 emissions, affect the exact value of carbon budgets high 2 train +270 AR6_WGI 115 14 In the same way that part of current anthropogenic net CO 2 emissions are taken up by land and ocean carbon stores, net CO 2 removal will be partially counteracted by CO 2 release from these stores very high 3 train +271 AR6_WGI 115 15 Asymmetry in the carbon cycle response to simultaneous CO 2 emissions and removals implies that a larger amount of CO 2 would need to be removed to compensate for an emission of a given magnitude to attain the same change in atmospheric CO 2 medium 1 train +272 AR6_WGI 115 16 CDR methods have wide-ranging side- effects that can either weaken or strengthen the carbon sequestration and cooling potential of these methods and affect the achievement of sustainable development goals high 2 train +273 AR6_WGI 115 25 In the same way part of current anthropogenic net CO 2 emissions are taken up by land and ocean carbon stores, net CO 2 removal will be partially counteracted by CO 2 release from these stores, such that the amount of CO 2 sequestered by CDR will not result in an equivalent drop in atmospheric CO 2 very high 3 train +274 AR6_WGI 115 26 The fraction of CO 2 removed from the atmosphere that is not replaced by CO 2 released from carbon stores – a measure of CDR effectiveness – decreases slightly with increasing amounts of removal (medium confidence) and decreases strongly if CDR is applied at lower atmospheric CO 2 concentrations medium 1 train +275 AR6_WGI 115 27 The reduction in global surface temperature is approximately linearly related to cumulative CO 2 removal high 2 train +276 AR6_WGI 115 28 Because of this near-linear relationship, the amount of cooling per unit CO 2 removed is approximately independent of the rate and amount of removal medium 1 train +277 AR6_WGI 116 7 For instance, sea level rise due to ocean thermal expansion would not reverse for several centuries to millennia high 2 train +278 AR6_WGI 116 9 Biophysical and biogeochemical side-effects of CDR methods are associated with changes in surface albedo, the water cycle, emissions of CH 4 and N 2O, ocean acidification and marine ecosystem productivity high 2 train +279 AR6_WGI 116 10 These side-effects and associated Earth system feedbacks can decrease carbon uptake and/or change local and regional climate and in turn limit the CO 2 sequestration and cooling potential of specific CDR methods medium 1 train +280 AR6_WGI 116 11 Deployment of CDR, particularly on land, can also affect water quality and quantity, food production and biodiversity high 2 train +281 AR6_WGI 116 12 These effects are often highly dependent on local context, management regime, prior land use, and scale high 2 train +282 AR6_WGI 116 13 The largest co-benefits are obtained with methods that seek to restore natural ecosystems or improve soil carbon sequestration medium 1 test +283 AR6_WGI 116 14 The climate and biogeochemical effects of terminating CDR are expected to be small for most CDR methods medium 1 train +284 AR6_WGI 117 2 Carbon-cycle responses are more robustly accounted for in emissions metrics compared to AR5 high 2 train +285 AR6_WGI 117 9 The methodology for doing this has been placed on a more robust scientific footing compared to AR5 high 2 train +286 AR6_WGI 117 10 Methane from fossil fuel sources has slightly higher emissions metric values than those from biogenic sources since it leads to additional fossil CO 2 in the atmosphere high 2 train +287 AR6_WGI 117 11 Updates to the chemical adjustments for CH 4 and N2O emissions (Section TS.3.1) and revisions in their lifetimes result in emissions metrics for GWP and GTP that are slightly lower than in AR5 medium 1 train +288 AR6_WGI 117 15 When GHGs are aggregated using standard metrics such as GWP or GTP, cumulative CO 2-e emissions are not necessarily proportional to future global surface temperature outcomes high 2 train +289 AR6_WGI 117 19 The warming evolution resulting from net zero GHG emissions defined in this way corresponds approximately to reaching net zero CO 2 emissions, and would thus not lead to declining temperatures after net zero GHG emissions are achieved but to an approximate temperature stabilization high 2 train +290 AR6_WGI 117 20 The choice of emissions metric hence affects the quantification of net zero GHG emissions, and therefore the resulting temperature outcome of reaching and sustaining net zero GHG emissions levels high 2 train +291 AR6_WGI 117 29 Fossil fuel combustion for energy, industry and land transportation are the largest contributing sectors on a 100-year time scale high 2 train +292 AR6_WGI 117 30 Current emissions of CO 2, N2O and SLCFs from East Asia and North America are the largest regional contributors to additional net future warming on both short (medium confidence) and long time scales (10 and 100 years, respectively) high 2 train +293 AR6_WGI 118 2 However, these reductions were lower than what would be expected from sustained implementation of policies addressing air quality and climate change medium 1 train +294 AR6_WGI 118 4 Consistent with this small net radiative forcing, and against a large component of internal variability, Earth system models show no detectable effect on global or regional surface temperature or precipitation high 2 train +295 AR6_WGI 119 2 This additional warming is stable after 2040 in SSPs associated with lower global air pollution as long as CH 4 emissions are also mitigated, but the overall warming induced by SLCF changes is higher in scenarios in which air quality continues to deteriorate (induced by growing fossil fuel use and limited air pollution control) high 2 train +296 AR6_WGI 119 3 Sustained CH 4 mitigation reduces global surface ozone, contributing to air quality improvements, and also reduces surface temperature in the longer term, but only sustained CO 2 emissions reductions allow long-term climate stabilization high 2 train +297 AR6_WGI 119 4 Future changes in air quality (near-surface ozone and particulate matter, or PM) at global and local scales are predominantly driven by changes in ozone and aerosol precursor emissions rather than climate high 2 train +298 AR6_WGI 119 5 Air quality improvements driven by rapid decarbonization strategies, as in SSP1-1.9 and SSP1-2.6, are not sufficient in the near term to achieve air quality guidelines set by the World Health Organization in some highly polluted regions high 2 train +299 AR6_WGI 119 20 Under the SSP3-7.0 scenario, PM levels are projected to increase until 2050 over large parts of Asia, and surface ozone pollution is projected to worsen over all continental areas through 2100 high 2 train +300 AR6_WGI 119 21 In SSP5-8.5, a scenario without climate change mitigation but with stringent air pollution control, PM levels decline through 2100, but high CH 4 levels hamper the decline in global surface ozone at least until 2080 high 2 train +301 AR6_WGI 120 11 These investigations have consistently shown that SRM could offset some of the effects of increasing greenhouse gases on global and regional climate, including the carbon and water cycles high 2 train +302 AR6_WGI 120 12 However, there would be substantial residual or overcompensating climate change at the regional scales and seasonal time scales high 2 train +303 AR6_WGI 120 13 The cooling caused by SRM would increase the global land and ocean CO 2 sinks (medium confidence), but this would not stop CO 2 from increasing in the atmosphere or affect the resulting ocean acidification under continued anthropogenic emissions high 2 train +304 AR6_WGI 120 15 A sudden and sustained termination of SRM in a high CO 2 emissions scenario would cause rapid climate change high 2 train +305 AR6_WGI 120 16 However, a gradual phase-out of SRM combined with emissions reduction and carbon dioxide removal (CDR) would avoid these termination effects medium 1 train +306 AR6_WGI 121 10 Since AR5, more modelling work has been conducted with more sophisticated treatment of aerosol-based SRM approaches, but the uncertainties in cloud–aerosol–radiation interactions are still large high 2 train +307 AR6_WGI 121 11 Modelling studies suggest that it is possible to stabilize multiple large-scale temperature indicators simultaneously by tailoring the deployment strategy of SRM options medium 1 train +308 AR6_WGI 121 13 In contrast, cirrus cloud thinning, targeting longwave radiation, is expected to cause an increase in global mean precipitation medium 1 train +309 AR6_WGI 121 14 If shortwave approaches are used to offset global mean warming, the magnitude of reduction in regional precipitation minus evapotranspiration (P–E) (Box TS.5), which is more relevant to freshwater availability, is smaller than precipitation decrease because of simultaneous reductions in both precipitation and evapotranspiration medium 1 train +310 AR6_WGI 121 16 If SRM is used to cool the planet, it would cause a reduction in plant and soil respiration and slow the reduction of ocean carbon uptake due to warming medium 1 train +311 AR6_WGI 121 17 The result would be an enhancement of the global land and ocean CO 2 sinks medium 1 train +312 AR6_WGI 121 18 However, SRM would not stop CO 2 from increasing in the atmosphere or affect the resulting ocean acidification under continued anthropogenic emissions high 2 train +313 AR6_WGI 121 20 A sudden and sustained termination of SRM in a high GHG emissions scenario would cause rapid climate change and a reversal of the SRM effects on the carbon sinks high 2 train +314 AR6_WGI 121 23 However, a gradual phase-out of SRM combined with emissions reductions and CDR would avoid larger rates of changes medium 1 train +315 AR6_WGI 122 4 For global climate indicators, evidence for abrupt change is limited, but deep ocean warming, acidification and sea level rise are committed to ongoing change for millennia after global surface temperatures initially stabilize and are irreversible on human time scales very high 3 train +316 AR6_WGI 122 5 At the regional scale, abrupt responses, tipping points and even reversals in the direction of change cannot be excluded high 2 train +317 AR6_WGI 122 10 Some processes suspected of having tipping points, such as the Atlantic Meridional Overturning Circulation (AMOC), have been found to often undergo recovery after temperature stabilization with a time delay low 0 train +318 AR6_WGI 122 13 It is likely that under stabilization of global warming at 1.5°C, 2.0°C or 3.0°C relative to 1850–1900, the AMOC will continue to weaken for several decades by about 15%, 20% and 30% of its strength and then recover to pre-decline values over several centuries medium 1 train +319 AR6_WGI 122 14 At sustained warming levels between 2°C and 3°C, there is limited evidence that the Greenland and West Antarctic ice sheets will be lost almost completely and irreversibly over multiple millennia; both the probability of their complete loss and the rate of mass loss increases with higher surface temperatures high 2 train +320 AR6_WGI 122 15 At sustained warming levels between 3°C and 5°C, near-complete loss of the Greenland Ice Sheet and complete loss of the West Antarctic Ice Sheet is projected to occur irreversibly over multiple millennia (medium confidence); with substantial parts or all of Wilkes Subglacial Basin in East Antarctica lost over multiple millennia low 0 train +321 AR6_WGI 122 17 For other hazards (e.g., ice-sheet behaviour, glacier mass loss and global mean sea level change, coastal floods, coastal erosion, air pollution, and ocean acidification) the time and/or scenario dimensions remain critical, and a simple and robust relationship with global warming level cannot be established high 2 train +322 AR6_WGI 122 19 For global warming up to 2°C above 1850–1900 levels, paleoclimate records do not indicate abrupt changes in the carbon cycle low 0 test +323 AR6_WGI 122 20 Despite the wide range of model responses, uncertainty in atmospheric CO 2 by 2100 is dominated by future anthropogenic emissions rather than uncertainties related to carbon–climate feedbacks high 2 train +324 AR6_WGI 122 22 The increase in global ocean heat content (Section TS.2.4) will likely continue until at least 2300 even for low emissions scenarios, and global mean sea level will continue to rise for centuries to millennia following cessation of emissions (Box TS.4) due to continuing deep ocean heat uptake and mass loss of the Greenland and Antarctic ice sheets high 2 train +325 AR6_WGI 123 2 Possible abrupt changes and tipping points in biogeochemical cycles lead to additional uncertainty in 21st century atmospheric GHG concentrations, but future anthropogenic emissions remain the dominant uncertainty high 2 train +326 AR6_WGI 123 5 Continued Amazon deforestation, combined with a warming climate, raises the probability that this ecosystem will cross a tipping point into a dry state during the 21st century low 0 train +327 AR6_WGI 123 11 A key methodology is distillation – combining lines of evidence and accounting for stakeholder context and values – which helps ensure the information is relevant, useful and trusted for decision-making (see Core Concepts Box) high 2 train +328 AR6_WGI 123 14 Since AR5, climate change information produced for climate services has increased significantly due to scientific and technological advancements and growing user awareness, requirements, and demand very high 3 train +329 AR6_WGI 123 20 Discarding models that fundamentally misrepresent relevant processes improves the credibility of regional climate information generated from these ensembles high 2 train +330 AR6_WGI 123 21 However, multi-model mean and ensemble spread are not a full measure of the range of projection uncertainty and are not sufficient to characterize low-likelihood, high-impact changes (Box TS.3) or situations where different models simulate substantially different or even opposite changes high 2 train +331 AR6_WGI 123 22 Large single-model ensembles are now available and provide a more comprehensive spectrum of possible changes associated with internal variability high 2 train +332 AR6_WGI 124 14 No-till farming, irrigation and crop expansion have similarly attenuated increases in summer hot extremes in some regions, such as central North America medium 1 train +333 AR6_WGI 124 22 In some cases, even the sign of a projected change in regional climate cannot be trusted if relevant regional processes are not represented, for example, for variables such as precipitation and wind speed medium 1 train +334 AR6_WGI 124 25 However, the performance of these techniques depends on that of the driving climate model: in particular, bias adjustment cannot overcome all consequences of unresolved or strongly misrepresented physical processes, such as large-scale circulation biases or local feedbacks medium 1 train +335 AR6_WGI 126 18 Distilling regional climate information from multiple lines of evidence and taking the user context into account increases fitness, usefulness, relevance and trust in that information for use in climate services (Box TS.11) and decision-making high 2 train +336 AR6_WGI 127 1 Since AR5, there has been a significant increase in the range and diversity of climate service activities very high 3 train +337 AR6_WGI 127 6 Since AR5, climate change information produced in climate service contexts has increased significantly due to scientific and technological advancements and growing user awareness, requirements and demand very high 3 train +338 AR6_WGI 127 9 They require different types of user–producer engagement depending on what the service aims to deliver high 2 train +339 AR6_WGI 127 10 Realization of the full potential of climate services is often hindered by limited resources for the co-design and co-production process, including sustained engagement between scientists, service providers and users high 2 train +340 AR6_WGI 129 2 An exception is the Southern Annular Mode (SAM), which has become systematically more positive (high confidence) and is projected to be more positive in all seasons, except for December–January–February (DJF), in high CO 2 emissions scenarios high 2 train +341 AR6_WGI 129 3 The influence of stratospheric ozone forcing on the SAM trend has been reduced since the early 2000s compared to earlier decades, contributing to the weakening of its positive trend as observed over 2000– 2019 medium 1 train +342 AR6_WGI 129 10 Along latitudes, it is more uniform, with strong amplification of the temperature response towards the Arctic medium 1 train +343 AR6_WGI 129 11 The decrease of SO 2 emissions since the 1980s reduces the damping effect of aerosols, leading to a faster increase in surface air temperature that is most pronounced at mid- and high latitudes of the Northern Hemisphere, where the largest emissions reductions have taken place medium 1 train +344 AR6_WGI 129 13 Multi-decadal variation in anthropogenic aerosol emissions are thought to be a major contributor medium 1 train +345 AR6_WGI 129 16 Future urbanization will amplify the projected air temperature under different background climates, with a strong effect on minimum temperatures that could be as large as the global warming signal very high 3 train +346 AR6_WGI 129 17 Irrigation and crop expansion have attenuated increases in summer hot extremes in some regions, such as central North America medium 1 train +347 AR6_WGI 130 2 The influence of stratospheric ozone forcing on the SAM trend has been reduced since the early 2000s compared to earlier decades, contributing to the weakening of its positive trend observed over 2000–2019 medium 1 train +348 AR6_WGI 130 6 This is because of the opposing influence in the near to mid-term from stratospheric ozone recovery and increases in other greenhouse gases on the Southern Hemisphere summertime mid-latitude circulation high 2 train +349 AR6_WGI 130 12 There is no consensus from models for a systematic change in amplitude of ENSO sea surface temperature (SST) variability over the 21st century in any of the SSP scenarios assessed medium 1 train +350 AR6_WGI 130 17 Internal variability is the main driver of Pacific Decadal Variability (PDV) observed since the start of the instrumental records high 2 train +351 AR6_WGI 133 2 Multiple lines of evidence, combining multi-model ensemble global projections with those coming from single-model initial-condition large ensembles, show that internal variability is largely contributing to the delayed or absent emergence of the anthropogenic signal in long-term regional mean precipitation changes high 2 train +352 AR6_WGI 133 3 Internal variability in ocean dynamics dominates regional patterns on annual to decadal time scales high 2 train +353 AR6_WGI 133 4 The anthropogenic signal in regional sea level change will emerge in most regions by 2100 medium 1 train +354 AR6_WGI 133 6 Time evolution of mechanisms operating at different time scales can modify the amplitude of the regional-scale response of temperature, and both the amplitude and sign of the response of precipitation, to anthropogenic forcing high 2 train +355 AR6_WGI 133 8 Land-use and aerosol forcings and land–atmosphere feedback play important roles in modulating regional changes, for instance in weather and climate extremes high 2 train +356 AR6_WGI 134 8 Multi-model mean and ensemble spread are not sufficient to characterize situations where different models simulate substantially different or even opposite changes high 2 train +357 AR6_WGI 134 10 In addition, single-model initial-condition large ensembles of many realizations of internal variability are required to separate internal variability from forced changes high 2 train +358 AR6_WGI 134 12 Northern Hemispheric anthropogenic aerosols weakened the regional monsoon circulations in South Asia, East Asia and West Africa during the second half of the 20th century, thereby offsetting the expected strengthening of monsoon precipitation in response to GHG-induced warming high 2 train +359 AR6_WGI 134 13 During the 21st century, global land monsoon precipitation is projected to increase in response to GHG warming in all time horizons and scenarios high 2 train +360 AR6_WGI 134 14 Over South and South East Asia, East Asia and the central Sahel, monsoon precipitation is projected to increase, whereas over North America and the far western Sahel it is projected to decrease medium 1 train +361 AR6_WGI 134 16 At global and regional scales, near-term monsoon changes will be dominated by the effects of internal variability medium 1 test +362 AR6_WGI 134 19 Contrary to the expected increase of precipitation under global warming, the Northern Hemisphere monsoon regions experienced declining precipitation from the 1950s to 1980s, which is partly attributable to the influence of anthropogenic aerosols medium 1 train +363 AR6_WGI 134 21 A slowdown of the tropical circulation with global warming can partly offset the warming-induced strengthening of precipitation in monsoon regions high 2 train +364 AR6_WGI 134 22 In the near term, global monsoon changes are likely to be dominated by the effects of internal variability and model uncertainties medium 1 train +365 AR6_WGI 134 23 In the long term, global monsoon rainfall change will feature a robust north–south asymmetry characterized by a greater increase in the Northern Hemisphere than in the Southern Hemisphere and an east–west asymmetry characterized by enhanced Asian– African monsoons and a weakened North American monsoon medium 1 train +366 AR6_WGI 136 2 The recent partial recovery and enhanced intensity of monsoon precipitation over West Africa is related to the growing influence of GHGs with an additional contribution due to the reduced cooling effect of anthropogenic aerosols, emitted largely from North America and Europe medium 1 train +367 AR6_WGI 136 5 The annual contrast between the wettest and driest month of the year is likely to increase by 3–5% per degree Celsius in most monsoon regions in terms of precipitation, precipitation minus evaporation, and runoff medium 1 train +368 AR6_WGI 136 6 For the North American monsoon, projections indicate a decrease in precipitation, whereas increased monsoon rainfall is projected over South and South East Asia and over East Asia medium 1 train +369 AR6_WGI 136 7 West African monsoon precipitation is projected to increase over the central Sahel and decrease over the far western Sahel medium 1 train +370 AR6_WGI 136 12 A collapse of the Atlantic Meridional Overturning Circulation could weaken the African and Asian monsoons but strengthen the Southern Hemisphere monsoons high 2 train +371 AR6_WGI 136 16 It is very likely that mean temperatures have increased in all land regions and will continue to increase at rates greater than the global average high 2 test +372 AR6_WGI 136 18 These changes are attributed to human influence in almost all regions (medium to high confidence) and will continue through the 21st century high 2 train +373 AR6_WGI 136 19 In particular, extreme heat would exceed critical thresholds for health, agriculture and other sectors more frequently by the mid 21st century with 2°C of global warming high 2 train +374 AR6_WGI 136 20 Relative sea level rise is very likely to virtually certain (depending on the region) to continue during the 21st century, contributing to increased coastal flooding in low- lying areas (high confidence) and coastal erosion along most sandy coasts high 2 train +375 AR6_WGI 136 21 Sea level will continue to rise beyond 2100 high 2 train +376 AR6_WGI 136 22 Every region of the world will experience concurrent changes in multiple CIDs by mid-century or at 2°C global warming and above high 2 train +377 AR6_WGI 137 2 Changes in many other regional CIDs have higher confidence later in the 21st century or at higher GWLs high 2 train +378 AR6_WGI 148 18 In tropical regions, recent past temperature distributions have already shifted to a range different to that of the early 20th century high 2 train +379 AR6_WGI 148 20 On regional-to-continental scales, trends of increased frequency of hot extremes and decreased frequency of cold extremes are generally consistent with the global-scale trends in mean temperature high 2 train +380 AR6_WGI 148 26 At increasing warming levels, extreme heat will exceed critical thresholds for health, agriculture and other sectors more frequently high 2 train +381 AR6_WGI 148 27 For example, by the end of the 21st century, dangerous humid heat thresholds, such as the National Oceanic and Atmospheric Administration (NOAA) heat index (HI) threshold of 41°C, will be exceeded much more frequently under the SSP5-8.5 scenario than under SSP1-2.6 and will affect many regions high 2 train +382 AR6_WGI 148 28 In many tropical regions, the number of days per year where a heat index of 41°C is exceeded would increase by more than 100 days relative to the recent past under SSP5-8.5, while this increase will be limited to less than 50 days under SSP1-2.6 high 2 train +383 AR6_WGI 148 29 The number of days per year where temperature exceeds 35°C would increase by more than 150 days in many tropical areas, such as the Amazon basin and South East Asia, by the end of century for the SSP5-8.5 scenario, while it is expected to increase by less than 60 days in these areas under SSP1-2.6 (except for the Amazon Basin) high 2 train +384 AR6_WGI 148 31 However, GHG forcing has driven increased contrasts in precipitation amounts between wet and dry seasons and weather regimes over tropical land areas (medium confidence), with a detectable precipitation increase in the northern high latitudes high 2 train +385 AR6_WGI 148 32 The frequency and intensity of heavy precipitation events have increased over a majority of land regions with good observational coverage high 2 train +386 AR6_WGI 148 33 A majority of land areas have experienced decreases in available water in dry seasons due to human-induced climate change associated with changes in evapotranspiration medium 1 train +387 AR6_WGI 148 34 Global hydrological models project a larger fraction of land areas to be affected by an increase rather than by a decrease in river floods medium 1 train +388 AR6_WGI 148 35 Extreme precipitation and pluvial flooding will increase in many regions around the world on almost all continents high 2 train +389 AR6_WGI 148 38 The proportion of intense TCs, average peak TC wind speeds, and peak wind speeds of the most intense TCs will increase on the global scale with increasing global warming high 2 train +390 AR6_WGI 150 17 Glaciers will continue to shrink and permafrost to thaw in all regions where they are present high 2 train +391 AR6_WGI 150 22 Regional sea level change has been the main driver of changes in extreme sea levels across the quasi-global tide gauge network over the 20th century high 2 train +392 AR6_WGI 150 23 With the exception of a few regions with substantial land uplift, relative sea level rise is very likely to virtually certain (depending on the region) to continue during the 21st century, contributing to increased coastal flooding in low-lying areas (high confidence) and coastal erosion along most sandy coasts high 2 train +393 AR6_WGI 150 24 In the open ocean, acidification, changes in sea ice, and deoxygenation have already emerged in many areas high 2 train +394 AR6_WGI 150 25 Marine heatwaves are also expected to increase around the globe over the 21st century high 2 train +395 AR6_WGI 151 3 The probability of compound events has increased in the past due to human-induced climate change and will likely continue to increase with further global warming, including for concurrent heatwaves and droughts, compound flooding, and the possibility of connected sectors experiencing multiple regional extreme events at the same time (for example, in multiple breadbaskets) high 2 train +396 AR6_WGI 151 6 In a number of regions (Southern Africa, the Mediterranean, North Central America, Western North America, the Amazon regions, South-Western South America, and Australia), increases in one or more of drought, aridity and fire weather high 2 train +397 AR6_WGI 151 7 In another group of regions (North-Western, Central and Eastern North America, Arctic regions, North- Western South America, Northern, Western and Central and Eastern Europe, Siberia, Central, South and East Asia, Southern Australia and New Zealand), decreases in snow and/or ice or increases in pluvial/river flooding high 2 train +398 AR6_WGI 151 9 Increases in heavy precipitation that can lead to pluvial floods high 2 train +399 AR6_WGI 151 13 North Eastern Africa, East Southern Africa and Central Africa have experienced a decline in rainfall since about 1980 and parts of West Africa an increase high 2 train +400 AR6_WGI 151 14 Increases in the frequency and/or the intensity of heavy rainfall have been observed in East and West Southern Africa, and the eastern Mediterranean region medium 1 train +401 AR6_WGI 151 15 Increasing trends in river flood occurrence can be identified beyond 1980 in East and West Southern Africa (medium confidence) and Western Africa high 2 train +402 AR6_WGI 151 17 Over West Africa, rainfall is projected to decrease in the western Sahel subregion and increase along the Guinea Coast subregion medium 1 train +403 AR6_WGI 151 18 Rainfall is projected to increase over Eastern Africa medium 1 train +404 AR6_WGI 151 20 Trends towards increased hydrological droughts have been observed in the Mediterranean (high confidence) and Western Africa medium 1 train +405 AR6_WGI 151 21 These trends correspond with projected regional increases in aridity and fire weather conditions high 2 train +406 AR6_WGI 151 23 Over Western Africa and Southern Africa, a future significant increase in wind speed and wind energy potential is projected medium 1 train +407 AR6_WGI 151 24 There is a projected decrease in the frequency of tropical cyclones making landfall over Madagascar, East Southern Africa and East Africa medium 1 train +408 AR6_WGI 154 1 Additional regional changes in Asia, besides those features described in Section TS.4.3.1, include historical trends of annual precipitation that show considerable regional differences high 2 train +409 AR6_WGI 154 2 East Asian Monsoon precipitation has changed, with drying in the north and wetting in the south since the 1950s, and annual mean precipitation totals very likely have increased over most territories of North Asia since the mid-1970s high 2 train +410 AR6_WGI 154 3 South Asian summer monsoon precipitation decreased over several areas since the mid-20th century high 2 train +411 AR6_WGI 154 5 Aridity in East and West Central Asia is projected to increase, especially beyond the middle of the 21st century and global warming levels beyond 2°C medium 1 train +412 AR6_WGI 154 6 Fire weather seasons are projected to lengthen and intensify everywhere except South East Asia, Tibetan Plateau and Arabian Peninsula medium 1 train +413 AR6_WGI 154 7 Surface wind speeds have been decreasing in Asia high 2 train +414 AR6_WGI 154 8 Over North Asia, increases in permafrost temperature and its thawing have been observed over recent decades high 2 train +415 AR6_WGI 154 9 Future projections indicate continuing decline in seasonal snow duration, glacial mass, and permafrost area by mid-century high 2 test +416 AR6_WGI 154 10 Snow-covered areas and snow volumes will decrease in most regions of the Hindu Kush Himalaya (HKH) during the 21st century, and snowline elevations will rise high 2 train +417 AR6_WGI 154 11 Heavy snowfall is increasing in East Asia and North Asia medium 1 train +418 AR6_WGI 154 16 Daily precipitation extremes have increased over part of the region high 2 train +419 AR6_WGI 154 17 Extreme hydrological drought frequency has increased in a region extending from south-west to north-east China, with projected increases of agricultural and ecological drought for 4°C GWL and fire weather for 2°C and above medium 1 test +420 AR6_WGI 154 19 Concurrently, total soil moisture is projected to decline extensively medium 1 train +421 AR6_WGI 154 21 The frequency of heavy precipitation and flood events has increased over several areas during the last few decades medium 1 train +422 AR6_WGI 154 23 Most of the region has experienced an increase in rainfall intensity but with a reduced number of wet days medium 1 train +423 AR6_WGI 154 24 Rainfall is projected to increase in the northern parts of South East Asia and decrease in areas in the Maritime Continent medium 1 train +424 AR6_WGI 155 1 Since the mid 1980’s, there has been an increase in the number and intensification rate of intense TCs (medium confidence), with a significant north-westward shift in tracks and a northward shift in their average latitude, increasing exposure over East China, the Korean Peninsula and the Japanese Archipelago medium 1 train +425 AR6_WGI 155 5 Heavy snowfall is projected to occur more frequently in some parts of Japan medium 1 train +426 AR6_WGI 155 9 More than 60% of glacier mass in the Caucasus is projected to disappear under RCP8.5 emissions by the end of the 21st century medium 1 train +427 AR6_WGI 155 13 This will contribute to more frequent coastal flooding and higher ETWL in low-lying areas and coastal erosion along sandy beaches high 2 train +428 AR6_WGI 155 16 Agricultural and ecological droughts and hydrological droughts have increased over Southern Australia (medium confidence), and meteorological droughts have decreased over Northern and Central Australia medium 1 train +429 AR6_WGI 155 17 Relative sea level has increased over the period 1993–2018 at a rate higher than GMSL around Australasia high 2 train +430 AR6_WGI 155 20 Agricultural and ecological droughts are projected to increase in Southern and Eastern Australia medium 1 train +431 AR6_WGI 155 21 Fire weather is projected to increase throughout Australia (high confidence) and New Zealand medium 1 train +432 AR6_WGI 155 22 Snowfall is expected to decrease throughout the region at high altitudes in both Australia (high confidence) and New Zealand (medium confidence), with glaciers receding in New Zealand high 2 train +433 AR6_WGI 155 26 Annual mean precipitation is projected to increase in the south and west of New Zealand (medium confidence) and is projected to decrease in south-west Southern Australia (high confidence), Eastern Australia (medium confidence), and in the north and east of New Zealand medium 1 train +434 AR6_WGI 155 28 Aridity is projected to increase with medium confidence in Southern Australia (high confidence in south-west Southern Australia), Eastern Australia (medium confidence) and in the north and east of New Zealand medium 1 train +435 AR6_WGI 156 1 TCs in north-eastern and north Australia are projected to decrease in number (high confidence) but increase in intensity except for ‘east coast lows’ low 0 train +436 AR6_WGI 156 6 The present day 1-in-100-year ETWL is between 0.5–2.5 m around most of Australia, except the north-western coast where 1-in-100-year ETWL can be as high as 6–7 m. {Box 9.1, 12.3.1.5, 12.4.3.5} TS.4.3.2.4 Central and South America Additional regional changes in Central and South America, besides those features described in Section TS.4.3.1, include increases in mean and extreme precipitation in South- Eastern South America since the 1960s high 2 train +437 AR6_WGI 156 7 Decreasing trends in mean precipitation and increasing trends in agricultural and ecological drought are observed over North-Eastern South America medium 1 train +438 AR6_WGI 156 8 The intensity and frequency of extreme precipitation and pluvial floods is projected to increase over South-Eastern South America, Southern South America, Northern South America, South American Monsoon and North-Eastern South America medium 1 train +439 AR6_WGI 156 9 Increases of agricultural and ecological drought are projected in South America Monsoon and Southern South America, and fire weather is projected to increase over several regions (Northern South America, the South American Monsoon, North-Eastern South America and South-Western South America) high 2 train +440 AR6_WGI 156 21 Increased mean precipitation amounts at high latitudes in boreal winter and reduced summer precipitation in southern Europe are projected starting from a 2°C GWL high 2 train +441 AR6_WGI 156 22 Aridity, agricultural and hydrological droughts and fire weather conditions will increase in the Mediterranean region starting from 2°C GWL high 2 train +442 AR6_WGI 156 23 Pluvial flooding will increase everywhere with high confidence except for medium confidence in the Mediterranean; in Western and Central Europe this also applies to river flooding starting from a 2°C GWL high 2 train +443 AR6_WGI 156 24 Most periglacial processes in Northern Europe are projected to disappear by the end of the 21st century, even for a low warming scenario medium 1 train +444 AR6_WGI 157 3 In the European Mediterranean, the magnitude and sign of observed land precipitation trends depend on time period and exact study region medium 1 train +445 AR6_WGI 157 7 The frequency of Medicanes (tropical-like cyclones in the Mediterranean) is projected to decrease medium 1 train +446 AR6_WGI 157 16 Severe wind storms, tropical cyclones and dust storms in North America are shifting toward more extreme characteristics (medium confidence), and both observations and projections point to strong changes in the seasonal and geographic range of snow and ice conditions in the coming decades very high 3 train +447 AR6_WGI 157 24 Mean wind speed and wind power potential are projected to decrease in Western North America high 2 train +448 AR6_WGI 157 26 At sustained GWLs between 3°C and 5°C, nearly all glacial mass in Western Canada and Western North America will disappear medium 1 train +449 AR6_WGI 158 4 Fewer but more intense tropical cyclones are projected starting from a 2°C GWL medium 1 train +450 AR6_WGI 158 8 Higher evapotranspiration under a warming climate can partially offset future increases or amplify future reductions in rainfall, resulting in increased aridity as well as more severe agricultural and ecological drought in the Caribbean medium 1 train +451 AR6_WGI 158 11 Shoreline retreat is projected along sandy coasts of most small islands high 2 train +452 AR6_WGI 158 13 An intensification of the polar water cycle will increase mean precipitation, with precipitation intensity becoming stronger and more likely to be rainfall rather than snowfall high 2 train +453 AR6_WGI 158 14 Permafrost warming, loss of seasonal snow cover, and glacier melt will be widespread high 2 train +454 AR6_WGI 158 16 Relative sea level and coastal flooding are projected to increase in areas other than regions with substantial land uplift medium 1 train +455 AR6_WGI 158 23 In the Arctic, this will result in higher river flood potential and earlier meltwater flooding, altering seasonal characteristics of flooding high 2 train +456 AR6_WGI 158 24 A lengthening of the fire season (medium confidence) and encroachment of fire regimes into tundra regions high 2 train +457 AR6_WGI 158 27 Permafrost warming and thawing have been widespread in the Arctic since the 1980s high 2 train +458 AR6_WGI 159 2 The Pacific and Southern Ocean are projected to freshen and the Atlantic to become more saline medium 1 train +459 AR6_WGI 159 3 Anthropogenic warming is very likely to further decrease ocean oxygen concentrations, and this deoxygenation is expected to persist for thousands of years medium 1 train +460 AR6_WGI 159 4 Arctic sea ice losses are projected to continue, leading to a practically ice-free Arctic in September by the end of the 21st century under high CO 2 emissions scenarios high 2 train +461 AR6_WGI 159 7 Global warming of 2°C above 1850–1900 levels would result in the exceedance of numerous hazard thresholds for pathogens, seagrasses, mangroves, kelp forests, rocky shores, coral reefs and other marine ecosystems medium 1 train +462 AR6_WGI 159 16 These include biodiversity hot spots that will very likely see even more extreme heat and droughts, mountain areas where a projected raising in the freezing level height will alter snow and ice conditions (high confidence), and tropical forests that are increasingly prone to fire weather medium 1 train +463 AR6_WGI 159 21 Extreme precipitation is projected to increase in major mountainous regions (medium to high confidence depending on location), with potential cascading consequences of floods, landslides and lake outbursts in all scenarios medium 1 train +464 AR6_WGI 159 23 Water cycle changes bring prolonged drought, longer dry seasons and increased fire weather to many tropical forests medium 1 train +465 AR6_WGI 160 2 With global warming, increasing relative sea level compounded by increasing tropical cyclone storm surge and rainfall intensity will increase the probability of coastal city flooding high 2 train +466 AR6_WGI 160 3 Arctic coastal settlements are particularly exposed to climate change due to sea ice retreat high 2 train +467 AR6_WGI 160 4 Improvements in urban climate modelling and climate monitoring networks have contributed to understanding the mutual interaction between regional and urban climate high 2 test +468 AR6_WGI 160 7 Compared to present day, large implications are expected from the combination of future urban development and more frequent occurrence of extreme climate events, such as heatwaves, with more hot days and warm nights adding to heat stress in cities very high 3 train +469 AR6_WGI 166 11 However, the NDCs submitted as of 2020 are insufficient to reduce greenhouse gas emissions enough to be consistent with trajectories limiting global warming to well below 2°C above pre-industrial levels high 2 train +470 AR6_WGI 166 24 There was likely a net anthropogenic forcing of 0.0 –0.3 W m–2 in 1850 –1900 relative to 1750 medium 1 train +471 AR6_WGI 166 27 At the current level of global warming, an observed signal of temperature change relative to the 1850–1900 baseline has emerged above the levels of background variability over virtually all land regions high 2 train +472 AR6_WGI 167 1 Accordingly, the signal of change is more apparent in tropical regions than in regions with greater warming but larger interannual variations high 2 train +473 AR6_WGI 167 6 Scientific knowledge interacts with pre-existing conceptions of weather and climate, including values and beliefs stemming from ethnic or national identity, traditions, religion or lived relationships to land and sea high 2 train +474 AR6_WGI 167 8 Social values may guide certain choices made during the construction, assessment and communication of information high 2 train +475 AR6_WGI 167 13 In addition, paleoclimate archives such as mid-latitude and tropical glaciers, as well as modern natural archives used for calibration (e.g., corals and trees), are rapidly disappearing due to a host of pressures, including increasing temperatures high 2 train +476 AR6_WGI 167 27 The broader availability of ensemble model simulations has contributed to better estimations of uncertainty in projections of future change high 2 train +477 AR6_WGI 184 21 These diverse, more local understandings can both contrast with and enrich the planetary-scale analyses of global climate science high 2 train +478 AR6_WGI 184 30 In summary, environmental and socio- altruistic values are the most significant influences on public opinion about climate change globally, while political views, political party affiliation, and corporate influence also had strong effects, especially in the USA high 2 train +479 AR6_WGI 186 11 Statement of factCertain/factLikelihood outcome66% 90% Likely range Very likely rangeProbability LikelyVery likelyExtremely likelyCumulative probability 66%90%95% Examples of assessments Assessed evidence and agreement Past projections of global temperature and the pattern of warming are broadly consistent with subsequent observations (limited evidence, high agreement) {1.3.6}.Assessed confidence The probability of low-likelihood, high impact outcomes increases with higher global warming levels high 2 train +480 AR6_WGI 186 15 The AR6 assessed best estimate is 3°C with a likely range of 2.5°C to 4°C high 2 train +481 AR6_WGI 187 10 In summary, the calibrated language cannot entirely prevent misunderstandings, including a tendency to systematically underestimate the probability of the IPCC’s higher-likelihood conclusions and overestimate the probability of the lower-likelihood ones high 2 train +482 AR6_WGI 188 26 These approaches are more effective when combined with other policies and tailored to the motivations, capabilities and resources of specific actors and contexts (high confidence).’ These extended dialogic co-production and education processes have thus been demonstrated to improve the quality of both scientific information and governance high 2 train +483 AR6_WGI 189 17 We thus assess that specific characteristics of media coverage play a major role in climate understanding and perception high 2 train +484 AR6_WGI 192 4 The AR5 WGI assessed that the pH of ocean surface water has decreased by 0.1 since the beginning of the industrial era high 2 train +485 AR6_WGI 193 6 This thermal expansion, along with glacier mass loss, were the dominant contributors to GMSL rise during the 20th century high 2 train +486 AR6_WGI 193 21 In summary, these data allowed AR5 WGI to assess that over the last two decades, the Greenland and Antarctic ice sheets have been losing mass, glaciers have continued to shrink almost worldwide, and Arctic sea ice and Northern Hemisphere spring snow cover have continued to decrease in extent high 2 train +487 AR6_WGI 194 19 The AR5 WGI (IPCC, 2013b) used paleoclimatic evidence to put recent warming and sea level rise in a multi-century perspective and assessed that 1983–2012 was likely to have been the warmest 30-year period of the last 1400 years in the Northern Hemisphere medium 1 train +488 AR6_WGI 194 20 The AR5 also assessed that the rate of sea level rise since the mid-19th century has been larger than the mean rate during the previous two millennia high 2 train +489 AR6_WGI 197 2 Overall, AR5 assessed that total aerosol effects, including cloud adjustments, resulted in a negative RF of –0.9 [–1.9 to −0.1] W m−2 (medium confidence), offsetting a substantial portion of the positive RF resulting from the increase in GHGs high 2 train +490 AR6_WGI 198 17 Similarly, over the period 1993–2010, when observations of all sea level components were available, AR5 WGI assessed the observed global mean sea level rise to be consistent with the sum of the observed contributions from ocean thermal expansion (due to warming) combined with changes in glaciers, the Antarctic and Greenland ice sheets, and land-water storage high 2 train +491 AR6_WGI 198 27 In response, AR5 WGI made a specific assessment for how global surface temperature was projected to evolve over the next two decades, concluding that the change for the period 2016–2035 relative to 1986–2005 will likely be in the range of 0.3°C–0.7°C medium 1 train +492 AR6_WGI 203 7 For the period 2006–2015, observed global mean surface temperature (GMST7) was 0.87°C ± 0.12°C higher than the average over the 1850–1900 period very high 3 train +493 AR6_WGI 203 8 Anthropogenic global warming was estimated to be increasing at 0.2 ± 0.1°C per decade high 2 train +494 AR6_WGI 203 10 This observed warming has already led to increases in the frequency and intensity of climate and weather extremes in many regions and seasons, including heat waves in most land regions (high confidence), increased droughts in some regions (medium confidence), and increases in the intensity of heavy precipitation events at the global scale medium 1 train +495 AR6_WGI 203 11 These climate changes have contributed to desertification and land degradation in many regions high 2 train +496 AR6_WGI 203 12 Increased urbanization can enhance warming in cities and their surroundings (heat island effect), especially during heat waves (high confidence), and intensify extreme rainfall medium 1 train +497 AR6_WGI 203 15 Over the period 1982–2016, marine heatwaves have very likely doubled in frequency and are increasing in intensity very high 3 test +498 AR6_WGI 203 16 In addition, the surface ocean acidified further (virtually certain) and loss of oxygen occurred from the surface to a depth of 1000 m medium 1 train +499 AR6_WGI 203 23 Feedbacks from the loss of summer sea ice and spring snow cover on land have contributed to amplified warming in the Arctic high 2 train +500 AR6_WGI 203 24 By contrast, Antarctic sea ice extent overall saw no statistically significant trend for the period 1979–2018 high 2 train +501 AR6_WGI 204 1 It also found evidence for an increase in the annual global proportion of Category 4 or 5 tropical cyclones in recent decades low 0 train +502 AR6_WGI 204 4 The natural response of land to human-induced environmental change – such as increasing atmospheric CO 2 concentration, nitrogen deposition and climate change – caused a net CO 2 sink equivalent of around 29% of total CO 2 emissions (medium confidence); however, the persistence of the sink is uncertain due to climate change high 2 train +503 AR6_WGI 204 7 However, the report estimated that the resulting net effect on globally averaged surface temperature was small over the historical period medium 1 train +504 AR6_WGI 204 8 The SROCC found that the carbon content of Arctic and boreal permafrost is almost twice that of the atmosphere medium 1 train +505 AR6_WGI 204 9 Projections of climate change The SR1.5 concluded that global warming is likely to reach 1.5°C between 2030 and 2052 if it continues to increase at the current rate high 2 train +506 AR6_WGI 204 11 The SR1.5 also found that reaching and sustaining net zero anthropogenic CO 2 emissions and reducing net non-CO 2 radiative forcing would halt anthropogenic global warming on multi-decadal time scales high 2 train +507 AR6_WGI 204 12 The maximum temperature reached is then determined by (i) cumulative net global anthropogenic CO 2 emissions up to the time of net zero CO 2 emissions (high confidence) and (ii) the level of non-CO 2 radiative forcing in the decades prior to the time that maximum temperatures are reached medium 1 train +508 AR6_WGI 204 13 Furthermore, climate models project robust differences in regional climate characteristics between the present day and a global warming of 1.5°C, and between 1.5°C and 2°C, including mean temperature in most land and ocean regions and hot extremes in most inhabited regions high 2 train +509 AR6_WGI 204 15 The SROCC projected that global-scale glacier mass loss, permafrost thaw, and decline in snow cover and Arctic sea ice extent will continue in the period 2031–2050 due to surface air temperature increases high 2 train +510 AR6_WGI 204 16 The Greenland and Antarctic ice sheets are projected to lose mass at an increasing rate throughout the 21st century and beyond high 2 train +511 AR6_WGI 204 19 For the RCP8.5 scenario, projections of GMSL rise by 2100 are higher by 0.1 m than in AR5 due to a larger contribution from the Antarctic Ice Sheet medium 1 train +512 AR6_WGI 204 20 Extreme sea level events that occurred once per hundred years in the recent past are projected to occur at least once per year at many locations by 2050, especially in tropical regions, under all RCP scenarios high 2 train +513 AR6_WGI 204 21 According to SR1.5, by 2100 GMSL rise would be around 0.1 m lower with 1.5°C global warming compared to 2°C medium 1 train +514 AR6_WGI 204 23 However, instability and/or irreversible loss of the Greenland and Antarctic ice sheets, resulting in a multi-metre rise in sea level over hundreds to thousands of years, could be triggered at 1.5°C–2°C of global warming medium 1 train +515 AR6_WGI 204 24 According to SROCC, sea level rise in an extended RCP2.6 scenario would be limited to around 1 m in 2300 (low confidence) while under RCP8.5 multi-metre sea level rise is projected by then medium 1 train +516 AR6_WGI 204 25 The SROCC projected that over the 21st century, the ocean will transition to unprecedented conditions, with increased temperatures (virtually certain), further acidification (virtually certain), and oxygen decline medium 1 train +517 AR6_WGI 204 26 Marine heatwaves are projected to become more frequent (very high confidence) as are extreme El Niño and La Niña events medium 1 train +518 AR6_WGI 205 20 Starting from year 2018, the remaining carbon budget for a one-in-two (50%) chance of limiting global warming to 1.5°C is about 580 GtCO 2, and about 420 GtCO 2 for a two-in-three (66%) chance medium 1 train +519 AR6_WGI 205 22 Using GMST instead of GSAT gives estimates of 770 GtCO 2 and 570 GtCO 2, respectively medium 1 train +520 AR6_WGI 209 1 It is likely that there was a net anthropogenic forcing of 0.0 –0.3 Wm–2 in 1850 –1900 relative to 1750 medium 1 train +521 AR6_WGI 209 11 Combining these different sources of evidence, we assess that from the period around 1750 to 1850–1900 there was a change in global temperature of around 0.1 [–0.1 to +0.3] °C (medium confidence), with an anthropogenic component in a likely range of 0.0°C–0.2°C medium 1 train +522 AR6_WGI 212 15 Overall, tropical regions show earlier emergence of temperature changes than at higher latitudes high 2 train +523 AR6_WGI 227 7 In summary, the observational coverage of ongoing changes to the climate system is improved at the time of AR6, relative to what was available for AR5 high 2 train +524 AR6_WGI 228 2 Overall, the number, temporal resolution and chronological accuracy of paleoclimate reconstructions have increased since AR5, leading to improved understanding of climate system processes (or Earth system processes) high 2 train +525 AR6_WGI 228 28 In summary, while the quantity, quality and diversity of climate system observations have grown since AR5, the loss or potential loss of several critical components of the observational network is also evident high 2 train +526 AR6_WGI 230 6 In summary, the improvements in atmospheric reanalyses, and the greater number of years since the routine ingestion of satellite data began, relative to AR5, mean that there is increased confidence in using atmospheric reanalysis products alongside more standard observation-based datasets in AR6 high 2 train +527 AR6_WGI 231 10 In summary, reanalyses have improved since AR5 and can increasingly be used as a line of evidence in assessments of the state and evolution of the climate system high 2 train +528 AR6_WGI 239 1 Overall, we assess that increases in computing power and the broader availability of larger and more varied ensembles of model simulations have contributed to better estimations of uncertainty in projections of future change high 2 train +529 AR6_WGI 297 25 The remainder is due to improved scientific understanding and changes in the assessment of aerosol forcing, which include decreases in concentration and improvement in its calculation high 2 train +530 AR6_WGI 298 24 The total 20th century rise is estimated to be 0.17 [0.12 to 0.22] m. The rate of sea level rise since the mid-19th century has been larger than the mean rate during the previous two millennia high 2 train +531 AR6_WGI 298 26 The average rate of sea level rise was 1.3 [0.6 to 2.1] mm yr–1 between 1901 and 1971, increasing to 1.9 [0.8 to 2.9] mm yr–1 between 1971 and 2006, and further increasing to 3.7 [3.2 to 4.2] mm yr–1 between 2006 and 2018 high 2 train +532 AR6_WGI 299 1 Such warming causes seawater to expand, contributing to sea level rise.Ocean warming dominates the increase in energy stored in the climate system, accounting for more than 90% of the energy accumulated between 1971 and 2010 high 2 train +533 AR6_WGI 299 5 The observed average rate of heating of the climate system increased from 0.50 [0.32 to 0.69] W m–2 for the period 1971–2006 to 0.79 [0.52 to 1.06] W m–2 for the period 2006– 2018 high 2 train +534 AR6_WGI 299 6 Ocean warming accounted for 91% of the heating in the climate system, with land warming, ice loss and atmospheric warming accounting for about 5%, 3% and 1%, respectively high 2 train +535 AR6_WGI 299 10 The ocean has absorbed about 30% of the emitted anthropogenic carbon dioxide, causing ocean acidification.In 2019, atmospheric CO 2 concentrations were higher than at any time in at least 2 million years (high confidence), and concentrations of CH 4 and N 2O were higher than at any time in at least 800,000 years very high 3 train +536 AR6_WGI 299 11 Since 1750, increases in CO 2 (47%) and CH 4 (156%) concentrations far exceed – and increases in N 2O (23%) are similar to – the natural multi-millennial changes between glacial and interglacial periods over at least the past 800,000 years very high 3 train +537 AR6_WGI 300 11 In the Northern Hemisphere, 1983–2012 was likely the warmest 30-year period of the last 1400 years medium 1 train +538 AR6_WGI 300 12 Global surface temperature has increased faster since 1970 than in any other 50-year period over at least the last 2000 years high 2 train +539 AR6_WGI 300 13 Temperatures during the most recent decade (2011–2020) exceed those of the most recent multi-century warm period, around 6500 years ago [0.2°C to 1°C relative to 1850–1900] medium 1 train +540 AR6_WGI 300 14 Prior to that, the next most recent warm period was about 125,000 years ago, when the multi-century temperature [0.5°C to 1.5°C relative to 1850–1900] overlaps the observations of the most recent decade medium 1 train +541 AR6_WGI 300 16 There is very high confidence that maximum global mean sea level during the last interglacial period (129,000 to 116,000 years ago) was, for several thousand years, at least 5 m higher than present, and high confidence that it did not exceed 10 m above present.Global mean sea level has risen faster since 1900 than over any preceding century in at least the last 3000 years high 2 train +542 AR6_WGI 300 17 The global ocean has warmed faster over the past century than since the end of the last deglacial transition (around 11,000 years ago) medium 1 train +543 AR6_WGI 300 18 A long- term increase in surface open ocean pH occurred over the past 50 million years high 2 train +544 AR6_WGI 300 19 However, surface open ocean pH as low as recent decades is unusual in the last 2 million years medium 1 train +545 AR6_WGI 302 4 Relative to 1995–2014, the likely global mean sea level rise by 2100 is 0.28–0.55 m under the very low GHG emissions scenario (SSP1-1.9); 0.32–0.62 m under the low GHG emissions scenario (SSP1-2.6); 0.44–0.76 m under the intermediate GHG emissions scenario (SSP2-4.5); and 0.63–1.01 m under the very high GHG emissions scenario (SSP5-8.5); and by 2150 is 0.37–0.86 m under the very low scenario (SSP1- 1.9); 0.46–0.99 m under the low scenario (SSP1-2.6); 0.66–1.33 m under the intermediate scenario (SSP2- 4.5); and 0.98–1.88 m under the very high scenario (SSP5-8.5) medium 1 train +546 AR6_WGI 302 5 Global mean sea level rise above the likely range – approaching 2 m by 2100 and 5 m by 2150 under a very high GHG emissions scenario (SSP5-8.5) low 0 train +547 AR6_WGI 306 9 Present-day global concentrations of atmospheric carbon dioxide (CO 2) are at higher levels than at any time in at least the past two million years high 2 train +548 AR6_WGI 306 10 Changes in ERF since the late 19th century are dominated by increases in concentrations of greenhouse gases and trends in aerosols; the net ERF is positive and changing at an increasing rate since the 1970s medium 1 test +549 AR6_WGI 306 12 Solar activity since 1900 was high but not exceptional compared to the past 9000 years high 2 train +550 AR6_WGI 306 13 The average magnitude and variability of volcanic aerosol forcing since 1900 have not been unusual compared to the past 2500 years medium 1 train +551 AR6_WGI 306 17 These changes are larger than those between glacial and interglacial periods over the last 800,000 years for CO 2 and CH 4 and of comparable magnitude for N 2O very high 3 train +552 AR6_WGI 306 21 Aerosol optical depth (AOD) has decreased since 2000 over Northern Hemisphere mid-latitudes and Southern Hemisphere mid-latitude continents, but increased over South Asia and East Africa high 2 train +553 AR6_WGI 306 25 Stratospheric ozone has declined between 60°S and 60°N by 2.2% from the 1980s to 2014–2017 high 2 train +554 AR6_WGI 306 26 Since the mid-20th century, tropospheric ozone has increased by 30–70% across the Northern Hemisphere medium 1 train +555 AR6_WGI 306 27 Since the mid-1990s, free tropospheric ozone increases were 2–7% per decade in the northern mid-latitudes (high confidence), 2–12% per decade in the tropics (high confidence) and <5% per decade in southern mid-latitudes medium 1 train +556 AR6_WGI 306 31 The best-estimate ERF from the increase in global albedo is –0.15 W m–2 since 1700 and –0.12 W m–2 since 1850 medium 1 test +557 AR6_WGI 306 33 Over the past several decades, key indicators of the climate system are increasingly at levels unseen in centuries to millennia, and are changing at rates unprecedented in at least the last 2000 years high 2 train +558 AR6_WGI 306 34 Temperatures as high as during the most recent decade (2011–2020) exceed the warmest centennial-scale range reconstructed for the present interglacial, around 6,500 years ago [0.2°C–1°C relative to 1850–1900] medium 1 train +559 AR6_WGI 306 35 The next older warm period is the last interglacial when the multi-centennial temperature range about 125,000 years ago [0.5°C–1.5°C relative to 1850–1900] encompassed the recent decade values medium 1 train +560 AR6_WGI 306 40 Over the last 50 years, observed GMST has increased at a rate unprecedented in at least the last 2000 years high 2 train +561 AR6_WGI 307 4 In the Tropics, the upper troposphere has warmed faster than the near-surface since at least 2001, the period over which new observational techniques permit more robust quantification medium 1 train +562 AR6_WGI 307 8 Global land precipitation has likely increased since 1950, with a faster increase since the 1980s medium 1 train +563 AR6_WGI 307 15 Global monsoon precipitation has likely increased since the 1980s, mainly in the Northern Hemisphere medium 1 train +564 AR6_WGI 307 20 Between 1979 and 2019, Arctic sea ice area has decreased in both summer and winter, with sea ice becoming younger, thinner and more dynamic very high 3 train +565 AR6_WGI 307 22 Antarctic sea ice area has experienced little net change since 1979 high 2 train +566 AR6_WGI 307 24 Reductions in spring snow cover extent have occurred across the Northern Hemisphere since at least 1978 very high 3 train +567 AR6_WGI 307 25 With few exceptions, glaciers have retreated since the second half of the 19th century and have continued to retreat at increased rates since the 1990s (very high confidence); this behaviour is unprecedented in at least the last 2000 years medium 1 train +568 AR6_WGI 307 26 Greenland Ice Sheet (GrIS) mass loss has increased substantially since 2000 high 2 train +569 AR6_WGI 307 27 The Greenland Ice Sheet was smaller than at present during the Last Interglacial period (high confidence) and the mid-Holocene high 2 train +570 AR6_WGI 307 28 The Antarctic Ice Sheet (AIS) lost mass between 1992 and 2020 (very high confidence), with an increasing rate of mass loss over this period medium 1 train +571 AR6_WGI 307 29 Although permafrost persists in areas of the Northern Hemisphere where it was absent prior to 3000 years ago, increases in temperatures in the upper 30 m over the past three to four decades have been widespread high 2 train +572 AR6_WGI 307 32 The average rate of sea level rise was 1.3 [0.6 to 2.1] mm yr –1 between 1901 and 1971, increasing to 1.9 [0.8 to 2.9] mm yr –1 between 1971 and 2006, and further increasing to 3.7 [3.2 to 4.2] mm yr –1 between 2006 and 2018 high 2 test +573 AR6_WGI 307 35 Since 1971, it is virtually certain that global ocean heat content has increased for the upper (0–700 m) layer, very likely for the intermediate (700–2000 m) layer and likely below 2000 m, and is currently increasing faster than at any point since at least the last deglacial transition (18 to 11 thousand years ago) medium 1 train +574 AR6_WGI 307 37 The Atlantic Meridional Overturning Circulation (AMOC) was relatively stable during the past 8000 years (medium confidence) but declined during the 20th century low 0 train +575 AR6_WGI 307 38 Ocean pH has declined globally at the surface over the past four decades (virtually certain) and in all ocean basins in the ocean interior high 2 train +576 AR6_WGI 307 39 A long-term increase in surface open ocean pH occurred over the past 50 million years (high confidence), and surface ocean pH as low as recent times is uncommon in the last 2 million years medium 1 train +577 AR6_WGI 308 1 Oxygen minimum zones are expanding at many locations high 2 train +578 AR6_WGI 308 3 The ranges of many marine organisms are shifting towards the poles and towards greater depths high 2 train +579 AR6_WGI 308 4 This mismatch in responses across species means that the species composition of ecosystems is changing medium 1 train +580 AR6_WGI 308 5 At multiple locations, various phenological metrics for marine organisms have changed in the last 50 years, with the nature of the changes varying with location and with species high 2 train +581 AR6_WGI 308 6 In the last two decades, the concentration of phytoplankton at the base of the marine food web, as indexed by chlorophyll concentration, has shown weak and variable trends in low and mid-latitudes and an increase in high latitudes medium 1 train +582 AR6_WGI 308 7 Global marine primary production decreased slightly from 1998–2018, with increasing production in the Arctic medium 1 train +583 AR6_WGI 308 9 Over the last century, there have been increases in species turnover within many ecosystems high 2 test +584 AR6_WGI 308 10 Over the past half century, climate zones have shifted poleward, accompanied by an increase in the length of the growing season in the Northern Hemisphere extratropics and an increase in the amplitude of the seasonal cycle of atmospheric CO 2 above 45°N high 2 train +585 AR6_WGI 308 11 Since the early 1980s, there has been a global-scale increase in the greenness of the terrestrial surface high 2 train +586 AR6_WGI 308 13 CO2 levels during the MPWP were similar to present for a sustained period, within a range of 360–420 ppm medium 1 test +587 AR6_WGI 308 14 Relative to the present, GMST, GMSL and precipitation rate were all higher, the Northern Hemisphere latitudinal temperature gradient was lower, and major terrestrial biomes were shifted northward very high 3 train +588 AR6_WGI 308 22 Both polar annular modes have exhibited strong positive trends toward increased zonality of mid-latitude circulation over multi-decadal periods, but these trends have not been sustained for the Northern Annular Mode since the early 1990s high 2 train +589 AR6_WGI 308 23 For tropical ocean modes, a sustained shift beyond multi-centennial variability has not been observed for El Niño–Southern Oscillation medium 1 train +590 AR6_WGI 308 24 Modes of decadal and multi-decadal variability over the Pacific and Atlantic oceans exhibit no significant trends over the period of observational records high 2 train +591 AR6_WGI 314 1 To conclude, solar activity since the late 19th century was relatively high but not exceptional in the context of the past 9 kyr high 2 train +592 AR6_WGI 314 25 However, the average magnitude and variability of SAOD and its associated volcanic aerosol forcing since 1900 are not unusual in the context of at least the past 2.5 kyr medium 1 train +593 AR6_WGI 315 13 Based on boron and carbon isotope data, supported by other proxies (Hollis et al., 2019), atmospheric CO2 during the EECO (50 Ma) was between 1150 and 2500 ppm medium 1 test +594 AR6_WGI 316 6 High-resolution sampling (about 1 sample per 3 kyr) with the boron-isotope proxy indicates mean CO 2 mixing ratios for the Marine Isotope Stage KM5c interglacial were 360–420 ppm medium 1 train +595 AR6_WGI 316 21 Although ice core records present low-pass filtered time series due to gas diffusion and gradual bubble close- off in the snow layer over the ice sheet (Fourteau et al., 2020), the rate of increase since 1850 CE (about 125 ppm increase over about 170 years) is far greater than implied for any 170-year period by ice core records that cover the last 800 ka very high 3 train +596 AR6_WGI 319 15 The last time CO 2 concentrations were similar to the present-day was over 2 Ma high 2 train +597 AR6_WGI 319 18 Between 1750–2019 mixing ratios increased by 131.6 ± 2.9 ppm (47%), 1137 ± 10 ppb (156%), and 62 ± 6 ppb (23%), for CO 2, CH 4, and N 2O, respectively very high 3 train +598 AR6_WGI 323 29 In summary, limited available isotopic evidence constrains the global tropospheric ozone increase to less than 40% between 1850 and 2005 low 0 train +599 AR6_WGI 323 30 Based on sparse historical surface/low altitude data tropospheric ozone has increased since the mid-20th century by 30–70% across the NH medium 1 train +600 AR6_WGI 323 31 Surface/low altitude ozone trends since the mid-1990s are variable at northern mid-latitudes, but positive in the tropics [2 to 17% per decade] high 2 train +601 AR6_WGI 323 32 Since the mid-1990s, free tropospheric ozone has increased by 2–7% per decade in most regions of the northern mid-latitudes, and 2–12% per decade in the sampled regions of the northern and southern tropics high 2 train +602 AR6_WGI 323 33 Limited coverage by surface observations precludes identification of zonal trends in the SH, while observations of tropospheric column ozone indicate increases of less than 5% per decade at southern mid-latitudes medium 1 test +603 AR6_WGI 326 1 Satellite data and ground-based records indicate that AOD exhibits predominantly negative trends since 2000 over NH mid-latitudes and SH continents, but increased over South Asia and East Africa high 2 train +604 AR6_WGI 326 30 In summary, biophysical effects from historical changes in land use have an overall negative ERF medium 1 train +605 AR6_WGI 326 31 The best-estimate ERF from the increase in global albedo is –0.15 W m–2 since 1700 and –0.12 W m–2 since 1850 medium 1 train +606 AR6_WGI 326 32 Biophysical effects of land-use change likely resulted in a net global cooling of about 0.1°C since 1750 medium 1 train +607 AR6_WGI 327 11 The net effect of aerosols (Sections 2.2.6 and 6.4) on the radiation budget, including their effect on clouds, and cloud adjustments, as well as the deposition of black carbon on snow (Section 7.3.4.3), was negative throughout the industrial period high 2 train +608 AR6_WGI 328 1 The relative importance of aerosol forcing compared to other forcing agents has decreased globally in the most recent 30 years medium 1 train +609 AR6_WGI 328 3 This effect increased since 1750, reaching current values of about –0.20 W m–2 medium 1 train +610 AR6_WGI 330 14 The AR5 concluded that the reconstructed GMST during the PETM was 4°C–7°C warmer than pre-PETM mean climate (low confidence), and that the EECO and the MPWP were 9°C–14°C and 1.9°C–3.6°C warmer than pre-industrial, respectively medium 1 train +611 AR6_WGI 330 15 The GMST during the LIG was assessed at 1°C–2°C warmer than pre-industrial medium 1 train +612 AR6_WGI 331 3 Together, these studies indicate that GMST was 4°C–10°C warmer during the MCO medium 1 train +613 AR6_WGI 331 15 In summary, GMST during the warmest millennia of the LIG (within the interval of around 129–125 ka) is estimated to have reached 0.5°C–1.5°C higher values than the 1850–1990 reference period medium 1 train +614 AR6_WGI 331 26 In summary, GMST is estimated to have been 5°C–7°C lower during the LGM (around 23–19 ka) compared with 1850–1900 medium 1 train +615 AR6_WGI 331 32 For average annual NH temperatures, the period 1983–2012 was assessed as very likely the warmest 30-year period of the past 800 years (high confidence) and likely the warmest 30-year period of the past 1.4 kyr (medium confidence); the warm multi-decadal periods prior to the 20th century were unsynchronized across regions, in contrast to the warming since the mid-20th century high 2 train +616 AR6_WGI 333 4 Taking all lines of evidence into account, the GMST averaged over the warmest centuries of the current interglacial period (sometime between around 6 and 7 ka) is estimated to have been 0.2°C–1.0°C higher than 1850–1900 medium 1 train +617 AR6_WGI 333 11 Moreover, the new proxy data compilation shows that the warming of the 20th century was more spatially uniform than any other century-scale temperature change of the CE medium 1 train +618 AR6_WGI 333 14 To conclude, following approximately 6 ka, GMST generally decreased, culminating in the coldest multi-century interval of the post-glacial period (since 8 ka), which occurred between around 1450 and 1850 high 2 train +619 AR6_WGI 333 16 Since around 1950, GMST has increased at an observed rate unprecedented for any 50-year period in at least the last 2000 years high 2 train +620 AR6_WGI 334 15 The effect of this change from trend-based to change-based metrics is currently relatively minor at –0.03°C (<5%) for the most recent decade, but this may not remain the case in future high 2 train +621 AR6_WGI 336 10 Given the projected future sea ice losses, the effect will grow in future low 0 test +622 AR6_WGI 337 19 The other half arises because, for central estimates of climate sensitivity, most scenarios show stronger warming over the near term than was assessed as ‘current’ in SR1.5 medium 1 train +623 AR6_WGI 339 16 The assessment of the implications of limiting global warming to 1.5°C compared to 2°C will also remain broadly unchanged by the updated estimate of historical warming, as this depends on the relative impacts rather than the absolute impacts at any specific definition of global temperature anomaly high 2 train +624 AR6_WGI 339 20 The SRCCL concluded that since the pre-industrial period, surface air temperature over land areas has risen nearly twice as much as the global mean surface temperature high 2 train +625 AR6_WGI 346 7 There are indications from multiple sources of a wetting trend during the Holocene, particularly for the NH and parts of the SH tropics medium 1 train +626 AR6_WGI 346 8 Hydroclimate during the CE is dominated by regional variability, generally precluding definitive statements on changes at continental and larger scales, with a general reduction of mega-drought occurrences over the last about 500 years medium 1 train +627 AR6_WGI 348 27 In summary, globally averaged land precipitation has likely increased since the middle of the 20th century medium 1 train +628 AR6_WGI 348 28 A faster increase in global land precipitation was observed since the 1980s medium 1 train +629 AR6_WGI 351 14 In summary, the sign of global streamflow trends remains uncertain, with slightly more globally gauged rivers experiencing significantly decreasing flows than significantly increasing flows since the 1950s low 0 train +630 AR6_WGI 353 10 This has been accompanied by a strengthening of the Hadley circulation, particularly in the NH medium 1 train +631 AR6_WGI 353 41 In summary, observed trends during the last century indicate that the GM precipitation decline reported in AR5 has reversed since the 1980s, with a likely increase mainly due to a significant positive trend in the NH summer monsoon precipitation medium 1 train +632 AR6_WGI 355 7 In summary, the total number of extratropical cyclones has likely increased since the 1980s in the NH low 0 train +633 AR6_WGI 355 8 The number of strong extratropical cyclones has likely increased in the SH medium 1 train +634 AR6_WGI 355 9 The extratropical jets and cyclone tracks have likely been shifting poleward in both hemispheres since the 1980s with marked seasonality in trends medium 1 train +635 AR6_WGI 357 32 This was confirmed by SROCC reporting the strongest reductions in September (12.8 ± 2.3% per decade; 1979–2018) and stating that these changes were likely unprecedented in at least 1 kyr medium 1 train +636 AR6_WGI 357 33 The spatial extent had decreased in all seasons, with the largest decrease for September high 2 train +637 AR6_WGI 357 34 The AR5 reported also that the average winter sea ice thickness within the Arctic Basin had likely decreased by between 1.3 m and 2.3 m from 1980 to 2008 high 2 train +638 AR6_WGI 358 1 Lower sea ice volume in 2010–2012 compared to 2003–2008 was documented in AR5 medium 1 train +639 AR6_WGI 359 14 In summary, over 1979–2019 Arctic SIA has decreased for all months, with the strongest decrease in summer very high 3 train +640 AR6_WGI 359 16 Arctic sea ice has become younger, thinner and faster moving very high 3 train +641 AR6_WGI 359 17 Snow thickness on sea ice has decreased in the western Arctic Ocean medium 1 train +642 AR6_WGI 359 19 Current pan-Arctic sea ice coverage levels (annual mean and late summer) are unprecedentedly low since 1850 high 2 train +643 AR6_WGI 359 22 The SROCC stated also that historical Antarctic sea ice data from different sources indicated a decrease in overall Antarctic sea ice cover since the early 1960s, but was too small to be separated from natural variability high 2 train +644 AR6_WGI 360 15 The 2020 September level (OSISAF) remains below the levels observed over 2012–2014.In summary, Antarctic sea ice has experienced both increases and decreases in SIA over 1979–2019, and substantively lower levels since 2016, with only minor differences between decadal means of SIA for the first (for February 2.04 million km2, for September 15.39 million km2) and last decades (for February 2.17 million km2, for September 15.75 million km2) of satellite observations high 2 train +645 AR6_WGI 361 5 In summary, substantial reductions in spring snow cover extent have occurred in the NH since 1978 very high 3 train +646 AR6_WGI 361 6 Since 1981 there has been a general decline in NH spring snow water equivalent high 2 train +647 AR6_WGI 361 9 The SROCC reported a globally coherent picture of continued glacier recession in recent decades very high 3 train +648 AR6_WGI 362 5 The current global character of glacier mass loss is highly unusual (almost all glaciers simultaneously receding) in the context of at least the last 2 kyr medium 1 train +649 AR6_WGI 362 6 Glacier mass loss rates have increased since the 1970s high 2 train +650 AR6_WGI 362 7 Although many surveyed glaciers are currently more extensive than during the MH high 2 train +651 AR6_WGI 362 11 It reported that the GrIS had lost ice during the prior two decades (very high confidence), that the ice loss had occurred in several sectors, and that high rates of mass loss had both expanded to higher elevations high 2 train +652 AR6_WGI 362 13 The SROCC also found that summer melting rate had increased since the 1990s to a rate unprecedented over the last 350 years (very high confidence), being two to five times greater than the pre-industrial rates medium 1 train +653 AR6_WGI 363 2 In summary, the GrIS was smaller than present during the MPWP (medium confidence), LIG (high confidence) and the MH high 2 train +654 AR6_WGI 363 3 GrIS mass loss began following a peak volume attained during the 1450–1850 period and the rate of loss has increased substantially since the turn of the 21st century high 2 train +655 AR6_WGI 363 8 The SROCC concluded that over 2006–2015, the AIS lost mass at an average rate of 155 ± 19 Gt yr –1 very high 3 train +656 AR6_WGI 363 21 Overall, during the MH, the AIS was retreating, but remained more extensive than present, while some parts of the ice sheet might have been smaller than now low 0 train +657 AR6_WGI 364 1 In summary, the AIS has lost mass between 1992 and 2020 very high 3 train +658 AR6_WGI 364 2 During the MPWP and LIG, the ice sheet was smaller than present medium 1 train +659 AR6_WGI 364 6 The AR5 also noted positive trends in active layer thickness (ALT; the seasonally thawed layer above the permafrost) since the 1990s for many high latitude sites medium 1 train +660 AR6_WGI 364 7 The SROCC concluded permafrost temperatures have increased to record high levels since the 1980s very high 3 train +661 AR6_WGI 365 2 In summary, increases in permafrost temperatures in the upper 30 m have been observed since the start of observational programs over the past three to four decades throughout the permafrost regions high 2 train +662 AR6_WGI 365 3 Limited evidence suggests that permafrost was less extensive during the MPWP low 0 train +663 AR6_WGI 365 4 Permafrost that formed after 3ka still persists in areas of the NH, but there are indications of thaw after the mid-1800s medium 1 test +664 AR6_WGI 366 6 In summary, current multi-decadal to centennial rates of OHC gain are greater than at any point since the last deglaciation medium 1 train +665 AR6_WGI 366 7 At multi-centennial timescales, changes in OHC based upon proxy indicators demonstrate a tight link with surface temperature changes during the last deglaciation (high confidence), as well as during the Holocene and CE low 0 train +666 AR6_WGI 368 18 The differences between high-salinity and low- salinity regions are linked to an intensification of the hydrological cycle medium 1 train +667 AR6_WGI 369 2 AR5 reported that GMSL during the LIG was, over several thousand years, between 5 and 10 m higher than 1985–2004 (medium confidence) whereas SROCC concluded it was virtually certain that GMSL exceeded current levels (high confidence), and reached a peak that was likely 6–9 m higher than today, but did not exceed 10 m medium 1 train +668 AR6_WGI 369 7 Given that GMSL change must be due to some combination of transient land ice growth and changes in terrestrial water storage, additional global mean thermosteric sea-level increase of 7 ± 2 m (Fischer et al., 2018) implies a peak EECO GMSL of 70–76 m low 0 train +669 AR6_WGI 369 10 Thus, consistent with SROCC, GMSL during the MPWP was higher than present by 5–25 m medium 1 train +670 AR6_WGI 369 14 It shows that GMSL during the Holocene was among the highest over this entire interval, and was surpassed only during the LIG (Marine Isotope Stage (MIS 5e)) and MIS 11 medium 1 train +671 AR6_WGI 369 32 The fastest rise occurred during Meltwater Pulse 1A, at about 14.6–14.3 ka (Deschamps et al., 2012; Sanborn et al., 2017), when GMSL rose by between 8 m and 15 m medium 1 test +672 AR6_WGI 371 16 In summary, GMSL is rising, and the rate of GMSL rise since the 20th century is faster than over any preceding century in at least the last three millennia high 2 train +673 AR6_WGI 371 21 SROCC assessed that there was emerging evidence in sustained observations, both in situ (2004–2017) and revealed from SST-based reconstructions, that the AMOC had weakened during the instrumental era relative to 1850–1900 medium 1 train +674 AR6_WGI 372 15 In summary, proxy-based reconstructions suggest that the AMOC was relatively stable during the past 8 kyr (medium confidence), with a weakening beginning since the late 19th century medium 1 train +675 AR6_WGI 372 16 From the mid-2000s to mid-2010s, the directly observed weakening in AMOC (high confidence) cannot be distinguished between decadal-scale variability or a long-term trend high 2 train +676 AR6_WGI 373 12 In summary, over the past 3–4 decades, the WBC strength is highly variable (high confidence), and WBCs and subtropical gyres have shifted poleward since 1993 medium 1 train +677 AR6_WGI 373 13 Net Arctic Ocean volume exchanges with the other ocean basins remained stable over the mid-1990s to the mid-2010s high 2 train +678 AR6_WGI 374 9 To conclude, it is virtually certain that surface open ocean pH has declined globally over the last 40 years by 0.003–0.026 pH per decade, and a decline in the ocean interior has been observed in all ocean basins over the past 2–3 decades high 2 train +679 AR6_WGI 374 10 A long-term increase in surface open ocean pH occurred over the past 50 Myr, and surface open ocean pH as low as recent times is uncommon in the last 2 Myr medium 1 test +680 AR6_WGI 374 14 Multidecadal rates of deoxygenation showed variability throughout the water column and across ocean basins high 2 train +681 AR6_WGI 375 17 In summary, episodes of widespread and long-lasting (100 ka scales) open-ocean deoxygenation were related to warm climate intervals of the Permian-Cretaceous, with conditions becoming generally better oxygenated as the climate cooled over the course of the Cenozoic high 2 train +682 AR6_WGI 375 18 The largest expansions of oxygen depleted waters over the past 25 ka were strongly linked to rapid warming rates medium 1 train +683 AR6_WGI 375 19 Open-ocean deoxygenation has occurred in most regions of the open ocean during the mid-20th to early 21st centuries (high confidence), and shows decadal variability medium 1 train +684 AR6_WGI 375 20 Evidence further confirms SROCC that OMZs are expanding at many locations high 2 train +685 AR6_WGI 376 8 Similarly, globally-integrated results from the SH also show an increase in seasonal amplitude of atmospheric CO 2 signal, from around 2009 to 2018 low 0 train +686 AR6_WGI 376 12 With respect to distributions of marine organisms, AR5 WGII reported range shifts of benthic, pelagic, and demersal species and communities high 2 train +687 AR6_WGI 378 12 A small decrease in productivity is evident globally for the period 1998–2015, but regional changes are larger and of opposing signs low 0 test +688 AR6_WGI 378 30 New in situ data as well as satellite observations strengthen AR5 and SROCC findings that various phenological metrics for many species of marine organisms have changed in the last half century high 2 test +689 AR6_WGI 378 31 The changes vary with location and with species high 2 train +690 AR6_WGI 378 32 There is a strong dependence of survival in higher trophic-level organisms (fish, exploited invertebrates, birds) on the availability of food at various stages in their life cycle, which in turn depends on phenologies of both high 2 train +691 AR6_WGI 380 10 Similarly, SRCCL assessed that many land species have experienced range size and location changes as well as altered abundances over recent decades high 2 train +692 AR6_WGI 380 11 SROCC noted that species composition and abundance have markedly changed in high mountain ecosystems in recent decades very high 3 train +693 AR6_WGI 381 12 The SRCCL subsequently concluded that greening had increased globally over the past 2–3 decades high 2 train +694 AR6_WGI 383 9 Many more integrative components of the climate system (e.g., glaciers, GMSL) are experiencing conditions unseen in millennia, whereas the most slowly responding components (e.g., ice-sheet extent, permafrost, tree line) are at levels unseen in centuries high 2 train +695 AR6_WGI 383 10 The rate at which several assessed climate indicators (e.g., GMSL, OHC, GSAT) have changed over recent decades is highly unusual in the context of preceding slower changes during the current post-glacial period high 2 train +696 AR6_WGI 386 3 Moreover, periods of persistent negative or positive NAO states observed during the latter part of the 20th century were not unusual, based on NAO reconstructions spanning the last half millennium high 2 train +697 AR6_WGI 386 18 In summary, positive trends for the NAM/NAO winter indices were observed between the 1960s and the early 1990s, but these indices have become less positive or even negative thereafter high 2 train +698 AR6_WGI 387 11 In summary, historical station-based reconstructions of the SAM show that there has been a robust positive trend in the SAM index, particularly since 1950 and for the austral summer high 2 train +699 AR6_WGI 387 14 It was also found high 2 train +700 AR6_WGI 390 7 Neither the IOD nor the IOB have exhibited behaviour outside the range implied by proxy records low 0 train +701 AR6_WGI 392 12 In summary, no sustained change in AMV indices has been observed over the instrumental period high 2 train +702 AR6_WGI 441 11 High-resolution models exhibit reduced biases in some but not all aspects of surface and ocean climate (medium confidence), and most Earth system models, which include biogeochemical feedbacks, perform as well as their lower-complexity counterparts medium 1 train +703 AR6_WGI 441 12 The multi-model mean captures most aspects of observed climate change well high 2 train +704 AR6_WGI 441 13 The multi-model mean captures the proxy-reconstructed global-mean surface air temperature (GSAT) change during past high- and low-CO 2 climates (high confidence) and the correct sign of temperature and precipitation change in most assessed regions in the mid-Holocene medium 1 train +705 AR6_WGI 441 14 The simulation of paleoclimates on continental scales has improved compared to AR5 (medium confidence), but models often underestimate large temperature and precipitation differences relative to the present day high 2 train +706 AR6_WGI 441 19 The likely ranges for human-induced GSAT and global mean surface temperature (GMST) warming are equal medium 1 train +707 AR6_WGI 441 26 CMIP6 models broadly reproduce surface temperature variations over the past millennium, including the cooling that follows periods of intense volcanism medium 1 train +708 AR6_WGI 441 28 The latest updates to satellite-derived estimates of stratospheric temperature have resulted in decreased differences between simulated and observed changes of global mean temperature through the depth of the stratosphere medium 1 train +709 AR6_WGI 441 31 All the observed estimates of the 1998–2012 GMST trend lie within the 10th–90th percentile range of CMIP6 simulated trends high 2 train +710 AR6_WGI 441 32 Internal variability, particularly Pacific Decadal Variability, and variations in solar and volcanic forcings partly offset the anthropogenic surface warming trend over the 1998–2012 period high 2 train +711 AR6_WGI 442 1 Since 2012, GMST has warmed strongly, with the past five years (2016–2020) being the warmest five-year period in the instrumental record since at least 1850 high 2 train +712 AR6_WGI 442 5 New attribution studies strengthen previous findings of a detectable increase in Northern Hemisphere mid- to high-latitude land precipitation high 2 train +713 AR6_WGI 442 6 Human influence has contributed to strengthening the zonal mean precipitation contrast between the wet tropics and dry subtropics medium 1 train +714 AR6_WGI 442 7 Yet, anthropogenic aerosols contributed to decreasing global land summer monsoon precipitation from the 1950s to 1980s medium 1 train +715 AR6_WGI 442 9 Despite improvements, models still have deficiencies in simulating precipitation patterns, particularly over the tropical ocean high 2 train +716 AR6_WGI 442 13 The magnitude and frequency of extreme precipitation simulated by CMIP6 models are similar to those simulated by CMIP5 models high 2 train +717 AR6_WGI 442 16 The causes of the observed strengthening of the Pacific Walker circulation since the 1980s are not well understood, and the observed strengthening trend is outside the range of trends simulated in the coupled models medium 1 train +718 AR6_WGI 442 17 While CMIP6 models capture the general characteristics of the tropospheric large-scale circulation (high confidence), systematic biases exist in the mean frequency of atmospheric blocking events, especially in the Euro-Atlantic sector, some of which reduce with increasing model resolution medium 1 train +719 AR6_WGI 442 19 There is new evidence that increases in anthropogenic aerosols have offset part of the greenhouse gas-induced Arctic sea ice loss since the 1950s medium 1 train +720 AR6_WGI 442 20 In the Arctic, despite large differences in the mean sea ice state, loss of sea ice extent and thickness during recent decades is reproduced in all CMIP5 and CMIP6 models high 2 train +721 AR6_WGI 442 23 The seasonal cycle in Northern Hemisphere snow cover is better reproduced by CMIP6 than by CMIP5 models high 2 train +722 AR6_WGI 442 29 Updated observations and model simulations show that warming extends throughout the entire water column high 2 train +723 AR6_WGI 442 30 The structure and magnitude of multi-model mean ocean temperature biases have not changed substantially between CMIP5 and CMIP6 medium 1 train +724 AR6_WGI 442 32 The associated pattern of change corresponds to fresh regions becoming fresher and salty regions becoming saltier high 2 train +725 AR6_WGI 442 33 Changes to the coincident atmospheric water cycle and ocean-atmosphere fluxes (evaporation and precipitation) are the primary drivers of the observed basin-scale salinity changes high 2 train +726 AR6_WGI 443 1 The basin-scale changes are consistent across models and intensify through the historical period high 2 train +727 AR6_WGI 443 2 The structure of the biases in the multi-model mean has not changed substantially between CMIP5 and CMIP6 medium 1 train +728 AR6_WGI 443 9 However, biases are apparent in the modelled circulation strengths (high confidence) and their variability medium 1 train +729 AR6_WGI 443 12 Earth system models simulate globally averaged land carbon sinks within the range of observation-based estimates high 2 train +730 AR6_WGI 443 18 The influence of ozone forcing on the SAM trend has been small since the early 2000s compared to earlier decades, contributing to a weaker SAM trend observed over 2000–2019 medium 1 train +731 AR6_WGI 443 19 Climate models reproduce the summertime SAM trend well, with CMIP6 models outperforming CMIP5 models medium 1 train +732 AR6_WGI 443 21 Models reproduce the observed spatial features and variance of the SAM and NAM very well high 2 train +733 AR6_WGI 443 23 Further assessment since AR5 confirms that climate and Earth system models are able to reproduce most aspects of the spatial structure and variance of the El Niño–Southern Oscillation and Indian Ocean Basin and Dipole modes medium 1 train +734 AR6_WGI 443 28 Uncertainties remain in quantification of the human influence on AMV and PDV due to brevity of the observational records, limited model performance in reproducing related sea surface temperature (SST) anomalies despite improvements from CMIP5 to CMIP6 medium 1 train +735 AR6_WGI 449 6 CMIP6 models broadly reproduce surface temperature variations over the past millennium, including the cooling that follows periods of intense volcanism medium 1 train +736 AR6_WGI 449 7 Simulated GMST anomalies are well within the uncertainty range of temperature reconstructions medium 1 test +737 AR6_WGI 449 8 Before the year 1300, larger disagreements between models and temperature reconstructions are expected because forcing and temperature reconstructions are increasingly uncertain further back in time, but specific causes have not been identified conclusively (Ljungqvist et al., 2019; PAGES 2k Consortium, 2019) medium 1 train +738 AR6_WGI 454 18 CMIP6 models also reproduce historical GSAT changes similarly to their CMIP5 counterparts medium 1 train +739 AR6_WGI 461 26 In summary, based on the latest updates to satellite observations of stratospheric temperature, we assess that simulated and observed trends in global mean temperature through the depth of the stratosphere are more consistent than based on previous datasets, but some differences remain medium 1 train +740 AR6_WGI 462 13 Therefore, the observed 1998–2012 trend is consistent with both the CMIP5 or CMIP6 multi-model ensemble of trends over the same period high 2 train +741 AR6_WGI 463 12 Although there is noticeable uncertainty among observational products (H. Su et al., 2017) and observation quality changes through time, global ocean heat content continued to increase during the slower surface warming period very high 3 test +742 AR6_WGI 463 20 Nonetheless, the heating of the climate system continued during this period, as reflected in the continued warming of the global ocean (very high confidence) and in the continued rise of hot extremes over land medium 1 train +743 AR6_WGI 463 23 The past five-year period (2016–2020) is the hottest five-year period in the instrumental record up to 2020 high 2 train +744 AR6_WGI 465 7 Records of tree ring width have provided evidence that recent prolonged dry spells in the Levant and Chile are unprecedented in the last millennium high 2 train +745 AR6_WGI 465 10 Likewise, tree rings indicate that the 2012–2014 drought in the south-western United States was exceptionally severe in the context of natural variability over the last millennium, and may have been exacerbated by the contribution of anthropogenic temperature rise medium 1 train +746 AR6_WGI 465 13 Nonetheless, tree rings also indicate the presence of prolonged megadroughts in western North America throughout the last millennium that were more severe than 20th and 21st century events high 2 train +747 AR6_WGI 466 4 However, prescribing changes in vegetation and dust was found to improve the match to the paleoclimate record (Pausata et al., 2016; Tierney et al., 2017) suggesting that vegetation feedbacks in the CMIP5 models may be too weak low 0 train +748 AR6_WGI 468 30 Based on these results we assess that despite some improvements, CMIP6 models still have deficiencies in simulating precipitation patterns, particularly over the tropical ocean high 2 train +749 AR6_WGI 470 11 Global land precipitation has likely increased since the middle of the 20th century medium 1 train +750 AR6_WGI 472 11 New attribution studies strengthen previous findings of a detectable increase in mid to high latitude land precipitation over the Northern Hemisphere high 2 train +751 AR6_WGI 472 36 Moreover, human interventions and water withdrawals, while affecting streamflow, cannot explain the observed spatio-temporal trends medium 1 train +752 AR6_WGI 473 17 The overall performance of CMIP6 models in simulating extreme precipitation intensity and frequency is similar to that of CMIP5 models high 2 train +753 AR6_WGI 477 9 For the strong ozone depletion period of 1981–2000, human influence is detectable in the summertime poleward expansion in the Southern Hemisphere medium 1 train +754 AR6_WGI 477 11 The causes of the observed strengthening of the Pacific Walker circulation over the 1980–2014 period are not well understood, since the observed strengthening trend is outside the range of variability simulated in the coupled models medium 1 train +755 AR6_WGI 477 18 AR5 assessed that CMIP5 models simulated monsoons better than CMIP3 models but that biases remained in domains and intensity high 2 train +756 AR6_WGI 479 15 In summary, while greenhouse gas increases acted to enhance the global land monsoon precipitation over the 20th century (medium confidence), consistent with projected future enhancement (Section 4.5.1.5), this tendency was overwhelmed by anthropogenic aerosols from the 1950s to the 1980s, which contributed to weakening of global land summer monsoon precipitation intensity for this period medium 1 train +757 AR6_WGI 479 19 While CMIP5 models can capture the domain and precipitation intensity of the global monsoon, biases remain in their regional representations, and they are unsuccessful in quantitatively reproducing changes in paleo reconstructions high 2 train +758 AR6_WGI 479 20 CMIP6 models reproduce the domain and precipitation intensity of the global monsoon observed over the instrumental period better than CMIP5 models medium 1 train +759 AR6_WGI 480 23 Section 2.3.1.4.3 assesses that the total number of extratropical cyclones has likely increased since the 1980s in the Northern Hemisphere low 0 train +760 AR6_WGI 480 28 Section 2.3.1.4.3 assesses that the extratropical jets and cyclone tracks have likely shifted poleward in both hemispheres since the 1980s with marked seasonality in trends medium 1 train +761 AR6_WGI 482 22 The recent Arctic sea ice loss during summer is unprecedented since 1850 high 2 train +762 AR6_WGI 483 10 Sea ice thermodynamic considerations indicate that the magnitude of sea ice variability and loss depends on ice thickness (Bitz, 2008; Massonnet et al., 2018) and hence the climatology simulated by different models may influence their simulated sea ice trends medium 1 train +763 AR6_WGI 483 11 An important consideration in comparing Arctic sea ice loss in models and observations is the role of internal variability medium 1 train +764 AR6_WGI 484 12 Increases in anthropogenic aerosols have offset part of the greenhouse gas induced Arctic sea ice loss since the 1950s medium 1 train +765 AR6_WGI 487 6 CMIP6 models better represent the seasonality and geographical distribution of snow cover than CMIP5 simulations high 2 train +766 AR6_WGI 488 9 Section 2.3.2.4.1 assesses that Greenland Ice Sheet mass loss began in the latter half of the 19th century and that the rate of loss has increased substantially since the turn of the 21st century high 2 train +767 AR6_WGI 488 25 It further assessed that Antarctic ice loss is dominated by acceleration, retreat and rapid thinning of the major West Antarctic Ice Sheet outlet glaciers (very high confidence), driven by melting of ice shelves by warm ocean waters high 2 train +768 AR6_WGI 492 1 Nevertheless, the CMIP6 models show a somewhat more realistic pattern of SST trends low 0 test +769 AR6_WGI 493 6 The multi-model means of both CMIP5 and CMIP6 historical simulations forced with time varying natural and anthropogenic forcing shows robust increases in ocean heat content in the upper (0–700 m) and intermediate (700–2000 m) ocean high 2 train +770 AR6_WGI 494 14 It is extremely likely that human influence was the main driver of the ocean heat content increase observed since the 1970s, which extends into the deeper ocean very high 3 train +771 AR6_WGI 494 15 Updated observations, like model simulations, show that warming extends throughout the entire water column high 2 train +772 AR6_WGI 495 14 The structure of the salinity biases in the multi-model mean has not changed substantially between CMIP5 and CMIP6 medium 1 test +773 AR6_WGI 497 5 All available multi-decadal assessments have confirmed that the associated pattern of change corresponds to fresh regions becoming fresher and salty regions becoming saltier high 2 train +774 AR6_WGI 497 6 CMIP5 and CMIP6 models are only able to reproduce these patterns in simulations that include greenhouse gas increases medium 1 test +775 AR6_WGI 497 7 Changes to the coincident atmospheric water cycle and ocean-atmosphere fluxes (evaporation and precipitation) are the primary drivers of the basin-scale observed salinity changes high 2 train +776 AR6_WGI 497 9 The basin-scale changes are consistent across models and intensify on centennial scales from the historical period through to the projections of future climate high 2 train +777 AR6_WGI 499 27 As reported in Section 2.3.3.4.1, estimates of AMOC since at least 1950, which are generated from observed surface temperatures or sea surface height, suggest the AMOC weakened through the 20th century low 0 train +778 AR6_WGI 501 21 In summary, while there have been improvements across successive CMIP phases (from CMIP3 to CMIP6) in the representation of the Southern Ocean circulation, such that the mean zonal and overturning circulations of the Southern Ocean are now broadly reproduced, substantial observational uncertainty and climate model challenges preclude attribution of Southern Ocean circulation changes high 2 train +779 AR6_WGI 501 24 That increase was attributed to direct land use and management changes, as well as to CO 2 fertilization, nitrogen deposition, increased diffuse radiation and climate change high 2 train +780 AR6_WGI 503 40 In summary, Earth system models simulate globally averaged land carbon sinks within the range of observation-based estimates high 2 train +781 AR6_WGI 504 15 The decrease in ocean oxygen content in the upper 1000 m, between 1970 and 2010, is further confirmed in SROCC medium 1 train +782 AR6_WGI 505 6 Section 5.2.1.3 assesses that both observational reconstructions based on the partial pressure of CO 2 and ocean biogeochemical models show a quasi-linear increase in the ocean sink of anthropogenic CO 2 from 1.0 ± 0.3 PgC yr–1 to 2.5 ± 0.6 PgC yr–1 between 1960–1969 and 2010–2019 in response to global CO 2 emissions high 2 train +783 AR6_WGI 509 9 CMIP5 and CMIP6 models are skilful in simulating the spatial features and the variance of the NAM/NAO and associated teleconnections high 2 test +784 AR6_WGI 510 20 However, despite the aforementioned limitations of the reconstructions, Section 2.4.1.2 assesses that the recent positive trend in the SAM is likely unprecedented in at least the past millennium medium 1 train +785 AR6_WGI 510 24 While ozone depletion contributed to the trend from the 1970s to the 1990s (medium confidence), its influence has been small since 2000, leading to a weaker summertime SAM trend over 2000–2019 medium 1 train +786 AR6_WGI 510 25 Climate models reproduce the spatial structure of the summertime SAM observed since the late 1970s well high 2 train +787 AR6_WGI 510 26 CMIP6 models reproduce the spatiotemporal features and recent multi-decadal trend of the summertime SAM better than CMIP5 models medium 1 train +788 AR6_WGI 510 27 However, there is a large spread in the intensity of the SAM response to ozone and greenhouse gas changes in both CMIP5 and CMIP6 models high 2 train +789 AR6_WGI 511 1 CMIP5 and CMIP6 models do not capture multicentennial variability of the SAM found in proxy reconstructions low 0 train +790 AR6_WGI 511 16 As assessed in Section 2.4.2, ENSO amplitude since 1950 is higher than over the pre-industrial period from 1850 as far back as 1400 medium 1 train +791 AR6_WGI 517 1 Nevertheless, CMIP5 models have medium overall performance in reproducing both the interannual IOB and IOD modes, with an apparently good performance in reproducing the IOB magnitude arising from compensation of biases in the formation process, and overly high IOD magnitude due to the mean state bias high 2 train +792 AR6_WGI 520 3 As such, PDV is an important driver of decadal internal climate variability which limits detection of human influence on various aspects of decadal climate change on global to regional scales high 2 train +793 AR6_WGI 520 5 Despite the limitations of these model-observation comparisons, CMIP5 models, on average, simulate broadly realistic spatial structures of the PDV, but with a clear bias in the South Pacific medium 1 train +794 AR6_WGI 522 5 In particular, anthropogenic and volcanic aerosols are thought to have played a role in the timing and intensity of the negative (cold) phase of AMV recorded from the mid-1960s to mid-1990s and subsequent warming medium 1 train +795 AR6_WGI 522 23 The pattern of ocean salinity changes indicate that fresh regions are becoming fresher and that salty regions are becoming saltier as a result of changes in ocean-atmosphere fluxes through evaporation and precipitation high 2 test +796 AR6_WGI 526 1 Earth System models characterized by additional biogeochemical feedbacks often perform at least as well as related more constrained, lower-complexity models lacking these feedbacks medium 1 train +797 AR6_WGI 526 2 In many cases, the models score similarly against both observational references, indicating that model errors are usually larger than observational uncertainties high 2 train +798 AR6_WGI 526 3 Moreover, synthesizing across Sections 3.3–3.7, we assess that the CMIP6 multi-model mean captures most aspects of observed climate change well high 2 train +799 AR6_WGI 528 27 Overall, the PMIP multi-model means agree very well (within 0.5°C of the assessed range) with GSAT reconstructed from proxies across multiple time periods, spanning a range from 6°C colder than pre-industrial (Last Glacial Maximum) to 14°C warmer than pre-industrial (Early Eocene Climate Optimum) high 2 train +800 AR6_WGI 528 28 During the orbitally-forced mid-Holocene, the CMIP6 multi-model mean captures the sign of the regional changes in temperature and precipitation in most regions assessed, and there have been some regional improvements compared to AR5 medium 1 train +801 AR6_WGI 528 29 The limited number of CMIP6 simulations of the LGM hinders model evaluation of the multi-model mean, but for both LGM and mid-Holocene, models tend to underestimate the magnitude of large changes high 2 train +802 AR6_WGI 528 30 Some long-standing model-data discrepancies, such as a dry bias in North Africa in the mid-Holocene, have not improved in CMIP6 compared with PMIP3 high 2 train +803 AR6_WGI 530 7 In comparison with standard resolution CMIP6 models, higher resolution probed under the HighResMIP activity (Haarsma et al., 2016) improves aspects of the simulation of climate (particularly concerning sea surface temperature) but discrepancies remain and there are some regions, such as parts of the Southern Ocean, where currently attainable resolution produces inferior performance high 2 train +804 AR6_WGI 571 16 The other half arises because for central estimates of climate sensitivity, most scenarios show stronger warming over the near term than was assessed as ‘current’ in SR1.5 medium 1 train +805 AR6_WGI 571 19 If climate sensitivity lies near the lower end of the assessed very likely range, crossing the 1.5°C warming threshold is avoided in scenarios SSP1-1.9 and SSP1-2.6 medium 1 train +806 AR6_WGI 571 21 Uncertainty in near-term projections of annual GSAT arises in roughly equal measure from natural internal variability and model uncertainty high 2 train +807 AR6_WGI 571 23 Forecasts initialized from recent observations simulate annual GSAT changes for the period 2019–2028 relative to the recent past that are consistent with the assessed very likely range high 2 train +808 AR6_WGI 571 26 The uncertainty ranges for the period 2081–2100 continue to be dominated by the uncertainty in ECS and TCR very high 3 train +809 AR6_WGI 571 27 Emissions-driven simulations for SSP5-8.5 show that carbon-cycle uncertainty is too small to change the assessment of GSAT projections high 2 train +810 AR6_WGI 571 30 In SSP1-2.6 and SSP2-4.5, ERF changes also explain about half of the changes in the range of warming medium 1 train +811 AR6_WGI 571 31 For SSP5-8.5, higher climate sensitivity is the primary reason behind the upper end of the warming being higher than in CMIP5 medium 1 train +812 AR6_WGI 571 33 For SSP1-2.6, such a high-warming storyline implies long-term (2081–2100) warming well above, rather than well below, 2°C high 2 train +813 AR6_WGI 571 34 Irrespective of scenario, high-warming storylines imply changes in many aspects of the climate system that exceed the patterns associated with the central estimate of GSAT changes by up to more than 50% high 2 train +814 AR6_WGI 572 2 The warming pattern likely varies across seasons, with northern high latitudes warming more during boreal winter than summer medium 1 train +815 AR6_WGI 572 11 As warming increases, a larger land area will experience statistically significant increases or decreases in precipitation medium 1 train +816 AR6_WGI 572 13 Interannual variability of precipitation over many land regions will increase with global warming medium 1 train +817 AR6_WGI 572 15 In the near term, no discernible differences in precipitation changes are projected between different SSPs high 2 train +818 AR6_WGI 572 16 The anthropogenic aerosol forcing decreases in most scenarios, contributing to increases in GSAT (medium confidence) and global mean land precipitation low 0 train +819 AR6_WGI 572 18 In the long term (2081–2100), monsoon rainfall change will feature a north–south asymmetry characterized by a greater increase in the Northern Hemisphere than in the Southern Hemisphere and an east–west asymmetry characterized by an increase in Asian-African monsoon regions and a decrease in the North American monsoon region medium 1 train +820 AR6_WGI 572 19 Near-term changes in global monsoon precipitation and circulation are uncertain due to model uncertainty and internal variability such as Atlantic Multi-decadal Variability and Pacific Decadal Variability medium 1 train +821 AR6_WGI 572 24 This is because of the opposing influence in the near- to mid-term from stratospheric ozone recovery and increases in other greenhouse gases on the Southern Hemisphere summertime mid-latitude circulation high 2 train +822 AR6_WGI 573 1 One exception is the expected decrease in frequency of atmospheric blocking events over Greenland and the North Pacific in boreal winter in SSP3-7.0 and SSP5-8.5 scenarios medium 1 train +823 AR6_WGI 573 5 AMV influences on the nearby regions can be predicted over lead times of 5–8 years medium 1 train +824 AR6_WGI 573 7 There is no model consensus for a systematic change in intensity of ENSO sea surface temperature variability over the 21st century in any of the SSP scenarios assessed medium 1 train +825 AR6_WGI 573 10 Arctic sea ice area in March, the month of annual maximum sea ice area, also decreases in the future under each of the considered scenarios, but to a much lesser degree (in percentage terms) than in September high 2 train +826 AR6_WGI 573 12 For the period 2081–2100 relative to 1995–2014, GMSL is likely to rise by 0.46–0.74 m under SSP3-7.0 and by 0.30–0.54 m under SSP1-2.6 medium 1 train +827 AR6_WGI 573 17 The fraction of emissions absorbed by land and ocean sinks will be smaller under high emissions scenarios than under low emissions scenarios high 2 train +828 AR6_WGI 573 18 Ocean surface pH will decrease steadily through the 21st century, except for SSP1-1.9 and SSP1-2.6 where values decrease until around 2070 and then increase slightly to 2100 high 2 train +829 AR6_WGI 573 20 However, the response of many other climate quantities to mitigation would be largely masked by internal variability during the near term, especially on the regional scale high 2 train +830 AR6_WGI 573 21 The mitigation benefits for these quantities would emerge only later during the 21st century high 2 train +831 AR6_WGI 573 22 During the near term, a small fraction of the surface can show cooling under all scenarios assessed here, so near-term cooling at any given location is fully consistent with GSAT increase high 2 train +832 AR6_WGI 573 23 Events of reduced and increased GSAT trends at decadal time scales will continue to occur in the 21st century but will not affect the centennial warming very high 3 train +833 AR6_WGI 573 27 The climate effect of a sudden and sustained CDR termination would depend on the amount of CDR-induced cooling prior to termination and the rate of background CO 2 emissions at the time of termination high 2 train +834 AR6_WGI 573 31 A sudden and sustained termination of SRM in a high-emissions scenario such as SSP5-8.5 would cause a rapid climate change high 2 train +835 AR6_WGI 574 2 Overshooting specific global warming levels such as 2°C has effects on the climate system that persist beyond 2100 medium 1 train +836 AR6_WGI 574 3 Under one scenario including a peak and decline in atmospheric CO 2 concentration (SSP5-3.4-OS), some climate metrics such as GSAT begin to decline but do not fully reverse by 2100 to levels prior to the CO 2 peak medium 1 train +837 AR6_WGI 574 9 GSAT projected for the end of the 23rd century under SSP5-8.5 (likely 6.6°C–14.1°C higher than over the period 1850–1900) overlaps with the range estimated for the Miocene Climatic Optimum (5°C–10°C higher) and Early Eocene Climatic Optimum (10°C–18°C higher), about 15 and 50 million years ago, respectively medium 1 train +838 AR6_WGI 584 25 The warming has increased in part because of models with higher ECS in CMIP6, compared to CMIP5 high 2 train +839 AR6_WGI 587 3 The AR5 further assessed that GSAT averaged over the period 2081–2100 are projected to likely exceed 1.5°C above 1850–1900 for RCP4.5, RCP6.0 and RCP8.5 (high confidence) and are likely to exceed 2°C above 1850–1900 for RCP6.0 and RCP8.5 high 2 train +840 AR6_WGI 587 4 Global surface temperature changes above 2°C under RCP2.6 were deemed unlikely medium 1 train +841 AR6_WGI 588 4 In summary, the CMIP6 models show a general tendency toward larger long-term globally averaged surface warming than did the CMIP5 models, for nominally comparable scenarios very high 3 train +842 AR6_WGI 588 5 In SSP1-2.6 and SSP2-4.5, the 5–95% ranges have remained similar to the ranges in RCP2.6 and RCP4.5, respectively, but the distributions have shifted upward by about 0.3°C high 2 train +843 AR6_WGI 588 6 For SSP5-8.5 compared to RCP8.5, the 5% bound of the distribution has hardly changed, but the 95% bound and the range have increased by about 20% and 40%, respectively high 2 train +844 AR6_WGI 588 22 This means that the CMIP6 spread in GSAT response to CO 2 emissions is dominated by climate sensitivity differences between ESMs more than by carbon cycle differences high 2 train +845 AR6_WGI 589 4 These differences due to experimental configuration would be smaller still under scenarios with lower CO 2 levels, and so we assess that results from concentration-driven and emissions- driven configurations do not affect the assessment of GSAT projections high 2 train +846 AR6_WGI 589 8 Based on these results, we conclude that global land precipitation is larger during the period 2081–2100 than during the period 1995–2014, under all scenarios considered here high 2 train +847 AR6_WGI 590 4 There is no change in subtropical precipitation in the North Atlantic following SSP1-1.9, SSP1-2.6, or SSP2-4.5 high 2 train +848 AR6_WGI 590 6 These range from 43% under RCP2.6 and 94% under RCP8.5 in September, and from 8% under RCP2.6 and 34% under RCP8.5 in March medium 1 train +849 AR6_WGI 591 3 Based on results from the CMIP6 models, we conclude that on average the Arctic will become practically ice-free in September by the end of the 21st century under SSP2-4.5, SSP3-7.0, and SSP5-8.5 high 2 train +850 AR6_WGI 591 4 Also, in the CMIP6 models, Arctic SIA in March decreases in the future, but to a much lesser degree, in percentage terms, than in September high 2 train +851 AR6_WGI 591 11 The individual model simulations, for which there are twenty for each stabilized temperature level, show that the probability of the Arctic becoming practically ice free at the end of the 21st century is significantly higher for 2°C warming than for 1.5°C warming above 1850–1900 levels high 2 train +852 AR6_WGI 591 13 Further, AR5 concluded that for the period 2081–2100, compared to 1986–2005, GMSL rise is likely medium 1 train +853 AR6_WGI 592 11 In summary, we assess from the CMIP6 models that AMOC weakening over the 21st century is very likely ; the rate of weakening is approximately independent of the emissions scenario high 2 train +854 AR6_WGI 592 15 In summary, in these model simulations the AMOC recovers over several centuries after the cessation of CO 2 emissions medium 1 train +855 AR6_WGI 593 3 We assess that the cumulative uptake of carbon by the ocean and by land will increase through the 21st century irrespective of the considered emissions scenarios except SSP1-1.9 very high 3 test +856 AR6_WGI 593 6 Based on results from the CMIP6 models we conclude that, except for the lower-emissions scenarios SSP1-1.9 and SSP1-2.6, ocean surface pH decreases monotonically through the 21st century high 2 train +857 AR6_WGI 594 7 Strong positive trends for the NAM/NAO indices were observed since 1960, which have weakened since the 1990s high 2 train +858 AR6_WGI 594 13 Based on CMIP6 model results displayed in Figure 4.9a, we conclude that the boreal wintertime surface NAM is more positive by the end of the 21st century under SSP3-7.0 and SSP5-8.5 high 2 train +859 AR6_WGI 594 15 On the other hand, under neither of the lowest emissions scenarios, SSP1-1.9 and SSP1-2.6, does the NAM show a robust change, by the end of the 21st century high 2 train +860 AR6_WGI 595 12 Over the instrumental period, there has been a robust positive trend in the SAM index, particularly since 1970 high 2 train +861 AR6_WGI 595 22 In summary, under the highest emissions scenarios in the CMIP6 models, the SAM in the austral summer becomes more positive through the 21st century high 2 train +862 AR6_WGI 596 14 Over the mid-term period 2041–2060, the very likely GSAT ranges of SSP1-1.9 and SSP5-8.5 are almost completely distinct high 2 train +863 AR6_WGI 596 15 CMIP6 models project a wider range of GSAT change than the assessed range high 2 train +864 AR6_WGI 598 10 Roughly half of this difference arises from a larger historical warming diagnosed in AR6, while the other half arises because for central estimates of climate sensitivity, most scenarios show stronger warming over the near term than was estimated as ‘current’ in SR1.5 medium 1 train +865 AR6_WGI 598 14 If ECS and TCR lie near the lower end of the assessed very likely range, crossing the 1.5°C warming threshold is avoided in scenarios SSP1-1.9 and SSP1-2.6 medium 1 train +866 AR6_WGI 599 5 The AR5 further assessed that it is more likely than not that the mean GSAT for the period 2016–2035 will be more than 1°C above the mean for 1850–1900, and it is very unlikely that it will be more than 1.5°C above the 1850–1900 mean medium 1 train +867 AR6_WGI 599 8 Averaged over the twenty years of the near term and across all scenarios, GSAT is very likely to be higher than over 1995–2014 by 0.4°C–1.0°C (Table 4.5), with most of the uncertainty arising from that in ECS and TCR high 2 train +868 AR6_WGI 599 13 For annual mean GSAT, uncertainty in near-term projections arises in roughly equal measure from internal variability and model uncertainty high 2 train +869 AR6_WGI 599 14 Forecasts initialized from recent observations simulate GSAT changes for the period 2019–2028 relative to the recent past that are consistent with the assessed very likely range in annual mean GSAT high 2 train +870 AR6_WGI 599 19 By 2030, GSAT in any individual year could exceed 1.5°C relative to 1850–1900 with a likelihood between 40 and 60 percent, across the scenarios considered here medium 1 train +871 AR6_WGI 601 9 Near-term projected changes in precipitation are uncertain mainly because of natural internal variability, model uncertainty, and uncertainty in natural and anthropogenic aerosol forcing medium 1 train +872 AR6_WGI 602 7 In summary, we assess that near-term changes in global monsoon precipitation and circulation will be affected by the combined effects of model uncertainty and internal variability, such as AMV and PDV, which together are larger than the forced signal medium 1 train +873 AR6_WGI 603 1 Based on results from CMIP6 models, we conclude that SSP2-4.5, SSP3-7.0, and SSP5-8.5 all clearly lead to increasing 10-, 20-, and 30-year trends in ocean carbon flux over the near term high 2 train +874 AR6_WGI 604 5 Considering these new results since AR5, in the near-term it is likely that any anthropogenic forced signal in the NAM will be of comparable magnitude or smaller than natural internal variability in the NAM medium 1 train +875 AR6_WGI 605 12 Hence, no robust change in amplitude of ENSO SST and rainfall variability is expected in the near term although the rainfall variability slightly increases medium 1 train +876 AR6_WGI 606 1 However, some studies show that despite severe model biases there are skilful predictions in the mean state of tropical Atlantic surface temperature several years ahead medium 1 train +877 AR6_WGI 607 11 The AR5 assessed that emission reductions aimed at decreasing local air pollution could have a near-term warming impact on climate high 2 train +878 AR6_WGI 608 15 Volcanic eruptions generally result in decreased global precipitation for up to a few years following the eruption (Iles and Hegerl, 2014, 2015; Man et al., 2014), with climatologically wet regions drying and climatologically dry regions wetting medium 1 train +879 AR6_WGI 608 22 In these simulations with multiple volcanic forcing futures there is: (i) an increase in the frequency of extremely cold individual years; (ii) an increased likelihood of decades with negative GSAT trend (decades with negative GSAT trends become 50% more commonplace); (iii) later anthropogenic signal emergence (the mean time at which the signal of global warming emerges from the noise of natural climate variability is delayed almost everywhere) high 2 train +880 AR6_WGI 609 20 Yet, remaining disagreements reflect differences in the volcanic forcing datasets used in the simulations medium 1 train +881 AR6_WGI 612 8 The ratio of land-to-ocean warming is greater than one for almost all regions high 2 train +882 AR6_WGI 612 22 Projected reduction in the strength of the AMOC over the 21st century is expected to reduce Arctic warming, but even a strong AMOC reduction would not eliminate Arctic amplification entirely medium 1 train +883 AR6_WGI 614 2 There is growing evidence that year-to-year and day-to-day temperature variability decreases in winter over northern mid- to high-latitudes (Fischer et al., 2011; De Vries et al., 2012; Screen, 2014; Schneider et al., 2015; Holmes et al., 2016; Borodina et al., 2017; Tamarin-Brodsky et al., 2020) which implies that the lowest temperatures rise more than the respective climatological mean temperatures medium 1 train +884 AR6_WGI 617 32 Consistent with the AR5, patterns of precipitation change are very likely to increase in the high latitudes especially during local winter and over tropical oceans under SSP3-7.0 high 2 train +885 AR6_WGI 617 33 CMIP6 projections show an increase in precipitation over larger parts of the monsoon regions and decreases in many subtropical regions including the Mediterranean, southern Africa and south-west Australia medium 1 train +886 AR6_WGI 618 5 CMIP6 models show greater increases in precipitation over land than either globally or over the ocean high 2 train +887 AR6_WGI 618 8 The patterns of precipitation change will exhibit substantial regional differences and seasonal contrast as GSAT increases over the 21st century high 2 train +888 AR6_WGI 619 20 The patterns of monsoon rainfall change in the mid- to long-term include a north–south asymmetry characterized by greater increase in the NH than the SH, and an East–West asymmetry characterized by enhanced Asian-African monsoons and weakened North American monsoon medium 1 train +889 AR6_WGI 619 23 Monsoon precipitation responses depend on region and emissions scenario high 2 train +890 AR6_WGI 620 4 The AR5 assessed that a poleward shift of the SH westerlies and storm track is likely by the end of the 21st century under RCP8.5 medium 1 train +891 AR6_WGI 624 13 The projected large-scale surface ocean acidification will be primarily determined by the pathway of atmospheric CO 2, with weak dependence on change in climate high 2 train +892 AR6_WGI 624 14 However, for a given atmospheric CO 2 scenario, uncertainty in projected ocean acidification increases with ocean depth because of model-simulated differences in ocean circulation that transports anthropogenic CO 2 from the surface to bottom ocean high 2 train +893 AR6_WGI 627 3 To conclude, the forced change in ENSO SST variability is highly uncertain in CMIP5 and CMIP6 models medium 1 train +894 AR6_WGI 629 22 In the Southern Hemisphere the strongest warming over land is to occur, at any given level of global warming, over the subtropical areas of South America, southern Africa and Australia high 2 train +895 AR6_WGI 629 24 Across the globe, in the tropics, subtropics, and mid- to high latitudes, temperatures tend to scale linearly with the level of increase in GSAT and patterns of change are largely scenario independent high 2 train +896 AR6_WGI 631 3 The increases and decreases in precipitation will amplify at higher levels of global warming high 2 train +897 AR6_WGI 631 6 Over the austral-winter rainfall regions of south-western South America, South Africa and Australia, projected decreases in mean annual rainfall show high agreement across models and a strong climate change signal even under 1.5°C of global warming, with further amplification of the signal at higher levels of global warming high 2 train +898 AR6_WGI 632 4 As warming increases, a larger global and land area will experience statistically significant increases or decreases in precipitation medium 1 train +899 AR6_WGI 634 11 Although these climate quantities are not fully reversible, the overshoot scenario results in reduced climate change compared with stabilisation or continued increase in greenhouse gases (Tsutsui et al., 2006; Palter et al., 2018; Tachiiri et al., 2019) high 2 train +900 AR6_WGI 634 16 TCRE remains a valuable concept to assess climate policy goals and how to achieve them but given the non-reversibility of different climate metrics with CO 2 and GSAT reductions, it has limitations associated with evaluating the climate response under overshoot scenarios and CO 2 removal medium 1 train +901 AR6_WGI 634 33 Compared with the differences between the CMIP5 and CMIP6 multi-model ensembles for the same scenario pairs (Table A6 in Tebaldi et al., 2021), the higher ERFs of the SSP scenarios contribute approximately half of the warmer CMIP6 SSP outcomes medium 1 train +902 AR6_WGI 635 18 Surface warming would likewise initially continue under scenarios of decreasing emissions, resulting in a substantial lag between a peak in CO 2 emissions and peak warming high 2 train +903 AR6_WGI 636 7 Among global quantities, emergence of the response to differing CO 2 emissions – representing differences between low- and high-emissions scenarios – is first expected to arise in global mean CO 2 concentrations, about 10 years after emissions pathways have started diverging high 2 train +904 AR6_WGI 637 12 This earlier diagnosed time of emergence compared to Marotzke (2019), while using a similar statistical approach, presumably arose because of the longer-period trends (20 rather than 15 years) and the larger difference between emissions trajectories considered medium 1 train +905 AR6_WGI 637 24 However, the response of many other climate quantities to mitigation would be largely masked by internal variability during the near term, especially on the regional scale high 2 train +906 AR6_WGI 637 25 The mitigation benefits for these quantities would emerge only later during the 21st century high 2 train +907 AR6_WGI 637 26 During the near term, a small fraction of the surface can show cooling under all scenarios assessed here, so near-term cooling at any given location is fully consistent with globally averaged surface warming high 2 train +908 AR6_WGI 638 8 Hence, the potential role that CDR will play in lowering the temperature in high-emissions scenarios is limited medium 1 train +909 AR6_WGI 640 3 The climate system response to the deployment of CDR is expected to be delayed by years (e.g., in temperature, precipitation, sea ice extent) to centuries (e.g., sea level and AMOC) high 2 train +910 AR6_WGI 640 4 The climate response to a sudden and sustained CDR termination would depend on the amount of CDR-induced cooling prior to termination and the rate of background CO 2 emissions at the time of termination high 2 train +911 AR6_WGI 643 24 By appropriately adjusting the amount, latitude, altitude, and timing of the aerosol injection, modelling studies suggest that SAI is conceptually able to achieve some desired combination of radiative forcing and climate response medium 1 train +912 AR6_WGI 644 13 Relative to the high-GHG climate, it is likely that MCB would increase precipitation over tropical land due to the inhomogeneous forcing pattern of MCB over ocean and land medium 1 train +913 AR6_WGI 645 20 For the same amount of global mean cooling, different SRM options would cause different patterns of climate change medium 1 train +914 AR6_WGI 646 28 There is an offset of continued warming following cessation of emissions by continued CO 2 removal by natural sinks high 2 train +915 AR6_WGI 647 24 Changes in climate at 2300 have impacts and commitments beyond this timeframe high 2 train +916 AR6_WGI 648 3 Non-CO 2 forcing and feedbacks remain important by 2300 high 2 train +917 AR6_WGI 649 1 GSAT differences between SSP5-3.4-overshoot and SSP1-2.6 peak during the 21st century but decline to less than about 0.25°C after 2150 medium 1 train +918 AR6_WGI 649 3 GSAT projected for the end of the 23rd century under SSP5-8.5 (likely 6.6°C–14.1°C higher than over the period 1850–1900) overlaps with the range estimated for the Miocene Climatic Optimum (5°C–10°C higher) and Early Eocene Climatic Optimum (10°C–18°C higher), about 15 and 50 million years ago, respectively medium 1 train +919 AR6_WGI 655 11 For SSP1-2.6, such a high-warming storyline implies warming well above rather than well below 2°C high 2 train +920 AR6_WGI 655 12 Irrespective of scenario, high-warming storylines imply changes in many aspects of the climate system that exceed the patterns associated with the best estimate of GSAT changes by up to more than 50% high 2 train +921 AR6_WGI 692 6 The Human Perturbation of the Carbon and Biogeochemical Cycles Global mean concentrations for well-mixed GHGs (CO 2, CH 4 and N 2O) in 2019 correspond to increases of about 47%, 156%, and 23%, respectively, above the levels in 1750 (representative of the pre-industrial era) high 2 train +922 AR6_WGI 692 7 Current atmospheric concentrations of the three GHGs are higher than at any point in the last 800,000 years, and in 2019 reached 409.9 parts per million (ppm) of CO2, 1866.3 parts per billion (ppb) of CH4, and 332.1 ppb of N2O very high 3 test +923 AR6_WGI 692 8 Current CO 2 concentrations in the atmosphere are also unprecedented in the last 2 million years high 2 train +924 AR6_WGI 692 9 In the past 60 million years, there have been periods in Earth’s history when CO 2 concentrations were significantly higher than at present, but multiple lines of evidence show that the rate at which CO 2 has increased in the atmosphere during 1900–2019 is at least 10 times faster than at any other time during the last 800,000 years (high confidence), and 4–5 times faster than during the last 56 million years low 0 train +925 AR6_WGI 692 12 During the last measured decade, global average annual anthropogenic emissions of CO 2, CH 4, and N 2O, reached the highest levels in human history at 10.9 ± 0.9 petagrams of carbon per year (PgC yr–1, 2010–2019), 335–383 teragrams of methane per year (TgCH 4 yr–1, 2008–2017), and 4.2–11.4 teragrams of nitrogen per year (TgN yr–1, 2007–2016), respectively high 2 train +926 AR6_WGI 692 16 The ocean and land sinks of CO 2 have continued to grow over the past six decades in response to increasing anthropogenic CO 2 emissions high 2 train +927 AR6_WGI 692 17 Interannual and decadal variability of the regional and global ocean and land sinks indicate that these sinks are sensitive to climate conditions and therefore to climate change high 2 train +928 AR6_WGI 692 19 However, the effects of these changes are not yet reflected in a weakening trend of the contemporary (1960–2019) ocean sink high 2 train +929 AR6_WGI 692 21 The multi-decadal growth trend in atmospheric CH 4 is dominated by anthropogenic activities (high confidence), and the growth since 2007 is largely driven by emissions from both fossil fuels and agriculture (dominated by livestock) medium 1 train +930 AR6_WGI 692 24 This increase is dominated by anthropogenic emissions, which have increased by 30% between the 1980s and the most recent observational decade (2007–2016) high 2 train +931 AR6_WGI 692 25 Increased use of nitrogen fertilizer and manure contributed to about two-thirds of the increase during the 1980–2016 period, with the fossil fuels/industry, biomass burning, and wastewater accounting for much of the rest high 2 train +932 AR6_WGI 693 7 This is attributed to the effect of larger surface warming in CMIP6 models, which increases ocean stratification and reduces ventilation medium 1 train +933 AR6_WGI 693 12 Despite the wide range of model responses, uncertainty in atmospheric CO 2 by 2100 is dominated by future anthropogenic emissions rather than uncertainties related to carbon–climate feedbacks high 2 train +934 AR6_WGI 693 16 More than half of the latest CMIP6 ESMs include nutrient limitations on the carbon cycle, but these models still project increasing tropical land carbon (medium confidence) and increasing global land carbon high 2 train +935 AR6_WGI 693 22 Under very high emissions scenarios such as SSP5-8.5, ecosystem carbon losses due to warming lead the land to transition from a carbon sink to a source (medium confidence), while the ocean is expected to remain a sink high 2 train +936 AR6_WGI 693 24 In scenarios with moderate net negative CO 2 emissions, and CO 2 concentrations declining during the 21st century (e.g., SSP1-2.6), the land sink transitions to a net source in decades to a few centuries after CO 2 emissions become net negative, while the ocean remains a sink low 0 train +937 AR6_WGI 693 25 Under scenarios with large net negative CO 2 emissions and rapidly declining CO 2 concentrations (e.g., SSP5-3.4-OS (overshoot)), both land and ocean switch from a sink to a transient source during the overshoot period medium 1 train +938 AR6_WGI 694 2 Models project that, over the 21st century, the combined feedback of 0.02–0.09 W m–2 °C–1 is comparable to the effect of a CO 2 release of 5–18 petagrams of carbon equivalent per °C (PgCeq °C–1) low 0 train +939 AR6_WGI 694 4 The probability of crossing uncertain regional thresholds (e.g., high severity fires, forest dieback) increases with climate change high 2 train +940 AR6_WGI 694 5 Possible abrupt changes and tipping points in biogeochemical cycles lead to additional uncertainty in 21st century GHG concentrations, but these are very likely to be smaller than the uncertainty associated with future anthropogenic emissions high 2 train +941 AR6_WGI 694 23 The fraction of CO2 removed that remains out of the atmosphere, a measure of CDR effectiveness, decreases slightly with increasing amount of removal (medium confidence) and decreases strongly if CDR is applied at lower CO 2 concentrations medium 1 test +942 AR6_WGI 694 27 The net effect of this asymmetry on the global surface temperature is poorly constrained due to low agreement between models low 0 train +943 AR6_WGI 695 1 These side effects and associated Earth system feedbacks can decrease carbon uptake and/or change local and regional climate, and in turn limit the CO 2 sequestration and cooling potential of specific CDR methods medium 1 train +944 AR6_WGI 695 2 Deployment of CDR, particularly on land, can also affect water quality and quantity, food production and biodiversity, with consequences for the achievement of related sustainable development goals high 2 train +945 AR6_WGI 695 3 These effects are often highly dependent on local context, management regime, prior land use, and scale of deployment high 2 train +946 AR6_WGI 695 4 A wide range of co-benefits are obtained with methods that seek to restore natural ecosystems or improve soil carbon high 2 train +947 AR6_WGI 695 5 The biogeochemical effects of terminating CDR are expected to be small for most CDR methods medium 1 train +948 AR6_WGI 695 7 SRM acts to cool the planet relative to unmitigated climate change, which would increase the land sink by reducing plant and soil respiration and slow the reduction of ocean carbon uptake due to warming medium 1 train +949 AR6_WGI 695 8 SRM would not counteract or stop ocean acidification high 2 train +950 AR6_WGI 695 9 The sudden and sustained termination of SRM would rapidly increase global warming, with the return of positive and negative effects on the carbon sinks very high 3 train +951 AR6_WGI 710 15 Increasing global net land CO 2 sink since the 1980s is mainly driven by the fertilization effect from rising atmospheric CO 2 concentrations (Schimel et al., 2015; Sitch et al., 2015; Fernández-Martínez et al., 2019; O’Sullivan et al., 2019; Tagesson et al., 2020; Walker et al., 2021) medium 1 train +952 AR6_WGI 713 7 Multiple lines of evidence suggest that WUE has increased in near proportionality to atmospheric CO 2 high 2 train +953 AR6_WGI 713 19 Increased CO 2 concentrations alleviate the effects of water deficits on plant productivity medium 1 train +954 AR6_WGI 715 1 In conclusion, extensive deployment of BECCS and afforestation/reforestation will require larger amounts of freshwater resources than used by the previous vegetation, altering the water cycle at regional scales high 2 train +955 AR6_WGI 715 2 Consequences of high water consumption on downstream uses, biodiversity, and regional climate depend on prior land cover, background climate conditions, and scale of deployment high 2 train +956 AR6_WGI 715 14 Over the industrial era (1750–2019), the total cumulative CO 2 fossil fuel and industry emissions were 445 ± 20 PgC, and the LULUCF flux (= net land-use change in Figure 5.12) was 240 ± 70 PgC medium 1 train +957 AR6_WGI 716 3 The net export of carbon from the terrestrial domain to the open oceans is estimated to be 0.80 PgC yr–1 medium 1 train +958 AR6_WGI 721 1 Although there is evidence that regional human activities and global warming both increase inland water CH 4 emissions (Beaulieu et al., 2019), the increase in the decadal emissions since AR5 (Ciais et al., 2013) rather reflect improvements in the estimate medium 1 train +959 AR6_WGI 723 5 These contrasting lines of evidence suggest that OH changes may have had a small moderating influence on methane growth since 2007 low 0 train +960 AR6_WGI 724 1 Isotopic composition observations and inventory data suggest that concurrent emissions changes from both fossil fuels and agriculture are playing roles in the resumed CH 4 growth since 2007 high 2 train +961 AR6_WGI 724 2 Shorter-term decadal variability is predominantly driven by the influence of El Niño–Southern Oscillation on emissions from wetlands and biomass burning (Cross-Chapter Box 5.2, Figure 2), and loss due to OH variations medium 1 train +962 AR6_WGI 725 4 Several studies using complementary methods indicate that agricultural N 2O emissions have increased by more than 45% since the 1980s high 2 train +963 AR6_WGI 728 3 Supported by multiple studies and extensive observational evidence (Sections 5.2.3.2–5.2.3.4 and Figure 5.17), anthropogenic emissions contributed about 40% (7.3; uncertainty range: 4.2 to 11.4 TgN yr–1) to the total N 2O source in 2007–2016 high 2 train +964 AR6_WGI 728 4 This estimate is larger than in AR5 (WGI, 6.4.3) due to a larger estimated effect of nitrogen deposition on soil N 2O emissions and the explicit consideration of the role of anthropogenic nitrogen in determining inland water and estuary emissions.Based on bottom-up estimates, anthropogenic emissions from agricultural nitrogen use, industry and other indirect effects have increased by 1.7 (1.0 to 2.7) TgN yr–1 between the decades 1980– 1989 and 2007–2016, and are the primary cause of the increase in the total N 2O source high 2 train +965 AR6_WGI 729 12 For North Asia, Europe, Temperate North America and West Asia, the most dominant GHG source is CO 2 high 2 train +966 AR6_WGI 729 15 Persistent emissions of CO 2 are observed for Tropical and South America, northern Africa, and South East Asia medium 1 train +967 AR6_WGI 730 7 Ocean warming decreases the solubility of dissolved oxygen in seawater, and it contributes to about 15% of the dissolved oxygen decrease in the oceans according to estimates based on solubility and the recent SROCC assessment medium 1 train +968 AR6_WGI 731 23 Western Pacific coral records show depleted δ13C trends since the late 19th century that are more prominent since the mid-20th century high 2 train +969 AR6_WGI 731 24 Overall, many of the records show a highly variable seawater pH underlaying strong imprints of internal climate variability (high confidence) and, in most instances, superimposed on a decreasing δ11B trend that is indicative of anthropogenic ocean acidification in recent decades medium 1 train +970 AR6_WGI 733 10 Observations over past decades of basin-wide and global syntheses of ocean interior carbon show that the extent of acidification due to anthropogenic CO 2 invasion tends to diminish with depth very high 3 train +971 AR6_WGI 733 11 The regions of deep convection such as subpolar North Atlantic and Southern Ocean present the deepest acidification detections below 2000 m medium 1 train +972 AR6_WGI 733 25 Its rates at depths are controlled by the ventilation of the ocean interior as well as anthropogenic CO 2 uptake at the surface, thereby diminishing with depth very high 3 train +973 AR6_WGI 735 2 It is concluded that the oxygen content of subsurface ocean is projected to transition to historically unprecedented condition with decline over the 21st century medium 1 train +974 AR6_WGI 735 7 Projected oxygen loss in the ocean is thought to result in an ocean-climate feedback through changes in the natural emissions of GHGs low 0 test +975 AR6_WGI 738 18 New syntheses since AR5 corroborate that the effect of elevated CO 2 on plant growth and ecosystem carbon storage is generally positive high 2 train +976 AR6_WGI 739 27 Recent studies (Katavouta et al., 2018; Jiang et al., 2019; Arora et al., 2020; Rodgers et al., 2020) suggest with medium confidence that the decrease in the ocean CO 2 uptake ratio to anthropogenic CO 2 emissions, under low to no mitigation scenarios over the 21st century, is predominantly attributable to the ocean carbon- concentration feedback through the reduction of the seawater CO 2 buffering capacity, but with contributions from physical drivers such as warming and wind stress (medium confidence) and biological drivers low 0 train +977 AR6_WGI 740 3 Overall, there is medium confidence on three outcomes in the ocean from projected CO 2 uptake under medium to high CO 2 concentration scenarios: (i) a weakening of the buffering capacity, which impacts the airborne fraction via the reduction of the ocean CO 2 buffering capacity due to cumulative ocean CO 2 uptake, which reduces the net ocean CO 2 uptake ratio to anthropogenic CO 2 emissions (Katavouta et al., 2018; Arora et al., 2020; Rodgers et al., 2020); (ii) an amplification of the seasonal cycle of CO 2 variables, which impacts both the ocean sink and ocean acidification (Hauck et al., 2015); (iii) a decrease in the aragonite and calcite saturation levels in the ocean, which negatively impacts the calcification rates of marine organisms high 2 train +978 AR6_WGI 741 7 Other disturbances such as tree mortality will increase across several ecosystems (medium agreement) with decreased vegetation carbon medium 1 train +979 AR6_WGI 742 7 The permafrost region was a historic carbon sink over centuries to millennia high 2 train +980 AR6_WGI 742 14 One study inferred a multi-year net CO2 source for the tundra in Alaska (Commane et al., 2017), which is equivalent to 0.3 PgC yr–1 when scaled up to the northern permafrost region low 0 test +981 AR6_WGI 743 1 There is robust evidence that some CH 4 emissions sources for some regions have increased over the past decades medium 1 train +982 AR6_WGI 743 14 This suggests that large emissions of CH 4 from old carbon sources will not occur in response to future warming medium 1 train +983 AR6_WGI 743 16 Near-surface permafrost is projected to decrease significantly under future global warming scenarios high 2 train +984 AR6_WGI 744 16 In conclusion, thawing terrestrial permafrost will lead to carbon release under a warmer world high 2 train +985 AR6_WGI 749 19 In summary, oceanic and terrestrial carbon sinks are projected to continue to grow with increasing atmospheric concentrations of CO 2, but the fraction of emissions taken up by land and ocean is expected to decline as the CO 2 concentration increases high 2 train +986 AR6_WGI 749 22 Despite the wide range of model responses, uncertainty in atmospheric CO 2 by 2100 is dominated by future anthropogenic emissions rather than carbon–climate feedbacks high 2 train +987 AR6_WGI 753 19 However, given the contemporary estimate for CH 4 from wildfires of no more than 16 TgCH 4 yr–1 (van der Werf et al., 2017; Saunois et al., 2020), this feedback is small, adding no more than 40 ppb to the atmospheric CH 4 by the end of the 21st century medium 1 train +988 AR6_WGI 755 22 Abrupt Change/ Tipping PointKey Region(s)Probability to Occur in the 21st CenturyMaximum CO 2 or CH 4 Release in the 21st CenturyPrincipal Development Time ScaleMaximum CO 2 or CH 4 Rate of Change Over the 21st Century(Ir)reversibility Tropical forests dieback (Section 5.4.9.1.1)Amazon watershed Low<200 PgC as CO 2 (Section 5.4.9.1.1; medium confidence)Multi-decadal CO 2: <0.5 ppm yr–1Irreversible at multi-decadal scale (medium confidence) Boreal forests dieback (Sections 5.4.9.1.1, 5.4.3.2)Boreal Eurasia and North AmericaLow<27 Pg (Section 5.4.9.1.2; medium confidence)Multi-decadal Small (low confidence)Irreversible at multi-decadal scale (medium confidence) Biogenic emissions from permafrost thaw (Section 5.4.9.1.2)Pan-Arctic Highup to 240 PgC of CO 2 and up to 5300 Tg of CH 4 (Section 5.4.8.1.2; low confidence)Multi-decadalCO 2: ≤1 ppm yr–1 CH 4: ≤10 ppb yr–1Irreversible at centennial time scales high 2 train +989 AR6_WGI 756 12 This implies an upper limit to the release of tropical land carbon of <200 PgC over the 21st century (assuming tropical warming of <4°C , and no CO 2-fertilization), which translates to dCO 2/dt <0.5 ppm yr–1.Boreal forest dieback is not expected to change the atmospheric CO 2 concentration substantially because forest loss at the south is partly compensated by: (i) temperate forest invasion into previously boreal areas; and (ii) boreal forest gain at the north (Friend et al., 2014; Kicklighter et al., 2014; Schaphoff et al., 2016) medium 1 train +990 AR6_WGI 757 8 However, these are very likely to be small compared to the uncertainty associated with future anthropogenic emissions high 2 train +991 AR6_WGI 759 18 Monteiro (South Africa), Jean-Baptiste Sallée (France), Piers Foster (United Kingdom), Baylor Fox-Kemper (United States of America), Helen T. Hewitt (United Kingdom), Masao Ishii (Japan), Joeri Rogelj (United Kingdom/Belgium), Kirsten Zickfeld (Canada/Germany) Context In the past 60 years, the ocean has taken up and stored 23 ± 5% of anthropogenic carbon emissions medium 1 train +992 AR6_WGI 760 10 These processes are typically simplified into widely verified expressions that link the flux to wind stress, the solubility and the gradient across the air–sea interface medium 1 train +993 AR6_WGI 760 13 The role of the biological carbon pump in influencing the ocean sink of anthropogenic carbon into the ocean interior is assessed to be minimal during the historical period, but this may change, particularly in regional contexts, by 2100 medium 1 train +994 AR6_WGI 760 14 Its role is important in the natural or pre- industrial carbon cycle medium 1 train +995 AR6_WGI 761 11 In contrast, in the Southern Ocean, the future 21st century projected increase in upper ocean overturning circulation low 0 train +996 AR6_WGI 762 1 In this way, the processes of the ocean carbon-heat nexus help understand the quasi-linear and path independence of properties of TCRE, which forms the basis for the zero emissions commitment (ZEC; Section 5.5) medium 1 train +997 AR6_WGI 762 3 Increasing ocean warming and stratification may decrease exchanges between the surface and subsurface ocean, which could reduce the path independence of TCRE, though this effect can be partially counterbalanced regionally by increasing circulation associated with increasing winds low 0 test +998 AR6_WGI 775 15 Whether the transition to source occurs at all, the timing of the transition and the magnitude of the CO 2 source are determined by the magnitude of the removal and the rate and amount of net CO 2 emissions prior to emissions becoming net negative medium 1 train +999 AR6_WGI 778 16 In response to increasing risks to permanence of carbon stocks of some types of afforestation practices and the competition for land, there has been an increasing recognition that secondary forest regrowth and restoration of degraded forests and non-forest ecosystems can play a large role in carbon sequestration high 2 train +1000 AR6_WGI 778 26 Forest- based methods can either raise or lower N 2O emissions, depending on tree species, previous land use, soil type and climatic factors low 0 train +1001 AR6_WGI 778 28 Afforestation will decrease biodiversity if native species are replaced by monocultures high 2 train +1002 AR6_WGI 779 8 Some soil carbon sequestration methods, such as cover crops and crop diversity, can increase biodiversity medium 1 train +1003 AR6_WGI 779 11 Biochar application improves many soil qualities and increases crop yield (medium confidence) (Ye et al., 2020; SRCCL, Chapter 4.9.5), particularly in already degraded or weathered soils (Woolf et al., 2010; Lorenz and Lal, 2014; Jeffery et al., 2016), increases soil water holding capacity medium 1 train +1004 AR6_WGI 779 12 Fischer et al., 2019; Verheijen et al., 2019) and evapotranspiration low 0 train +1005 AR6_WGI 779 14 The use of biochar reduces nutrient losses low 0 test +1006 AR6_WGI 779 15 Biochar addition may decrease methane (CH 4) emissions in inundated and acid soils such as rice fields low 0 train +1007 AR6_WGI 779 16 In non-inundated, neutral soils, CH 4 uptake from the atmosphere is suppressed after biochar application (low confidence) (Jeffery et al., 2016), and soil N2O emissions decline medium 1 train +1008 AR6_WGI 779 21 High water level and anoxic conditions are prerequisites for restoring by returning drained and/or degraded peatlands back to their natural state as CO 2 sinks, but restoration also results in enhanced CH 4 emissions which are similar to or higher than the pre-drainage fluxes high 2 train +1009 AR6_WGI 779 23 Rewetting drained peatlands will decrease N2O emissions medium 1 train +1010 AR6_WGI 779 25 Peatland restoration can also recover much of the original biodiversity medium 1 train +1011 AR6_WGI 779 33 BECCS has several trade- offs to deal with, including possible threats to water supply and soil nutrient deficiencies medium 1 train +1012 AR6_WGI 780 1 Highest co-benefits are obtained with methods that seek to restore natural ecosystems and improve soil carbon sequestration (Figure 5.36) while highest trade- off possibilities (symmetry with the highest co-benefits) occur for reforestation or afforestation with monocultures and BECCS, again with strong dependence on scale and context medium 1 train +1013 AR6_WGI 780 15 While coastal habitat restoration potentially provides significant mitigation of national emissions for some countries (Taillardat et al., 2018; Serrano et al., 2019), the global sequestration potential of blue carbon approaches is <0.02 PgC yr–1 medium 1 train +1014 AR6_WGI 780 20 EW can also contribute to freshwater salinization as a result of increased salt inputs and cation exchange in watersheds, and so adversely affecting drinking water quality low 0 train +1015 AR6_WGI 780 26 Ocean alkalinization ameliorates surface ocean acidification high 2 train +1016 AR6_WGI 784 21 Despite the large uncertainty in modelled NPP response, existing modelling studies consistently show that SRM would increase the global land carbon sink relative to a high-CO 2 world without SRM high 2 train +1017 AR6_WGI 784 22 Based on available evidence, SRM with elevated CO 2 would increase global mean NPP and carbon storage on land relative to an unperturbed climate, mainly because of CO 2 fertilization of photosynthesis high 2 train +1018 AR6_WGI 784 34 As a result of enhanced global carbon uptake, SRM would reduce the burden of atmospheric CO 2 high 2 train +1019 AR6_WGI 785 9 Compared to a high-CO 2 world without SRM, SRM would enhance the net uptake of CO 2 by the terrestrial biosphere and ocean, thus acting to reduce atmospheric CO 2 high 2 train +1020 AR6_WGI 835 8 Recent Evolution in Short-lived Climate Forcer (SLCF) Emissions and Abundances Over the last decade (2010–2019), strong shifts in the geographical distribution of emissions have led to changes in atmospheric abundances of highly variable SLCFs high 2 train +1021 AR6_WGI 835 9 Evidence from satellite and surface observations shows strong regional variations in trends of ozone (O 3), aerosols and their precursors high 2 train +1022 AR6_WGI 835 10 In particular, tropospheric columns of nitrogen dioxide (NO 2) and sulphur dioxide (SO 2) continued to decline over North America and Europe (high confidence), and to increase over Southern Asia (medium confidence), but have declined over Eastern Asia high 2 train +1023 AR6_WGI 835 11 Global carbon monoxide (CO) abundance has continued to decline high 2 train +1024 AR6_WGI 835 12 The concentrations of hydrofluorocarbons (HFCs) are increasing high 2 train +1025 AR6_WGI 835 21 Radiative forcings induced by aerosol changes lead to both local and remote temperature responses high 2 train +1026 AR6_WGI 835 22 The temperature response preserves the south to north gradient of the aerosol ERF – hemispherical asymmetry – but is more uniform with latitude and is strongly amplified towards the Arctic medium 1 train +1027 AR6_WGI 835 25 The spatial and temporal distribution of the net aerosol ERF from 1850 to 2014 is highly heterogeneous, with stronger magnitudes in the Northern Hemisphere high 2 test +1028 AR6_WGI 835 27 Near its maximum, the response slows down but will then take centuries to reach equilibrium high 2 train +1029 AR6_WGI 835 28 For SLCFs with longer lifetimes (e.g., a decade), a delay equivalent to their lifetimes is appended to the delay due to thermal inertia high 2 train +1030 AR6_WGI 835 30 The net global emissions‑based ERF of NO x is negative and that of non‑methane volatile organic compounds (NMVOCs) is positive, in agreement with the AR5 Assessment high 2 train +1031 AR6_WGI 835 31 For methane, the emissions‑based ERF is twice as high as the abundance‑based ERF high 2 train +1032 AR6_WGI 836 1 Over the 1750–2019 period, the contributions from the emitted compounds to changes in global surface air temperature (GSAT) broadly match their contributions to the ERF high 2 train +1033 AR6_WGI 836 5 As such, these effects are assessed to be of second order in comparison to the direct CO 2 forcing (high confidence), but effects of ozone on terrestrial vegetation could add a substantial (positive) forcing compared with the direct ozone forcing low 0 train +1034 AR6_WGI 836 10 A warmer climate is expected to reduce surface ozone in regions remote from pollution sources high 2 train +1035 AR6_WGI 836 11 Future climate change is expected to have mixed effects, positive or negative, with an overall low effect, on global surface PM and more generally on the aerosol global burden (medium confidence), but stronger effects are not excluded in regions prone to specific meteorological conditions low 0 train +1036 AR6_WGI 836 17 Under the SSP3‑7.0 scenario, PM levels are projected to increase until 2050 over large parts of Asia, and surface ozone pollution is projected to worsen over all continental areas through 2100 high 2 test +1037 AR6_WGI 836 18 Without climate change mitigation but with stringent air pollution control (SSP5‑8.5), PM levels decline through 2100, but high methane levels hamper the decline in global surface ozone at least until 2080 high 2 train +1038 AR6_WGI 837 6 Sectors producing the largest SLCF-induced warming are those dominated by methane emissions: fossil fuel production and distribution, agriculture and waste management high 2 train +1039 AR6_WGI 837 7 On these time scales, SLCFs with cooling effects can significantly mask the CO 2 warming in the case of fossil fuel combustion for energy and land transportation, or completely offset the CO 2 warming and lead to an overall net cooling in the case of industry and maritime shipping (prior to the implementation of the revised fuel‑sulphur limit policy for shipping in 2020) medium 1 train +1040 AR6_WGI 837 8 Ten years after a one‑year pulse of present‑day aviation emissions, SLCFs induce strong but short‑lived warming contributions to the GSAT response medium 1 train +1041 AR6_WGI 837 10 Consequently, on time scales longer than about 30 years, the net long-term global temperature effects of sectors and regions are dominated by CO 2 high 2 train +1042 AR6_WGI 837 13 About 30 years or more after a one‑year emission pulse occurs, the sectors contributing the most to global warming are industry, fossil fuel combustion for energy and land transportation, essentially through CO2 high 2 test +1043 AR6_WGI 837 14 Current emissions of SLCFs, CO 2 and N 2O from Eastern Asia and North America are the largest regional contributors to additional net future warming on both short (medium confidence) and long time scales high 2 train +1044 AR6_WGI 837 17 Energy and industry are important PM 2.5 contributors in most regions, except Africa high 2 train +1045 AR6_WGI 837 20 Further improvements in the efficiency of refrigeration and air ‑conditioning equipment during the transition to low‑global‑ warming‑potential refrigerants would bring additional greenhouse gas reductions medium 1 train +1046 AR6_WGI 837 22 This warming is stable after 2040 in scenarios leading to lower global air pollution as long as methane emissions are also mitigated, but the overall warming induced by SLCF changes is higher in scenarios in which air quality continues to deteriorate (induced by growing fossil fuel use and limited air pollution control) high 2 train +1047 AR6_WGI 837 26 Sustained methane mitigation, wherever it occurs, stands out as an option that combines near‑ and long‑term gains on surface temperature (high confidence) and leads to air ‑quality benefits by reducing surface ozone levels globally high 2 train +1048 AR6_WGI 838 1 Additional methane and BC mitigation would contribute to offsetting the additional warming associated with SO 2 reductions that would accompany decarbonization high 2 train +1049 AR6_WGI 838 2 Strong air pollution control as well as strong climate change mitigation, implemented separately, lead to large reductions in exposure to air pollution by the end of the century high 2 train +1050 AR6_WGI 838 3 Implementation of air pollution controls, relying on the deployment of existing technologies, leads more rapidly to air quality benefits than climate change mitigation high 2 train +1051 AR6_WGI 838 4 However, in both cases, significant parts of the population are projected to remain exposed to air pollution exceeding the WHO guidelines high 2 train +1052 AR6_WGI 838 8 Global anthropogenic NO x emissions decreased by a maximum of about 35% in April 2020 medium 1 train +1053 AR6_WGI 838 10 Global fossil CO 2 emissions decreased by 7% (with a range of 5.8–13.0%) in 2020 relative to 2019, largely due to reduced emissions from the transportation sector medium 1 train +1054 AR6_WGI 838 12 Consistent with this small net radiative forcing, and against a large component of internal variability, Earth system model simulations show no detectable effect on global or regional surface temperature or precipitation high 2 train +1055 AR6_WGI 841 30 The SRCCL concluded that: (i) there is no agreement about the direction of future changes in mineral dust emissions; (ii) fossil fuel and biomass burning, and secondary organic aerosols (SOA) from natural BVOC emissions are the main global sources of carbonaceous aerosols whose emissions are expected to increase in the near future due to possible increases in open biomass burning and increase in SOA from oxidation of BVOCs medium 1 train +1056 AR6_WGI 843 27 However, Chinese emissions declined by nearly 70% between about 2006 and 2017 high 2 train +1057 AR6_WGI 843 32 Since about 2011, global NO x emissions appear to have stabilized or slightly declined medium 1 train +1058 AR6_WGI 843 34 Liu et al., 2016; Miyazaki et al., 2017; Silver et al., 2018): a strong decline of NO 2 column over eastern China high 2 train +1059 AR6_WGI 844 3 Efforts to control transport emissions (i.e., increasing stringency of vehicle emissions limits) were largely offset by the fast growth of emissions from chemical industries and solvent use, as well as from fossil fuel production and distribution, resulting in continued growth of global anthropogenic NMVOC emissions since 1900 high 2 train +1060 AR6_WGI 844 5 Increasing (since 2008) oil‑ and gas‑extraction activities in North America lead to a strong growth of NMVOC emissions high 2 train +1061 AR6_WGI 844 13 Emissions of carbonaceous aerosols (BC, OC) have been steadily increasing and their emissions have almost doubled since 1950 medium 1 train +1062 AR6_WGI 844 14 Before 1950, North America and Europe contributed about half of the global total but successful introduction of diesel particulate filters on road vehicles (Fiebig et al., 2014; Robinson et al., 2015; Klimont et al., 2017a) and declining reliance on solid fuels for heating brought in large reductions high 2 train +1063 AR6_WGI 844 15 Currently, global carbonaceous aerosol emissions originate primarily from Asia and Africa (Bond et al., 2013; Hoesly et al., 2018; Elguindi et al., 2020; McDuffie et al., 2020), representing about 80% of the global total high 2 train +1064 AR6_WGI 844 20 Overall, a factor two uncertainty in global estimates of BC and OC emissions remains, with post‑ 2005 emissions overestimated in Asia (high confidence) and Africa medium 1 train +1065 AR6_WGI 844 21 Bottom‑up global emissions estimates of methane (Lamarque et al., 2010; Hoesly et al., 2018; Janssens‑Maenhout et al., 2019; Höglund‑Isaksson et al., 2020) for the last two decades are higher than top‑down assessments (e.g., Saunois et al., 2016, 2020) but trends from the two methods are similar and indicate continued growth high 2 train +1066 AR6_WGI 844 24 NH 3 emissions are estimated to have grown strongly since 1850, especially since 1950, driven by continuously increasing livestock production, widespread application of mineral nitrogen fertilizers, and lack of action to control ammonia high 2 train +1067 AR6_WGI 847 25 Overall, we assess that historical global isoprene emissions declined between the pre‑industrial period and the present day by 10–25% low 0 train +1068 AR6_WGI 854 12 This increase is assessed to be 109 ± 25 Tg medium 1 train +1069 AR6_WGI 855 16 In summary, global tropospheric NO x abundance has increased from 1850–2015 high 2 train +1070 AR6_WGI 855 19 NO 2 trends have reversed (declining) over China beginning in 2012 and NO 2 has increased over Southern Asia by 50% since 2005 medium 1 train +1071 AR6_WGI 858 7 In summary, after a decline between 1980 and 2008, abundances of light NMVOCs have increased again over the Northern Hemisphere due to the extraction of oil and gas in North America high 2 train +1072 AR6_WGI 858 8 Trends in satellite HCHO observations, used as a proxy of anthropogenic NMVOC over industrialized areas, show a significant positive trend over eastern China high 2 train +1073 AR6_WGI 864 27 That would give BrC a burden similar to that of BC low 0 train +1074 AR6_WGI 867 6 In summary, global mean tropospheric OH does not show a significant trend from 1850 up to around 1980 low 0 train +1075 AR6_WGI 869 4 In summary, the spatial and temporal distribution of the net aerosol ERF from 1850–2014 is highly heterogeneous high 2 train +1076 AR6_WGI 869 5 Globally, there has been a shift from increase to decrease of the negative net aerosol ERF driven by trends in aerosol and their precursor emissions high 2 train +1077 AR6_WGI 871 10 In summary, emissions of SLCFs, especially methane, NO x and SO 2, have substantial effects on effective radiative forcing (ERF) high 2 train +1078 AR6_WGI 871 11 The net global emissions‑based ERF of NO x is negative and that of NMVOCs is positive, in agreement with the AR5 assessment high 2 train +1079 AR6_WGI 871 12 For methane, the emissions‑based ERF is twice as high as the abundance‑based ERF high 2 train +1080 AR6_WGI 871 13 SO 2 emissions make the dominant contribution to the ERF associated with the aerosol–cloud interaction high 2 train +1081 AR6_WGI 871 14 The contributions from the emitted compounds to GSAT broadly follow their contributions to the ERF high 2 train +1082 AR6_WGI 871 21 The overall effect of surface cooling from anthropogenic aerosols is to reduce global precipitation and alter large‑ scale atmospheric circulation patterns high 2 train +1083 AR6_WGI 872 2 The zonal‑mean temperature response is negative at all latitudes high 2 train +1084 AR6_WGI 872 7 The asymmetric historical radiative forcing due to aerosols led to a southward shift in the tropical rain belt (high confidence) and contributed to the Sahel drought from the 1970s to the 1980s high 2 train +1085 AR6_WGI 872 8 Furthermore, the asymmetry of the forcing led to contrasting effects in monsoon precipitation changes over West Africa, Southern Asia and Eastern Asia over much of the mid‑20th century due to GHG‑induced precipitation increases counteracted by anthropogenic aerosol‑ induced decreases high 2 train +1086 AR6_WGI 873 1 In summary, over the historical period, changes in aerosols and their ERF have primarily contributed to cooling, partly masking the human‑induced warming high 2 train +1087 AR6_WGI 873 2 Radiative forcings induced by aerosol changes lead to both local and remote changes in temperature high 2 train +1088 AR6_WGI 873 3 The temperature response preserves hemispheric asymmetry of the ERF but is more latitudinally uniform with strong amplification of the temperature response towards the Arctic medium 1 train +1089 AR6_WGI 873 11 The new modelling studies tend to focus on ozone effects on plant productivity rather than the land carbon storage and agree that ozone‑induced gross‑primary productivity (GPP) losses are largest today in the eastern USA, Europe and eastern China, ranging from 5–20% on the regional scale low 0 train +1090 AR6_WGI 873 23 At large scales the dominant effect of aerosols on the carbon cycle is likely a global cooling effect of the climate medium 1 train +1091 AR6_WGI 873 25 In summary, reactive nitrogen, ozone and aerosols affect terrestrial vegetation and the carbon cycle through deposition and effects on large‑scale radiation high 2 train +1092 AR6_WGI 873 26 As such, we assess the effects to be of second order in comparison to the direct CO 2 forcing (high confidence) but, at least for ozone, it could add a substantial (positive) forcing compared with its direct forcing low 0 train +1093 AR6_WGI 877 6 Modelling studies have been published with more sophisticated treatment of SRM since AR5, but the uncertainties, such as cloud–aerosol radiation interactions, remain large high 2 train +1094 AR6_WGI 879 7 Overall, warmer climate is expected to reduce surface ozone in unpolluted regions as a result of greater water vapour abundance accelerating ozone chemical loss high 2 train +1095 AR6_WGI 879 10 The response of surface ozone to future climate change through stratosphere–troposphere exchange, soil NO x emissions and wildfires is positive medium 1 train +1096 AR6_WGI 881 21 For SLCFs with short lifetime (e.g., months), the response in surface temperature occurs strongly as soon as a sustained change in emissions is implemented and continues to grow for a few years, primarily due to thermal inertia in the climate system high 2 test +1097 AR6_WGI 881 22 Near its maximum, the response slows down but will then take centuries to reach equilibrium high 2 train +1098 AR6_WGI 881 23 For SLCFs with longer lifetimes (e.g., a decade), a delay equivalent to their lifetimes comes in addition to the delay due the thermal inertia high 2 train +1099 AR6_WGI 882 21 Residential sector emissions are an important source of indoor and outdoor air pollution in Asia and globally high 2 train +1100 AR6_WGI 883 25 In summary, the net aviation ERF is assessed to be +0.1 W m–2 (±0.045) for the year 2018 low 0 train +1101 AR6_WGI 883 27 The climate response to SLCF‑related aviation terms exhibits substantial spatio‑temporal heterogeneity in characteristics high 2 train +1102 AR6_WGI 883 28 Overall, cirrus and contrail cirrus warming, as well as NO x‑induced ozone increase, induce strong but short‑lived warming contributions to the GSAT response 10 years after a one‑year pulse of present‑day aviation emissions (medium confidence), while CO 2 both gives a warming effect in the near term and dominates the long‑term warming impact high 2 train +1103 AR6_WGI 883 39 In summary, a year’s worth of present‑day global shipping emissions (i.e., without the implementation of the 2020 clean fuel standards) cause a net global cooling (–0.0024 ± 0.0025°C) on 10–20 year time horizons high 2 train +1104 AR6_WGI 884 12 In summary, the present‑day global land‑based transport pulse emissions cause a net global warming on all time scales (high confidence) and are detrimental to air quality high 2 train +1105 AR6_WGI 884 29 By far the largest 10‑year GSAT effects are from the energy production (fossil fuel mining and distribution), agriculture and waste management sectors high 2 train +1106 AR6_WGI 884 31 On the 10‑year time horizon, other net warming sectors are residential fossil fuel and energy combustion (dominated by CO 2) and aviation and residential biofuel (dominated by SLCFs and cloud) medium 1 train +1107 AR6_WGI 885 5 In terms of source regions, the largest contributions to net short‑ term warming are caused by emissions in Eastern Asia, Latin America and North America, followed by Africa, Eastern Europe, West‑Central Asia and South East Asia medium 1 train +1108 AR6_WGI 885 9 Overall, the global sectors that contribute the largest warming on short time scales are the methane‑dominated sources, that is energy production (fossil fuel mining and distribution), and agriculture and waste management high 2 train +1109 AR6_WGI 885 10 On short time scales, other net warming sectors are residential fossil fuel and energy combustion (dominated by CO 2), and aviation and residential biofuel (dominated by SLCFs) medium 1 train +1110 AR6_WGI 885 11 On short time scales, global emissions from industry and shipping cause a net cooling effect despite a considerable warming from CO 2 emissions high 2 train +1111 AR6_WGI 885 12 On longer time horizons, the sectors that contribute the largest warming are energy combustion and industry due to the large CO 2 emissions high 2 train +1112 AR6_WGI 885 26 Residential and commercial cooking and heating are among the most important anthropogenic sources of ambient PM 2.5, except in the Middle East and Asia‑Pacific Developed (high confidence) and agriculture is the dominant source in Europe and North America medium 1 train +1113 AR6_WGI 885 27 Energy and industry are important PM 2.5 contributors in most regions, except Africa high 2 train +1114 AR6_WGI 891 1 Additional policies (e.g., access to clean energy, waste management) envisaged to attain SDGs bring complementary SLCF reduction high 2 train +1115 AR6_WGI 891 2 Sustained methane mitigation, wherever it occurs, stands out as an option that combines near ‑ and long‑term gains on surface temperature (high confidence) and leads to an air pollution benefit by reducing ozone levels globally high 2 train +1116 AR6_WGI 891 21 Emissions Global fossil CO 2 emissions are estimated to have declined by 7% medium 1 train +1117 AR6_WGI 892 8 Anthropogenic NO x emissions, which are largely from the transport sector, are estimated to have decreased by a maximum of 35% in April medium 1 train +1118 AR6_WGI 892 18 Except for ozone, temporary improvement of air quality during lockdown periods was observed in most regions of the world (high confidence), resulting from a combination of interannual meteorological variability and the impact of COVID‑19 containment measures high 2 train +1119 AR6_WGI 892 19 Estimated air pollution reductions associated with lockdown periods are lower than what can be expected from integrated mitigation policy leading to lasting reductions medium 1 train +1120 AR6_WGI 893 17 The range in peak global mean ERF in spring 2020 was [0.025 to 0.2] W m–2 medium 1 train +1121 AR6_WGI 893 18 By the end of 2020, the ERF was at half the peak value medium 1 train +1122 AR6_WGI 893 19 Climate responses Changes in atmospheric composition due to COVID‑19 emissions reductions are not thought to have caused a detectable change in global temperature or rainfall in 2020 high 2 train +1123 AR6_WGI 894 22 Simulated GSAT and rainfall changes are unlikely to be detectable in observations high 2 train +1124 AR6_WGI 898 16 Consequently, SSPs span a wider range of SLCF emissions than considered in the RCPs, better covering the diversity of future options in air pollution management and SLCF‑induced climate effects high 2 train +1125 AR6_WGI 899 1 However, this decrease is twice as large when methane emissions are abated simultaneously (SSP3‑7.0‑lowSLCF‑lowCH 4), underlying the importance of methane emissions reduction as an important lever to reduce ozone pollution high 2 train +1126 AR6_WGI 900 1 In the latter case, PM levels are estimated to increase until 2050 over large parts of Asia and surface ozone pollution worsens over all continental areas throughout the whole century high 2 train +1127 AR6_WGI 900 2 In scenarios without climate change mitigation but with strong air pollution control (SSP5‑8.5), high methane levels hamper the decline in global surface ozone in the near term and only PM levels decrease high 2 train +1128 AR6_WGI 904 29 Furthermore, under SSP5‑8.5, HFCs induce a warming of 0.06°C with a very likely range of [0.04 to 0.08] °C in 2050 and 0.2 [0.1 to 0.3] °C by the end of the 21st century, relative to 2019, while under SSP1‑2.6, warming due to HFCs is negligible (below 0.02°C) high 2 train +1129 AR6_WGI 905 2 Methane mitigation that also reduces tropospheric ozone, stands out as an option that combines near‑ and long‑term gains on surface temperature high 2 train +1130 AR6_WGI 905 14 The assessment shows that both strong air pollution control and strong climate change mitigation, implemented independently, lead to a large reduction of exposure to PM 2.5 and ozone by the end of the century high 2 train +1131 AR6_WGI 905 15 However, implementation of air pollution control, relying on the deployment of existing technologies, leads to benefits more rapidly than climate change mitigation high 2 train +1132 AR6_WGI 905 16 Notably, under the underlying SSP3 context, significant parts of the population remain exposed to air quality exceeding the WHO guidelines for PM 2.5 over the whole century high 2 train +1133 AR6_WGI 908 1 In summary, the warming induced by SLCF changes is stable after 2040 in the WGI core set of SSP scenarios associated with lower global air pollution as long as methane emissions are also mitigated, but the overall warming induced by SLCF changes is higher in scenarios in which air quality continues to deteriorate (caused by growing fossil fuel use and limited air pollution control) high 2 train +1134 AR6_WGI 908 5 Sustained methane mitigation, wherever it occurs, stands out as an option that combines near ‑ and long‑term gains on surface temperature (high confidence) and leads to air pollution benefits by reducing surface ozone level globally high 2 train +1135 AR6_WGI 908 6 Strong air pollution control as well as strong climate change mitigation, implemented independently, lead to large reduction of the exposure to air pollution by the end of the century high 2 train +1136 AR6_WGI 908 7 Implementation of air pollution control, relying on the deployment of existing technologies, leads more rapidly to air ‑quality benefits than climate change mitigation which requires systemic changes but, in both cases, significant parts of the population remain exposed to air pollution exceeding the WHO guidelines high 2 train +1137 AR6_WGI 941 13 Energy will continue to accumulate in the Earth system until at least the end of the 21st century, even under strong mitigation scenarios, and will primarily be observed through ocean warming and associated with continued sea level rise through thermal expansion high 2 train +1138 AR6_WGI 941 19 More comprehensive analysis of inventory components and cross-validation of global heating rates from satellite and in situ observations lead to a strengthened assessment relative to AR5 high 2 train +1139 AR6_WGI 941 23 Both estimates are consistent with an independent observation-based assessment of the global energy increase of 284 [96 to 471] ZJ, (very likely range) expressed relative to the estimated 1850–1900 Earth energy imbalance high 2 train +1140 AR6_WGI 941 25 These trends are neither a local phenomenon nor a measurement artefact high 2 train +1141 AR6_WGI 941 26 Multi-decadal variation in anthropogenic aerosol emissions are thought to be a major contributor medium 1 train +1142 AR6_WGI 941 27 The downward and upward thermal radiation at the surface has increased in recent decades, in line with increased greenhouse gas concentrations and associated surface and atmospheric warming and moistening medium 1 train +1143 AR6_WGI 941 31 For carbon dioxide, the adjustments include the physiological effects on vegetation high 2 train +1144 AR6_WGI 942 3 The 0.59 W m–2 increase in ERF from greenhouse gases is partly offset by a better-constrained assessment of total aerosol ERF that is more strongly negative than in AR5, based on multiple lines of evidence high 2 train +1145 AR6_WGI 942 4 Changes in surface reflectance from land-use change, deposition of light-absorbing particles on ice and snow, and contrails and aviation-induced cirrus have also contributed to the total anthropogenic ERF over the industrial era, with –0.20 [–0.30 to –0.10] W m–2 (medium confidence), +0.08 [0 to 0.18] W m–2 (low confidence) and +0.06 [0.02 to 0.10] W m–2 low 0 train +1146 AR6_WGI 942 10 The estimated ERF for methane has slightly increased due to a combination of increases from improved spectroscopic treatments being somewhat offset by accounting for adjustments high 2 train +1147 AR6_WGI 942 12 The ERF due to aerosol–cloud interactions (ERFaci) contributes most to the magnitude of the total aerosol ERF (high confidence) and is assessed to be –1.0 [–1.7 to –0.3] W m–2 (medium confidence), with the remainder due to aerosol–radiation interactions (ERFari), assessed to be –0.3 [–0.6 to 0.0] W m–2 medium 1 train +1148 AR6_WGI 942 17 It is also assessed to be smaller in magnitude at –1.1 [–1.7 to –0.4] W m–2, primarily due to recent emissions changes medium 1 train +1149 AR6_WGI 942 22 A net negative cloud feedback is very unlikely high 2 train +1150 AR6_WGI 942 27 Furthermore, on long time scales the ice-sheet feedback parameter is very likely positive, promoting additional warming on millennial time scales as ice sheets come into equilibrium with the forcing high 2 train +1151 AR6_WGI 942 29 This new understanding, along with updated estimates of historical temperature change, ERF, and Earth’s energy imbalance, reconciles previously disparate ECS estimates high 2 train +1152 AR6_WGI 942 31 Warming over the instrumental record provides robust constraints on the lower end of the ECS range high 2 train +1153 AR6_WGI 943 7 These higher ECS and TCR values can, in some models, be traced to changes in extra-tropical cloud feedbacks that have emerged from efforts to reduce biases in these clouds compared to satellite observations medium 1 train +1154 AR6_WGI 943 11 The CMIP models with the highest ECS and TCR values provide insights into low-likelihood, high-impact outcomes, which cannot be excluded based on currently available evidence high 2 train +1155 AR6_WGI 943 14 The calculated GSAT change is composed of a well-mixed greenhouse gas warming of 1.58 [1.17 to 2.17] °C (high confidence), a warming from ozone changes of 0.23 [0.11 to 0.39] °C (high confidence), a cooling of –0.50 [–0.22 to –0.96] °C from aerosol effects (medium confidence), and a –0.06 [–0.15 to +0.01] °C contribution from surface reflectance changes from land-use change and light-absorbing particles on ice and snow medium 1 train +1156 AR6_WGI 943 15 Changes in solar and volcanic activity are assessed to have together contributed a small change of –0.02 [–0.06 to +0.02] °C since 1750 medium 1 train +1157 AR6_WGI 943 18 Global ocean heat uptake is a smaller source of uncertainty in centennial-time scale surface warming high 2 train +1158 AR6_WGI 943 22 When used for multi-scenario experiments, calibrated physically based emulators can adequately reflect assessments regarding future GSAT from Earth system models and/or other lines of evidence high 2 train +1159 AR6_WGI 943 24 The causes of this polar amplification are well understood, and the evidence is stronger than at the time of AR5, supported by better agreement between modelled and observed polar amplification during warm paleo time periods high 2 train +1160 AR6_WGI 943 26 The rate of Arctic surface warming will continue to exceed the global average over this century high 2 train +1161 AR6_WGI 943 31 Metrics for methane from fossil fuel sources account for the extra fossil CO 2 that these emissions contribute to the atmosphere and so have slightly higher emissions metric values than those from biogenic sources high 2 train +1162 AR6_WGI 944 1 By contrast, if emissions are weighted by their 100-year GWP or GTP values, different multi-gas emissions pathways with the same aggregated CO 2 equivalent emissions rarely lead to the same estimated temperature outcome high 2 train +1163 AR6_WGI 944 5 In contrast, reaching net zero GHG emissions when quantified using new emissions metrics such as CGTP or GWP* would lead to approximate temperature stabilization high 2 train +1164 AR6_WGI 951 31 In summary, since AR5, the magnitudes of the global mean energy budget components have been quantified more accurately, not only at the TOA, but also at the Earth’s surface, where independent estimates of the radiative components have converged high 2 train +1165 AR6_WGI 953 4 In summary, variations in the energy exchange between Earth and space can be accurately tracked since the advent of improved observations since the year 2000 (high confidence), while reconstructions indicate that the Earth’s energy imbalance was larger in the 2000s than in the 1985–1999 period high 2 train +1166 AR6_WGI 953 34 The assessed changes in the global energy inventory (Box 7.2, Figure 1, and Table 7.1) yields an average value for Earth’s energy imbalance (N in Box 7.1, Equation 7.1) of 0.57 [0.43 to 0.72] W m–2 for the period 1971–2018, expressed relative to Earth’s surface area high 2 train +1167 AR6_WGI 953 35 The estimates for the periods 1993–2018 and 2006–2018 yield substantially larger values of 0.72 [0.55 to 0.89] W m–2 and 0.79 [0.52 to 1.06] W m–2, respectively, consistent with the increased radiative forcing from GHGs high 2 train +1168 AR6_WGI 954 1 To put these numbers in context, the 2006–2018 average Earth system heating is equivalent to approximately 20 times the annual rate of global energy consumption in 2018.1 Consistent with AR5 (Rhein et al., 2013), this Report finds that ocean warming dominates the changes in the global energy inventory high 2 train +1169 AR6_WGI 954 2 The contributions from the other components across all periods are approximately 5% from land heating, 3% for cryosphere heating and 1% associated with warming of the atmosphere high 2 train +1170 AR6_WGI 955 23 The origin of these trends is not fully understood, although there is evidence that anthropogenic aerosols have made a substantial contribution medium 1 train +1171 AR6_WGI 957 4 For the period 1970–2011, AR5 reported that the global energy budget was closed within uncertainties high 2 train +1172 AR6_WGI 961 10 The physiological adjustments are therefore assessed to make a substantial contribution to the overall tropospheric adjustment for CO 2 high 2 train +1173 AR6_WGI 961 15 Due to the agreement between the studies and the understanding of the physical mechanisms there is medium confidence in the mechanisms underpinning the tropospheric adjustment, but low confidence in its magnitude.The ERF from doubling CO 2 (2×CO 2) from the 1750 level (278 ppm; Section 2.2.3.3) is assessed to be 3.93 ± 0.47 W m–2 high 2 train +1174 AR6_WGI 963 9 This does not correspond to the division between ozone production and ozone depletion and is sensitive to the choice of tropopause high 2 train +1175 AR6_WGI 963 11 This small positive (but with uncertainty encompassing negative values) stratospheric ozone SARF is due to contributions from ozone precursors to lower stratospheric ozone and some of the CMIP6 models showing ozone depletion in the upper stratosphere, where depletion contributes a positive radiative forcing medium 1 train +1176 AR6_WGI 963 12 As there is insufficient evidence to quantify adjustments, for total ozone the assessed central estimate for ERF is assumed to be equal to SARF low 0 train +1177 AR6_WGI 963 32 It has increased by 0.49 W m–2 compared to AR5 (reference year 2011) high 2 train +1178 AR6_WGI 963 34 Changes in the radiative efficiencies (including adjustments) of CO 2, CH 4, N 2O and halogenated compounds have increased the ERF by an additional 0.15 W m–2 compared to the AR5 values high 2 train +1179 AR6_WGI 963 38 The combined ERF from ozone and stratospheric water vapour has increased since AR5 by 0.10 ± 0.50 W m–2 high 2 train +1180 AR6_WGI 966 22 Based on the above, IRFari is assessed to be –0.25 ± 0.2 W m–2 medium 1 train +1181 AR6_WGI 966 24 Adding this small adjustment to our assessed IRFari estimate of –0.25 W m–2, and accounting for additional uncertainty in the adjustments, ERFari is assessed to –0.3 ± 0.3 medium 1 train +1182 AR6_WGI 968 27 Taking the average across the studies providing IRFaci estimates discussed above and considering the general agreement among estimates (Table 7.7), IRFaci is assessed to be –0.7 ± 0.5 W m–2 medium 1 train +1183 AR6_WGI 969 9 These three studies together suggest a global Cf adjustment that augments ERFaci relative to IRFaci by –0.5 ± 0.4 W m–2 medium 1 train +1184 AR6_WGI 969 18 Combining IRFaci and the associated adjustments in Cf and LWP (adding uncertainties in quadrature), considering only liquid-water clouds and evidence from satellite observations alone, the central estimate and very likely range for ERFaci is assessed to be –1.0 ± 0.7 W m–2 medium 1 train +1185 AR6_WGI 969 37 There is thus limited evidence and medium agreement for a small negative contribution to ERFaci from anthropogenic INP-induced cirrus modifications low 0 test +1186 AR6_WGI 969 41 From model-based evidence, ERFaci is assessed to –1.0 ± 0.8 W m–2 medium 1 train +1187 AR6_WGI 970 1 The strong agreement between the two largely independent lines of evidence increases confidence in the overall assessment of the central estimate and very likely range for ERFaci of –1.0 ± 0.7 W m–2 medium 1 train +1188 AR6_WGI 970 38 Based solely on energy balance considerations or other observational constraints, it is extremely likely that the total aerosol ERF is negative (high confidence), but extremely unlikely that the total aerosol ERF is more negative than –2.0 W m–2 high 2 train +1189 AR6_WGI 971 18 Combining the lines of evidence and adding uncertainties in quadrature, the ERFari+aci estimated for 2014 relative to 1750 is assessed to be –1.3 [–2.0 to –0.6] W m–2 medium 1 train +1190 AR6_WGI 971 26 However, based on a general reduction in global mean AOD over this period (Section 2.2.6 and Figure 2.9), combined with a reduction in emissions of aerosols and their precursors in updated emissions inventories (Hoesly et al., 2018), the aerosol ERF is assessed to have decreased in magnitude from about 2014 to 2019 medium 1 train +1191 AR6_WGI 971 28 Aerosols are therefore assessed to have contributed an ERF of –1.1 [–1.7 to –0.4] W m–2 over 1750–2019 medium 1 train +1192 AR6_WGI 972 35 Adding the irrigation effect to this gives an overall assessment of the ERF from land-use change of –0.20 ± 0.10 W m–2 medium 1 train +1193 AR6_WGI 974 6 The best estimate solar ERF is assessed to be 0.01 W m–2, using the 14C reconstruction from SATIRE-M, with a likely range of –0.06 to +0.08 W m–2 medium 1 train +1194 AR6_WGI 974 8 The Lockwood and Ball (2020) full uncertainty range is halved as the period of reduced solar activity in the Maunder Minimum had ended by 1750 medium 1 train +1195 AR6_WGI 974 38 The ERF for volcanic stratospheric aerosols is assessed to be –20 ± 5 W m–2 per unit SAOD medium 1 train +1196 AR6_WGI 975 7 The stratospheric-temperature-adjusted radiative efficiency for CH 4 is increased by approximately 25% high 2 train +1197 AR6_WGI 975 8 The tropospheric adjustment is tentatively assessed to be –14% low 0 train +1198 AR6_WGI 975 9 A +7% tropospheric adjustment has been added to the radiative efficiency for N 2O and +12% to CFC-11 and CFC-12 low 0 train +1199 AR6_WGI 976 4 However, the total aerosol ERF is assessed to be more negative compared to AR5, due to revised estimates rather than trends high 2 train +1200 AR6_WGI 976 6 Carbon dioxide continues to contribute the largest part (56 ± 16%) of this GHG ERF high 2 train +1201 AR6_WGI 976 7 As discussed in Section 7.3.3, aerosols have in total contributed an ERF of –1.1 [–1.7 to –0.4] W m–2 over 1750–2019 medium 1 train +1202 AR6_WGI 977 15 The total human forced GSAT change from 1750 to 2019 is calculated to be 1.29 [1.00 to 1.65] °C high 2 train +1203 AR6_WGI 978 4 Nonetheless, simple physically basedof a WMGHG warming of 1.58 [1.17 to 2.17] °C (high confidence), a warming from ozone changes of 0.23 [0.11 to 0.39] °C (high confidence), and a cooling of –0.50 [–0.22 to –0.96] °C from aerosol effects medium 1 train +1204 AR6_WGI 978 6 There is also a –0.06 [–0.15 to +0.01] °C contribution from surface reflectance changes which is dominated by land-use change medium 1 train +1205 AR6_WGI 978 7 Changes in solar and volcanic activity are assessed to have together contributed a small change of –0.02 [–0.06 to +0.02] °C since 1750 medium 1 train +1206 AR6_WGI 978 17 However, since 1980 the aerosol cooling trend has stabilized and may have started to reverse, so that over the last few decades the long-term warming has been occurring at a faster rate than would be expected due to CO 2 alone high 2 train +1207 AR6_WGI 992 15 Since AR5, several studies have confirmed that a shift from tundra to boreal forests and the associated albedo change leads to increased warming in Northern Hemisphere high latitudes high 2 train +1208 AR6_WGI 992 24 Given the limited number of studies, we take the full range of estimates discussed above for the biogeophysical feedback parameter, and assess the very likely range to be from 0.0 to +0.3 W m–2 °C–1, with a central estimate of +0.15 W m–2 °C–1 low 0 train +1209 AR6_WGI 993 25 As such, overall, on multi- centennial time scales the feedback parameter associated with ice sheets is likely negative (medium confidence), but on multi-millennial time scales by the time the ice sheets reach equilibrium, the feedback parameter is very likely positive high 2 train +1210 AR6_WGI 1003 13 For both the MPWP and EECO, models are more consistent with the temperature and CO 2 proxies than at the time of AR5 high 2 train +1211 AR6_WGI 1003 14 For the LGM Northern Hemisphere, which is the region with the most data and the time period with the least uncertainty in model boundary conditions, polar amplification in the PMIP4 ensemble mean is in good agreement with the proxies, especially for SAT medium 1 train +1212 AR6_WGI 1003 18 It is very likely that the warming in the Arctic will be more pronounced than the global average over the 21st century high 2 train +1213 AR6_WGI 1003 19 This is supported by models’ improved ability to simulate polar amplification during past time periods, compared with at the time of AR5 high 2 train +1214 AR6_WGI 1008 7 The value of α’ is larger if quantified based on the observed pattern of warming since 1980 (Figure 2.11b) which is more distinct from the equilibrium warming pattern expected under CO 2 forcing high 2 train +1215 AR6_WGI 1011 3 In summary, the process-based estimate of TCR is assessed to have the central value of 2.0°C with the likely range from 1.6 to 2.7 °C and the very likely range from 1.3 to 3.1 °C high 2 train +1216 AR6_WGI 1014 33 Regression-based feedbacks thus provide estimates of the radiative feedbacks that are associated with internal climate variability (e.g., Brown et al., 2014), and do not provide a direct estimate of ECS high 2 train +1217 AR6_WGI 1015 8 Estimates based on the response to volcanic eruptions agree with other lines of evidence (Knutti et al., 2017), but they do not constitute a direct estimate of ECS high 2 train +1218 AR6_WGI 1015 11 However, there is robust evidence and high agreement across the lines of evidence that ECS is extremely likely greater than 1.6°C high 2 train +1219 AR6_WGI 1015 12 There is robust evidence and medium agreement across the lines of evidence that ECS is very likely greater than 1.8°C and likely greater than 2.2°C high 2 train +1220 AR6_WGI 1015 13 These ranges of ECS correspond to estimates based on historical global energy budget constraints (Section 7.5.2.1) under the assumption of no feedback dependence on evolving SST patterns (i.e., α’ = 0) and thus represent an underestimate of the true ECS ranges that can be inferred from this line of evidence high 2 train +1221 AR6_WGI 1015 15 Global energy budget constraints indicate a central estimate (median) TCR value of 1.9°C and that TCR is likely in the range 1.5 to 2.3 °C and very likely in the range 1.3 to 2.7 °C high 2 train +1222 AR6_WGI 1018 1 As such, based solely on the paleoclimate record, it is assessed to be very likely that ECS is greater than 1.5°C high 2 train +1223 AR6_WGI 1019 2 Despite the large variation in individual studies at the extreme upper end, all except two studies (both of which are from glacial–interglacial time periods associated with low confidence) have central estimates that are below 6°C; overall we assess that it is extremely likely that ECS is below 8°C high 2 train +1224 AR6_WGI 1025 3 The high ECS and TCR values can in some CMIP6 models be traced to improved representation of extratropical cloud feedbacks medium 1 train +1225 AR6_WGI 1025 5 Solely based on its ECS or TCR values an individual ESM cannot be ruled out as implausible, though some models with high (greater than 5°C) and low (less than 2°C) ECS are less consistent with past climate change high 2 train +1226 AR6_WGI 1027 30 Global ocean heat uptake is a smaller source of uncertainty in long-term surface warming high 2 train +1227 AR6_WGI 1030 22 In summary, GWPs and GTPs for methane and nitrous oxide are slightly lower than in AR5 medium 1 train +1228 AR6_WGI 1030 29 Estimates of the yield are 61% (Boucher et al., 2009) and 88% (Shindell et al., 2017), so the assessed range is 50–100% with a central value of 75% low 0 train +1229 AR6_WGI 1030 34 Methane from fossil fuel sources has therefore slightly higher emissions metric values than those from biogenic sources high 2 train +1230 AR6_WGI 1035 10 Hence, net zero CO 2 and net zero GHG, quantified using these new approaches, would both lead to approximately stable contributions to temperature change after net zero emissions are achieved high 2 train +1231 AR6_WGI 1073 8 The overall effect of anthropogenic aerosols is to reduce global precipitation and alter large-scale atmospheric circulation patterns through their well- understood surface radiative cooling effect high 2 train +1232 AR6_WGI 1073 9 Land- use and land-cover changes also drive regional water cycle changes through their influence on surface water and energy budgets high 2 train +1233 AR6_WGI 1073 11 An increase in near-surface atmospheric water holding capacity of about 7% per 1°C of warming explains a similar magnitude of intensification of heavy precipitation events (from sub-daily up to seasonal time scales) that increases the severity of flood hazards when these extremes occur high 2 train +1234 AR6_WGI 1073 12 The severity of very wet and very dry events increases in a warming climate high 2 train +1235 AR6_WGI 1073 13 A slowdown of tropical circulation with global warming partly offsets the warming- induced strengthening of precipitation in monsoon regions high 2 train +1236 AR6_WGI 1073 15 Greater warming over land than over the ocean alters atmospheric circulation patterns and, on average, reduces continental near-surface relative humidity, which contributes to regional drying high 2 train +1237 AR6_WGI 1073 18 Global warming has contributed to an overall increase in atmospheric moisture and precipitation intensity (high confidence), increased terrestrial evapotranspiration (medium confidence), influenced global patterns in aridity (very likely), and enhanced contrasts in surface salinity and precipitation minus evaporation patterns over the oceans high 2 train +1238 AR6_WGI 1073 21 Earlier onset of spring snowmelt and increased melting of glaciers have already contributed to seasonal changes in streamflow in high-latitude and low-elevation mountain catchments high 2 train +1239 AR6_WGI 1073 23 Shifts in the tropical rain belt are associated with the inter-hemispheric temperature response to the time-evolving radiative influence of anthropogenic aerosols and the ongoing warming influence of greenhouse gases high 2 train +1240 AR6_WGI 1073 24 Cooling in the Northern Hemisphere by sulphate aerosols explained a southward shift in the tropical rain belt and contributed to the Sahel drought from the 1970s to the 1980s (high confidence), subsequent recovery from which has been linked with greenhouse gas warming medium 1 train +1241 AR6_WGI 1073 25 Observed changes in regional monsoon precipitation, especially over South Asia, East Asia and West Africa, have been limited over much of the 20th century due to increases driven by warming from greenhouse gases being counteracted by decreases due to cooling from anthropogenic aerosols high 2 train +1242 AR6_WGI 1073 28 Urbanization has increased local precipitation (medium confidence) and resulting runoff intensity high 2 train +1243 AR6_WGI 1073 29 Increased precipitation intensities have enhanced groundwater recharge, most notably in tropical regions medium 1 train +1244 AR6_WGI 1074 8 Projected increases in precipitation amount and intensity will be associated with increased runoff in the northern high latitudes high 2 train +1245 AR6_WGI 1074 10 Runoff from small glaciers will typically decrease through loss of ice mass, while runoff from large glaciers is likely to increase with increasing global warming until glacier mass becomes depleted high 2 train +1246 AR6_WGI 1074 12 In the Mediterranean, south-western South America, and western North America, future aridification will far exceed the magnitude of change seen in the last millennium high 2 train +1247 AR6_WGI 1074 13 Some tropical regions are also projected to experience increased aridity, including the Amazon basin and Central America high 2 train +1248 AR6_WGI 1074 15 In the tropics and in the extratropics of both hemispheres during summer/warm season, interannual variability of precipitation and runoff over land is projected to increase at a faster rate than changes in seasonal mean precipitation amount medium 1 train +1249 AR6_WGI 1074 17 Sub-seasonal precipitation variability is also projected to increase, with fewer rainy days but increased daily mean precipitation intensity over many land regions high 2 train +1250 AR6_WGI 1074 18 Precipitation extremes will increase in almost all regions (high confidence), even where seasonal mean precipitation is projected to decrease medium 1 train +1251 AR6_WGI 1074 21 Summer monsoon precipitation is projected to increase for the South, South East and East Asian monsoon domains, while North American monsoon precipitation is projected to decrease medium 1 train +1252 AR6_WGI 1074 22 West African monsoon precipitation is projected to increase over the Central Sahel and decrease over the far western Sahel medium 1 train +1253 AR6_WGI 1074 26 A continued poleward shift of storm tracks in the Southern Hemisphere (likely) and the North Pacific medium 1 train +1254 AR6_WGI 1074 29 The annual contrast between the wettest and driest month of the year is likely to increase by 3–5% per 1°C in most monsoon regions in terms of precipitation, precipitation minus evaporation, and runoff medium 1 train +1255 AR6_WGI 1075 1 Increasing horizontal resolution in global climate models improves the representation of small-scale features and the statistics of daily precipitation high 2 train +1256 AR6_WGI 1075 2 High-resolution climate and hydrological models provide a better representation of land surfaces, including topography, vegetation and land-use change, which improve the accuracy of simulations of regional changes in the water cycle high 2 train +1257 AR6_WGI 1075 6 Water cycle changes that have already emerged from natural variability will become more pronounced in the near term, but the occurrence of volcanic eruptions (either single large events or clustered smaller ones) can alter the water cycle for several years, decreasing global mean land precipitation and altering monsoon circulation high 2 train +1258 AR6_WGI 1075 8 Non-linear water cycle responses are explained by the interaction of multiple drivers, feedbacks and time scales high 2 train +1259 AR6_WGI 1075 9 Non-linear responses of regional runoff, groundwater recharge and water scarcity highlight the limitations of simple pattern-scaling techniques medium 1 test +1260 AR6_WGI 1075 10 Water resources fed by melting glaciers are particularly exposed to non-linear responses high 2 train +1261 AR6_WGI 1075 14 Continued Amazon deforestation, combined with a warming climate, raises the probability that this ecosystem will cross a tipping point into a dry state during the 21st century low 0 train +1262 AR6_WGI 1075 15 The paleoclimate records show that a collapse in the Atlantic Meridional Overturning Circulation (AMOC) causes abrupt shifts in the water cycle high 2 train +1263 AR6_WGI 1075 19 The impact of SRM can affect different regions in potentially disruptive ways low 0 test +1264 AR6_WGI 1078 17 Detectable human influ ence on changes to the water cycle were found in atmospheric moisture content (medium confidence), global-scale changes of precipitation over land (medium confidence), intensification of heavy precipitation events over land regions where sufficient data networks exist medium 1 train +1265 AR6_WGI 1079 8 Further evidence that anthropogenic global warming has caused an increase in the frequency, intensity and/or amount of heavy precipitation events at the global scale (medium confidence), as well as in drought occurrence in the Mediterranean region medium 1 train +1266 AR6_WGI 1079 11 A key finding is that ‘limiting global warming to 1.5°C compared to 2°C would approximately halve the proportion of the world population expected to suffer water scarcity, although there is considerable variability between regions medium 1 train +1267 AR6_WGI 1079 24 This Report indicates that since 1850–1900, land surface temperature has risen nearly twice as much as global surface temperature (high confidence), with an increase in dry climates high 2 train +1268 AR6_WGI 1079 25 Land surface processes modulate the likelihood, intensity and duration of many extreme events including droughts (medium confidence) and heavy precipitation medium 1 train +1269 AR6_WGI 1079 26 The direction and magnitude of hydrological changes induced by land use change and land surface feedbacks vary with location and season high 2 train +1270 AR6_WGI 1079 27 Desertification exacerbates climate change through feedbacks involving vegetation cover, greenhouse gases and mineral dust aerosol high 2 train +1271 AR6_WGI 1079 28 Urbanization increases extreme rainfall events over or downwind of cities medium 1 train +1272 AR6_WGI 1083 25 Global precipitation increases due to GHGs are offset by the well-understood overall surface radiative cooling effect by aerosols high 2 train +1273 AR6_WGI 1083 26 Over land, the average warming-related increase in precipitation is expected to be smaller than over the ocean due to increasing land –ocean thermal contrast and surface feedbacks, but the overall precipitation increase over land is generally reinforced by fast atmospheric responses to GHGs that strengthens convergence of winds medium 1 train +1274 AR6_WGI 1083 27 Global mean precipitation and evaporation increase at a lower rate than atmospheric moisture per 1°C of global warming high 2 train +1275 AR6_WGI 1085 2 To summarize, increased moisture transport from evaporative oceans to high precipitation regions of the atmospheric circulation will drive amplified P–E and salinity patterns over the ocean high 2 test +1276 AR6_WGI 1085 3 Greater warming over land than ocean alters atmospheric circulation patterns and on average reduces continental near-surface relative humidity which along with vegetation feedbacks can contribute to regional decreases in precipitation high 2 train +1277 AR6_WGI 1087 9 Mean tropical circulation is expected to slow with global warming (high confidence) but temporary multi-decadal strengthening is possible due to internal variability medium 1 train +1278 AR6_WGI 1087 10 Slowing of the tropical circulation reduces the meridional P–E gradient over the Pacific and can partly offset thermodynamic amplification of P–E patterns and strengthening of monsoons high 2 train +1279 AR6_WGI 1088 8 Melting of snowpack or glaciers can increase streamflow in high-latitude and high-altitude catchments until frozen water reserves are depleted high 2 train +1280 AR6_WGI 1089 31 On a global level, warmer temperatures increase evaporative demand in the atmosphere, and thus (assuming sufficient soil moisture is available) increase moisture loss from evapotranspiration high 2 train +1281 AR6_WGI 1091 16 In snow-dominated regions, high temperatures increase the fraction of precipitation falling as rain instead of snow and advance the timing of spring snowmelt high 2 train +1282 AR6_WGI 1091 22 While potentially buffering water resource capacity, in some cases these interventions may unexpectedly increase vulnerability medium 1 train +1283 AR6_WGI 1091 29 Beyond a lack of precipitation, changes in evapotranspiration are critical components of drought, because these can lead to soil moisture declines high 2 train +1284 AR6_WGI 1091 30 Under very dry soil conditions, evapotranspiration becomes restricted and plants experience water stress in response to increased atmospheric demand medium 1 train +1285 AR6_WGI 1091 31 Human activities and decision-making have a critical impact on drought severity high 2 train +1286 AR6_WGI 1092 27 Urbanization can increase local precipitation (medium confidence) and resulting runoff intensity high 2 train +1287 AR6_WGI 1092 28 A warming climate combined with direct human demand for water is expected to deplete groundwater resources in dry regions high 2 train +1288 AR6_WGI 1093 5 These processes partially explain medium 1 train +1289 AR6_WGI 1093 6 These processes also explain high 2 train +1290 AR6_WGI 1094 1 In particular, aerosols induce intense convection at the Himalaya foothills during the pre-monsoon season, which generates a regional convergence there medium 1 train +1291 AR6_WGI 1096 6 It is likely that the global land P–E variations observed since the late 1970s were dominated by internal variability, mostly linked to ENSO teleconnections medium 1 train +1292 AR6_WGI 1097 8 There is further evidence of a faster increase since the 1980s medium 1 train +1293 AR6_WGI 1100 1 There is, however, a detectable increase in northern high-latitude annual precipitation over land which has been primarily driven by human-induced global warming high 2 train +1294 AR6_WGI 1100 2 Human influence has strengthened the zonal mean precipitation contrast between the wet tropics and dry subtropics since the 1980s medium 1 train +1295 AR6_WGI 1100 15 In contrast to AR5, there are now consistent trends in pan evaporation and evapotranspiration at the global scale, given the recent increase in both variables since the mid-1990s medium 1 train +1296 AR6_WGI 1100 26 Plant water use efficiency (WUE) is expected to rise with CO 2 levels high 2 train +1297 AR6_WGI 1103 44 Both central and north-eastern Africa have experienced a decline in rainfall since about 1980 high 2 train +1298 AR6_WGI 1105 23 In some regions of western North America and the Mediterranean, paleoclimate evidence suggests that recent warming has resulted in droughts that are of similar or greater intensity than those reconstructed over the last millennium medium 1 train +1299 AR6_WGI 1106 7 In summary, human-induced global warming has been the primary driver of a global glacier recession since the early 20th century high 2 train +1300 AR6_WGI 1106 8 Most glaciers have lost mass more rapidly since the 1960s and in an unprecedented way over the last decade, thereby contributing to increased glacier runoff, especially from larger glaciers until a maximum is reached, which tends to occur later in basins with larger glaciers and higher ice-cover fractions high 2 train +1301 AR6_WGI 1106 11 AR6 assesses that NH spring snow cover has been decreasing since 1978 (very high confidence) and that this trend extends back to 1950 high 2 train +1302 AR6_WGI 1106 29 In summary, a decline in the spring NH snow cover extent, snow depth and duration has been observed since the late 1960s and has been attributed to human influence high 2 train +1303 AR6_WGI 1109 21 Decreasing precipitation trends in the NH during the 1950s –1980s have been attributed to anthropogenic aerosol emissions from North America and Europe high 2 train +1304 AR6_WGI 1109 23 In AR6, Chapter 2 (Section 2.3.1.4.1) states that the HC has very likely widened and strengthened since at least the 1980s, mostly in the NH medium 1 test +1305 AR6_WGI 1110 6 As assessed in Section 3.3.3.1, GHG increases and stratospheric ozone depletion have contributed to the expansion of the zonal mean HC in the SH since around 1980, and the expansion of the NH HC has not exceeded the range of internal variability medium 1 train +1306 AR6_WGI 1110 11 GHG increases and stratospheric ozone depletion have contributed to expansion of the zonal mean HC in the SH since around 1970, while the expansion of the NH HC has not exceeded the range of internal variability medium 1 train +1307 AR6_WGI 1110 14 The causes of the observed strengthening of the WC during 1980 –2014 are not well understood due to competing influences from individual external forcings and since this strengthening is outside the range of variability simulated in coupled models medium 1 train +1308 AR6_WGI 1112 3 The SAsiaM strengthened during past periods of enhanced summer insolation in the NH, such as the early-to-mid Holocene warm period around 9000 to 6000 years before the present (BP) (Masson-Delmotte et al., 2013; Mohtadi et al., 2016; Braconnot et al., 2019) and weakened during cold periods high 2 train +1309 AR6_WGI 1112 16 Results from climate models indicate that anthropogenic aerosol forcing has dominated the recent decrease in summer monsoon precipitation, as opposed to the expected intensification due to GHG forcing high 2 train +1310 AR6_WGI 1112 17 On paleoclimate time scales, the SAsiaM strengthened in response to enhanced summer warming in the NH during the early-to-mid Holocene, while it weakened during cold intervals high 2 train +1311 AR6_WGI 1112 18 These changes are tightly linked to orbital forcing and changes in high-latitude climate medium 1 train +1312 AR6_WGI 1113 16 The transition towards a positive PDV phase has been one of the main drivers of the EAsiaM weakening since the 1970s high 2 train +1313 AR6_WGI 1113 18 On paleoclimate time scales, enhanced summer insolation in the Northern Hemisphere (NH) intensified the WAfriM precipitation during the early-to-mid Holocene high 2 train +1314 AR6_WGI 1113 21 The WAfriM experienced the wettest decade of the 20th century during the 1950s and early 1960s high 2 train +1315 AR6_WGI 1113 25 Wetter conditions of the WAfriM prevailed later from the mid-to-late 1990s, although the positive trend in precipitation started since the late 1980s (see also Section 10.4.2.1) over the Sahel (high confidence) and in the Guinean coastal region medium 1 train +1316 AR6_WGI 1114 12 In summary, most regions of West Africa experienced a wet period in the mid-20th century followed by a very dry period in the 1970s and 1980s that is attributed to aerosol cooling of the NH high 2 train +1317 AR6_WGI 1114 13 Recent estimates provide evidence of a WAfriM recovery from the mid-to-late 1990s, with more intense extreme events partly due to the combined effects of increasing GHG and decreasing anthropogenic aerosols over Europe and North America high 2 train +1318 AR6_WGI 1114 17 The NAmerM strengthened until the mid-Holocene period, in response to ice-emsheet retreat and rising summer insolation, but probably did not exceed the strength of the modern system low 0 train +1319 AR6_WGI 1115 1 The intensification recorded since about the 1970s has been partly driven by GHG emissions medium 1 train +1320 AR6_WGI 1115 27 Paleoclimate reconstructions and simulations suggest a weaker SAmerM during warmer epochs such as the Mid-Holocene or the 900–1100 period, and stronger monsoon during colder epochs such as the LGM or the 1400–1600 period high 2 train +1321 AR6_WGI 1117 23 In summary, no robust trend in atmospheric blocking has been detected in modern reanalyses and in CMIP6 historical simulations medium 1 train +1322 AR6_WGI 1117 24 The lack of trend is explained by strong internal variability and/or the competing effects of low-level Arctic amplification and upper-level tropical amplification of the equator- to-pole temperature gradient medium 1 train +1323 AR6_WGI 1120 22 Since the end of the 19th century, synchronous hydroclimate changes medium 1 train +1324 AR6_WGI 1120 31 Over equatorial East Africa the IOD affects the short rain season medium 1 train +1325 AR6_WGI 1121 1 The strength and frequency of the MJO have increased over the past century medium 1 train +1326 AR6_WGI 1121 17 In summary, multiple water cycle changes related to ENSO and IOD teleconnections have been observed across the 20th century high 2 train +1327 AR6_WGI 1121 18 The MJO amplitude has increased in the second half of the 20th century partly because of anthropogenic global warming medium 1 train +1328 AR6_WGI 1122 4 In summary, while the attribution of 20th century variations of the NAM/NAO is still unclear, there is a strong relationship with precipitation changes over Europe and in the Mediterranean region high 2 train +1329 AR6_WGI 1122 5 SAM teleconnections are associated with changes in moisture transport and extend to South America, Australia and Antarctica high 2 train +1330 AR6_WGI 1122 17 In this Report, Chapter 4 provides an updated assessment of global annual precipitation (Section 4.3.1), finding that it is very likely that annual precipitation averaged over all land regions continuously increases as global surface temperatures increase in the 21st century high 2 train +1331 AR6_WGI 1125 12 In general, there will be increases in moisture transport into storm systems, monsoons and high latitudes medium 1 train +1332 AR6_WGI 1125 16 The AR5 assessed that the contrast of mean precipitation amount between dry and wet regions and seasons is expected to increase over most of the globe as temperatures increase high 2 test +1333 AR6_WGI 1127 6 Beyond annual or seasonal mean precipitation amounts, an implication of the parallel intensification of the global water cycle and of the increased residence time of atmospheric water vapour (Section 8.2.1) is that the distribution of daily and sub-daily precipitation intensities will experience significant changes (Pendergrass and Hartmann, 2014b; Pendergrass et al., 2015; Bador et al., 2018; Douville and John, 2021), with fewer but potentially stronger events high 2 train +1334 AR6_WGI 1129 8 Daily mean precipitation intensities, including extremes, are projected to increase over most regions high 2 train +1335 AR6_WGI 1129 9 The number of dry days is projected to increase over the subtropics, Amazonia, and Central America medium 1 train +1336 AR6_WGI 1131 5 In summary, the annual range of precipitation, water availability and streamflow will increase with global warming over subtropical regions and the Amazon (medium confidence), especially around the Mediterranean and across southern Africa high 2 train +1337 AR6_WGI 1131 6 The contrast between the wettest and driest month of the year is likely to increase by 3–5% °C–1 with global warming in most monsoon regions, in terms of precipitation, water availability (P–E) and runoff medium 1 train +1338 AR6_WGI 1132 7 CMIP5 models also project an increase in evapotranspiration over most land areas medium 1 train +1339 AR6_WGI 1133 8 Evapotranspiration increases in most land regions, except in areas that are projected to become moisture-limited (due to reduced precipitation and increased evaporative demand), such as the Mediterranean, South Africa, and the Amazonian basin medium 1 train +1340 AR6_WGI 1133 9 The patterns of change increase in magnitude from low to high-emissions SSP scenarios medium 1 train +1341 AR6_WGI 1133 10 In summary, future projections indicate that anthropogenic forcings will drive an increase in global mean evaporation over most oceanic areas (high confidence) (Figure 8.17), an increase in global atmospheric demand (virtually certain) and an increase in evapotranspiration over most land areas, with the exception of moisture-limited regions medium 1 train +1342 AR6_WGI 1135 14 There is a likely increase in drought occurrence medium 1 test +1343 AR6_WGI 1135 18 Consistent with the coherent nature of warming in future projections, increases in vapour pressure deficit and evaporative demand are widespread and consistent across regions, seasons, and models, increasing in magnitude in accordance with the emissions scenario high 2 train +1344 AR6_WGI 1135 19 Even under a low-emissions scenario (SSP1-2.6), projections of soil moisture show significant decreases in the Mediterranean, southern Africa, and the Amazonian basin high 2 train +1345 AR6_WGI 1135 20 Under mid- and high-emissions scenarios (SSP2-4.5 and SSP5-8.5), coherent declines emerge across Europe, westernmost North Africa, south- western Australia, Central America, south-western North America, and south-western South America high 2 train +1346 AR6_WGI 1135 27 The percentage of land area experiencing drying is slightly lower when runoff is used as an aridity metric instead (20–30%); taking this into consideration, it is estimated that about a third of global land areas will experience at least moderate drying in response to anthropogenic emissions, even under SSP1-2.6 medium 1 train +1347 AR6_WGI 1136 2 The CORDEX South Asia multi-model ensemble projections indicate an increase in the frequency and severity of droughts over central and northern India during the 21st century, under the RCP4.5 and RCP8.5 scenarios medium 1 train +1348 AR6_WGI 1138 1 In general, these regions are expected to become drier both due to reduced precipitation (medium confidence) and increases in evaporative demand high 2 train +1349 AR6_WGI 1138 2 These same regions are likely to experience increases in drought duration and/or severity high 2 train +1350 AR6_WGI 1138 3 The magnitude of expected change scales with emissions scenarios (high confidence) but even under low-emissions trajectories, large changes in drought and aridity are expected to occur high 2 train +1351 AR6_WGI 1138 4 In the Mediterranean, Central Chile, and western North America, future aridification will far exceed the magnitude of change seen over the last millennium high 2 train +1352 AR6_WGI 1138 6 The SROCC noted that these declines are projected to continue almost everywhere over the 21st century (high confidence), with complete glacier loss expected in regions with only small glaciers very high 3 train +1353 AR6_WGI 1138 8 The SROCC concluded that cryosphere changes had already altered the seasonal timing and volume of runoff (very high confidence), which in turn had affected water resources and agriculture medium 1 train +1354 AR6_WGI 1138 14 Because of their lagged response to warming, glaciers will continue to lose mass for decades even if global temperature is stabilized very high 3 train +1355 AR6_WGI 1138 30 In summary, glaciers are projected to continue to lose mass under all emissions scenarios very high 3 train +1356 AR6_WGI 1138 31 Runoff from glaciers is projected to peak at different times in different places, with maximum rates of glacier mass loss in low latitude regions taking place in the next few decades in all scenarios high 2 train +1357 AR6_WGI 1138 32 While runoff from small glaciers will typically decrease because of glacier mass depletion, runoff from larger glaciers will increase with increasing global warming until glacier mass is similarly depleted, after which runoff peaks and then declines and which tends to occurs later in basins with larger glaciers and higher ice-cover fractions high 2 train +1358 AR6_WGI 1138 33 Glaciers in the Arctic and Antarctic will continue to lose mass through the latter half of the century and beyond high 2 train +1359 AR6_WGI 1140 19 It is likely that the zonal mean of the ITCZ will narrow and strengthen in the core region with projected surface warming high 2 train +1360 AR6_WGI 1140 20 Distinct regional shifts in the ITCZ will be associated with regional changes in precipitation amount and seasonality medium 1 train +1361 AR6_WGI 1141 4 The Hadley cells are projected to expand polewards with global warming, most notably in the SH high 2 train +1362 AR6_WGI 1141 20 Discrepancies between observed and simulated changes in SSTs in the tropics indicate that a temporary strengthening of the Walker Circulation can arise from internal variability medium 1 test +1363 AR6_WGI 1143 5 Since AR5, most studies have confirmed projected increases in South Asian monsoon precipitation high 2 train +1364 AR6_WGI 1146 6 Changes in seasonality (Box 8.2) are projected with a later monsoon onset (high confidence) over the Sahel and a late cessation medium 1 train +1365 AR6_WGI 1146 9 In summary, post-AR5 studies and newly available CMIP6 results indicate projected rainfall increases in the eastern-central WAfriM region but decreases in the west (high confidence), with a delayed wet season medium 1 train +1366 AR6_WGI 1146 10 Overall, WAfriM summer precipitation is projected to increase during the 21st century but with larger uncertainty noted under high-emissions scenarios medium 1 train +1367 AR6_WGI 1147 27 There is a projected increase in rainfall variability over northern Australia, with increased intensity of rainfall during the active or ‘burst’ phase medium 1 train +1368 AR6_WGI 1147 31 Section 11.7.1.5 assesses that the average tropical cyclone rain-rate is projected to increase with warming (high confidence), and peak rain rates are projected to increase at greater than the Clausius–Clapeyron scaling rate of 7% °C–1 warming in some regions due to increased low-level moisture convergence medium 1 train +1369 AR6_WGI 1150 16 This is projected to increase the intensity of heavy precipitation events on the west coast of the USA and in western Europe high 2 train +1370 AR6_WGI 1151 6 In summary, even though there is low confidence in how the tropical MoVs will change in the future (Sections 4.3.3.2 and 4.5.3.3), their regional hydrological consequences, in terms of precipitation, are projected to intensify medium 1 train +1371 AR6_WGI 1151 7 For example, the ENSO influence on precipitation over the Indo–Pacific sector is projected to strengthen and shift eastward medium 1 train +1372 AR6_WGI 1151 8 The MJO is projected to intensify in a warmer climate, with increased associated precipitation medium 1 train +1373 AR6_WGI 1151 16 In summary, projected changes in the intensity, frequency and phase of extratropical MoVs (see also Sections 4.3 and 4.5) may amplify regional changes in precipitation and contribute to an increase in their intra-seasonal and interannual variability medium 1 train +1374 AR6_WGI 1151 17 Regionally, there are potentially significant precipitation and atmospheric circulation changes associated with changes in extratropical dynamics low 0 train +1375 AR6_WGI 1159 2 For example, internal variability will continue to play an important role in the variability of river flows over France in coming decades medium 1 train +1376 AR6_WGI 1160 8 Major volcanic eruptions temporarily reduce total global and wet tropical region precipitation (high confidence) (Iles and Hegerl, 2014), can weaken or shift the ITCZ (Iles and Hegerl, 2014; Colose et al., 2016; Liu et al., 2016), and reduce summer monsoon rainfall medium 1 train +1377 AR6_WGI 1160 9 Monsoon precipitation in one hemisphere can be enhanced by the remote volcanic forcing occurring in the other hemisphere medium 1 train +1378 AR6_WGI 1160 16 The occurrence of volcanic eruptions in the coming century, either as single large events or clustered smaller ones, can alter the water cycle (see also Cross-Chapter Box 4.1), and regional drought events may be enhanced by co-occurring volcanic (Liu et al., 2016; Gao and Gao, 2017; Zambri et al., 2017) and GHG (e.g., Cook et al., 2018) forcing low 0 train +1379 AR6_WGI 1160 17 Volcanic eruptions may also lead to widespread precipitation anomalies up to several years following an eruption through their potential influence on the El Niño Southern Oscillation low 0 train +1380 AR6_WGI 1160 18 In summary, large volcanic eruptions reduce global mean precipitation, as well as precipitation in tropical wet regions high 2 train +1381 AR6_WGI 1164 10 In summary, there is both numerical and process-based evidence that terrestrial water cycle changes can be non-linear at the regional scale high 2 train +1382 AR6_WGI 1164 11 Non-linear regional responses of runoff, groundwater recharge and water scarcity have been documented based on both CMIP5 and CMIP6 models, and highlight the limitations of simple pattern-scaling techniques medium 1 train +1383 AR6_WGI 1164 12 Water resources fed by melting glaciers are particularly exposed to such non-linearities high 2 train +1384 AR6_WGI 1166 12 Observed transitions into and out of Green Sahara states are always faster than the underlying forcing, in agreement with theoretical considerations high 2 train +1385 AR6_WGI 1166 14 Both paleoclimate data and modelling experiments suggest that the timing and speed of the transition was spatially heterogeneous high 2 train +1386 AR6_WGI 1166 17 CMIP5 and CMIP6 models, some of which include dynamic vegetation schemes, cannot simulate the magnitude, nor the spatial extent, of greening and precipitation change associated with the last Green Sahara under standard mid-Holocene (6,000 years ago) boundary conditions high 2 train +1387 AR6_WGI 1167 17 In contrast, cirrus cloud thinning, a longwave radiation technique, results in increased global precipitation as it causes enhanced radiative cooling in the troposphere medium 1 train +1388 AR6_WGI 1167 25 The additional global warming caused by SRM termination may result in a rapid increase in global mean precipitation medium 1 train +1389 AR6_WGI 1230 8 Since the 1950s, the fastest surface warming has occurred in the Indian Ocean and in western boundary currents, while ocean circulation has caused slow warming or surface cooling in the Southern Ocean, equatorial Pacific, North Atlantic, and coastal upwelling systems very high 3 train +1390 AR6_WGI 1230 13 The long time scale also implies that the amount of deep-ocean warming is irreversible over centuries to millennia very high 3 test +1391 AR6_WGI 1230 14 On annual to decadal time scales, the redistribution of heat by the ocean circulation dominates spatial patterns of temperature change high 2 train +1392 AR6_WGI 1230 15 At longer time scales, the spatial patterns are dominated by additional heat, primarily stored in water masses formed in the Southern Ocean, and by weaker warming in the North Atlantic where heat redistribution caused by changing circulation counteracts the additional heat input through the surface high 2 train +1393 AR6_WGI 1230 17 Since the 1980s, they have approximately doubled in frequency (high confidence) and have become more intense and longer medium 1 train +1394 AR6_WGI 1230 19 The largest changes will occur in the tropical ocean and the Arctic medium 1 train +1395 AR6_WGI 1230 22 Based on recent refined analyses of the available observations, the global 0–200 m stratification is now assessed to have increased about twice as much as reported by SROCC, with a 4.9 ± 1.5% increase from 1970 to 2018 high 2 train +1396 AR6_WGI 1230 32 Western boundary currents have shifted poleward since 1993 medium 1 train +1397 AR6_WGI 1231 1 In the 21st century, consistent with projected changes in the surface winds, the East Australian Current Extension and Agulhas Current Extension will intensify, while the Gulf Stream and Indonesian Throughflow will weaken medium 1 train +1398 AR6_WGI 1231 2 Eastern boundary upwelling systems will change, with a dipole spatial pattern within each system of reduction at low latitude and enhancement at high latitude high 2 train +1399 AR6_WGI 1231 4 There is no tipping point for this loss of Arctic summer sea ice high 2 train +1400 AR6_WGI 1231 5 The practically ice-free state is projected to occur more often with higher greenhouse gas concentrations, and it will become the new normal for high-emissions scenarios by the end of this century high 2 train +1401 AR6_WGI 1231 6 Based on observational evidence, Coupled Model Intercomparison Project Phase 6 (CMIP6) models and conceptual understanding, the substantial satellite-observed decrease of Arctic sea ice area over the period 1979 –2019 is well described as a linear function of global mean surface temperature, and thus of cumulative anthropogenic carbon dioxide (CO 2) emissions, with superimposed internal variability high 2 train +1402 AR6_WGI 1231 9 The regionally opposing trends result primarily from changing regional wind forcing medium 1 train +1403 AR6_WGI 1231 13 This mass loss is driven by both discharge and surface melt, with the latter increasingly becoming the dominating component of mass loss with high interannual variability in the last decade high 2 train +1404 AR6_WGI 1231 14 The largest mass losses occurred in the north-west and the south-east of Greenland high 2 train +1405 AR6_WGI 1231 18 Mass losses from West Antarctic outlet glaciers outpaced mass gain from increased snow accumulation on the continent and dominated the ice-sheet mass losses since 1992 very high 3 train +1406 AR6_WGI 1231 19 These mass losses from the West Antarctic outlet glaciers were mainly induced by ice-shelf basal melt (high confidence) and locally by ice-shelf disintegration preceded by strong surface melt high 2 train +1407 AR6_WGI 1231 20 Parts of the East Antarctic Ice Sheet have lost mass in the last two decades high 2 train +1408 AR6_WGI 1231 23 The loss of ice from Greenland will become increasingly dominated by surface melt, as marine margins retreat and the ocean-forced dynamic response of ice-sheet margins diminishes high 2 train +1409 AR6_WGI 1231 24 In the Antarctic, dynamic losses driven by ocean warming and ice-shelf disintegration will likely continue to outpace increasing snowfall this century medium 1 train +1410 AR6_WGI 1231 25 Beyond 2100, total mass loss from both ice sheets will be greater under high-emissions scenarios than under low-emissions scenarios high 2 train +1411 AR6_WGI 1231 29 During the decade 2010–2019, glaciers lost more mass than in any other decade since the beginning of the observational record very high 3 train +1412 AR6_WGI 1231 30 For all regions with long-term observations, glacier mass in the decade 2010 –2019 is the smallest since at least the beginning of the 20th century medium 1 train +1413 AR6_WGI 1231 31 Because of their lagged response, glaciers will continue to lose mass at least for several decades even if global temperature is stabilized very high 3 train +1414 AR6_WGI 1232 2 Complete permafrost thaw in recent decades is a common phenomenon in discontinuous and sporadic permafrost regions medium 1 train +1415 AR6_WGI 1232 3 Permafrost warmed globally by 0.29 [0.17 to 0.41, likely range] °C between 2007 and 2016 medium 1 train +1416 AR6_WGI 1232 4 An increase in the active layer thickness is a pan-Arctic phenomenon medium 1 train +1417 AR6_WGI 1232 5 The volume of perennially frozen soil within the upper 3 m of the ground will decrease by about 25% per 1°C of global surface air temperature change (up to 4°C above pre-industrial temperature) medium 1 train +1418 AR6_WGI 1232 9 It is virtually certain that Northern Hemisphere snow cover extent will continue to decrease as global climate continues to warm, and process understanding strongly suggests that this also applies to Southern Hemisphere seasonal snow cover high 2 train +1419 AR6_WGI 1232 10 Northern Hemisphere spring snow cover extent will decrease by about 8% per 1°C of global surface air temperature change (up to 4°C above pre-industrial temperature) medium 1 train +1420 AR6_WGI 1232 12 GMSL rise has accelerated since the late 1960s, with an average rate of 2.3 [1.6 to 3.1] mm yr –1 over the period 1971–2018 increasing to 3.7 [3.2 to 4.2] mm yr –1 over the period 2006–2018 high 2 train +1421 AR6_WGI 1232 15 The contribution of Greenland and Antarctica to GMSL rise was four times larger during 2010–2019 than during 1992 –1999 high 2 train +1422 AR6_WGI 1232 16 Because of the increased ice-sheet mass loss, the total loss of land ice (glaciers and ice sheets) was the largest contributor to global mean sea level rise over the period 2006–2018 high 2 train +1423 AR6_WGI 1232 19 Temporal variability in ocean dynamics dominates regional patterns on annual to decadal time scales high 2 train +1424 AR6_WGI 1232 20 The anthropogenic signal in regional sea level change will emerge in most regions by 2100 medium 1 train +1425 AR6_WGI 1232 22 Observations show that high-tide flooding events that occurred five times per year during the period 1960–1980 occurred, on average, more than eight times per year during the period 1995–2014 high 2 train +1426 AR6_WGI 1232 23 Under the assumption that other contributors to extreme sea levels remain constant (e.g., stationary tides, storm-surge, and wave climate), extreme sea levels that occurred once per century in the recent past will occur annually or more frequently at about 19–31% of tide gauges by 2050 and at about 60% (SSP1-2.6) to 82% (SSP5-8.5) of tide gauges by 2100 medium 1 train +1427 AR6_WGI 1232 24 In total, such extreme sea levels will occur about 20 to 30 times more frequently by 2050 and 160 to 530 times more frequently by 2100 compared to the recent past, as inferred from the median amplification factors for SSP1-2.6, SSP2-4.5, and SSP5-8.5 medium 1 train +1428 AR6_WGI 1232 25 Over the 21st century, the majority of coastal locations will experience a median projected regional sea level rise within ±20% of the median projected GMSL change medium 1 train +1429 AR6_WGI 1233 5 By 2300, GMSL will rise between 0.3 m and 3.1 m under SSP1-2.6, between 1.7 m and 6.8 m under SSP5-8.5 in the absence of marine ice cliff instability, and by up to 16 m under SSP5-8.5 considering marine ice cliff instability low 0 train +1430 AR6_WGI 1239 6 Common regional biases in SST or historical SST trends are not exclusively linked to the representation of the ocean high 2 train +1431 AR6_WGI 1239 10 In summary, CMIP6 models show persistent regional biases in representing the climatological SST state very high 3 train +1432 AR6_WGI 1239 15 Warming is projected at varying rates in all regions by 2050, except the North Atlantic Subpolar Region, the equatorial Pacific, and the Southern Ocean where models disagree high 2 train +1433 AR6_WGI 1239 24 Similarly, the SST change pattern observed in the tropical Pacific Ocean will transition on centennial time scales to a mean pattern resembling the El Niño pattern medium 1 train +1434 AR6_WGI 1241 7 In summary, globally integrated and large-scale fluxes are more reliably inferred from heat content and salinity change, while regional trends are rarely robust in observations; where they are robust, they tend to be underestimated or in disagreement in models very high 3 train +1435 AR6_WGI 1241 9 The AR5 (Rhein et al., 2013) assessed with medium confidence that zonal wind stress over the Southern Ocean increased from the early 1980s to the 1990s medium 1 train +1436 AR6_WGI 1242 6 In summary, there is limited observational evidence that the mixed layer is globally deepening, while models show no emergence of a trend until later in the 21st century under strong emissions.The SROCC assessed that upper-ocean stratification will continue to increase in the 21st century under increased radiative forcing high 2 train +1437 AR6_WGI 1243 17 The SROCC highlighted that future change of MHWs will not be globally uniform, with the largest changes in the frequency of marine heatwaves being projected to occur in the western tropical Pacific and the Arctic Ocean medium 1 train +1438 AR6_WGI 1243 19 Moderate increases are projected for mid-latitudes, and only small increases are projected for the Southern Ocean medium 1 train +1439 AR6_WGI 1243 21 The resolution of current climate models (CMIP5 and CMIP6) capture the broad features of MHWs, but they may have a bias towards weaker and longer MHWs in the historical period medium 1 train +1440 AR6_WGI 1244 4 Section 2.3.3.1 reports that current multi-decadal to centennial rates of OHC gain are greater than at any point since the last deglaciation medium 1 train +1441 AR6_WGI 1244 9 Section 3.5.1.3 assessed that it is extremely likely that human influence was the main driver of the ocean heat content increase observed since the 1970s, which extends into the deeper ocean very high 3 train +1442 AR6_WGI 1246 1 In summary, in the upper 2000 m since the 1970s, the subpolar North Atlantic has been slowly warming, and the Southern Ocean has stored a disproportionally large amount of anthropogenic heat medium 1 train +1443 AR6_WGI 1247 4 In summary, and strengthening SROCC assessment, ocean warming is not globally uniform due to patterns of uptake predominantly along known water mass pathways, and due to changing ocean circulation redistributing heat within the ocean high 2 train +1444 AR6_WGI 1247 11 Despite a decrease of AMOC northward heat (0.17 PW) and mass (2.5 Sverdrup (Sv); 1 Sv = 109 kg s–1) transport, OHT has increased toward the Arctic through increased upper northern North Atlantic temperatures and stronger wind-driven gyres medium 1 train +1445 AR6_WGI 1247 12 In summary, OHT has increased toward the Arctic in recent decades, which at least partially explains the recent OHC change in the Arctic medium 1 train +1446 AR6_WGI 1248 12 In summary, climate models have more skill in representing OHC change from added heat than from ocean circulation change high 2 train +1447 AR6_WGI 1249 2 In summary, on decadal time scales, redistribution will dominate regional patterns of OHC change without affecting the globally integrated OHC; however, by 2100, particularly under strong warming scenarios, there is high confidence that regional patterns of OHC change will be dominated by added heat entering the sea surface, primarily in water mass formation regions in the subtropics; and reduced aerosols will increase the relative rate of Northern Hemisphere heat uptake medium 1 train +1448 AR6_WGI 1250 2 Section 2.3.3.2 strengthens evidence that subsurface salinity trends are connected to surface trends (very likely), which are, in turn, linked to an intensifying hydrological cycle medium 1 train +1449 AR6_WGI 1250 21 Section 3.5.2.1 reports, however, that the fidelity of ocean salinity simulation has improved in CMIP6, and near-surface and subsurface biases have been reduced medium 1 train +1450 AR6_WGI 1250 26 Projections confirm SROCC assessment that fresh ocean regions will continue to get fresher and salty ocean regions will continue to get saltier in the 21st century medium 1 test +1451 AR6_WGI 1251 5 Consistently, we assess that STMW have deepened worldwide, with greatest deepening in the Southern Hemisphere high 2 train +1452 AR6_WGI 1251 12 The SROCC connected SAMW and AAIW to Southern Ocean temperature changes as the large Southern Ocean surface heat uptake is circulated and mixed along with these water masses high 2 train +1453 AR6_WGI 1253 18 The SROCC (Collins et al., 2019) assessed that in situ observations (2004–2017) and sea surface temperature reconstructions indicate that AMOC has weakened relative to 1850–1900 medium 1 train +1454 AR6_WGI 1254 10 This suggests that the observed AMOC-shift between 2007 and 2011 may be part of a longer-term decrease medium 1 train +1455 AR6_WGI 1255 4 Projected AMOC decline by 2100 ranges from 24 [4 to 46] % in SSP1-2.6 to 39 [17–55] % in SSP5-8.5 medium 1 train +1456 AR6_WGI 1255 11 Tuning towards stability and model biases (Valdes, 2011; Liu et al., 2017; Mecking et al., 2017; Weijer et al., 2019) provides CMIP models a tendency toward unrealistic stability medium 1 train +1457 AR6_WGI 1256 2 It also reported that, instead of increasing the mean ACC transport, additional energy input associated with increased wind stress cascades into the eddy field medium 1 train +1458 AR6_WGI 1256 22 For the lower cell overturning circulation, SROCC assessed that a slowdown of its transport is consistent with the observed decrease in volume medium 1 train +1459 AR6_WGI 1257 9 Section 2.4 concludes that a sustained shift beyond multi-centennial variability has not been observed for El Niño–Southern Oscillation (ENSO) medium 1 train +1460 AR6_WGI 1257 17 In summary, while future changes in tropical modes of variability remain unclear, change in atmospheric and ocean circulation will drive continued change in tropical ocean temperature in the 21st century (medium confidence), with part of the region experiencing drastic marine heat wave conditions high 2 train +1461 AR6_WGI 1257 23 Section 2.3.3.4 assesses that, while WBC strength is highly variable at multi-decadal scale (high confidence), WBCs and subtropical gyres have shifted poleward since 1993 medium 1 train +1462 AR6_WGI 1258 3 In the North Pacific, there has been an increasing trend in the Alaska Gyre from 1993 to 2017 (Cummins and Masson, 2018), which might be attributed to Pacific Decadal Oscillation low 0 train +1463 AR6_WGI 1258 5 All climate models reproduce WBCs and gyres, but eddy- present or eddy-rich models (roughly 10–25 km and about 10 km resolution, respectively) represent these currents more realistically than eddy-parameterized models very high 3 train +1464 AR6_WGI 1258 7 Increased resolution admits mesoscale eddies, and also improves simulation of the strength and position of WBCs such as the Kuroshio Current, Gulf Stream, and East Australian Current very high 3 train +1465 AR6_WGI 1258 9 The wind-current feedback, implemented by considering relative velocity of currents and wind, realistically dampens mesoscale eddies and WBCs, through mesoscale air–sea interaction (Ma et al., 2016; Renault et al., 2016, 2019), even though sub-mesoscale wind- current damping feedback is missing in these models medium 1 train +1466 AR6_WGI 1258 10 As eddies potentially play a role in determining the strength of gyre circulations and their low- frequency variability (Fox-Kemper and Pedlosky, 2004; Berloff et al., 2007), it is expected that eddy-present and eddy-rich models will differ in their decadal variability and sensitivity to changes in the wind stress of gyres from eddy-parameterized models medium 1 test +1467 AR6_WGI 1259 5 The SROCC (Collins et al., 2019) concluded with high confidence that Indonesian Throughflow (ITF) transport from the Pacific Ocean to the Indian Ocean has increased in the past two decades as a result medium 1 train +1468 AR6_WGI 1260 6 Projected future annual cumulative upwelling wind changes at most locations, and seasons remain within ±10–20% of present-day values in the 21st century, even in the context of high-end emissions scenarios (4×CO 2 or RCP8.5) medium 1 train +1469 AR6_WGI 1260 10 Change in upper-ocean stratification (Section 9.2.1.3) is projected to increase confinement of upwelling vertical velocities to near the ocean surface high 2 train +1470 AR6_WGI 1260 11 The California Current system has experienced some upwelling-favourable wind intensification since the 1980s high 2 test +1471 AR6_WGI 1260 13 New evidence reinforces our confidence in SROCC assessment that, under increased radiative forcing, EBUS winds will change with a dipole spatial pattern within each EBUS of reduction (weaker and/or shorter) at low latitude, and enhancement (stronger and/or longer) at high latitude high 2 train +1472 AR6_WGI 1261 19 Thus, while conversions between OHC, mean ocean temperature and GMTSL across applications are within uncertainty ranges medium 1 train +1473 AR6_WGI 1263 3 Patterns of change are consistent between model simulations and observations medium 1 train +1474 AR6_WGI 1263 15 Projections of dynamic sea level variability require fully three-dimensional ocean models, and only high-resolution ocean models are statistically consistent on short time scales with satellite altimeter observations very high 3 train +1475 AR6_WGI 1263 17 The SROCC (Meredith et al., 2019) assesses that sea ice extent, which is the total area of all grid cells with at least 15% sea ice concentration, has declined since 1979 in each month of the year very high 3 train +1476 AR6_WGI 1263 21 Sea ice area has decreased in every month of the year from 1979 to the present very high 3 train +1477 AR6_WGI 1263 22 The absolute and the relative ice losses are highest in late summer-early autumn high 2 train +1478 AR6_WGI 1263 23 Averaged over the decade 2010–2019, the monthly Arctic sea ice area from August to October has been around 2 million km² (or about 25%) smaller than during 1979–1988 high 2 train +1479 AR6_WGI 1263 26 In the Bering Sea, expanding winter sea ice cover was observed until 2017 (Frey et al., 2015; Onarheim et al., 2018; Peng and Meier, 2018), but a marked reduction in sea ice concentration has occurred since then high 2 train +1480 AR6_WGI 1264 7 Since 1953, the years 2015 to 2018 had the four lowest values of maximum Arctic sea ice area, which usually occurs in March high 2 train +1481 AR6_WGI 1264 11 These records and other proposed paleo proxies, including bromine in ice cores (Spolaor et al., 2016), dinocyst assemblages (e.g., De Vernal et al., 2013b) and driftwood (e.g., Funder et al., 2011), provide evidence of sea ice fluctuations that exceed internal variability high 2 train +1482 AR6_WGI 1265 2 The SROCC assessed that approximately half of the satellite-observed Arctic summer sea ice loss is driven by increased concentrations of atmospheric greenhouse gases medium 1 train +1483 AR6_WGI 1265 10 In addition to changes in the external forcing, internal variability substantially affects Arctic sea ice, evidenced from both paleorecords (e.g., Chan et al., 2017; Hörner et al., 2017; Kolling et al., 2018) and satellites after 1979 (e.g., Notz and Stroeve, 2018; Roberts et al., 2020) high 2 train +1484 AR6_WGI 1265 20 In examining temperature thresholds for the loss of Arctic summer sea ice, the Special Report on Global Warming of 1.5°C (SR1.5; Hoegh-Guldberg et al., 2018) and SROCC assess that a reduction of September mean sea ice area to below 1 million km2, practically a sea ice-free Arctic Ocean, is more probable for a global mean warming of 2°C compared to global mean warming of 1.5°C high 2 train +1485 AR6_WGI 1265 22 Quantitatively, existing studies (Screen and Williamson, 2017; Jahn, 2018; Ridley and Blockley, 2018; Sigmond et al., 2018; Notz and SIMIP Community, 2020) also show that, for a warming between 1.5 and 2°C, the Arctic will only be practically sea ice free in September in some years, while at 3°C warming, the Arctic is practically sea ice free in September in most years, with longer practically sea ice-free periods at higher warming levels medium 1 train +1486 AR6_WGI 1267 8 In addition, there is no tipping point or critical threshold in global mean temperature beyond which the loss of summer sea ice becomes self-accelerating and irreversible high 2 train +1487 AR6_WGI 1267 11 The loss of winter sea ice is reversible as well, but the loss of winter sea ice area per degree of warming in CMIP5 and CMIP6 projections increases as the ice retreats from the continental shore lines, because these limit the possible areal fluctuations high 2 train +1488 AR6_WGI 1267 29 As assessed by SROCC, the evolution of mean Antarctic sea ice area is the result of opposing regional trends high 2 train +1489 AR6_WGI 1268 5 The changes in stratification result partly from surface freshening (De Lavergne et al., 2014), associated with increased northward sea ice advection (Haumann et al., 2020) and/or melting of the Antarctic ice sheet medium 1 test +1490 AR6_WGI 1268 6 In the Amundsen Sea, strong ice shelf melting can cause local sea ice melt next to the ice shelf front by entraining warm circumpolar deep water to the ice shelf cavity and surface ocean medium 1 train +1491 AR6_WGI 1269 14 Paleo-proxy data indicate that, on multi-decadal to multi-centennial time scales, sea ice coverage of the Southern Ocean follows large-scale temperature trends (e.g., Crosta et al., 2018; Chadwick et al., 2020; Lamping et al., 2020), for example linked to fluctuations in the El Niño–Southern Oscillation and Southern Annular Mode (Crosta et al., 2021), and that during the Last Glacial Maximum, Antarctic sea ice extended to about the polar front latitude in most regions during winter, whereas the extent during summer is less well understood (e.g., Benz et al., 2016; Xiao et al., 2016; Nair et al., 2019).Regionally, proxy data from ice cores consistently indicate that the increase of sea ice area in the Ross Sea and the decrease of sea ice area in the Bellingshausen Sea are part of longer centennial trends and exceed internal variability on multi-decadal time scales medium 1 train +1492 AR6_WGI 1269 25 Data from ICESat-1 laser altimetry (Kurtz and Markus, 2012), from Operation IceBridge (Kwok and Kacimi, 2018), and long-term shipboard observations collected in the Antarctic Sea Ice Processes and Climate (ASPeCt) dataset (Worby et al., 2008) suggest that sea ice thicker than 1 m prevails in regions of multi-year ice along the eastern coast of the Antarctic Peninsula in the Weddell Sea, in the high-latitude embayment of the Weddell Sea, and along the coast of the Amundsen Sea, with remaining regions dominated by thinner first-year sea ice high 2 train +1493 AR6_WGI 1272 6 In summary, the detailed regional records show an increase in mass loss in all regions after the 1980s, caused by both increases in discharge and decreases in SMB high 2 train +1494 AR6_WGI 1272 7 The largest mass loss occurred in the north-west and the south-east of Greenland high 2 train +1495 AR6_WGI 1276 1 The SROCC stated that surface processes, rather than ice discharged into the ocean, will dominate Greenland ice loss over the 21st century, regardless of the emissions scenario high 2 train +1496 AR6_WGI 1276 3 The projected mass loss of Greenland is predominantly due to increased surface meltwater and loss in refreezing capacity resulting in decreasing SMB high 2 train +1497 AR6_WGI 1277 22 The SROCC adopted the AR5 assessment that complete loss of Greenland ice, contributing about 7 m to sea level, over a millennium or more would occur for a sustained global mean surface temperature (GMST) between 1°C (low confidence) and 4°C medium 1 train +1498 AR6_WGI 1279 5 Mass loss of the West Antarctic and Antarctic Peninsula ice sheets has increased since about 2000 very high 3 train +1499 AR6_WGI 1279 11 In summary, WAIS losses, through acceleration, retreat and thinning of the principal outlet glaciers, dominated the AIS mass losses over the last decades very high 3 train +1500 AR6_WGI 1280 1 Furthermore, parts of the EAIS have lost mass in the last two decades high 2 train +1501 AR6_WGI 1280 2 As stated in SROCC, snowfall and glacier flow are the largest components determining AIS mass changes, with glacier flow acceleration (dynamic thinning) on the WAIS and the Antarctic Peninsula driving total loss trends in recent decades (very high confidence), and a partial offset of the dominating dynamic-thinning losses by increased snowfall high 2 train +1502 AR6_WGI 1281 6 In summary, ice-shelf thinning, mainly driven by basal melt, is widespread around the Antarctic coast and particularly strong around the WAIS high 2 train +1503 AR6_WGI 1281 28 In summary, the observed evolution of the ASE glaciers is compatible with, but not unequivocally indicating an ongoing MISI medium 1 train +1504 AR6_WGI 1285 28 The SROCC assessed that ice-sheet interactions with the solid Earth are not expected to substantially slow sea level rise from marine-based ice in Antarctica over the 21st century medium 1 train +1505 AR6_WGI 1285 32 Grounding line stabilization by the solid Earth response may therefore occur over the 21st century in the Amundsen Sea Embayment, where most mass loss is occurring (Section 9.4.2.1), but more generally occurs over multi-centennial to millennial time scales medium 1 train +1506 AR6_WGI 1288 12 In summary, it is likely that the AIS will continue to lose mass throughout this century under all emissions scenarios – that is, dynamic losses driven by ocean warming and ice-shelf disintegration will likely continue to outpace increasing snowfall medium 1 train +1507 AR6_WGI 1288 32 The SR1.5 (Hoegh-Guldberg et al., 2018) assessed that a threshold for WAIS instability may be close to 1.5°C–2°C medium 1 test +1508 AR6_WGI 1289 9 Overall, increased evidence and agreement on the time scales and drivers of mass loss confirm the SR1.5 assessment that a threshold for WAIS instability may be close to 1.5°C–2°C (medium confidence), and that the probability of passing a threshold is larger for 2°C warming than for 1.5°C medium 1 train +1509 AR6_WGI 1289 10 New projections agree with previous studies that only part of WAIS would be lost on multi-century time scales if warming remains less than 2°C medium 1 train +1510 AR6_WGI 1289 12 Under around 2°C–3°C peak warming, complete or near-complete loss of the WAIS is projected in most studies after multiple millennia, with continent-wide mass losses of around 2–5 m SLE or more; this could occur on multi-century time scales under very high basal melting (Lipscomb et al., 2021) or widespread ice-shelf loss and/or MICI low 0 test +1511 AR6_WGI 1289 14 If warming exceeds around 3°C above pre-industrial, part of the EAIS (typically the Wilkes Subglacial Basin) is projected to be lost on multi-millennial time scales low 0 train +1512 AR6_WGI 1289 17 The SROCC (Meredith et al., 2019; Oppenheimer et al., 2019) assessed that Antarctic mass losses could be irreversible over decades to millennia low 0 train +1513 AR6_WGI 1289 22 New research therefore confirms SROCC assessment that mass loss from the AIS is irreversible on decadal to millennial time scales (low confidence) (FAQ 9.1), and suggests that reducing atmospheric CO 2 concentrations or temperatures to pre-industrial levels may not be sufficient to prevent or reverse substantial Antarctic mass losses low 0 train +1514 AR6_WGI 1291 15 Including the peripheral glaciers of the ice sheets, the global glacier mass loss rate in the period 2000–2019 is very likely 266 ± 16 Gt yr –1 (4 [3 to 6] % of glacier mass in 2000) with an increase in the mass loss rate from 240 ± 9 Gt yr –1 in 2000–2009 to 290 ± 10 Gt yr –1 in 2010–2019 high 2 train +1515 AR6_WGI 1291 17 In summary, new evidence published since SROCC shows that, during the decade 2010–2019, glaciers lost more mass than in any other decade since the beginning of the observational record very high 3 train +1516 AR6_WGI 1292 7 For all regions with long-term observations, glacier mass in the decade 2010–2019 was the smallest since at least the beginning of the 20th century medium 1 train +1517 AR6_WGI 1292 12 There is limited evidence to assess whether the Karakoram anomaly will persist in coming decades but, due to the projected increase in air temperature throughout the region, its long-term persistence is unlikely high 2 train +1518 AR6_WGI 1295 3 The SROCC assessed that global glacier mass loss by 2100, relative to 2015 will be 18 [likely range 11 to 25] % for scenario RCP2.6 and 36 [likely range 26 to 47] % for RCP8.5, and that many glaciers will disappear regardless of the emissions scenario very high 3 train +1519 AR6_WGI 1295 6 Glaciers will lose 29,000 [9000 to 49,000] Gt and 58,000 [28,000 to 88,000] Gt over the period 2015–2100 for RCP2.6 and RCP8.5, respectively medium 1 train +1520 AR6_WGI 1296 4 In summary, both global and regional studies agree that glacier mass loss will continue in all regions, with larger mass loss for high-emissions scenarios high 2 train +1521 AR6_WGI 1296 19 It is also found in high-altitude areas of mountain ranges in both hemispheres – estimated in SROCC (Hock et al., 2019b) as representing about 27–29% of the global permafrost area medium 1 train +1522 AR6_WGI 1300 7 Analysis of paleoclimate records (Pederson et al., 2011; Belmecheri et al., 2016) suggests that recent snowpack reductions in western North America are exceptional on a millennial time scale medium 1 train +1523 AR6_WGI 1303 10 In summary, consistent projections from all generations of global climate models, elementary process understanding and strong covariance between snow cover and temperature on several time scales make it virtually certain that future Northern Hemisphere snow cover extent and duration will continue to decrease as global climate continues to warm, and process understanding strongly suggests that this also applies to Southern Hemisphere seasonal snow cover high 2 train +1524 AR6_WGI 1303 18 Based on these updated data and methods, the GMSL change over the (pre-satellite) period 1901–1990 is assessed to be 0.12 [0.07 to 0.17, very likely range] m with an average rate of 1.35 [0.78 to 1.92, very likely range] mm yr –1 high 2 train +1525 AR6_WGI 1303 20 The SROCC found that four of the five available tide gauge reconstructions that extend back to at least 1902 showed a robust acceleration high 2 train +1526 AR6_WGI 1304 16 For the periods 1901–1990 and 1901–2018, the assessed very likely range for the sum of components is found to be consistent with the assessed very likely range of observed GMSL change medium 1 train +1527 AR6_WGI 1304 19 While the combined mass loss for Greenland and glaciers is consistent with SROCC, updates in the underlying datasets lead to differences in partitioning of the mass loss.9.6.1.2 Global Mean Sea Level Change Budget in the Satellite Era The SROCC (Oppenheimer et al., 2019) concluded that GMSL increased at a rate of 3.16 [2.79 to 3.53, very likely range] mm yr –1 in the period 1993–2015 (the satellite altimetry era), and a rate of 3.58 [3.10 to 4.06, very likely range] mm yr –1 in the period 2006–2015 – the Gravity Recovery and Climate Experiment (GRACE)/Argo data era high 2 train +1528 AR6_WGI 1304 20 An updated assessment for the periods 1993–2018 and 2006–2018 yields values of 3.25 [2.88 to 3.61] and 3.69 [3.21 to 4.17] mm yr –1 high 2 train +1529 AR6_WGI 1304 21 Based on the GMSL assessed time series presented in Section 2.3.3.3, GMSL acceleration is estimated as 0.075 [0.066 to 0.080] mm yr –2 for 1971–2018 and 0.094 [0.082–0.115] mm yr –2 for 1993–2018 high 2 train +1530 AR6_WGI 1304 26 For both periods in the satellite era – that is, 1993–2018 and 2006–2018 – the sum of contributions is consistent with the total observed GMSL change high 2 train +1531 AR6_WGI 1304 31 Because of the increased ice-sheet mass loss, the total loss of land ice (glaciers and ice sheets) was the largest contributor to GMSL rise over the period 2006–2018 high 2 train +1532 AR6_WGI 1306 17 In summary, the regional sea level budget, using either observations or models, can currently only be closed on basin scales medium 1 train +1533 AR6_WGI 1307 14 Adding the projected sea level change from land ice mass loss and groundwater extraction strengthens and modifies the forced signal, leading to times of emergence 10 to 20 years earlier in most parts of the ocean, except in regions close to sources of mass loss, with emergence over 50% of the ocean area by 2020, and nearly everywhere by 2100 medium 1 train +1534 AR6_WGI 1307 15 In summary, detection of forced regional changes for some ocean areas in recent decades is possible medium 1 train +1535 AR6_WGI 1308 6 The SROCC assessment that past multi-metre sea level changes have resulted from significant ice-sheet changes beyond those presently observed is confirmed (very high confidence).Cross-Chapter Box 9.1 (continued) The Earth system gained substantial energy over the period 1971–2018 high 2 train +1536 AR6_WGI 1308 7 Ocean warming dominates the energy inventory change high 2 train +1537 AR6_WGI 1308 14 A related assessment presented in Chapter 7 demonstrates closure of the global energy budget high 2 train +1538 AR6_WGI 1309 3 Partly on the basis of these studies, SROCC proposed a ‘plausible’ upper bound on GMSL of 25 m low 0 train +1539 AR6_WGI 1309 4 Ice-sheet model simulations indicate that Northern Hemisphere glaciation was limited to high-elevation regions in eastern and southern Greenland medium 1 train +1540 AR6_WGI 1309 5 Southern Hemisphere glaciation was characterized by an Antarctic Ice Sheet (AIS) reduced in volume from the present medium 1 train +1541 AR6_WGI 1309 7 In summary, under a past warming level of around 2.5°C–4°C, ice sheets in both hemispheres were reduced in extent compared to present high 2 train +1542 AR6_WGI 1309 8 Proxy-based evidence (Section 2.3.3.3) combined with numerical modelling indicates that, on millennial time scales, the GMSL contribution arising from ice sheets was >5 m (high confidence) or >10 m medium 1 train +1543 AR6_WGI 1309 12 The GMSL was 6–13 m above present medium 1 train +1544 AR6_WGI 1309 16 In summary, geological reconstructions and numerical simulations consistently show that past warming levels of <2°C (GMST) are sufficient to trigger multi- metre mass loss from both the Greenland and Antarctic ice sheets if maintained for millennia high 2 train +1545 AR6_WGI 1309 21 Since AR5, information has improved about the LIG, when GMST was about 0.5°C–1.5°C above 1850–1900 medium 1 train +1546 AR6_WGI 1309 24 It is virtually certain that GMSL was higher than today, likely by 5–10 m medium 1 train +1547 AR6_WGI 1310 15 Of the residual, up to about 1.4 m may be ascribed to groundwater, leaving a shortfall of 16 ± 10 m yet to be allocated among land ice reservoirs or lakes.9.6.2.5 Last Deglacial Transition: Meltwater pulse 1A During Meltwater pulse 1A (MWP-1A), GMSL very likely medium 1 train +1548 AR6_WGI 1310 22 In summary, MWP-1A appears to have been driven by a combination of melt in North America (high confidence), Eurasia (low confidence), and Antarctica low 0 train +1549 AR6_WGI 1310 32 At 6 ka, GMSL was –3.5 to +0.5 m medium 1 train +1550 AR6_WGI 1311 1 Geochronological and numerical modelling studies indicate that it is likely medium 1 train +1551 AR6_WGI 1311 11 In summary, higher-than-pre-industrial GMST during the mid-Holocene coincided with recession of the Greenland Ice Sheet to a smaller-than-present extent high 2 train +1552 AR6_WGI 1311 14 In summary, both proxies and model simulations indicate that GMSL changes during the early to mid-Holocene were the result of episodic pulses, due to drainage of meltwater lakes, superimposed on a trend of steady rise due to continued ice-sheet retreat high 2 train +1553 AR6_WGI 1311 15 The combination of tide gauge observations and geological reconstructions indicates that a sustained increase of GMSL began between 1820–1860 and led to a 20th-century GMSL rise that was very likely high 2 train +1554 AR6_WGI 1311 19 In summary, rates of RSL rise exceeding the pre-industrial background rate of rise are apparent in parts of the North Atlantic in the 19th century (medium confidence), and in most of the world in the 20th century high 2 train +1555 AR6_WGI 1313 2 Considering also projections incorporating MICI or SEJ low 0 train +1556 AR6_WGI 1316 21 RSL rise associated with GRD is very likely to be largest in the Pacific, due to the combined effects of projected GrIS, AIS and glacier mass loss high 2 train +1557 AR6_WGI 1316 24 In parts of north-eastern North America and north-western Europe, GRD effects associated with mass loss from southern Greenland will lead to an RSL fall, whereas mass loss from northern Greenland will lead to an RSL rise high 2 train +1558 AR6_WGI 1318 13 In particular, under SSP5-8.5, low-confidence processes could lead to a total GMSL rise of 0.6–1.6 m over this time period (17th–83rd percentile range of p-box, including SEJ- and MICI-based projections), with 5th–95th percentile projections extending to 0.5–2.3 m low 0 train +1559 AR6_WGI 1319 5 Approximately 60% (SSP1-1.9) to 70% (SSP5-8.5) of the global coastline has a projected median 21st century regional RSL rise within ±20% of the global mean m 1950 2000 2050 2100 2150 00.511.522.5 SSP5-8.5 SSP3-7.0 SSP2-4.5 SSP1-2.6 SSP1-1.9 HistoricalMedian (medium confidence) Likely range medium 1 train +1560 AR6_WGI 1320 1 Consistent with AR5, loss of land ice mass will be an important contributor to spatial patterns in RSL change (high confidence), with ocean dynamic sea level being particularly important as a dipolar contributor in the north-west Atlantic, a positive contributor in the Arctic Ocean, and a negative contributor in the Southern Ocean south of the Antarctic Circumpolar Current medium 1 train +1561 AR6_WGI 1320 2 As today, VLM will remain a major driver of RSL change high 2 train +1562 AR6_WGI 1320 3 Uncertainty in RSL projections is greatest in tectonically active areas in which VLM varies over short distances (e.g., Alaska) and in areas potentially subject to large ocean dynamic sea level change (e.g., the north-western Atlantic) high 2 train +1563 AR6_WGI 1321 12 The SROCC projected 0.6–1.0 m under RCP2.6 and 2.3–5.3 m under RCP8.5 low 0 train +1564 AR6_WGI 1322 2 Under SSP5-8.5, GMSL will rise between 1.7 and 6.8 m by 2300 in the absence of MICI and by up to 16 m considering MICI, a wider range than AR5 or SROCC assessments, but consistent with published projections low 0 train +1565 AR6_WGI 1323 3 On the basis of modelling studies, paleo constraints, single-ice-sheet studies finding multimillennial nonlinear responses from both the Greenland and Antarctic ice sheets (Sections 9.4.1.4 and 9.4.2.6), and the underlying physics, we conclude that GMSL commitment is nonlinear in peak warming on time scales of both 2,000 and 10,000 years medium 1 train +1566 AR6_WGI 1323 4 Although thermosteric sea level will start to decline slowly about 2,000 years after emissions cease, the slower responses from the Greenland and Antarctic ice sheets mean that GMSL will continue to rise for 10,000 years under most scenarios medium 1 train +1567 AR6_WGI 1323 10 Carbon dioxide removal, solar radiation modification, and local ice-shelf engineering may be effective at reducing the yet-to-be- realized sea level commitment, but ineffective at reversing GMSL rise low 0 test +1568 AR6_WGI 1324 6 In doing so, we note upfront that the main uncertainty related to high-end sea level rise is ‘when’ rather than ‘if’ it arises: the upper limit of 1.01 m of likely sea level range by 2100 for the SSP5-8.5 scenario will be exceeded in any future warming scenario on time scales of centuries to millennia high 2 train +1569 AR6_WGI 1325 6 As in AR5 and SROCC, tide gauge observations show that RSL rise (Section 9.6.1.3) is the primary driver of changes in ESWL at most locations and, across tide gauges, has led to a median 165% increase in high-tide flooding over 1995–2014 relative to those over 1960–1980 high 2 train +1570 AR6_WGI 1325 7 Some locations exhibit substantial differences between long-term RSL trends and ESWL high 2 train +1571 AR6_WGI 1325 11 Failing to consider the non-linear interactions between tide, surge and RSL may overestimate trends in ESWL low 0 train +1572 AR6_WGI 1326 3 Combined, observations and models indicate RSL rise and direct anthropogenic factors are the primary drivers of observed tidal changes at tide gauge stations medium 1 train +1573 AR6_WGI 1326 17 On the basis of satellite altimeter observations, SROCC reported increasing extreme wave heights in the Southern and North Atlantic oceans of around 1.0 and 0.8 cm yr –1, respectively, over the period 1985–2018 medium 1 train +1574 AR6_WGI 1326 18 The SROCC (Collins et al., 2019) also identified sea ice loss in the Arctic as leading to increased wave heights over the period 1992–2014 medium 1 train +1575 AR6_WGI 1326 26 However, sensitivity of processing techniques, inadequate spatial distribution of observations, and homogeneity issues in available records limit confidence in reported trends medium 1 train +1576 AR6_WGI 1326 28 Detection and attribution of the human influence on climatic changes in surges, and waves remains a challenge (Ceres et al., 2017), with limited evidence to suggest in some instances – for example, poleward migration of tropical cyclones in the Western North Pacific (Section 11.7.1.2), changes in surges and waves can be attributed to anthropogenic climate change low 0 train +1577 AR6_WGI 1326 30 The close relationship between local ESL and long-term RSL change, combined with the robust attribution of GMSL change (Section 9.6.1.4), implies that observed global changes in ESL can be attributed, at least in part, to human-caused climate change medium 1 train +1578 AR6_WGI 1328 8 The SROCC showed that currently rare ESL events (e.g., with an average return period of 100 years) will occur annually or more frequently at most available locations for RCP4.5 by the end of the century high 2 train +1579 AR6_WGI 1328 17 By 2100, the median frequency amplification factor is projected to be 163 for SSP1-2.6, 325 for SSP2-4.5, and 532 for SSP5-8.5, with respectively 60%, 71%, and 82% of the stations experiencing a currently 1% annual probability event at least yearly medium 1 train +1580 AR6_WGI 1328 28 In summary, despite waves and surges being non-negligible contributors to projected ETWL and ECWL changes (Vousdoukas et al., 2018; Melet et al., 2020), RSL change is expected to be the main driver in changes in future ESL return periods in most areas medium 1 train +1581 AR6_WGI 1328 32 At local and regional scales, anthropogenic factors such as major land reclamation efforts, as in the East China Sea (Song et al., 2013) or differing national coastal management strategies (maintaining the present coastline position or managed retreat) will locally modulate the influence of GMSL rise on tidal amplitude medium 1 train +1582 AR6_WGI 1330 11 Thus, projections of ESL that do not consider correlations between and among sea level forced and atmospherically forced drivers can differ strongly from coupled projections medium 1 train +1583 AR6_WGI 1330 12 The SROCC (Collins et al., 2019) highlighted compound events, or coincident occurrence of multiple hazards, as an example of deep uncertainty, and noted that failing to account for multiple factors contributing to extreme events will lead to underestimation of the probabilities of occurrence high 2 train +1584 AR6_WGI 1382 10 Distilling regional climate information from multiple lines of evidence and taking the user context into account will increase the fitness, usefulness and relevance for decision-making and enhances the trust users will have in applying it high 2 train +1585 AR6_WGI 1382 13 Taking the values of the relevant actors into account when co-producing climate information, and translating this information into the broader user context, improves the usefulness and uptake of this information high 2 train +1586 AR6_WGI 1382 15 The availability of multiple observational records, including reanalyses, that are fit for evaluating the phenomena of interest and account for observational uncertainty, are fundamental for both understanding past regional climate change and assessing climate model performance at regional scales high 2 train +1587 AR6_WGI 1382 16 Employing tailored, process-oriented and potentially multivariate diagnostics to evaluate whether a climate model realistically simulates relevant aspects of present-day regional climate increases trust in future projections of these aspects high 2 train +1588 AR6_WGI 1382 18 Precipitation measurements in mountainous areas, especially of solid precipitation, are strongly affected by gauge location and setup very high 3 train +1589 AR6_WGI 1382 23 Regional reanalyses represent the distributions of precipitation, surface air temperature, and surface wind, including the frequency of extremes, better than global reanalyses high 2 train +1590 AR6_WGI 1382 26 Global models by themselves provide a useful line of evidence for the construction of regional climate information through the attribution or projection of forced changes or the quantification of the role of the internal variability high 2 train +1591 AR6_WGI 1382 27 Dynamical downscaling using regional climate models adds value in representing many regional weather and climate phenomena, especially over regions of complex orography or with heterogeneous surface characteristics very high 3 test +1592 AR6_WGI 1382 28 Increasing climate model resolution improves some aspects of model performance high 2 train +1593 AR6_WGI 1382 29 Some local-scale phenomena such as land–sea breezes and mountain wind systems can only be realistically represented by simulations at a resolution of the order of 10 km or finer high 2 train +1594 AR6_WGI 1382 30 Simulations at kilometre-scale resolution add value in particular to the representation of convection, sub-daily precipitation extremes (high confidence) and soil-moisture– precipitation feedbacks medium 1 train +1595 AR6_WGI 1383 4 Including the relevant forcings (e.g., aerosols, land-use change and stratospheric ozone concentrations) and representing the relevant feedbacks (e.g., snow–albedo, soil-moisture–temperature, soil-moisture–precipitation) in global and regional models is a prerequisite for reproducing historical regional trends and ensuring fitness for future projections high 2 train +1596 AR6_WGI 1383 5 The sign of projected regional changes of variables such as precipitation and wind speed is in some cases only simulated in a trustworthy manner if relevant regional processes are represented medium 1 train +1597 AR6_WGI 1383 7 Statistical downscaling methods with carefully chosen predictors and an appropriate model structure for a given application realistically represent many statistical aspects of present-day daily temperature and precipitation high 2 train +1598 AR6_WGI 1383 8 Bias adjustment has proven beneficial as an interface between climate model projections and impact modelling in many different contexts high 2 train +1599 AR6_WGI 1383 9 Weather generators realistically simulate many statistical characteristics of present-day daily temperature and precipitation, such as extreme temperatures and wet- and dry-day transition probabilities high 2 train +1600 AR6_WGI 1383 12 Bias adjustment cannot overcome all consequences of unresolved or strongly misrepresented physical processes, such as large-scale circulation biases or local feedbacks, and may instead introduce other biases and implausible climate change signals medium 1 train +1601 AR6_WGI 1383 13 Using bias adjustment as a method for statistical downscaling, particularly for coarse-resolution global models, may lead to substantial misrepresentations of regional climate and climate change medium 1 train +1602 AR6_WGI 1383 15 The performance of statistical approaches and their fitness for future projections depends on predictors and change factors taken from the driving dynamical models high 2 train +1603 AR6_WGI 1383 17 Multi-model ensembles enable the assessment of regional climate response uncertainty very high 3 train +1604 AR6_WGI 1383 18 Discarding models that fundamentally misrepresent processes relevant for a given purpose improves the fitness of multi-model ensembles for generating regional climate information high 2 train +1605 AR6_WGI 1383 19 At the regional scale, multi-model mean and ensemble spread are not sufficient to characterize low-likelihood, high-impact changes or situations where different models simulate substantially different or even opposing changes high 2 train +1606 AR6_WGI 1383 21 Since AR5, the availability of multiple single-model initial-condition large ensembles (SMILEs) allows for a more robust separation of model uncertainty and internal variability in regional-scale projections and provides a more comprehensive spectrum of possible changes associated with internal variability high 2 train +1607 AR6_WGI 1383 25 Multiple attribution approaches, including optimal fingerprinting, grid-point detection, pattern recognition and dynamical adjustment methods, as well as multi-model, single-forcing large ensembles and multi-centennial paleoclimate records, support the contribution of human influence to several regional multi-decadal mean precipitation changes high 2 train +1608 AR6_WGI 1383 27 Multiple lines of evidence, combining multi-model ensemble global projections with those coming from SMILEs, show that internal variability is largely contributing to the delayed or absent emergence of the anthropogenic signal in long-term regional mean precipitation changes high 2 train +1609 AR6_WGI 1384 6 Annual mean daily minimum temperature is more affected by urbanization than annual mean daily maximum temperature very high 3 train +1610 AR6_WGI 1384 7 The global annual mean surface air temperature response to urbanization is, however, negligible very high 3 train +1611 AR6_WGI 1384 9 A large effect is expected from the combination of future urban development and more frequent occurrence of extreme climatic events, such as heatwaves very high 3 train +1612 AR6_WGI 1384 14 For example, the apparent contradiction between the observed decrease in Indian monsoon rainfall over the second half of the 20th century and the projected long-term increase is explained by attribution of the trends to different forcings, with aerosols dominating recently and greenhouse gases in the future high 2 train +1613 AR6_WGI 1393 10 Further, the chapter identified that the regional climate information was limited by the paucity of comprehensive observations and their analysis along with the different levels of confidence in projections high 2 train +1614 AR6_WGI 1394 2 Part of the projected increase in heatwaves and droughts can be attributed to soil moisture feedbacks in regions where evapotranspiration is limited by moisture availability medium 1 train +1615 AR6_WGI 1394 3 Vegetation changes can also amplify or dampen extreme events through changes in albedo and evapotranspiration, which will influence future trends in extreme events medium 1 train +1616 AR6_WGI 1394 8 Urbanization increases the risks associated with extreme events high 2 train +1617 AR6_WGI 1394 9 Urbanization suppresses evaporative cooling and amplifies heatwave intensity (high confidence) with a strong influence on minimum temperatures high 2 train +1618 AR6_WGI 1394 12 The SROCC assessed that it is virtually certain that Antarctica and Greenland have lost mass over the past decade and observed glacier mass loss over the last decades is attributable to anthropogenic climate change high 2 train +1619 AR6_WGI 1395 3 Averaged over the decade 2010–2019, monthly average sea ice area in August, September and October has been about 25% smaller than during 1979–1988 high 2 train +1620 AR6_WGI 1398 13 Their observations are valuable high 2 train +1621 AR6_WGI 1398 20 Availability of active sensors on LEO satellites enables measurement of microphysical properties of aerosol, cloud and precipitation, which can advance regional climate studies and process evaluation studies to improve regional climate models high 2 train +1622 AR6_WGI 1400 4 The regional reanalyses represent the frequencies of extremes and the distributions of precipitation, surface air temperature, and surface wind better than global reanalyses high 2 train +1623 AR6_WGI 1400 7 Regional reanalyses can provide estimates that are more consistent with observations than dynamical downscaling approaches, due to the assimilation of additional local observations high 2 train +1624 AR6_WGI 1400 23 Quality-controlled, high-resolution observational datasets are especially needed at regional and local scales to assess models as their resolution increases (Di Luca et al., 2016; Zittis and Hadjinicolaou, 2017), although the awareness and appropriate use of the QC information is challenging (Tapiador et al., 2017) when generating regional climate information high 2 train +1625 AR6_WGI 1402 13 Gridded products of climate variables, including temperature and precipitation, are strongly affected high 2 train +1626 AR6_WGI 1404 16 This technological trend could prove very useful high 2 train +1627 AR6_WGI 1423 14 Realistically representing local-scale phenomena such as land–sea breezes requires simulations at a resolution of the order of 10 km high 2 train +1628 AR6_WGI 1423 15 Simulations at kilometre-scale resolution add value in particular to the representation of convection, sub-daily summer precipitation extremes (high confidence) and soil-moisture–precipitation feedbacks medium 1 train +1629 AR6_WGI 1423 16 Resolving regional processes may be required to correctly represent the sign of regional climate change medium 1 train +1630 AR6_WGI 1423 22 Bias adjustment has proven beneficial as an interface between climate model projections and impact modelling in many different contexts high 2 test +1631 AR6_WGI 1423 23 Weather generators realistically simulate many statistical aspects of present-day daily temperature and precipitation high 2 train +1632 AR6_WGI 1423 24 The performance of these approaches and their fitness for future projections also depends on predictors and change factors taken from the driving dynamical models high 2 train +1633 AR6_WGI 1443 17 An accurate estimation of the delay in regional-scale emergence caused by internal variability remains challenging due to global model biases in their representation of internal variability as well as methodological difficulties to precisely estimate these biases high 2 train +1634 AR6_WGI 1449 11 This process enhances trust in the information as well its usefulness, relevance, and uptake, especially when the communication involves complex, contextual details high 2 train +1635 AR6_WGI 1449 13 Drawing upon multiple lines of evidence in the construction of climate information increases the fitness of this information and creates a stronger foundation high 2 train +1636 AR6_WGI 1453 8 Projections There is high confidence that by the end of century most regions in South America will undergo extreme heat stress conditions much more often than in the recent past, with about 50–100 more days per year under SSP1-2.6 and more than 200 additional days per year under SSP5-8.5 high 2 train +1637 AR6_WGI 1453 15 There is an increase in probability and intensity of agricultural and ecological droughts (medium confidence) and there is an increase in frequency and severity of hydrological droughts high 2 train +1638 AR6_WGI 1453 16 Attribution Global warming has contributed to drying in dry summer climates including the Mediterranean high 2 train +1639 AR6_WGI 1453 18 Multiple lines of evidence suggest that anthropogenic forcings are causing increased aridity and drought severity in the Mediterranean region high 2 train +1640 AR6_WGI 1454 24 These factors contribute to attribution of historical climate change signals (Section 10.4), recognizing that attribution must account for the interplay between externally forced Cross-Chapter Box 10.3 (continued) An increasing trend towards agricultural and ecological droughts has been attributed to human-induced climate change in the Mediterranean medium 1 train +1641 AR6_WGI 1454 28 Increased evapotranspiration due to growing atmospheric water demand will decrease soil moisture high 2 train +1642 AR6_WGI 1454 29 The seasonality of runoff and streamflow (the annual difference between the wettest and driest months of the year) is expected to increase with global warming high 2 train +1643 AR6_WGI 1454 32 The paleoclimate record provides context for these future expected changes: climate change will shift soil moisture outside the range of observed and reconstructed values spanning the last millennium high 2 train +1644 AR6_WGI 1463 10 In summary, long-term future scenarios dominated by GHG increases (such as the RCPs) suggest increases in Indian summer monsoon rainfall high 2 train +1645 AR6_WGI 1472 5 The SROCC assessed that HKH (named High Mountain Asia) surface-air temperature has warmed more rapidly than the global mean over recent decades high 2 train +1646 AR6_WGI 1474 6 Glacier trends The SROCC assessed that snow cover has declined in duration, depth and accumulated mass at lower elevations in mountain regions, including the HKH high 2 train +1647 AR6_WGI 1474 7 Glaciers are losing mass (very high confidence) and permafrost is warming high 2 train +1648 AR6_WGI 1474 21 The SR1.5 (IPCC, 2018b) stated that heavy precipitation risk in high-elevation regions is projected to be higher at 2°C compared to 1.5°C of global warming medium 1 train +1649 AR6_WGI 1474 27 The SROCC assessed that glaciers will lose substantial mass (high confidence) and permafrost will undergo increasing thaw and degradation very high 3 train +1650 AR6_WGI 1474 33 Glacier mass in HKH will decline through the 21st century high 2 train +1651 AR6_WGI 1533 12 New evidence strengthens the conclusion from the IPCC Special Report on Global Warming of 1.5°C (SR1.5) that even relatively small incremental increases in global warming (+0.5°C) cause statistically significant changes in extremes on the global scale and for large regions high 2 train +1652 AR6_WGI 1533 13 In particular, this is the case for temperature extremes (very likely), the intensification of heavy precipitation (high confidence) including that associated with tropical cyclones (medium confidence), and the worsening of droughts in some regions high 2 train +1653 AR6_WGI 1533 17 Projected percentage changes in frequency are higher for the rarer extreme events high 2 train +1654 AR6_WGI 1533 19 There have been improvements in some observation-based datasets, including reanalysis data high 2 train +1655 AR6_WGI 1533 20 Climate models can reproduce the sign (direction) of changes in temperature extremes observed globally and in most regions, although the magnitude of the trends may differ high 2 train +1656 AR6_WGI 1533 21 Models are able to capture the large-scale spatial distribution of precipitation extremes over land high 2 train +1657 AR6_WGI 1533 22 The intensity and frequency of extreme precipitation simulated by Coupled Model Intercomparison Project Phase 6 (CMIP6) models are similar to those simulated by CMIP5 models high 2 train +1658 AR6_WGI 1533 23 Higher horizontal model resolution improves the spatial representation of some extreme events (e.g., heavy precipitation events), in particular in regions with highly varying topography high 2 train +1659 AR6_WGI 1533 30 Irrigation and crop expansion have attenuated increases in summer hot extremes in some regions, such as the Midwestern USA medium 1 train +1660 AR6_WGI 1534 5 In most regions, future changes in the intensity of temperature extremes will very likely be proportional to changes in global warming, and up to two to three times larger high 2 train +1661 AR6_WGI 1534 6 The highest increase of temperature of hottest days is projected in some mid-latitude and semi-arid regions and in the South American Monsoon region, at about 1.5 times to twice the rate of global warming high 2 train +1662 AR6_WGI 1534 7 The highest increase of temperature of coldest days is projected in Arctic regions, at about three times the rate of global warming high 2 train +1663 AR6_WGI 1534 14 Evidence of a human influence on heavy precipitation has emerged in some regions high 2 train +1664 AR6_WGI 1534 18 At the global scale, the intensification of heavy precipitation will follow the rate of increase in the maximum amount of moisture that the atmosphere can hold as it warms high 2 test +1665 AR6_WGI 1534 19 The increase in the frequency of heavy precipitation events will be non-linear with more warming and will be higher for rarer events high 2 train +1666 AR6_WGI 1534 20 Increases in the intensity of extreme precipitation at regional scales will vary, depending on the amount of regional warming, changes in atmospheric circulation and storm dynamics high 2 test +1667 AR6_WGI 1534 23 The seasonality of river floods has changed in cold regions where snow-melt is involved, with an earlier occurrence of peak streamflow high 2 train +1668 AR6_WGI 1534 31 Increases in evapotranspiration have been driven by increases in atmospheric evaporative demand induced by increased temperature, decreased relative humidity and increased net radiation high 2 train +1669 AR6_WGI 1534 32 Trends in precipitation are not a main driver in affecting global-scale trends in drought medium 1 train +1670 AR6_WGI 1535 2 Regional- scale attribution shows that human-induced climate change has contributed to increased agricultural and ecological droughts (MED, WNA), and increased hydrological drought (MED) in some regions medium 1 train +1671 AR6_WGI 1535 4 Several regions will be affected by more severe agricultural and ecological droughts even if global warming is stabilised at 2°C, including MED, WSAF, SAM and SSA (high confidence), and ESAF, MDG, EAU, SAU, SCA, CAR, NSA, NES, SWS, WCE, NCA, WNA and CNA medium 1 train +1672 AR6_WGI 1535 6 At 4°C of global warming, about 50% of all inhabited AR6 regions would be affected by increases in agricultural and ecological droughts (WCE, MED, CAU, EAU, SAU, WCA, EAS, SCA, CAR, NSA, NES, SAM, SWS, SSA, NCA, CNA, ENA, WNA, WSAF , ESAF , MDG: medium confidence or higher), and only two regions (NEAF , SAS) would experience decreases in agricultural and ecological drought medium 1 train +1673 AR6_WGI 1535 14 Peak TC rain rates increase with local warming at least at the rate 2 Six-hourly intensity estimates during the lifetime of each TC.of mean water vapour increase over oceans (about 7% per 1°C of warming) and in some cases exceeding this rate due to increased low-level moisture convergence caused by increases in TC wind intensity medium 1 train +1674 AR6_WGI 1535 19 None of these changes can be explained by natural variability alone medium 1 train +1675 AR6_WGI 1535 20 The proportion of intense TCs, average peak TC wind speeds, and peak wind speeds of the most intense TCs will increase on the global scale with increasing global warming high 2 train +1676 AR6_WGI 1535 21 The total global frequency of TC formation will decrease or remain unchanged with increasing global warming medium 1 train +1677 AR6_WGI 1535 23 Future wind speed changes are expected to be small, although poleward shifts in the storm tracks could lead to substantial changes in extreme wind speeds in some regions medium 1 train +1678 AR6_WGI 1535 25 The frequency of spring severe convective storms is projected to increase in the USA, leading to a lengthening of the severe convective storm season medium 1 train +1679 AR6_WGI 1535 28 Concurrent heatwaves and droughts have become more frequent, and this trend will continue with higher global warming high 2 train +1680 AR6_WGI 1535 29 Fire weather conditions (compound hot, dry and windy events) have become more probable in some regions medium 1 train +1681 AR6_WGI 1535 30 The probability of compound flooding (storm surge, extreme rainfall and/or river flow) has increased in some locations (medium confidence), and will continue to increase due to sea level rise and increases in heavy precipitation, including changes in precipitation intensity associated with tropical cyclones high 2 train +1682 AR6_WGI 1535 31 The land area affected by concurrent extremes has increased high 2 train +1683 AR6_WGI 1536 3 Compound events, including concurrent extremes, are a factor increasing the probability of low- likelihood, high-impact events high 2 train +1684 AR6_WGI 1536 4 With increasing global warming, some compound events with low likelihood in past and current climates will become more frequent, and there is a higher chance of occurrence of historically unprecedented events and surprises high 2 train +1685 AR6_WGI 1536 5 However, even extreme events that do not have a particularly low probability in the present climate (at more than 1°C of global warming) can be perceived as surprises because of the pace of global warming high 2 train +1686 AR6_WGI 1543 9 Anthropogenic forcing (e.g., increases in greenhouse gas concentrations) directly affects thermodynamic variables, including overall increases in high temperatures and atmospheric evaporative demand, and regional changes in atmospheric moisture, which intensify heatwaves, droughts and heavy precipitation events when they occur high 2 train +1687 AR6_WGI 1544 23 Land use can affect regional extremes, in particular hot extremes, in several ways high 2 train +1688 AR6_WGI 1544 28 Aerosol forcing also has a strong regional footprint associated with regional emissions, which affects temperature and precipitation extremes high 2 train +1689 AR6_WGI 1545 14 These can also lead to a higher warming of extreme temperatures compared to mean temperature (high confidence), and possibly cooling in some regions medium 1 train +1690 AR6_WGI 1549 5 Phenomenon and Direction of TrendProjected Changes at +1.5ºC Global WarmingProjected Changes at +2°C Global WarmingProjected Changes at +4°C Global Warming Warmer and/or more frequent hot days and nights over most land areas Warmer and/or fewer cold days and nights over most land areas Warm spells/heatwaves; increases in frequency or intensity over most land areas Cold spells/cold waves: decreases in frequency or intensity over most land areasVirtually certain on global scale Extremely likely on all continents Highest increase of temperature of hottest days is projected in some mid-latitude and semi-arid regions, and the South American Monsoon region, at about 1.5 times to twice the rate of global warming (high confidence) {11.3, Figure 11.3} Highest increase of temperature of coldest days is projected in Arctic regions, at about three times the rate of global warming (high confidence) {11.3} Continental-scale projections: Extremely likely : Africa, Asia, Australasia, Central and South America, Europe, North America {11.3, 11.9}Virtually certain on global scale Virtually certain on all continents Highest increase of temperature of hottest days is projected in some mid-latitude and semi-arid regions, and the South American Monsoon region, at about 1.5 times to twice the rate of global warming (high confidence) {11.3, Figure 11.3} Highest increase of temperature of coldest days is projected in Arctic regions, at about three times the rate of global warming (high confidence) {11.3} Continental-scale projections: Virtually certain: Africa, Asia, Australasia, Central and South America, Europe, North America {11.3, 11.9}Virtually certain on global scale Virtually certain on all continents Highest increase of temperature of hottest days is projected in some mid-latitude and semi-arid regions, and the South American Monsoon region, at about 1.5 times to twice the rate of global warming (high confidence) {11.3, Figure 11.3} Highest increase of temperature of coldest days is projected in Arctic regions, at about three times the rate of global warming (high confidence) {11.3} Continental-scale projections: Virtually certain: Africa, Asia, Australasia, Central and South America, Europe, North America {11.3, 11.9} Heavy precipitation events: increase in the frequency, intensity, and/or amount of heavy precipitation High confidence that increases take place in most land regions {11.4} Very likely: Asia, North America Likely: Africa, Europe High confidence: Central and South America Medium confidence: Australasia {11.4, 11.9}Likely that increases take place in most land regions {11.4} Extremely likely: Asia, North America Very likely: Africa, Europe Likely: Australasia, Central and South America {11.4, 11.9}Very likely that increases take place in most land regions {11.4} Virtually certain: Africa, Asia, North America Extremely likely: Central and South America, Europe Very likely Australasia {11.4, 11.9} Agricultural and ecological droughts: increases in intensity and/or duration of drought events More regions affected by increases in agricultural and ecological droughts compared to observed changes high 2 train +1691 AR6_WGI 1552 7 Compound events, including concurrent extremes, are a factor increasing the probability of LLHI events high 2 train +1692 AR6_WGI 1552 8 With increasing global warming, some compound events with low likelihood in past and current climate will become more frequent, and there is a higher chance of historically unprecedented events and surprises high 2 train +1693 AR6_WGI 1552 9 However, even extreme events that do not have a particularly low probability in the present climate (at more than 1°C of global warming) can be perceived as surprises because of the pace of global warming high 2 train +1694 AR6_WGI 1563 14 Land–atmosphere feedbacks strongly modulate regional- and local-scale changes in temperature extremes high 2 train +1695 AR6_WGI 1563 21 Regional external forcings, including land-use changes and emissions of anthropogenic aerosols, play an important role in the changes of temperature extremes in some regions high 2 train +1696 AR6_WGI 1564 1 This leads to heterogeneity in regional changes and their associated uncertainties high 2 train +1697 AR6_WGI 1564 3 Irrigation and crop expansion have attenuated increases in summer hot extremes in some regions, such as the Midwestern USA medium 1 train +1698 AR6_WGI 1565 5 The intensity and frequency of cold spells in all Asian regions have been decreasing since the beginning of the 20th century high 2 train +1699 AR6_WGI 1566 25 Annual minimum temperatures on land have increased about three times more than global surface temperature since the 1960s, with particularly strong warming in the Arctic high 2 train +1700 AR6_WGI 1566 32 In general, CMIP5 and CMIP6 historical simulations are similar in their performance in simulating the observed climatology of extreme temperatures high 2 train +1701 AR6_WGI 1569 25 Irrigation and crop intensification have been shown to lead to a cooling in some regions, in particular in North America, Europe, and India high 2 train +1702 AR6_WGI 1569 33 Crop intensification, irrigation and no-till farming have attenuated increases in summer hot extremes in some regions, such as Central North America medium 1 train +1703 AR6_WGI 1571 20 Exceptions include lower confidence in the projected decrease in the intensity and frequency of cold extremes compared with the 1995–2014 baseline under 1.5°C of global warming medium 1 train +1704 AR6_WGI 1572 26 In most regions, changes in the magnitude of temperature extremes are proportional to global warming levels high 2 train +1705 AR6_WGI 1573 1 The highest increase of temperature of coldest days is projected in Arctic regions, at about three times the rate of global warming high 2 train +1706 AR6_WGI 1573 2 The probability of temperature extremes generally increases nonlinearly with increasing global warming levels high 2 train +1707 AR6_WGI 1573 28 Urbanization intensifies extreme precipitation, especially in the afternoon and early evening, over the urban area and its downwind region medium 1 train +1708 AR6_WGI 1573 32 Warming-induced thermodynamic change results in an increase in extreme precipitation, at a rate that closely follows the C-C relationship at the global scale high 2 train +1709 AR6_WGI 1573 34 Precipitation extremes are also affected by forcings other than changes in greenhouse gases, including changes in aerosols, land- use and land-cover change, and urbanization medium 1 train +1710 AR6_WGI 1574 6 Daily mean precipitation intensities have increased since the mid-20th century in a majority of land regions high 2 train +1711 AR6_WGI 1574 32 In Asia (Table 11.8), there is robust evidence that extreme precipitation has increased since the 1950s high 2 train +1712 AR6_WGI 1575 7 In Australasia (Table 11.11), available evidence has not shown an increase or a decrease in heavy precipitation over Australasia as a whole medium 1 train +1713 AR6_WGI 1576 7 There were increases in extreme precipitation in Tobago from 1985–2015 (Stephenson et al., 2014; Dookie et al., 2019) and decreases in south-western French Polynesia and the southern subtropics low 0 train +1714 AR6_WGI 1578 16 The magnitude and frequency of extreme precipitation simulated by CMIP6 models are similar to those simulated by CMIP5 models high 2 train +1715 AR6_WGI 1579 27 Human influence has contributed to the intensification of heavy precipitation in three continents where observational data are more abundant high 2 train +1716 AR6_WGI 1581 10 Heavy precipitation will likely increase by the end of the century under RCP8.5 in West Africa (Diallo et al., 2016; Dosio, 2016; Sylla et al., 2016; Abiodun et al., 2017; Akinsanola and Zhou, 2019; Dosio et al., 2019) and is projected to increase high 2 train +1717 AR6_WGI 1581 11 In north-east and central east Africa, extreme precipitation intensity is projected to increase across CMIP5, CMIP6 and CORDEX-CORE high 2 train +1718 AR6_WGI 1581 26 Agreement in projected changes by different models is low in regions of complex topography such as Hindu-Kush Himalayas (Roy et al., 2019), but CMIP5, CMIP6 and CORDEX-CORE simulations consistently project an increase in heavy precipitation in higher latitude areas, such as West and East Siberia, and Russian Far East high 2 train +1719 AR6_WGI 1581 31 Daily rainfall extremes are projected to increase at the 2.0°C global warming level medium 1 train +1720 AR6_WGI 1583 1 The increase in the frequency of heavy precipitation events will be non-linear with more warming and will be higher for rarer events high 2 train +1721 AR6_WGI 1583 2 Increases in the intensity of extreme precipitation events at regional scales will depend on the amount of regional warming as well as changes in atmospheric circulation and storm dynamics leading to regional differences in the rate of heavy precipitation changes high 2 train +1722 AR6_WGI 1584 1 Changes in extreme precipitation may be used as a proxy to infer changes in some types of flash floods that are more directly related to extreme precipitation high 2 train +1723 AR6_WGI 1584 17 In summary, the seasonality of floods has changed in cold regions where snowmelt dominates the flow regime in response to warming high 2 train +1724 AR6_WGI 1586 3 In summary, global hydrological models project a larger fraction of land areas to be affected by an increase in river floods than by a decrease in river floods medium 1 train +1725 AR6_WGI 1586 23 Thermodynamic processes are thus the main driver of drought changes in a warming climate high 2 train +1726 AR6_WGI 1588 32 Drought events are the result of dynamic and/or thermodynamic processes, with thermodynamic processes being the main driver of drought changes under human-induced climate change high 2 train +1727 AR6_WGI 1595 9 Several meteorological and agricultural and ecological drought events have been attributed to human-induced climate change, even in regions where no long-term changes are detected medium 1 train +1728 AR6_WGI 1595 10 However, a lack of attribution to human-induced climate change has also been shown for some events medium 1 train +1729 AR6_WGI 1595 11 In summary, human influence has contributed to increases in agricultural and ecological droughts in the dry season in some regions due to increases in evapotranspiration medium 1 train +1730 AR6_WGI 1595 12 The increases in evapotranspiration have been driven by increases in atmospheric evaporative demand induced by increased temperature, decreased relative humidity and increased net radiation over affected land areas high 2 train +1731 AR6_WGI 1595 15 Human- induced climate change has contributed to global-scale change in low flow, but human water management and land-use changes are also important drivers medium 1 train +1732 AR6_WGI 1595 22 Projected changes show increases in drought frequency and intensity in several regions as function of global warming high 2 train +1733 AR6_WGI 1595 23 There are also substantial increases in drought hazard probability from 1.5°C to 2°C global warming and for further additional increments of global warming high 2 train +1734 AR6_WGI 1596 3 In Asia, most AR6 regions show low confidence in projected changes in meteorological droughts at 1.5°C and 2°C of global warming, with a few regions displaying a decrease in meteorological droughts at 4°C of global warming (RAR, ESB, RFE, ECA; medium confidence), although there is a projected increase in meteorological droughts in South East Asia at 4°C medium 1 train +1735 AR6_WGI 1596 16 The overall evidence suggests an increase in meteorological drought at 4°C in the WCE region medium 1 train +1736 AR6_WGI 1599 3 The assessment shows that several regions will be affected by more severe agricultural and ecological droughts even if global warming is stabilized at 2°C, including MED, WSAF, SAM and SSA (high confidence), and ESAF, MDG, EAU, SAU, SCA, CAR, NSA, NES, SWS, WCE, NCA, WNA and CNA medium 1 train +1737 AR6_WGI 1599 5 NEAF, SAS are also projected to experience less agricultural and ecological drought with global warming medium 1 train +1738 AR6_WGI 1599 11 In summary, more regions are affected by increases in agricultural and ecological droughts with increasing global warming high 2 train +1739 AR6_WGI 1599 12 New evidence strengthens the SR1.5 conclusion that even relatively small incremental increases in global warming (+0.5°C) cause a worsening of droughts in some regions high 2 train +1740 AR6_WGI 1599 14 A larger number of regions are projected to be affected by more severe agricultural and ecological droughts at 2°C of global warming, including MED, WSAF, SAM and SSA (high confidence), and ESAF, MDG, EAU, SAU, SCA, CAR, NSA, NES, SWS, WCE, NCA, WNA and CNA medium 1 train +1741 AR6_WGI 1599 18 Enhanced atmospheric CO 2 concentrations lead to enhanced water-use efficiency in plants medium 1 train +1742 AR6_WGI 1601 9 The SR1.5 (Chapter 3, Hoegh-Guldberg et al., 2018) essentially confirmed the AR5 assessment of TCs and ETCs, adding that heavy precipitation associated with TCs is projected to be higher at 2°C compared to 1.5°C global warming medium 1 train +1743 AR6_WGI 1602 18 Global changes (blue shading) from top to bottom: (i) Increased mean and maximum rain rates in TCs, ETCs, and ARs [past ( low confidence due to lack of reliable data) and projected (high confidence)]; (ii) Increased proportion of stronger TCs [past ( medium confidence) and projected (high confidence)]; (iii) Decrease or no change in global frequency of TC genesis [past ( low confidence due to lack of reliable data) and projected (medium confidence)]; and (iv) Increased and decreased ETC wind speed, depending on the region, as storm tracks change [past ( low confidence due to lack of reliable data) and projected medium 1 train +1744 AR6_WGI 1602 19 Regional changes, from left to right: (i) Poleward TC migration in the western North Pacific and subsequent changes in TC exposure [past ( medium confidence) and projected (medium confidence)]; (ii) Slowdown of TC forward translation speed over the contiguous USA and subsequent increase in TC rainfall [past ( medium confidence) and projected (low confidence due to lack of directed studies)]; and (iii) Increase in mean and maximum SCS rain rate and increase in spring SCS frequency and season length over the contiguous USA [past ( low confidence due to lack of reliable data) and projected medium 1 train +1745 AR6_WGI 1604 24 However, higher-resolution models generally capture TC properties more realistically high 2 train +1746 AR6_WGI 1604 27 Models with realistic atmosphere– ocean interactions are generally better than atmosphere-only models at reproducing realistic TC evolutions high 2 train +1747 AR6_WGI 1610 6 Subsequent literature assessed in the Climate Science Special Report (Kossin et al., 2017) led to the assessment of the observed tornado activity over the 2000s in the USA, with a decrease in the number of days per year with tornadoes and an increase in the number of tornadoes on these days medium 1 train +1748 AR6_WGI 1610 8 Climate models consistently project environmental changes that would support an increase in the frequency and intensity of severe thunderstorms that combine tornadoes, hail, and winds high 2 train +1749 AR6_WGI 1611 13 There is an upward trend in the frequency and intensity of extreme precipitation events in the USA high 2 train +1750 AR6_WGI 1611 37 There is medium confidence that the mean annual number of tornadoes in the USA has remained relatively constant, but their variability of occurrence has increased since the 1970s, particularly over the 2000s, with a decrease in the number of days per year, and an increase in the number of tornadoes on these days high 2 train +1751 AR6_WGI 1613 12 In summary, the average and maximum rain rates associated with severe convective storms increase in a warming world in some regions, including the USA high 2 train +1752 AR6_WGI 1613 14 The frequency of severe convective storms in spring is projected to increase in the USA, leading to a lengthening of the severe convective storm season medium 1 train +1753 AR6_WGI 1613 15 There is significant uncertainty about projected regional changes in tornadoes, hail, and lightning due to limited analysis of simulations using convection-permitting models high 2 train +1754 AR6_WGI 1614 12 Based on the consistency of these studies, it is likely that medicanes will decrease in frequency, while the strongest medicanes become stronger under warming scenario projections medium 1 train +1755 AR6_WGI 1614 13 In summary, the observed intensity of extreme winds is becoming less severe in the low to mid-latitudes, while becoming more severe in high latitudes poleward of 60 degrees low 0 train +1756 AR6_WGI 1614 14 Projected changes in the frequency and intensity of extreme winds are associated with projected changes in the frequency and intensity of TCs and ETCs medium 1 train +1757 AR6_WGI 1615 9 The land area affected by concurrent extremes has increased high 2 train +1758 AR6_WGI 1615 10 Concurrent extreme events at different locations, but possibly affecting similar sectors (e.g., breadbaskets) in different regions, will become more frequent with increasing global warming, in particular above +2°C of global warming high 2 train +1759 AR6_WGI 1617 13 Both the ENSO amplitude and the frequency of high-magnitude events since 1950 is higher than over the pre-industrial period medium 1 train +1760 AR6_WGI 1620 18 Both the ENSO amplitude and the frequency of high-magnitude events since 1950 is higher than over the pre-industrial period medium 1 test +1761 AR6_WGI 1720 11 Some evidence medium 1 train +1762 AR6_WGI 1786 10 Climate change has already altered CID profiles and resulted in shifts in the magnitude, frequency, duration, seasonality and spatial extent of associated indices high 2 train +1763 AR6_WGI 1786 11 Changes in temperature-related CIDs such as mean temperatures, growing season length, extreme heat and frost have already occurred and many of these changes have been attributed to human activities medium 1 train +1764 AR6_WGI 1786 12 Mean temperatures and heat extremes have emerged above natural variability in all land regions high 2 train +1765 AR6_WGI 1786 13 In tropical regions, recent past temperature distributions have already shifted to a range different to that of the early 20th century high 2 train +1766 AR6_WGI 1786 14 Ocean acidification and deoxygenation have already emerged over most of the global open ocean, as has reduction in Arctic sea ice high 2 train +1767 AR6_WGI 1786 17 Increasing precipitation is projected to emerge before the middle of the century in the high latitudes of the Northern Hemisphere high 2 train +1768 AR6_WGI 1786 18 Decreasing precipitation will emerge in a very few regions (Mediterranean, Southern Africa, south-western Australia) (medium confidence) by mid-century medium 1 train +1769 AR6_WGI 1786 19 The anthropogenic forced signal in near-coast relative sea level rise will emerge by mid-century RCP8.5 in all regions with coasts, except in the West Antarctic region where emergence is projected to occur before 2100 medium 1 train +1770 AR6_WGI 1786 20 The signal of ocean acidification in the surface ocean is projected to emerge before 2050 in every ocean basin high 2 train +1771 AR6_WGI 1786 29 Apart from a few regions with substantial land uplift, relative sea level rise is very likely to virtually certain (depending on the region) to continue in the 21st century, contributing to increased coastal flooding in most low-lying coastal areas (high confidence) and coastal erosion along most sandy coasts high 2 train +1772 AR6_WGI 1786 31 Glaciers will continue to shrink and permafrost to thaw in all regions where they are present high 2 train +1773 AR6_WGI 1787 3 Extreme precipitation and pluvial flooding will increase in many regions around the world high 2 train +1774 AR6_WGI 1787 4 Increases in river flooding are also expected in Western and Central Europe and in polar regions (high confidence), most of Asia, Australasia, North America, the South American Monsoon region and South-Eastern South America medium 1 train +1775 AR6_WGI 1787 5 Mean winds are projected to slightly decrease by 2050 over much of Europe, Asia and Western North America, and increase in many parts of South America except Patagonia, West and South Africa and the eastern Mediterranean medium 1 train +1776 AR6_WGI 1787 6 Extratropical storms are expected to have a decreasing frequency but increasing intensity over the Mediterranean, increasing intensity over most of North America, and an increase in both intensity and frequency over most of Europe medium 1 train +1777 AR6_WGI 1787 7 Enhanced convective conditions are expected in North America medium 1 train +1778 AR6_WGI 1787 8 Tropical cyclones are expected to increase in intensity despite a decrease in frequency in most tropical regions medium 1 train +1779 AR6_WGI 1787 11 Dangerous humid heat thresholds, such as the US National Oceanic and Atmospheric Administration Heat Index (NOAA HI) of 41°C, will be exceeded much more frequently under the SSP5-8.5 scenario than under SSP1-2.6 and will affect many more regions high 2 train +1780 AR6_WGI 1787 12 In many tropical regions, the number of days per year where an HI of 41°C is exceeded will increase by more than 100 days relative to the recent past under SSP5-8.5, while this increase will be limited to less than 50 days under SSP1-2.6 high 2 train +1781 AR6_WGI 1787 16 The frequency of the present-day 1-in- 100 year extreme sea level event (represented here by extreme total water level) is also projected to increase substantially in most regions high 2 train +1782 AR6_WGI 1787 22 These include both mean and extreme heat, cold, wet and dry hazards; cryospheric hazards (snow cover, ice extent, permafrost); and oceanic hazards (marine heatwaves) high 2 train +1783 AR6_WGI 1787 23 For some of these, a quantitative relation can be drawn high 2 train +1784 AR6_WGI 1787 25 For other hazards (e.g., ice-sheet behaviour, glacier mass loss, global mean sea level rise, coastal floods and coastal erosion) the time and/or scenario dimensions remain critical and a simple relation with GWLs cannot be drawn (high confidence), but still quantitative estimates assuming specific time frames and/ or stabilized GWLs can be derived medium 1 train +1785 AR6_WGI 1787 26 Model uncertainty challenges the link between specific GWLs and tipping points and irreversible behaviour, but their occurrence cannot be excluded and their chances increase with warming levels medium 1 train +1786 AR6_WGI 1806 20 Section 6.5 notes that climate change will have a small burden on particulate matter (PM) pollution (medium confidence) while the main controlling factor in determining future concentrations will be future emissions policy for PM and their precursors high 2 train +1787 AR6_WGI 1807 5 Over Asia, the CMIP5 multi-model mean response shows that solar radiation will decrease in South Asia and increase in East Asia medium 1 train +1788 AR6_WGI 1807 6 Projected solar resources show an increasing trend throughout the 21st century in East Asia under RCP2.6 and RCP8.5 scenarios in CMIP5 simulations medium 1 train +1789 AR6_WGI 1807 8 More sunshine is projected over Australia in winter and spring by the end of the century medium 1 train +1790 AR6_WGI 1807 9 In Central and South America, there is medium confidence of increasing solar radiation over the Amazon basin and the northern part of South America medium 1 train +1791 AR6_WGI 1807 13 Increasing radiation trends are also found over southern and eastern USA, and decreasing trends over North-Western North America (Wild et al., 2015; Losada Carreño et al., 2020), despite large differences between responses from regional climate models (RCMs) and general circulation models (GCMs) over southern and eastern USA low 0 train +1792 AR6_WGI 1807 20 Mean annual temperatures have increased at a high rate since the mid-20th century, reaching 0.2°C–0.5°C per decade in some regions such as north, north- eastern, west and south-western Africa high 2 train +1793 AR6_WGI 1807 23 A very likely warming with ranges between 0.5°C and 2.5°C is projected by the mid-century for all scenarios depending on the region high 2 train +1794 AR6_WGI 1807 25 Extreme heat: Warm extremes have increased in most of the regions (high confidence), NEAF, SEAF and MDG medium 1 train +1795 AR6_WGI 1807 27 A substantial increase in heatwave magnitude and frequency over most of the Africa domain is projected for even 2°C global warming high 2 train +1796 AR6_WGI 1809 11 Heating degree days will have a substantial decrease by the end of the century for up to about one month under RCP8.5 in North and Southern Africa high 2 train +1797 AR6_WGI 1809 13 Heatwaves and deadly heat stress and the frequency of exceedance of hot temperature thresholds (e.g., 35°C) will drastically increase by the end of the century high 2 train +1798 AR6_WGI 1809 17 The Western Africa region features a gradient in which precipitation decreases in the west and increases in the east and increase is also projected over Eastern Africa medium 1 train +1799 AR6_WGI 1809 19 A change in monsoon seasonality is also reported in Western Africa and Sahel low 0 train +1800 AR6_WGI 1809 24 In particular, over Western Africa, upward trends in hydrological extremes such as maximum peak discharge have likely occurred during the last few decades (i.e., after 1980) and have caused increased flood events in riparian countries of rivers such as Niger, Senegal and Volta high 2 train +1801 AR6_WGI 1809 25 In Southern Africa, trends in flood occurrences were decreasing prior to 1980 and increasing afterwards medium 1 train +1802 AR6_WGI 1810 7 In addition, East Africa has experienced strong precipitation variability and intense wet spells leading to widespread pluvial flooding events hitting most countries including Ethiopia, Somalia, Kenya and Tanzania medium 1 train +1803 AR6_WGI 1810 8 Finally, with respect to Southern Africa, heavy precipitations events have increased in frequency medium 1 train +1804 AR6_WGI 1810 10 Extreme precipitation intensity is also increasing in several other regions, such as SAH, NEAF, SEAF, ESAF and MDG high 2 test +1805 AR6_WGI 1810 17 Hydrological drought: Section 11.9 noted observed decreases in hydrological drought over the Mediterranean (high confidence) and diminished summer river flows in West Africa medium 1 train +1806 AR6_WGI 1810 28 Section 11.9 assesses increases in agricultural and ecological drought at 2°C GWL for North Africa and West Southern Africa (high confidence) and for East Southern Africa and Madagascar medium 1 train +1807 AR6_WGI 1810 32 Days prone to fire conditions are going to increase in all extratropical Africa until the end of the century and fire weather indices are projected to largely increase in North and Southern Africa, where increasing aridity trends occur high 2 train +1808 AR6_WGI 1811 1 Most African regions will undergo an increase in heavy precipitation that can lead to pluvial floods (high confidence), even as increasing dry climatic impact-drivers (aridity, hydrological, agricultural and ecological droughts, fire weather) are generally projected in the North Africa and Southern African regions (high confidence) and western portions of West Africa medium 1 train +1809 AR6_WGI 1811 4 Over Western Africa and Southern Africa a future significant increase in wind speeds and wind energy potential is expected medium 1 train +1810 AR6_WGI 1811 10 The frequency of Mediterranean wind storms reaching North Africa, including Medicanes, is projected to decrease, but their intensities are projected to increase, by the mid-century and beyond under SRES A1B, SRES A2 and RCP8.5 medium 1 train +1811 AR6_WGI 1811 12 However, there is a projected decrease in the frequency of tropical cyclones making landfall over Madagascar, South Eastern Africa and East Southern Africa in a 1°C, 2°C and 3°C warmer world medium 1 train +1812 AR6_WGI 1811 19 Dust loadings and related air pollution hazards (from fine particles that affect health) are projected to generally decrease in many regions of the Sahara and Sahel due to the changing winds (Evan et al., 2016) and slightly increase over the Guinea coast and West Africa low 0 train +1813 AR6_WGI 1811 23 Observation and future projection of African glacier mass changes are assessed in Section 9.5.1 within the low-latitude glacier region, which is one of the regions with the largest mass loss even under low-emissions scenarios (assessment of this region is dominated by glaciers in the South American Andes, however) high 2 train +1814 AR6_WGI 1811 27 Snow water equivalent and snow cover season duration also decline in the East African mountains, Ethiopian Highlands and Atlas Mountains with climate change high 2 train +1815 AR6_WGI 1812 7 Extreme total water level (ETWL) magnitude and occurrence frequency are expected to increase throughout the region high 2 train +1816 AR6_WGI 1812 16 Projections indicate that a vast majority of sandy coasts in the region will experience shoreline retreat throughout the 21st century high 2 train +1817 AR6_WGI 1812 31 Relative sea level rise is virtually certain to continue around Africa, contributing to increased coastal flooding in low-lying areas (high confidence) and shoreline retreat along most sandy coasts high 2 train +1818 AR6_WGI 1812 32 Marine heatwaves are also expected to increase around the region over the 21st century high 2 train +1819 AR6_WGI 1815 7 Records also indicate a higher rate of warming in minimum temperatures than maximum temperatures in Asia, leading to more frequent warm nights and warm days, and less frequent cold days and cold nights high 2 train +1820 AR6_WGI 1815 8 Projections show continued warming over Asia in the future with contrasted regional patterns across the continent high 2 train +1821 AR6_WGI 1815 9 For RCP8.5/SSP5-8.5 at the end of the century, the mean estimated warming exceeds 5°C in WSB, ESB and RFE and 7°C in some parts high 2 train +1822 AR6_WGI 1815 19 In these regions, the increase in number of days with exceedance of 35°C of high heat stress is also expected to increase substantially for the mid- century under SSP5-8.5 (typically by 10–50 days except in Arctic and Siberian regions), and by more than 60 days in areas of SEA, and a large difference is found between low- and high-end scenarios in the end of the century high 2 train +1823 AR6_WGI 1815 20 Over WSB, ESB and RFE also, an increase of extreme heat durations and frequency is expected in all scenarios high 2 train +1824 AR6_WGI 1815 21 Cold spell and frost: Cold spells intensity and frequency, as well as the number of frost days, in most Asian regions have been decreasing since the beginning of the 20th century (high confidence) (Chapter 11; Sheikh et al., 2015; Donat et al., 2016; Erlat and Türkeş, 2016; S. Dong et al., 2018; Liao et al., 2018, 2020; Lu et al., 2018; van Oldenborgh et al., 2019), except for the central Eurasian regions, where there was a cooling trend during 1995–2014, which is linked to sea ice loss in the Barents–Kara Seas medium 1 train +1825 AR6_WGI 1815 23 In Asia, temperatures have warmed during the last century high 2 train +1826 AR6_WGI 1815 24 Extreme heat episodes have become more frequent in most regions high 2 train +1827 AR6_WGI 1816 3 Mean precipitation is likely to increase in most areas of northern (WSB, ESB, RFE), southern (ECA, TIB, SAS) and East Asia (EAS) in different scenarios high 2 train +1828 AR6_WGI 1816 6 River flood: Flood risk has grown in many places in China from 1961 to 2017 (Kundzewicz et al., 2019) low 0 train +1829 AR6_WGI 1816 7 In SAS, the numbers of flood events and human fatalities have increased in India during 1978–2006 (Singh and Kumar, 2013), whereas the average country-wide inundation depth has been decreasing during 2002–2010 in Bangladesh, attributed to improved flood management low 0 train +1830 AR6_WGI 1816 9 Over China floods will increase with different levels under different warming scenarios medium 1 train +1831 AR6_WGI 1816 10 Monsoon floods will be more intense in SAS medium 1 test +1832 AR6_WGI 1816 12 A changing snowmelt regime in the mountains may contribute to a shift of spring floods to earlier periods in Central Asia in future medium 1 train +1833 AR6_WGI 1816 13 The annual maximum river discharge can almost double by the mid-21st century in major Siberian rivers, and annual maximum flood area is projected to increase across Siberia mostly by 2–5% relative to the baseline period (1990–1999) under RCP8.5 scenario medium 1 train +1834 AR6_WGI 1816 16 Heavy precipitation is very likely to become more intense and frequent in all areas of Asia except in ARP medium 1 train +1835 AR6_WGI 1816 22 Aridity: Aridity in West Central Asia and parts of South Asia increased in recent decades medium 1 train +1836 AR6_WGI 1816 25 There was a drying tendency in the dry season and significant wetting in the wet season in the Philippines during 1951–2010 (Villafuerte et al., 2014), and slight wetting in Vietnam during 1980–2017 (Stojanovic et al., 2020) low 0 train +1837 AR6_WGI 1817 2 Hydrological drought: Section 11.9 indicates that limited evidence and inconsistent regional trends gives low confidence to observed and projected changes in hydrological drought in all Asian regions at a 2°C GWL (approximately mid-century), although West Central Asia hydrological droughts increase at the 4°C GWL (approximately end- of-century under higher emissions scenarios) medium 1 train +1838 AR6_WGI 1817 6 Higher future temperatures are expected to alter the seasonal profile of hydrologic droughts given reduced summer snowmelt medium 1 test +1839 AR6_WGI 1817 12 Studies examining a 2°C GWL give low confidence for projected broad changes to agricultural and ecological drought across all Asia regions, although at 4°C GWL agricultural and ecological drought increases are projected for West Central Asia and East Asia along with a decrease in South Asia medium 1 train +1840 AR6_WGI 1817 15 Fire weather: Under the global warming scenario of 2°C, the magnitude of length and frequency of fire seasons are projected to increase with strong effects in India, China and Russia medium 1 train +1841 AR6_WGI 1817 19 The potential burned areas in five Central Asian countries (Kazakhstan, Kyrgyzstan, Tajikistan, Uzbekistan and Turkmenistan) will increase by 2–8% in the 2030s and 3–13% in the 2080s compared with the baseline medium 1 train +1842 AR6_WGI 1817 22 Fire weather seasons are projected to lengthen and intensify particularly in the northern regions medium 1 train +1843 AR6_WGI 1817 24 But a short-term strengthening in SWS was observed during the winter since 2000 in eastern China medium 1 train +1844 AR6_WGI 1817 30 Tropical cyclone: There was an increase in the number and intensification rate of intense tropical cyclones (TC), such as Category 4–5 (wind speeds >58 m s–1), in the Western North Pacific (WNP) and Bay of Bengal since the mid-1980s medium 1 train +1845 AR6_WGI 1818 2 However, while the analysis shows fewer typhoons, more extreme TCs have affected the Philippines low 0 train +1846 AR6_WGI 1818 3 The frequency and duration of tropical cyclones has significantly increased over time over the Arabian Sea and insignificantly decreased over the Bay of Bengal during 1977–2018 low 0 train +1847 AR6_WGI 1818 7 As a consequence, the intensity of TCs affecting the Japan Islands would increase in the future under the RCP8.5 scenario (Yoshida et al., 2017), whereas the frequency of TCs affecting the Philippine region and Vietnam is projected to decrease (Kieu-Thi et al., 2016; C. Wang et al., 2017; Gallo et al., 2019) medium 1 train +1848 AR6_WGI 1818 11 Dust storm frequency in most regions of northern China show a decreasing trend since the 1960s due to the decrease in surface wind speed medium 1 train +1849 AR6_WGI 1818 14 In conclusion, surface wind speeds have been decreasing in Asia high 2 train +1850 AR6_WGI 1818 17 Observations do show significant changes in the seasonal timing of Eurasian snow cover extent (especially for earlier spring snowmelt) since the 1970s, with seasonal changes expected to continue in the future high 2 train +1851 AR6_WGI 1818 20 All regions show continuing decline in glacier mass and area in the coming century high 2 train +1852 AR6_WGI 1818 24 Although enhanced meltwater from snow and glaciers largely offsets hydrological drought-like conditions (Pritchard, 2019), this effect is unsustainable and may reverse as these cryospheric buffers disappear medium 1 train +1853 AR6_WGI 1818 25 In the Himalayas and the TIB region higher temperatures will lead to higher glacier melt rates and significant glacier shrinkage and a summer runoff decrease medium 1 train +1854 AR6_WGI 1818 26 Glacier runoff in the Asian high mountains will increase up to mid-century, and after that runoff might decrease due to the loss of glacier storage medium 1 train +1855 AR6_WGI 1818 29 As many of these lakes will develop at the immediate foot of steep icy peaks with degrading permafrost and decreasing slope stability, the risk of glacier lake outburst floods and floods from landslides into moraine-dammed lakes is increasing in Asian high mountains high 2 train +1856 AR6_WGI 1818 30 Permafrost : Permafrost is thawing in Asia high 2 train +1857 AR6_WGI 1819 2 The permafrost area is projected to decline by 13.4–27.7% and 60–90% in TIB (L. Zhao et al., 2020) and 32% ± 11% and 76% ± 12% in Russia (Guo and Wang, 2016) by the end of the 21st century under the RCP2.6 and RCP8.5 scenarios respectively high 2 train +1858 AR6_WGI 1819 3 Lake and river ice: Lake ice cover duration got shorter in many lakes in TIB (Yao et al., 2016; Cai et al., 2019; Guo et al., 2020) and some other areas such as north-west China (Cai et al., 2020) and north-east China (Yang et al., 2019) in the last two decades high 2 train +1859 AR6_WGI 1819 5 Climate warming also leads to a significant reduction in the period with ice phenomena and the decrement of ice regime hazard in Russian lowland rivers (Agafonova et al., 2017), and the Inner Mongolia reach of the Yellow River in northern China (Wan et al., 2020) high 2 train +1860 AR6_WGI 1819 6 Lake ice and river ice in Asia are expected to decline with projected increases in surface air temperature towards the end of this century high 2 train +1861 AR6_WGI 1819 9 Observational results generally show a decrease in the frequency and an increase in the mean intensity of snowfalls in most Chinese regions medium 1 train +1862 AR6_WGI 1819 10 Because of the decrease in the snow frequency, the occurrence of large-scale snow disasters in TIB decreased low 0 train +1863 AR6_WGI 1819 11 Large parts of northern high-latitude continents (including Siberia and RFE) have experienced cold snaps and heavy snowfalls in the past few winters, and the reduction of Arctic sea ice would increase the chance of heavy snowfall events in those regions in the coming decades medium 1 train +1864 AR6_WGI 1819 12 Heavy snowfall is projected to occur more frequently in Japan’s Northern Alps, the inland areas of Honshu Island and Hokkaido Island (Kawase et al., 2016, 2020; MOE et al., 2018), and the heavy wet snowfall can be enhanced over the mountainous regions in central Japan and northern part of Japan (Ohba and Sugimoto, 2020) medium 1 train +1865 AR6_WGI 1819 17 Tree-ring-based snow avalanche reconstructions in the Indian Himalayas show an increase in avalanche occurrence and runout distances in recent decades (Ballesteros-Cánovas et al., 2018).In summary, snowpack and glaciers are projected to continue decreasing and permafrost to continue thawing in Asia high 2 train +1866 AR6_WGI 1819 25 RSL change in many coastal areas in Asia, especially in EAS, is affected by land subsidence due to sediment compaction under building mass and groundwater extraction high 2 train +1867 AR6_WGI 1819 31 Sea level rise and land subsidence will jointly lead to more flooding in delta areas in Asia high 2 train +1868 AR6_WGI 1819 32 Extreme total water level magnitude and occurrence frequency are expected to increase throughout the region high 2 train +1869 AR6_WGI 1821 8 Projections indicate that a majority of sandy coasts in the Asia region will experience shoreline retreat high 2 train +1870 AR6_WGI 1821 17 Relative sea level rise is very likely to continue around Asia, contributing to increased coastal flooding in low- lying areas (high confidence) and shoreline retreat along most sandy coasts high 2 train +1871 AR6_WGI 1821 18 Marine heatwaves are also expected to increase around the region over the 21st century high 2 train +1872 AR6_WGI 1821 25 The SROCC also projects an increase of mean significant wave height across the Southern Ocean high 2 train +1873 AR6_WGI 1823 17 Heat thresholds potentially affecting agriculture and health, such as 35°C or 40°C, are projected to be exceeded more frequently over the 21st century in Australia under all RCPs high 2 train +1874 AR6_WGI 1823 22 The projected frequency of exceeding dangerous humid heat thresholds is increasing in Australia, with a strong increase in Northern Australia for RCP8.5 high 2 train +1875 AR6_WGI 1823 25 Cold spell and frost: Excepting parts of Southern Australia, the Australasian region has a significant trend of decreasing frequency in cold extremes since the 1950s high 2 train +1876 AR6_WGI 1823 33 The mean temperature in Australasia is virtually certain to continue to rise through the 21st century, accompanied by less frequent cold extremes (virtually certain) and frost days high 2 train +1877 AR6_WGI 1823 34 Heat stress is projected to increase in Australia high 2 train +1878 AR6_WGI 1824 4 Annual mean precipitation is projected to increase in Central and north-east Australia (low confidence) and in the south and west of New Zealand medium 1 train +1879 AR6_WGI 1824 13 While median annual runoff is projected to decrease in most of Australia (Chiew et al., 2017), consistent with projected decreases in average rainfall (CSIRO and BOM, 2015; Alexander and Arblaster, 2017), river floods are projected to increase due to more intense extreme rainfall events and associated increase in runoff medium 1 train +1880 AR6_WGI 1824 27 Aridity: In terms of dry climatic impact-drivers, a substantial decrease in precipitation has been observed across Southern Australia during the cool season (April–October) medium 1 train +1881 AR6_WGI 1825 26 Fire weather indices are projected to increase in most of Australia (high confidence) and many parts of New Zealand medium 1 train +1882 AR6_WGI 1826 3 Aridity is projected to increase with medium confidence in Southern Australia (high confidence in south-west Western Australia), Eastern Australia (medium confidence), and in the north and east of New Zealand medium 1 train +1883 AR6_WGI 1826 4 Hydrological droughts are projected to increase in Southern Australia medium 1 train +1884 AR6_WGI 1826 5 Fire weather is projected to increase throughout Australia (high confidence) and New Zealand medium 1 train +1885 AR6_WGI 1826 9 In New Zealand, mean wind patterns are projected to become more north-easterly in summer, and westerlies to become more intense in winter low 0 test +1886 AR6_WGI 1826 11 CMIP5 projections of severe winds indicate a general increase in north-eastern Australia, and decreases in some parts in Southern and Central Australia medium 1 train +1887 AR6_WGI 1826 15 Due to the intensification and the shift of the austral storm track by the end of the century (Yin, 2005), increases in extreme wind speed in New Zealand are projected over the South Island and the southern part of the North Island by mid- and end-century for all RCPs low 0 train +1888 AR6_WGI 1826 16 Tropical cyclone: In Australia, the number of TCs has generally declined since 1982, and the frequency of intense TCs that make landfall in north-eastern Australia has declined significantly since the 19th century medium 1 train +1889 AR6_WGI 1826 21 While projections suggest a decrease in severe winds in Central and Southern Australia, changes in vegetation due to increased aridity and hydrological drought could be expected to result in increased wind erosion and dust emission across the country medium 1 train +1890 AR6_WGI 1826 23 Tropical cyclones in north-eastern and North Australia are projected to decrease in number (high confidence) while their intensity is projected to increase low 0 train +1891 AR6_WGI 1826 28 Projections for Southern Australia and New Zealand show a continuing reduction in snowfall during the 21st century high 2 train +1892 AR6_WGI 1827 2 Glacier: Glacier mass and areal extent in New Zealand is projected to continue to decease over the 21st century high 2 train +1893 AR6_WGI 1827 5 In summary, snowfall is expected to decrease throughout the region at high altitudes in both Australia (high confidence) and New Zealand medium 1 train +1894 AR6_WGI 1827 6 In New Zealand, glacier ice mass and extent are expected to decrease over the 21st century for all scenarios high 2 train +1895 AR6_WGI 1827 15 Extreme total water level magnitude and occurrence frequency are expected to increase throughout the region high 2 train +1896 AR6_WGI 1827 22 Projections indicate that a majority of sandy coasts in the region will experience shoreline retreat, throughout the 21st century high 2 train +1897 AR6_WGI 1828 14 Relative sea level rise is virtually certain to continue in the oceans around Australasia, contributing to increased coastal flooding in low-lying areas (high confidence) and shoreline retreat along most sandy coasts high 2 train +1898 AR6_WGI 1828 15 Marine heatwaves are also expected to increase around the region over the 21st century high 2 train +1899 AR6_WGI 1828 21 IPCC AR5 projections (IPCC, 2014b) pointed to increases in mean temperature between 2°C and 6°C by the end of the century (high confidence) and increases in the occurrence of warm days and nights under various future climate scenarios medium 1 train +1900 AR6_WGI 1828 22 The AR5 also pointed to patterns of changes in precipitation (medium confidence), changes in the duration of dry spells (medium confidence) and decreases in water supply high 2 train +1901 AR6_WGI 1831 2 Climate change projections point to major increases in several heat indices across the region for all scenarios high 2 train +1902 AR6_WGI 1831 10 Cold spells and frost days will have a decreasing trend high 2 train +1903 AR6_WGI 1831 12 Projections indicate a drying signal for SCA (medium confidence) (Coppola et al., 2014a; Nakaegawa et al., 2014), NES and SWS (high confidence) (Atlas.7.2.5) and the well-known dipole for South America, meaning increasing precipitation over subtropical regions like the Río de La Plata basin (SES) (high confidence) and decreasing precipitation in the Amazon (NSA) medium 1 train +1904 AR6_WGI 1831 21 Available projections for the region show increases in river floods in SES and SAM medium 1 train +1905 AR6_WGI 1832 6 Heavy precipitation and pluvial flood: Table 11.14 indicated that there is low confidence due to limited evidence of extreme precipitation trends in almost all Central and South America, except in SES where increases in the magnitude and frequency of heavy precipitation have been observed high 2 train +1906 AR6_WGI 1832 9 In NWS, a wide range of changes is projected low 0 test +1907 AR6_WGI 1832 18 Sections 8.3.2.4 and 8.4.1.6 point to two important drying hotspots in South America with long-term soil moisture decline and precipitation declines: the Amazon basin (SAM and NSA) and SWS medium 1 train +1908 AR6_WGI 1832 22 Regional projections for Central and South America also indicate an increase in dryness in SCA and NES by mid- to end-century medium 1 train +1909 AR6_WGI 1832 30 Agricultural and ecological drought: Section 11.9 assessed low confidence in observed changes in agricultural and ecological drought across Central and South America due to regional heterogeneity and differences depending on the drought metrics used, except in NES, which has seen a dominant increase in drought severity medium 1 train +1910 AR6_WGI 1833 5 Projections indicate that the Amazon will be one of the regions in the world with the highest increase in fire weather indices over the 21st century and under all RCPs high 2 train +1911 AR6_WGI 1833 8 Projections of fire weather indices also show an increased risk in SWS (high confidence), SSA and SCA medium 1 train +1912 AR6_WGI 1833 10 Mean precipitation is projected to change in a dipole pattern with increases in NWS and SES and decreases in NES and SWS (high confidence) with further decreases in NSA and SCA medium 1 train +1913 AR6_WGI 1833 14 The strongest signal of future increase in agricultural and ecological drought, aridity and fire weather is over the Amazon region high 2 train +1914 AR6_WGI 1833 16 Global climate models project an increase in wind speeds, under all future scenarios, augmenting wind power potential in most parts of Central and South America, especially in NES, where changes lie in the range 0–20% by 2050 under RCP8.5 and 0–40% under RCP8.5 medium 1 train +1915 AR6_WGI 1833 23 Tropical cyclone: CMIP5 and CMIP6 simulations, including the new High Resolution Model Intercomparison Project (HighResMIP), project a decrease in the frequency of tropical cyclones in the Atlantic and Pacific coasts of Central America for the mid-century or under a 2°C GWL, accompanied with an increased frequency of intense cyclones medium 1 train +1916 AR6_WGI 1833 25 Climate projections indicate a decrease in frequency of tropical cyclones in Central America accompanied with an increased frequency of intense cyclones, and an increase in mean wind speed and wind power potential in most parts of Central and South America medium 1 train +1917 AR6_WGI 1833 31 Projections (based on process understanding) in Section 9.5.3.3 point to decreases in seasonal snow cover extent and duration across South America as global climate continues to warm high 2 train +1918 AR6_WGI 1834 2 Glaciers across South America are expected to continue to lose mass and glacier area in the coming century high 2 train +1919 AR6_WGI 1834 9 In conclusion, glacier volume loss and permafrost thawing will continue in the Andes Cordillera under all climate scenarios high 2 train +1920 AR6_WGI 1834 16 ETWL magnitude and occurrence frequency are expected to increase throughout the region high 2 train +1921 AR6_WGI 1834 26 Projections indicate that a majority of sandy coasts in the region will experience shoreline retreat throughout the 21st century high 2 train +1922 AR6_WGI 1836 10 In summary, relative sea level rise is extremely likely to continue in the oceans around Central and South America, contributing to increased coastal flooding in low-lying areas (high confidence) and shoreline retreat along most sandy coasts high 2 train +1923 AR6_WGI 1836 11 Marine heatwaves are also expected to increase around the region over the 21st century high 2 train +1924 AR6_WGI 1836 26 Future warming leads to the exceedance of different temperature thresholds relevant for vector-borne diseases (medium confidence) (Caminade et al., 2012; Medlock et al., 2013), invasive allergens medium 1 train +1925 AR6_WGI 1836 27 Future warming is also projected to lead to the exceedance of cooling degree day index (>22°C) thresholds, characterizing a potential increase in energy demand for cooling in southern Europe with increases likely exceeding 40% in some areas (Spinoni et al., 2015) by 2050 under RCP8.5 high 2 train +1926 AR6_WGI 1838 6 Extreme heat will exceed critical thresholds for health, agriculture and other sectors more frequently high 2 train +1927 AR6_WGI 1839 18 Chapter 11 assesses that agricultural and ecological droughts will increase in the Mediterranean regions (high confidence) and Western and Central Europe medium 1 train +1928 AR6_WGI 1839 27 Fire weather conditions have been increasing since about 1980 over a few regions in Europe including Mediterranean areas low 0 test +1929 AR6_WGI 1839 29 An increase in fire weather is projected for most of Europe, especially western, eastern and central regions, by 2080 (current 100-year events will occur every 5–50 years), with a progressive increase in confidence and model agreement along the 21st century medium 1 train +1930 AR6_WGI 1839 30 With increased drying and heat combined, in Mediterranean areas, an increase in fire weather indices is projected under RCP4.5 and RCP8.5, or SRES A1B, as early as by mid-century high 2 train +1931 AR6_WGI 1840 5 The declining trend has induced a corresponding decline in wind power potential indices across Europe low 0 train +1932 AR6_WGI 1840 8 Due to changes in mean surface wind speed patterns (C. Li et al., 2018) and the poleward shift of the North Atlantic jet stream exit, mean surface wind speeds are projected to decrease in the Mediterranean areas under RCP4.5 and RCP8.5 by the middle of the century and beyond, or for GWLs of 2°C and higher (high confidence), with a subsequent decrease in wind power potential medium 1 train +1933 AR6_WGI 1840 9 However, sub-regional patterns of change are shown in regional climate models, such as an increase in wind speeds in the Aegean Sea and in the northern Adriatic Sea, where a reduction of Bora events and an increase of Scirocco events are projected for mid-century and beyond under RCP4.5 and RCP8.5 medium 1 train +1934 AR6_WGI 1840 11 Daily and interannual wind variability is projected to increase under RCP8.5 only in Northern Europe (low confidence) (Moemken et al., 2018), which can influence electrical grid management and wind energy production low 0 train +1935 AR6_WGI 1840 13 Wind stagnation events may become more frequent in future climate scenarios in some areas of Europe in the second half of the 21st century (Horton et al., 2014; Vautard et al., 2018), with potential consequences on air quality low 0 test +1936 AR6_WGI 1840 19 Strong winds and extratropical storms are projected to have a slightly increasing frequency and amplitude in the future in northern, western and Central Europe (Outten and Esau, 2013; Feser et al., 2015; Forzieri et al., 2016; Mölter et al., 2016; Ruosteenoja et al., 2019a; Vautard et al., 2019) under RCP8.5 and SRES A1B by the end of the century medium 1 train +1937 AR6_WGI 1840 20 The frequency of storms, including Medicanes, is projected to decrease in Mediterranean regions, and their intensities are projected to increase, by the middle of the century and beyond for SRES A1B, A2 and RCP8.5 medium 1 train +1938 AR6_WGI 1841 1 The frequency of Medicanes is projected to decrease medium 1 train +1939 AR6_WGI 1841 2 Proxies of intense convection indicate that the large-scale conditions conducive to severe convection will tend to increase in the future climate low 0 train +1940 AR6_WGI 1841 29 The change in snowpack in the Alps is expected to lead to a possible reduction in overall avalanche activity by end of the century low 0 train +1941 AR6_WGI 1841 31 Heavy snowfalls have decreased in frequency in the past decades and this is expected to continue in the future climate low 0 train +1942 AR6_WGI 1841 32 Freezing rain is projected to increase in western, central and southern Europe by the end of the century under RCP4.5 and RCP8.5 low 0 train +1943 AR6_WGI 1841 34 In summary, future snow cover extent and seasonal duration will reduce high 2 train +1944 AR6_WGI 1841 35 A reduction of glacier ice volume is projected in the European Alps and Scandinavia high 2 train +1945 AR6_WGI 1841 37 Most of the Northern Europe periglacial will disappear by the end of the century even for a lower emissions scenario (medium confidence) and the debris-flow season may last longer in a warming climate medium 1 train +1946 AR6_WGI 1842 21 Projections indicate that sandy coasts throughout the continent (except those bordering the northern Baltic Sea) will experience shoreline retreat through the 21st century high 2 train +1947 AR6_WGI 1844 3 Relative sea level rise is extremely likely to continue around Europe (except in the northern Baltic Sea), contributing to increased coastal flooding in low-lying areas and shoreline retreat along most sandy coasts high 2 train +1948 AR6_WGI 1844 4 Marine heatwaves are also expected to increase around the region over the 21st century high 2 train +1949 AR6_WGI 1844 25 Warmer temperatures reduce heating degree days and increase cooling degree days (high confidence) (Bartos et al., 2016; US EPA, 2016; Craig et al., 2018; X. Zhang et al., 2019; Coppola et al., 2021b) Extreme heat: Section 11.9 assessed that extreme temperatures in North America have increased in recent decades (medium evidence , medium agreement) other than in Central and Eastern North America (low confidence), and extreme heat in all regions is projected to increase with climate change high 2 train +1950 AR6_WGI 1846 16 Cold spells are projected to decrease over North America under climate change, with the largest decreases most common in the winter season high 2 train +1951 AR6_WGI 1846 21 Frost: An expansion of the frost-free season is underway and projections for North America indicate a continuation of this trend in the future high 2 train +1952 AR6_WGI 1846 25 Frosts are projected to persist as an episodic hazard in many regions given natural variability and cold air outbreaks even as mean temperature rises high 2 train +1953 AR6_WGI 1847 7 Section 11.4 found that high precipitation is projected to increase across North America high 2 train +1954 AR6_WGI 1847 15 Landslide frequency has increased in British Columbia (Canada; Geertsema et al., 2006) and is expected to increase in North-Western North America given the combination of these factors medium 1 train +1955 AR6_WGI 1848 4 Fire weather: Climatic conditions conducive to wildfire have increased in Mexico, Western and North-Western North America, primarily due to warming high 2 train +1956 AR6_WGI 1848 11 Climate change drives future increases in North American fire weather, particularly in the south-west high 2 test +1957 AR6_WGI 1848 19 The mean wind speeds decline in North America – as in other Northern Hemisphere areas – over the past four decades has reversed in the last decade low 0 test +1958 AR6_WGI 1848 22 Mean wind speeds are expected to decline over much of North America (Figure 12.4m–o), but the only broad signal of consistent change across model types is a reduction in wind speed in Western North America high 2 train +1959 AR6_WGI 1849 12 Tropical cyclones, severe wind and dust storms in North America are shifting towards more extreme characteristics, with a stronger signal towards heightened intensity than increased frequency, although specific regional patterns are more uncertain medium 1 train +1960 AR6_WGI 1849 13 Mean wind speed and wind power potential are projected to decrease in Western North America medium 1 train +1961 AR6_WGI 1849 18 Climate change is expected to reduce the total snow amount and the length of the snow cover season over most of North America, with a corresponding decrease in the proportion of total precipitation falling as snow and a reduction in end-of-season snowpack high 2 train +1962 AR6_WGI 1849 22 Glacier: Section 9.5.1 assessed that glaciers in Alaska, western Canada and the western USA are expected to continue to lose mass and areal extent high 2 train +1963 AR6_WGI 1849 26 Continued shrinkage of glaciers is projected to create further glacial lakes medium 1 train +1964 AR6_WGI 1849 27 Permafrost: Warmer ground temperatures are expected to extend the geographical extent and depth of permafrost thaw across northern North America very high 3 train +1965 AR6_WGI 1849 33 Lake, river and sea ice: Anthropogenic warming reduces the seasonal extent of lake and river ice over many North American freshwater systems, with ice-free winter conditions pushing further north with rising temperatures high 2 train +1966 AR6_WGI 1850 25 Observations and projections agree that snow and ice CIDs over North America are characterized by reduction in glaciers and the seasonality of snow and ice formation, loss of shallow permafrost, and shifts in the rain/snow transition line that alters the seasonal and geographic range of snow and ice conditions in the coming decades very high 3 train +1967 AR6_WGI 1852 16 Projections indicate that sandy coasts in most of the region will experience shoreline retreat through the 21st century high 2 train +1968 AR6_WGI 1852 32 An observed increase in relative sea level rise is virtually certain to continue in North America (other than around the Hudson Bay and southern Alaska) contributing to more frequent and severe coastal flooding in low-lying areas (high confidence) and shoreline retreat along most sandy coasts high 2 train +1969 AR6_WGI 1852 33 Marine heatwaves are also expected to increase all around the region over the 21st century high 2 train +1970 AR6_WGI 1853 10 Extreme heat: Observational records indicate warming trends in the temperature extremes since the 1950s in CAR and the Pacific small islands high 2 train +1971 AR6_WGI 1853 11 A detectable anthropogenic increase in summer heat stress has been identified over a number of island regions in CAR, western tropical Pacific, and tropical Indian Ocean, based on wet bulb globe temperature (WBGT) index trends for 1973–2012 medium 1 train +1972 AR6_WGI 1853 21 Although there are spatial variations, annual rainfall trends in the western Indian Ocean are mostly decreasing, with generally non-significant trends in the western tropical Pacific since the 1950s low 0 train +1973 AR6_WGI 1853 22 Significant drying trends are noted in the southern Pacific subtropics and south-western French Polynesia during the 1951–2015 period (McGree et al., 2019), and in some areas of Hawaii during the 1920–2012 period medium 1 train +1974 AR6_WGI 1853 23 Atlas.10.4 projects precipitation reduction over the Caribbean high 2 train +1975 AR6_WGI 1854 1 Heavy precipitation and pluvial flood: Heavy precipitation days in CAR have increased in magnitude, and have been more frequent in the northern part during the latter part of the 20th century low 0 train +1976 AR6_WGI 1854 2 The direction of change in extreme precipitation varies across the Pacific and depends on the season low 0 train +1977 AR6_WGI 1854 11 From 1950 to 2016, a heterogeneous but prevalent drying trend is found in CAR low 0 train +1978 AR6_WGI 1854 13 Increased aridity is projected for the majority of the small islands, such as in CAR, southern Pacific and western Indian Ocean, by 2041–2059 relative to 1981–1999 under RCP8.5 or at 1.5°C and 2°C GWLs, which will further intensify by 2081–2099 medium 1 train +1979 AR6_WGI 1854 22 Agricultural and ecological droughts are projected to increase in frequency, duration, magnitude, and extent in small islands, such as in CAR (medium confidence) and parts of the Pacific low 0 train +1980 AR6_WGI 1854 31 Higher evapotranspiration under a warming climate are projected to partially offset future increases or amplify future reductions in rainfall, resulting in drier conditions and increased water stress in the small islands medium 1 train +1981 AR6_WGI 1855 3 Projections estimate up to 0.4 m s–1 (8%) increase in annual winds in CAR under RCP8.5, which is associated with changes in the extension of the North Atlantic Subtropical High that enhances the Caribbean low-level jet during the wet season, and stronger local easterlies due to enhanced land–ocean temperature differences in the dry season (Costoya et al., 2019) low 0 train +1982 AR6_WGI 1855 6 Other, less data-sensitive tropical cyclone features, such as the poleward migration of where tropical cyclones reach peak intensity in the western North Pacific since the 1940s (medium confidence) and the slowdown in tropical cyclone translational speed over contiguous USA since 1900 medium 1 train +1983 AR6_WGI 1855 7 Projections of global changes in tropical cyclones indicate more frequent Category 4–5 storms (high confidence) and increased rain rates high 2 train +1984 AR6_WGI 1855 9 Over CAR, tropical cyclone intensity is expected to increase by the end of the century under RCP8.5 due to higher sea surface temperatures but can be inhibited by increases in vertical wind shear in the region medium 1 train +1985 AR6_WGI 1855 13 Given projected reductions to the overall number of tropical cyclones but increases in storm intensity, total rainfall and storm surge potential, we assess medium confidence of overall changes to tropical cyclones affecting the Caribbean and Pacific small islands.Projections indicate that small islands will generally face fewer but more intense tropical cyclones medium 1 train +1986 AR6_WGI 1855 24 As relative sea levels increase, the potential for coastal flooding increases in the small islands high 2 train +1987 AR6_WGI 1857 3 Coastal erosion: Recent studies have indicated variable and dynamic changes in shorelines of reef islands medium 1 train +1988 AR6_WGI 1857 10 Projections indicate that shoreline retreat will occur over most of the small islands in the Pacific and CAR throughout the 21st century with spatial variability high 2 train +1989 AR6_WGI 1857 23 Shoreline retreat is projected along sandy coasts of most small islands high 2 train +1990 AR6_WGI 1858 3 Global warming of 2°C above pre-industrial levels is projected to increase SST, resulting in the exceedance of numerous hazard thresholds for pathogens, seagrasses, mangroves, kelp forests, rocky shores, coral reefs and other marine ecosystems medium 1 train +1991 AR6_WGI 1858 4 It is virtually certain that upper- ocean stratification has increased at a rate of 4.9 ± 1.5% during 1970–2018 and that this will continue to increase in the 21st century (Section 9.2.1.3), potentially leading to reduced nutrient supply and total productivity low 0 train +1992 AR6_WGI 1858 5 Marine heatwave: Marine heatwaves (MHWs) have increased in frequency over the 20th century, with an approximate doubling since the 1980s (high confidence), and their intensity and duration have also increased medium 1 train +1993 AR6_WGI 1858 13 Ocean acidity: With the increasing CO 2 concentration, the global mean ocean surface pH is decreasing and is now the lowest it has been for at least a thousand years very high 3 train +1994 AR6_WGI 1858 16 Declining ocean pH will exacerbate negative impacts on marine species medium 1 train +1995 AR6_WGI 1858 26 In recent decades, low-oxygen zones in ocean ecosystems have expanded, and projections indicate an acceleration with global warming medium 1 train +1996 AR6_WGI 1859 7 It is virtually certain that global mean SST will continue to increase throughout the 21st century, resulting in the exceedance of numerous climatic impact-driver thresholds relevant to marine ecosystems medium 1 train +1997 AR6_WGI 1860 3 Future projections also indicate freshening of the Pacific, Southern and Indian oceans and a saltier Atlantic Ocean medium 1 train +1998 AR6_WGI 1860 20 Higher Arctic precipitable water totals are also connected with observed increases in heavy precipitation and convective activity high 2 train +1999 AR6_WGI 1860 24 Aridity and drought: Recent decades have seen a general decrease in Arctic aridity, with projections indicating a continuing trend towards reduced aridity high 2 train +2000 AR6_WGI 1861 2 Trends towards more frequent fires in tundra regions are expected to continue, driven in particular by increasing potential evapotranspiration and changes in vegetation high 2 train +2001 AR6_WGI 1861 10 In the Arctic, mid-winter snowpack increases in some of the coldest and high- elevation locations given higher precipitation totals medium 1 test +2002 AR6_WGI 1861 11 Higher temperatures result in a higher percentage of Arctic precipitation falling as rain (particularly in autumn and spring) high 2 train +2003 AR6_WGI 1861 12 Glacier and ice sheet: Section 9.5.1 and Section 2.3.2.3 found that glaciers have lost mass in all polar regions since 2000 high 2 train +2004 AR6_WGI 1861 17 Permafrost: Observations from recent decades (assessed in Section 9.5.2 and Section 2.3.2.5) show increases in permafrost temperature (very high confidence) and active layer thickness medium 1 train +2005 AR6_WGI 1863 22 Projections show increases in MHW intensity, frequency and duration will be larger over the Arctic Ocean than mid-latitude oceans due in part to low interannual variability under current sea ice high 2 train +2006 AR6_WGI 1863 25 Climate change has caused and will continue to induce an enhanced warming trend, increasing heat-related extremes and decreasing cold spells and frosts in the Arctic high 2 train +2007 AR6_WGI 1863 26 The water cycle is projected to intensify in polar regions, leading to more rainfall, higher river flood potential and more intense precipitation high 2 train +2008 AR6_WGI 1863 27 Projections indicate reductions in glaciers at both poles, with sea ice loss, enhanced permafrost warming, decreasing permafrost extent, and decreasing seasonal duration and extent of snow cover in the Arctic (high confidence) even as some of the coldest regions will see higher total snowfall given increased precipitation medium 1 train +2009 AR6_WGI 1863 28 Projections indicate relative sea level rises in polar regions (high confidence), with the exception of regions with substantial land uplift including North-Eastern North America high 2 train +2010 AR6_WGI 1863 29 Higher sea levels also contribute to high confidence for projected increases of Arctic coastal flooding and higher coastal erosion (aided by sea ice loss) medium 1 train +2011 AR6_WGI 1864 5 Marine ecoregions will experience ocean acidification and temperatures that increase faster in high latitudes high 2 train +2012 AR6_WGI 1864 13 Longer dry seasons also extend the seasonal length and geographical extent of fire weather in all future scenarios medium 1 train +2013 AR6_WGI 1864 18 Box 10.3 identified a continuing strong role of the urban heat island in amplifying heat extremes in cities, although changes in the urban heat island are an order of magnitude smaller than projected localized warming trends very high 3 train +2014 AR6_WGI 1864 19 Coastal cities’ proximity to the sea somewhat mitigates the effect of urban heat islands high 2 train +2015 AR6_WGI 1864 22 Such threshold exceedances are projected to increase for many coastal areas high 2 train +2016 AR6_WGI 1864 23 Climate change-related variations in oceanic drivers (e.g., relative sea level, storm surge, ocean waves), combined with tropical cyclones, extreme precipitation and river flooding, are expected to lead to more frequent and more intense coastal flooding and erosion very high 3 test +2017 AR6_WGI 1864 25 Compound flooding due to simultaneous storm surges and high river flows have been found to be increasingly frequent in several cities and/or low- lying areas in Europe and the USA high 2 train +2018 AR6_WGI 1864 26 Chapter 11 found that the frequency of such compound flood events is projected to increase high 2 train +2019 AR6_WGI 1865 5 These include increases in extreme heat, pluvial floods, coastal erosion and coastal flood high 2 train +2020 AR6_WGI 1865 6 Increasing relative sea level, compounding with increasing tropical cyclone storm surge and rainfall intensity, will increase the probability of coastal city flooding high 2 train +2021 AR6_WGI 1865 7 Arctic coastal settlements are particularly exposed to climate change due to sea ice retreat high 2 train +2022 AR6_WGI 1865 19 In general, droughts have increased in several arid and semi-arid areas over the last decades medium 1 train +2023 AR6_WGI 1865 24 Dust loadings are expected to decrease over most of the Sahara and Sahel (low confidence) (Section 12.4.1), increase over Mexico and the south-west USA medium 1 train +2024 AR6_WGI 1865 29 These included an observed general decline in low-elevation snow cover, glaciers and permafrost (high confidence), which induced changes in natural hazards such as decrease in slope stability high 2 train +2025 AR6_WGI 1866 4 Warming is also affecting mountain lake surface temperatures, increasing probabilities of ice-free winters and the frequency and duration of ‘lake heatwaves’ high 2 train +2026 AR6_WGI 1866 5 Elevation-dependent warming could speed up the observed, rapid upward shifts of the freezing level height (FLH) in several mountainous regions of the world and lead to faster changes in the snowline, the glacier equilibrium-line altitude and the snow/rain transition height high 2 train +2027 AR6_WGI 1866 13 These events are projected to increase in major mountainous regions (Alps, parts of the Andes, British Columbia, North- Western North America, Calabria, Carpathian, Hindu-Kush-Himalaya, Rocky Mountains, Umbria; medium to high confidence depending on location), with potential cascading consequences of floods, landslides and lake outbursts in mountainous areas in all scenarios medium 1 train +2028 AR6_WGI 1866 14 Declines in low-elevation snow depth and seasonal extent are projected for all SSP-RCPs (see Sections 12.4.1–12.4.6), along with reductions in mountain glacier surface area, increases in permafrost temperature, decreases in permafrost thickness, changes in lake and river ice, changes in the amount and seasonality of streamflows and hydrologic droughts in snow-dominated and glacier-fed river basins (e.g., in Central Asia; Sorg et al., 2014; Reyer et al., 2017b) medium 1 train +2029 AR6_WGI 1866 15 Glacier recession could lead to the creation of new glacial lakes in places like the Himalaya-Karakoram region (Linsbauer et al., 2016) and in Alaska and Canada (Carrivick and Tweed, 2016; Harrison et al., 2018) medium 1 train +2030 AR6_WGI 1866 16 With increasing temperature and precipitation these can increase the occurrence of glacier lake outburst floods and landslides over moraine-dammed lakes high 2 train +2031 AR6_WGI 1866 17 In conclusion, mountains face complex challenges from specific climatic impact-drivers drastically influenced by climate change: regional elevation-dependent warming (high confidence), low-to-mid-altitude snow cover and snow-season decrease even as some high elevations see more snow (high confidence), glacier mass reduction and permafrost thawing (high confidence), and increases in extreme precipitation and floods in most parts of major mountain ranges medium 1 train +2032 AR6_WGI 1866 20 The SRCCL assessed an enhanced risk and severity of wildfires in tropical rainforests high 2 train +2033 AR6_WGI 1867 5 In conclusion, most tropical forests are challenged by a mix of emerging warming trends that are particularly large in comparison to historical variability medium 1 train +2034 AR6_WGI 1867 6 Water cycle changes bring prolonged drought, longer dry seasons, and increased fire weather to many tropical forests, with plants also responding to CO 2 increases medium 1 train +2035 AR6_WGI 1867 14 All regions will experience, before 2050, increased warming, an increase of extreme heat and a decrease in cold spells, regardless of the emissions trajectory high 2 train +2036 AR6_WGI 1867 17 Several global-scale studies have shown that high temperature extremes will increase everywhere high 2 train +2037 AR6_WGI 1867 23 Increases in temperatures will result in reductions in heating degree days (Arnell et al., 2019; Coppola et al., 2021b) and a widespread reduced frequency of cold extremes high 2 train +2038 AR6_WGI 1869 7 One cluster of regions – East Southern and West Southern Africa regions, the Mediterranean, Northern Central America, Western North America, several regions in South America and Australia – will experience, in addition to the aforementioned globally changing CIDs, increases in either drought/ aridity or fire weather high 2 train +2039 AR6_WGI 1869 9 A second cluster of regions including mountainous areas or regions with seasonal snow cover will experience (in addition to increases in heat extremes, more intense short-duration rainfall, and increases in coastal hazards where coasts exist) reductions in snow and ice cover and/or increases in river flooding in many cases (Western, North- Western, Central and Eastern North America, Arctic regions, Andes regions, Europe, Siberia, Central and East Asia, Southern Australia and New Zealand) high 2 train +2040 AR6_WGI 1870 6 For instance, emergence is reached by 2050 under RCP8.5 in most areas of Europe, Australia or East Asia, but it does not occur within the 21st century under RCP2.6 medium 1 train +2041 AR6_WGI 1870 8 However, even under RCP2.6, mean temperatures in tropical regions that have not already emerged are projected to emerge before 2050 medium 1 train +2042 AR6_WGI 1870 9 Extreme heat and cold: An increase in heat extremes has emerged or will emerge in the coming three decades in most land regions high 2 train +2043 AR6_WGI 1870 11 In other regions emergence is projected at the latest in the first half of the 21st century under RCP8.5 high 2 train +2044 AR6_WGI 1870 12 Relative to the end of 20th-century conditions, changes in humid heat stress as characterized by wet bulb temperature, indicates a ToE as early as in the first two decades of the 21st century in RCP8.5 at least in many tropical regions (most of Africa in the band 20°S–20°N, South Asia and South East Asia) medium 1 train +2045 AR6_WGI 1870 15 Mean precipitation: Mean precipitation changes only emerged over a few regions in the historical period (increase in Northern and Eastern Europe and decrease in West Africa and Amazonia) from observations with an S/N ratio larger than one low 0 train +2046 AR6_WGI 1870 16 The emergence of increasing precipitation before the middle of the 21st century is found across scenarios in Siberian regions, Russian Far East, Northern Europe, Arctic regions and the northernmost parts of North America high 2 test +2047 AR6_WGI 1870 21 In climate projections, the emergence of increase in heavy precipitation strongly depends on the scale of aggregation (Kirchmeier-Young et al., 2019), with, in general, no emergence before a 1.5°C or 2°C warming level, and before the middle of the century medium 1 train +2048 AR6_WGI 1870 22 Emergent increases in heavy precipitation are found in several regions when aggregated at a regional scale in Northern Europe, Northern Asia and East Asia, at latest by the end of the century in SRES A1B or RCP8.5 scenarios or when considering the decadal variability as a reference medium 1 train +2049 AR6_WGI 1870 27 Even though significant drought trends are observed in several regions with at least medium confidence (Sections 11.6 and 12.4), agricultural and ecological drought indices have interannual variability that dominates trends, as can be seen from their time series medium 1 train +2050 AR6_WGI 1871 6 The snow cover duration period is projected to emerge over large parts of Eastern and Western North America and Europe by the mid-century both in spring and autumn, and emergence is expected in the second half of the 21st century in the Arctic regions in the high RCP8.5 scenario medium 1 train +2051 AR6_WGI 1871 10 Sea ice: Sea ice area decrease in the Arctic in all seasons has already emerged from the interannual variability high 2 train +2052 AR6_WGI 1871 13 Relative sea level, coastal flood and coastal erosion: Near- coast RSLR will emerge before 2050 for RCP4.5 along the coasts of all AR6 regions (with coasts) except East Asia, the Russian Far East, Madagascar, the southern part of Eastern North America and the Antarctic regions medium 1 train +2053 AR6_WGI 1871 14 Under RCP8.5, emergence of near-coast RSLR is projected by mid-century along the coasts of all AR6 regions (with coasts), except WAN where emergence is projected to occur before 2100 (Section 9.6.1.4; Lyu et al., 2014) medium 1 train +2054 AR6_WGI 1871 26 Heat and cold CIDs (excluding frost) that have not already emerged will emerge by 2050 whatever the scenario in almost all land regions medium 1 train +2055 AR6_WGI 1871 27 The emergence of increasing precipitation before the middle of the century is also projected in Siberian regions, Russian Far East, Northern Europe and the northernmost parts of North America and Arctic regions across scenarios with the various methods and emergence definitions used high 2 train +2056 AR6_WGI 1871 31 In all ocean basins, the signal of ocean acidification in the surface ocean is projected to emerge before 2050 high 2 train +2057 AR6_WGI 1874 6 Regional-to-continental scale trends generally consistent with global-scale trends high 2 train +2058 AR6_WGI 1874 10 Extreme Precipitation EventsAll RKRs; RFC2, RFC3Frequency and intensity of heavy precipitation events increased at the global scale over a majority of land regions with good observational coverage high 2 train +2059 AR6_WGI 1874 11 Larger percentage increases in heavy precipitation observed in the northern high latitudes in all seasons, and in the mid-latitudes in the cold season high 2 train +2060 AR6_WGI 1874 12 Regional increases in the frequency and/or intensity of heavy rainfall also observed in most parts of Asia, north-west Australia, northern Europe, South-Eastern South America, and most of the USA (high confidence), and West and Southern Africa, Central Europe, the eastern Mediterranean region, Mexico, and North-Western South America medium 1 train +2061 AR6_WGI 1874 18 Increase in frequency of heavy precipitation events accelerates with warming, higher for rarer events high 2 train +2062 AR6_WGI 1874 19 DroughtAll RKRs; RFC2, RFC3Increased atmospheric evaporative demand in dry seasons over a majority of land areas due to human-induced climate change medium 1 train +2063 AR6_WGI 1874 20 Especially observed in dry summer climates in Europe, North America and Africa (high confidence).Upward trend with GSAT high 2 train +2064 AR6_WGI 1874 22 Increase in the frequency and magnitude of pluvial floods high 2 train +2065 AR6_WGI 1874 23 Increasing flood potential in urban areas where heavy precipitation projected to increase, especially at high GWLs high 2 train +2066 AR6_WGI 1874 24 Tropical Cyclones (TCs)All RKRs; RFC2, RFC3Human contribution to extreme rainfall amount from specific TC events high 2 train +2067 AR6_WGI 1874 25 Global proportion of major TC intensities likely increased over the past four decades.Increase in precipitation from TC with GSAT; average peak TC wind speeds, proportion of intense TCs, and peak wind speeds of most intense TCs increase globally with GSAT high 2 train +2068 AR6_WGI 1874 26 Decrease or lack of change in global frequency of TCs (all categories) with GSAT medium 1 train +2069 AR6_WGI 1874 28 Spatial heterogeneity with larger changes in the tropical oceans and Arctic Ocean medium 1 train +2070 AR6_WGI 1875 2 Increasing trend in more frequent concurrent heatwaves and droughts with GSAT high 2 train +2071 AR6_WGI 1875 3 More frequent concurrent (in time) extreme events at different locations with increasing GSAT, for GWLs > 2°C high 2 train +2072 AR6_WGI 1875 4 Compound flooding risk (storm surge, extreme rainfall and/or river flow) increasing with GSAT high 2 train +2073 AR6_WGI 1875 5 Trends Fire Weather TrendsRKR-B, C; RFC1,2,3Weather conditions that promote wildfire (compound hot, dry and windy events) more probable in some regions over the last century medium 1 train +2074 AR6_WGI 1875 9 Patterns of Mean WarmingRKR-B, D, F, RFC1,3,4Spatial patterns of temperature changes associated with the 0.5°C difference in GMST warming between 1991–2010 and 1960–1970 consistent with projected changes under 1.5°C and 2°C of global warming.Temperatures scale approximately linearly with GSAT, largely independently of scenario high 2 train +2075 AR6_WGI 1875 11 Antarctic polar amplification smaller than Arctic high 2 train +2076 AR6_WGI 1875 13 In the Southern Hemisphere relatively high rates of warming in subtropical continental areas of South America, Southern Africa and Australia high 2 train +2077 AR6_WGI 1875 15 Arctic Warming TrendsRKR-A,C,G,H; RFC1, RFC3 Emerged already from internal variability.Very likely more pronounced (2–2.4 times faster) than the global average over the 21st century high 2 train +2078 AR6_WGI 1875 17 Patterns of Precipitation ChangeRKR-B, D, F, RFC1, RFC3Regional patterns of recent trends, over at least the past three decades, consistent with documented increase in precipitation over tropical wet regions and decrease over dry areas.Changes in large-scale atmospheric circulation and precipitation with each 0.5°C of warming high 2 train +2079 AR6_WGI 1875 19 Some departures from linearity possible at regional scale medium 1 train +2080 AR6_WGI 1875 21 Precipitation increases in large parts of the monsoon regions, tropics and high latitudes, decreases in the Mediterranean and large parts of the subtropics high 2 train +2081 AR6_WGI 1875 25 Ocean Acidification/ pHRKR-A,B; RFC1, RFC4Virtually certain decline of surface pH globally over the last 40 years at a rate of 0.017–0.027 pH units per decade; decline also in the subsurface over the past 2–3 decades medium 1 train +2082 AR6_WGI 1875 26 Surface pH now the lowest of at least the last 26,000 years (very high confidence).Increase of net ocean carbon flux throughout the century irrespective of the emissions scenario considered high 2 train +2083 AR6_WGI 1875 27 Decrease of ocean surface pH through the 21st century, except for SSP1-1.9 and SSP1-2.6 where values increase slightly starting from 2070–2100 high 2 train +2084 AR6_WGI 1876 1 No clear evidence shifts in ENSO or associated features or its teleconnections.No change in the amplitude of ENSO variability (medium confidence); enhanced ENSO-related variability of precipitation under SSP2-4.5 and higher high 2 train +2085 AR6_WGI 1876 3 Sea Ice LossRKR-A, B, H; RFC1,3,5Arctic sea ice area decreased for all months since 1970s; strongest decrease in summer very high 3 train +2086 AR6_WGI 1876 4 Arctic sea ice younger, thinner and faster moving very high 3 train +2087 AR6_WGI 1876 5 Current pan-Arctic sea ice levels unprecedented since 1850 high 2 train +2088 AR6_WGI 1876 7 Antarctic sea ice area experienced little net change since 1979 high 2 train +2089 AR6_WGI 1876 8 The Arctic Ocean will likely become sea ice-free in September before 2050 in all considered SSP scenarios; such disappearance including several months in most years at 3°C–5°C high 2 test +2090 AR6_WGI 1876 9 Permafrost ThawRKR-A,C; RFC3,5Increases in permafrost temperatures in the upper 30 m over the past three to four decades throughout the permafrost regions high 2 train +2091 AR6_WGI 1876 11 Relative to 1995–2014: at 1.5°C and 2°C decreasing by less than 40% (medium confidence), at 2°C and 3°C by less than 75% (medium confidence), at 3°C and 5°C by more than 60% loss medium 1 train +2092 AR6_WGI 1876 12 Sea Level ChangeRKR-A,C,D,E,F, G,H; RFC1,3,4Gobal mean sea level (GMSL) is rising at an accelerated rate since the 19th century high 2 train +2093 AR6_WGI 1876 13 GMSL increase over the 20th century faster than over any preceding century in at least the last three millennia high 2 train +2094 AR6_WGI 1876 14 By 2100, likely GMSL rise with respect to 1995–2014 of 0.51 (0.40–0.69) m, 0.62 (0.50–0.81) m and 0.70 (0.58–0.91) m for, respectively, GWLs of 2.0°C, 3.0°C, and 4.0°C medium 1 train +2095 AR6_WGI 1876 18 Spring Snow Cover has seen substantial reductions in spring snow cover extent in the Northern Hemisphere since 1978 very high 3 test +2096 AR6_WGI 1876 19 Since 1981, general decline in NH spring snow water equivalent high 2 train +2097 AR6_WGI 1876 20 Relative to 1995–2014: at 1.5°C–2°C NH spring snow cover extent likely decreases by less than 20% medium 1 train +2098 AR6_WGI 1876 22 Current global glacier mass loss highly unusual over at least the last 2000 years medium 1 train +2099 AR6_WGI 1876 23 Increased rate of glacier mass loss over the last 3 to 4 decades high 2 train +2100 AR6_WGI 1876 25 For 1.5°C–2°C about 50–60% low 0 train +2101 AR6_WGI 1876 26 At 2°C–3°C about 40–50% low 0 train +2102 AR6_WGI 1876 27 At sustained 3°C–5°C 25–40% low 0 train +2103 AR6_WGI 1877 3 Ice Sheets RFC5Greenland Ice Sheet mass-loss rate increased substantially since the turn of the 21st century high 2 train +2104 AR6_WGI 1877 4 The Antarctic Ice Sheet has lost mass between 1992 and 2017 (very high confidence), with an increasing mass-loss rate over this period (medium confidence).At sustained warming levels between 1.5°C and 2°C, the ice sheets will continue to lose mass (high confidence); on time scales of multiple centuries, the Greenland and West Antarctic ice sheets will partially be lost (medium confidence); there is limited evidence that the Greenland and West Antarctic ice sheets will be lost almost completely and irreversibly over multiple millennia; at sustained warming levels between 2°C and 3°C, there is limited evidence that the Greenland and West Antarctic ice sheets will be lost almost completely and irreversibly over multiple millennia, and high confidence in increasing risk of complete loss and increasing rate of mass loss for higher warming; At sustained warming levels between 3°C and 5°C, near- complete loss of the Greenland Ice Sheet and complete loss of the West Antarctic Ice Sheet will occur irreversibly over multiple millennia (medium confidence); substantial parts or all of Wilkes Subglacial Basin in East Antarctica will be lost over multiple millennia low 0 train +2105 AR6_WGI 1877 13 Observed decline since the mid-2000s cannot be distinguished from internal variability high 2 train +2106 AR6_WGI 1877 18 Arctic Sea Ice RFC5 Abrupt change already observed.Reversible within years to decades; no tipping point or threshold beyond which loss of ice becomes irreversible high 2 train +2107 AR6_WGI 1877 21 Global Monsoon RFC5Has likely increased over the last 40 years medium 1 train +2108 AR6_WGI 1879 11 User needs and decision-making contexts are very diverse and there is no ‘one size fits all’ solution to climate services very high 3 train +2109 AR6_WGI 1879 17 Climate services require user engagement and can take various forms in which climate information and data are delivered or communicated to the users very high 3 train +2110 AR6_WGI 1880 8 Three specific examples that elaborate in more detail on specific practices and products related to those general categories are provided in Cross-Chapter Box 12.2.12.6.3 Challenges Climate services set new scientific challenges to physical climate research high 2 train +2111 AR6_WGI 1881 5 Climate services require a sustained engagement between scientists, service providers and users that is often hindered by limited resources for the co-design and co-production process high 2 train +2112 AR6_WGI 1947 8 Observed Trends and Projections in Regional Climate Most land areas have warmed faster than the global average high 2 train +2113 AR6_WGI 1947 12 Significant negative trends have been observed in the Horn of Africa and south-west of the state of Western Australia (high confidence), parts of the Russian Far East, some parts of the Mediterranean and of the Caribbean, south-east and north-east Brazil, and southern Africa medium 1 train +2114 AR6_WGI 1947 13 In many other land areas there are no significant trends in annual precipitation over the period 1960–2015 though increases in average precipitation intensity have been observed in the Sahel and South East Asia medium 1 train +2115 AR6_WGI 1947 16 Across each of the continents, higher warming is likely to occur in northern Africa, the central interior of southern and Western Africa; in North Asia; in Central Australia; in Amazonia; in Northern Europe and northern North America high 2 test +2116 AR6_WGI 1947 23 Changes in monsoons are likely to result in increased precipitation in eastern and northern China and in South Asia in summer high 2 train +2117 AR6_WGI 1947 24 Precipitation intensity will increase in many areas, including in some where annual mean reductions are likely (e.g., southern Africa) high 2 train +2118 AR6_WGI 1948 4 Snow cover has declined over Australia as has annual maximum snow mass over North America medium 1 train +2119 AR6_WGI 1948 8 Antarctic precipitation and surface mass balance showed a significant positive trend over the 20th century, while strong interannual variability masks any existing trend over recent decades1 medium 1 train +2120 AR6_WGI 1948 9 Significant warming trends are observed in other West Antarctic regions and at selected stations in East Antarctica since the 1950s medium 1 train +2121 AR6_WGI 1948 20 An equivalent approximate description using specific years would be ‘since the 1980s’.Model Evaluation, Technical Infrastructure and the Interactive Atlas The regional performance of CMIP6 global climate models (GCMs) has improved overall compared to CMIP5 in simulating mean temperature and precipitation, though large errors still exist in some regions high 2 train +2122 AR6_WGI 1948 26 In particular, regional climate models (RCMs) with polar-optimized physics are important for estimating the regional and local surface mass balance and are improved compared to reanalyses and GCMs when evaluated with observations high 2 train +2123 AR6_WGI 1948 27 There is still a lack of high-quality and high-resolution observational data to assess observational uncertainty in climate studies, and this compromises the ability to evaluate models high 2 train +2124 AR6_WGI 1984 5 Atlas.4.1.2 Findings From Previous IPCC Assessments The most recent IPCC reports, AR5 and SR1.5 (Christensen et al., 2013; Hoegh-Guldberg et al., 2018), state that over most parts of Africa, minimum temperatures have warmed more rapidly than maximum temperatures during the last 50 to 100 years medium 1 train +2125 AR6_WGI 1984 6 In the same period, minimum and maximum temperatures have increased by more than 0.5°C relative to 1850–1900 high 2 train +2126 AR6_WGI 1984 12 According to SROCC, eastern Africa like other regions with smaller glaciers is projected to lose more than 80% of its glaciers by 2100 under RCP8.5 medium 1 train +2127 AR6_WGI 1984 18 The consequence of increased temperature and evapotranspiration, and decreased precipitation amount, in interaction with climate variability and human activities, have contributed to desertification in dryland areas in sub-Saharan Africa (medium confidence) as reported in SRCCL (Mirzabaev et al., 2019).Atlas.4.2 Assessment and Synthesis of Observations, Trends and Attribution Figure Atlas.11 shows observed trends in annual mean surface temperature and indicates it has been rising rapidly over Africa from 1961 to 2015 and with significant increases in all regions of 0.1°C–0.2°C per decade and higher over some northern, eastern and south-western regions high 2 train +2128 AR6_WGI 1985 1 Southern African rainfall shows a significant downtrend of –0.013 mm day–1 year –1 in recent decades and –0.003 mm day–1 year –1 for longer periods during 1900–2010 low 0 train +2129 AR6_WGI 1985 6 Enhanced rainfall intensity since the mid-1980s over the Sahel (Maidment et al., 2015; Sanogo et al., 2015) is associated with increased greenhouse gases suggesting an anthropogenic influence medium 1 train +2130 AR6_WGI 1985 8 Moreover, later onset and earlier cessation of eastern Africa rainfall is associated with a delayed and then faster movement of the tropical rainband northwards during the boreal spring and northward shift of the Saharan heat low (Wainwright et al., 2019), driven by anthropogenic carbon emissions and changing aerosol forcings medium 1 train +2131 AR6_WGI 1985 10 However, decadal natural variability from SST variations over the Pacific Ocean has also been associated with the drying trend of East Africa (Wang et al., 2014; Hoell et al., 2017) with an anthropogenic-forced rapid warming of Indian Ocean SSTs medium 1 train +2132 AR6_WGI 1987 14 The rate of surface temperature increase has generally been more rapid in Africa than the global average and by at least 0.1°C–0.2°C during 1961–2015 high 2 test +2133 AR6_WGI 1987 15 Minimum temperatures have increased more rapidly than maximum temperatures over inland southern Africa medium 1 train +2134 AR6_WGI 1987 16 Since 1970, mean temperature over East Africa has shown an increasing trend but showed a decreasing trend in the previous 40 years (medium confidence).The Horn of Africa has experienced significantly decreased rainfall during the long rains season from March to May (high confidence) and drying trends in this and other parts of Africa are attributable to oceanic influences high 2 train +2135 AR6_WGI 1987 19 The enhanced rainfall intensity over the Sahel in the last two decades is associated with increased greenhouse gases indicating an anthropogenic influence medium 1 train +2136 AR6_WGI 1987 20 Relative to the late 20th century, annual mean temperature over Africa is projected to rise faster than the global average very high 3 train +2137 AR6_WGI 1987 21 The central interiors of southern and northern Africa are likely to warm faster than equatorial and tropical regions high 2 train +2138 AR6_WGI 1987 22 There are contrasting signals in the projections of rainfall over some parts of Africa until the end of the 21st century high 2 train +2139 AR6_WGI 1987 24 However, northern Africa and the south-western parts of South Africa are likely to have a reduction in precipitation under higher warming levels high 2 train +2140 AR6_WGI 1987 25 Over Western Africa, rainfall is projected to decrease in the western Sahel sub-region (medium confidence) and increase in the central Sahel sub-region (low confidence) and along the Guinea coast sub-region medium 1 train +2141 AR6_WGI 1987 26 Rainfall amounts are projected to increase over Eastern Africa medium 1 train +2142 AR6_WGI 1987 27 Southern Africa is projected to have a reduction in annual mean rainfall but increases in rainfall intensity by 2100 medium 1 train +2143 AR6_WGI 1989 33 Precipitation trends over East Asia show considerable regional differences medium 1 train +2144 AR6_WGI 1991 20 Atlas.5.1.5 Summary In East Asia annual mean temperature has been increasing since the 1950s high 2 train +2145 AR6_WGI 1991 22 Trends of annual precipitation show considerable regional differences with areas of both increases and decreases (medium confidence), and with increases over north-west China and South Korea high 2 train +2146 AR6_WGI 1991 23 Agricultural intensification through oasis expansion in Xinjiang region has increased summer precipitation in the Tian Shan mountains high 2 train +2147 AR6_WGI 1992 10 The SR1.5 noted that future, higher levels of warming lead to greater impacts in key systems such as the Siberian ecosystems, identified as one of the threatened systems (‘Reason for Concern 1 – RFC1’; Hoegh-Guldberg et al., 2018) with impacts at 2°C expected to be greater than those at 1.5°C medium 1 train +2148 AR6_WGI 1992 15 In addition, significant warming in the last decade has halved the cooling trend in southern WSB from –0.6°C per decade during 1976–2012 to –0.3°C per decade during 1976–2018 high 2 train +2149 AR6_WGI 1993 9 The latter reanalysis also underestimates summer precipitation and shows large wet biases over north-east Asia during spring and underestimates mean seasonal temperature over north-east Asia in spring (MAM), autumn (SON), and winter (DJF), but overestimates it in summer (JJA) compared with the CRU dataset medium 1 train +2150 AR6_WGI 1993 16 For CMIP5, models with higher resolution do not always perform better than those with lower resolutions medium 1 train +2151 AR6_WGI 1994 6 The highest warming has been found in spring in ESB and RFE, strengthening from south to north with linear trends of 0.8°C–1.2°C per decade over the 1976–2014 period high 2 train +2152 AR6_WGI 1994 7 A temperature decrease was identified just in winter in the southern part of WSB and ESB as a result of natural variability, but halved from –0.6°C per decade in 1976–2012 to –0.3°C per decade for the longer 1976–2018 period due to recent warmer winters high 2 train +2153 AR6_WGI 1994 8 Over North Asia annual precipitation increases with estimated trends of 5–15 mm per decade in the 1976–2014 period have been recorded with an exception over the Kamchatka and the Chukchi peninsulas, where decreases of up to –20 mm per decade in the same period have been found medium 1 train +2154 AR6_WGI 1994 10 Most of the CMIP5 and some CMIP6 GCMs overestimate the annual mean air temperature and precipitation over the North Asia region medium 1 train +2155 AR6_WGI 1994 11 GCMs generally represent the observed decadal temperature trend (medium confidence) and biases primarily come from the winter (DJF) season high 2 train +2156 AR6_WGI 1994 13 Sparsity of observational data particularly in the northern part of ESB and the whole of the RFE results in low confidence in the assessments of model performance in North Asia.Surface air temperature and precipitation in North Asia are projected to increase further high 2 train +2157 AR6_WGI 1994 24 CMIP5 models projected for the 21st century a significant increase in temperature over South Asia (high confidence from robust evidence) and in projections of increased summer monsoon precipitation medium 1 train +2158 AR6_WGI 1995 2 The Indian summer monsoon circulation was found to have weakened, but this was compensated by increased local atmospheric moisture content leading to more rainfall medium 1 train +2159 AR6_WGI 1995 6 The annual average of daily maximum and minimum temperatures has increased over almost all Pakistan with a faster increasing trend in the south high 2 train +2160 AR6_WGI 1995 9 Concurrently, the frequency of heavy precipitation events has increased over India, while the frequency of moderate rain events has decreased since 1950 high 2 train +2161 AR6_WGI 1995 12 A global modelling study with high resolution over South Asia (Sabin et al., 2013) indicated that a juxtaposition of regional land-use changes, anthropogenic-aerosol forcing and the rapid warming signal of the Equatorial Indian Ocean was crucial to simulate the observed Indian summer monsoon weakening in recent decades medium 1 train +2162 AR6_WGI 1995 13 A dipole-like structure in summer monsoon rainfall trends is observed over the northern Indo-Pakistan area with significant increases over Pakistan and decreases over central north India resulting from strengthening (weakening) of vertically integrated meridional moisture transport over the Arabian Sea (Bay of Bengal) low 0 train +2163 AR6_WGI 1995 23 Observations and model simulations showed that the increasing temperature of frozen grounds is leading to thawing and reduced depth of permafrost, with further significant reductions projected under future global warming scenarios medium 1 train +2164 AR6_WGI 1996 21 With continued global warming and anticipated reductions in anthropogenic aerosol emissions in the future, CMIP5 models project an increase in the mean and variability of summer monsoon precipitation over India by the end of the 21st century, together with substantial increases in daily precipitation extremes medium 1 train +2165 AR6_WGI 1997 2 Atlas.5.3.5 Summary Mean, minimum and maximum daily temperatures in South Asia are increasing and winters are getting warmer faster than summers high 2 train +2166 AR6_WGI 1997 10 Over the same time periods CMIP6 models project an increase in annual precipitation in the range 14–36% under SSP5-8.5 and 0.4–16% under SSP1-2.6 medium 1 train +2167 AR6_WGI 1998 2 Over South East Asia, there has been a significant increase in the amount of precipitation and its extremes with La Niña episodes in the past decades, especially during the winter monsoon period high 2 train +2168 AR6_WGI 1998 3 Figure Atlas.11 shows trends in mean temperature and precipitation during 1961–2015 for two global datasets, indicating a significant overall warming over South East Asia (high confidence), with higher rates of warming in Malaysia, Indonesia, and the southern areas of mainland South East Asia low 0 train +2169 AR6_WGI 1998 4 Annual mean precipitation trends (Atlas.1.4.1 and the Interactive Atlas, which includes the regional dataset Aphrodite) over the region are mostly not significant except for increases over parts of Malaysia, Vietnam and the southern Philippines medium 1 train +2170 AR6_WGI 1998 7 Atlas.5.4.3 Assessment of Model Performance Performance in simulating rainfall over South East Asia varies among CMIP5 GCMs high 2 train +2171 AR6_WGI 1998 26 Projections of future rainfall changes are highly variable among sub-regions of South East Asia and among the models high 2 train +2172 AR6_WGI 1999 7 Atlas.5.4.5 Summary It is virtually certain that annual mean temperature has been increasing in South East Asia in the past decades while changes in annual mean precipitation are less spatially coherent though with some increasing trends over parts of Malaysia, Vietnam and the southern Philippines medium 1 train +2173 AR6_WGI 1999 10 Projected changes in rainfall over South East Asia vary, depending on model, sub-region and season (high confidence), with consistent projections of increases in annual mean rainfall from CMIP5 and CMIP6 over most land areas (medium confidence) and decreases in summer rainfall from CORDEX projections over much of Indonesia medium 1 train +2174 AR6_WGI 2000 6 WCA includes high mountains with enhanced warming above 500 m where, regardless of the emissions scenario, decreases in snow cover are projected due to increased winter snowmelt and more precipitation falling as rain high 2 train +2175 AR6_WGI 2000 10 This increases the uncertainty in both temperature and precipitation trends, particularly for elevated areas high 2 train +2176 AR6_WGI 2000 13 A strong increase in annual surface air temperature of 0.27°C–0.47°C per decade has been found over WCA between 1960 and 2013 very high 3 train +2177 AR6_WGI 2000 17 The plateau of Iran has experienced significant increases in the average monthly values of daily maximum and minimum temperatures with spatially varying rates of 0.1°C–0.3°C up to 0.3°C–0.4°C per decade and greater spatial variation in minimum temperatures high 2 train +2178 AR6_WGI 2000 24 Over the elevated part of eastern WCA precipitation increases in the range of 1.3–4.8 mm per decade during 1960–2013 were observed very high 3 train +2179 AR6_WGI 2000 27 A decreasing trend of precipitation is reported for ARP with the mean value of –6.3 mm per decade (range of –30 mm–16 mm) for the period 1978–2019 low 0 train +2180 AR6_WGI 2000 28 The same decreasing trend in precipitation totals and an increasing trend in the number of consecutive dry days are found for most of the Iranian Plateau medium 1 train +2181 AR6_WGI 2001 5 RCM simulations using the CORDEX-MENA domain reproduce the main features of the mean surface climatology over ARP with moderate biases high 2 train +2182 AR6_WGI 2002 7 Annual precipitation change over ARP since 1970 is estimated at –6.3 mm per decade (and in the range of –30 to 16 mm per decade) and over WCA is generally not significant except over the elevated part of eastern WCA where increases between 1.3 mm and 4.8 mm per decade during 1960–2013 have been observed very high 3 train +2183 AR6_WGI 2002 8 In mountainous areas, the scarcity and decline of the number of observation sites since the end of the former Soviet Union in 1991 increase the uncertainty of the long-term temperature and precipitation estimates high 2 train +2184 AR6_WGI 2002 9 Mean temperature biases in RCMs are within ±3°C in South West Asia, and annual precipitation biases are positive in almost all parts of the region, except over the ARP where they are negative in the wet season (November to April) and over WCA in winter and spring (from December to May) medium 1 train +2185 AR6_WGI 2002 12 Further warming over South West Asia is projected in the 21st century to be greater than the global average, with rates varying from 0.25°C to 0.8°C per decade depending on the season and scenario, and the maximum rates found in the northern part of the region in summer high 2 train +2186 AR6_WGI 2002 14 CMIP6 projected changes in annual precipitation totals are in the range of –3 to 29% (SSP1-2.6) and 12–107% (SSP5-8.5) in ARP medium 1 train +2187 AR6_WGI 2002 15 Strong spatio-temporal differences with overall precipitation decreases are projected in the central and northern parts of WCA in summer (JJA) with increases in winter (DJF) medium 1 train +2188 AR6_WGI 2002 24 For example, while annual rainfall has been significantly increasing in north-western Australia since the 1950s (very high confidence), it has been decreasing in the north-east of the South Island of New Zealand over 1950–2004 very high 3 train +2189 AR6_WGI 2003 1 Likewise, however, there is a projected increase in the frequency of drought in southern Australia (medium confidence) and in many parts of New Zealand medium 1 train +2190 AR6_WGI 2003 3 The AR5 reported mean sea levels have also increased in Australia and New Zealand at average rates of relative sea level rise of 1.4 ± 0.6 mm yr –1 from 1900 to 2011, and 1.7 ± 0.1 mm yr –1 from 1900 to 2009, respectively very high 3 train +2191 AR6_WGI 2003 4 The assessment found that the volume of ice in New Zealand has declined by 36–61% from the mid- to late 1800s to the late 1900s, while late-season significant snow depth has also declined in three out of four Snowy Mountain sites in Australia between 1957 and 2002 high 2 test +2192 AR6_WGI 2003 6 On the other hand, the volume of winter snow and the number of days with low-elevation snow cover in New Zealand are projected to decrease in the future (very high confidence), while both snow depth and area are projected to decline in Australia very high 3 train +2193 AR6_WGI 2003 8 It also reports on the vulnerability of some Australian communities and ecosystems to sea level rise, increases in the intensity and duration of marine heatwaves driven by human influence high 2 train +2194 AR6_WGI 2003 15 For a longer-term perspective based on high-quality regional datasets, Figure Atlas.20 shows Australasia has warmed over the last century very high 3 train +2195 AR6_WGI 2007 1 Snow cover is likely to decrease throughout the region at high altitudes in both Australia and New Zealand high 2 train +2196 AR6_WGI 2007 20 According to AR5 (Christensen et al., 2013), significant positive trends of temperature have been observed in Central America high 2 test +2197 AR6_WGI 2007 21 In addition, changes in climate variability and in extreme events have severely affected the region medium 1 train +2198 AR6_WGI 2007 23 El Niño and La Niña teleconnections are projected to move eastwards in the future (medium confidence), while changes in their effects on other regions, including Central America and the Caribbean is uncertain medium 1 train +2199 AR6_WGI 2007 26 Precipitation change is projected to vary between +10% and –25% medium 1 train +2200 AR6_WGI 2007 29 Jones et al., 2016a; Hidalgo et al., 2017), with the largest increases in the North American Monsoon region high 2 train +2201 AR6_WGI 2009 7 Positive trends in the duration of the MSD have been found in this region over the past four decades low 0 train +2202 AR6_WGI 2009 10 Particularly relevant for this region are increased model resolution and a better representation of the land surface processes high 2 train +2203 AR6_WGI 2009 11 Regional climate models (RCMs) forced with reanalyses and atmosphere-only global climate models provide simulations with a reasonably good performance over the core North American Monsoon region, mostly in NCA high 2 train +2204 AR6_WGI 2009 12 RCMs also reproduce the seasonal spatial patterns of temperature and the bimodal rainfall characteristics of the NCA, SCA and CAR high 2 train +2205 AR6_WGI 2009 13 RCM simulations in the region do not necessarily improve with the size of the domain, as important features of the regional circulation and key rainfall climate features, such as the CLLJ and MSD, are well represented for a variety of domains of different sizes high 2 train +2206 AR6_WGI 2009 22 Global and regional models consistently project warming in the whole region for the end of the century, under RCP4.5 and RCP8.5 for CMIP5 projections with greater warming for continental compared to insular territories, likely reaching values between 2°C and 4°C high 2 train +2207 AR6_WGI 2010 15 Atlas.7.1.5 Summary Significant warming trends between 0.2°C and 0.3°C per decade have been observed in the three reference regions of Central America in the last 30 years, with the largest increases in the North American Monsoon region high 2 train +2208 AR6_WGI 2010 20 Under moderate future emissions overall negative but non-significant precipitation trends are projected for the 21st century low 0 test +2209 AR6_WGI 2011 4 Atlas.7.2.1.2 Findings From Previous IPCC Assessments According to AR5 WGII Chapter 27 (Magrin et al., 2014), during the last decades of the 20th century, observational studies identified significant trends in precipitation and temperature in South America high 2 train +2210 AR6_WGI 2011 7 The AR5 WGI (Flato et al., 2013) noted that climate simulations from CMIP3 and CMIP5 models were able to represent well the main climatological features, such as seasonal mean and annual cycle high 2 train +2211 AR6_WGI 2011 10 Atlas.7.2.2 Assessment and Synthesis of Observations, Trends and Attribution Studies on climatic trends in South America indicate that mean temperature and extremely warm maximum and minimum temperatures have shown an increasing trend high 2 train +2212 AR6_WGI 2011 12 Regionally, analyses of temperatures point to an increased warming trend high 2 train +2213 AR6_WGI 2011 14 Andean temperatures showed significant warming trends, especially at inland and higher-elevation sites, while trends are non-significant or negative at coastal sites high 2 train +2214 AR6_WGI 2011 17 In general, the spatial patterns of observed trends in temperature are more consistent than for precipitation across the whole of South America medium 1 train +2215 AR6_WGI 2011 19 The most consistent evidence of positive rainfall trend occurs in the southern part of the La Plata basin high 2 train +2216 AR6_WGI 2011 23 Over the subtropical Andes, central Chile shows a robust signal of declining precipitation since 1970 high 2 train +2217 AR6_WGI 2011 24 Observational studies show that the dry-season length over southern Amazonia has increased significantly since 1979 high 2 train +2218 AR6_WGI 2011 27 In general, these changes are attributed mainly to decadal climate fluctuations high 2 train +2219 AR6_WGI 2012 1 Recent publications based on observational and modelling evidence assessed that anthropogenic forcing in CMIP5 models explains the overall warming high 2 train +2220 AR6_WGI 2012 5 In summary, analyses of historical temperature time series point strongly to an increased warming trend high 2 train +2221 AR6_WGI 2012 6 Annual rainfall has increased over South-Eastern South America and decreased in most tropical land regions, particularly in central Chile high 2 train +2222 AR6_WGI 2012 7 The number and strength of extreme events, such as extreme temperatures, droughts and floods, have already increased medium 1 train +2223 AR6_WGI 2012 11 Atlas.7.2.3 Assessment of Model Performance Since AR5 the number of publications on climate model performance and their projections in South America has increased, particularly for regional climate modelling studies (Giorgi et al., 2009; Boulanger et al., 2016; Ambrizzi et al., 2019) and the understanding of their strengths and weaknesses high 2 train +2224 AR6_WGI 2012 13 However, significant biases persist mainly at regional scales high 2 train +2225 AR6_WGI 2012 16 Most models show a dry bias over SES (Díaz and Vera, 2017; Barros and Doyle, 2018; Solman and Blázquez, 2019; Díaz et al., 2021) associated with an underestimation of the northern flow that brings water vapour into the region medium 1 train +2226 AR6_WGI 2012 17 The biases in seasonal precipitation, annual precipitation and climate extremes over several regions of South America were reduced, including the Amazon, central South America, Bolivia, eastern Argentina and Uruguay, in the CMIP5 models when compared to those of CMIP3 medium 1 train +2227 AR6_WGI 2012 21 Regional climate model (RCM) simulations over South America can reproduce the main features of temperature and precipitation in terms of both spatial distributions (Solman et al., 2013; Falco et al., 2019) and seasonal cycles over the different climate regimes, including the main SAmerM features high 2 train +2228 AR6_WGI 2012 22 However, RCMs showed systematic biases such as precipitation overestimations and temperature underestimations along the Andes throughout the year high 2 train +2229 AR6_WGI 2012 24 Temperature overestimation and precipitation underestimation over La Plata basin (in SES) are also RCM common biases, with the warm bias amplified for austral summer and the dry bias amplified for the rainy season high 2 train +2230 AR6_WGI 2012 25 Despite their relevance, RCM simulations at very high resolution (less than 10 km) are still few in South America high 2 train +2231 AR6_WGI 2013 1 Precipitation simulations based on ESD models are able to reproduce mean precipitation over tropical and subtropical South American regions, especially over maximum precipitation areas in western Colombia, south-eastern Peru, central Bolivia, Chile and the La Plata basin medium 1 train +2232 AR6_WGI 2013 3 Overall, climate modelling has made some progress in the past decade but there is no model that performs well in simulating all aspects of the present climate over South America high 2 train +2233 AR6_WGI 2013 6 Finally, observational reference datasets, such as reanalysis products, used in the calibration and validation of climate models can also be quite uncertain and may explain part of the apparent biases present in climate models high 2 train +2234 AR6_WGI 2013 9 The largest warmings over the South American continent are projected for the Amazon basin (SAM and NSA) and the central Andes range (southern SAM, northern SWS and south-eastern NWS; Figure Atlas.22), especially during the dry and dry-to-wet transition seasons (austral winter and spring) high 2 train +2235 AR6_WGI 2013 15 Under high RCPs, the CMIP5 ensemble projects that all Brazilian regions will experience more rainfall variability in the future, so drier dry periods and wetter wet periods on daily, weekly, monthly and seasonal time scales, despite the future changes in mean rainfall being currently uncertain medium 1 test +2236 AR6_WGI 2013 23 Long-term observed precipitation trends show an increase over South-Eastern South America and decreases in most tropical land regions high 2 train +2237 AR6_WGI 2013 26 On the other hand, there is still a lack of high-quality and high- resolution observational data that may explain part of the important biases present in climate models high 2 train +2238 AR6_WGI 2013 27 Climate model projections show a general increase in annual mean surface temperature over the coming century for all emissions scenarios (RCPs and SSPs) high 2 train +2239 AR6_WGI 2014 29 For a 2°C global warming level, an increase in runoff is projected for north-eastern Europe while decreases are projected in the Mediterranean region, where runoff differences between 1.5°C and 2°C global warming will be most prominent medium 1 train +2240 AR6_WGI 2014 30 According to SROCC (Hock et al., 2019b) the RCP8.5 projections lead to a loss of more than 80% of the ice mass from small glaciers by the end of century in Central Europe high 2 train +2241 AR6_WGI 2015 25 In the European Mediterranean, observed land precipitation trends show pronounced variability within the region, with magnitude and sign of trend in the past century depending on time period and exact study region medium 1 train +2242 AR6_WGI 2018 24 Regional warming is virtually certain to extend the observed downward trends in snow accumulation, snow water equivalent and length of the snow cover season in NEU and at low altitudes in mountainous areas in the Alps and Pyrenees very high 3 train +2243 AR6_WGI 2019 3 The representation of mean European climate features by GCMs and RCMs is improved compared to previous IPCC assessments medium 1 train +2244 AR6_WGI 2019 4 The added value of regional downscaling of GCMs by RCM projections for summer mean temperature, precipitation and shortwave radiation is constrained by the representation of processes that lead to a systematic difference between RCM and driving GCM, such as aerosol forcing medium 1 train +2245 AR6_WGI 2019 6 In the European Mediterranean trends in annual mean precipitation contain substantial spatial and temporal variability medium 1 train +2246 AR6_WGI 2019 10 At high latitudes and low-altitude mountain areas in Europe strong declines in snow accumulation are virtually certain to occur with further increasing regional temperatures very high 3 train +2247 AR6_WGI 2020 9 Atlas.9.1.2 Findings From Previous IPCC Assessments The IPCC AR5 (Bindoff et al., 2013; Hartmann et al., 2013) found that the climate of North America has changed due to anthropogenic causes (high confidence), in particular with primarily increasing annual precipitation and annual temperature very high 3 train +2248 AR6_WGI 2020 20 There is a detectable anthropogenic influence medium 1 train +2249 AR6_WGI 2020 21 Compared to temperature, trends in annual precipitation over 1961–2015 are generally non-significant though there are consistent positive trends over parts of ENA and CNA (Figure Atlas.11 and Daymet, Interactive Atlas) high 2 train +2250 AR6_WGI 2021 5 There is evidence of a recent decline in the overall North American annual maximum snow mass, with a trend for non-alpine regions above 40°N during 1980–2018 estimated from the bias-corrected GlobSnow 3.0 data medium 1 train +2251 AR6_WGI 2021 10 However, reported snow-decline trends are statistically significant only for a fraction of the concerned areas or locations low 0 train +2252 AR6_WGI 2023 1 The northern regions and ENA all show steady increases with the global warming levels very high 3 train +2253 AR6_WGI 2023 10 In summary, NEN, NWN and most of ENA will very likely experience increased annual mean precipitation, with greater increases at higher levels of warming very high 3 train +2254 AR6_WGI 2024 1 Projected changes in summer are highly uncertain throughout other regions apart from the far northern parts of NEN and NWN which will likely experience increases high 2 train +2255 AR6_WGI 2024 6 These investigations demonstrate the potential of very-high-resolution simulations to add important dimensions to our understanding of regional climate change, though not necessarily to reduce uncertainty high 2 train +2256 AR6_WGI 2024 16 Across near-Arctic latitudes of North America, increases are exceptionally pronounced, greater than 0.5°C per decade high 2 train +2257 AR6_WGI 2025 17 There is very high confidence that global mean sea level rise has accelerated in recent decades which, combined with increases in tropical cyclone winds and rainfall and increases in extreme waves, has exacerbated extreme sea level events and coastal hazards high 2 train +2258 AR6_WGI 2025 18 It is virtually certain that during the 21st century, the ocean will transition to unprecedented conditions with further warming and acidification virtually certain, increased upper ocean stratification very likely and continued oxygen decline medium 1 train +2259 AR6_WGI 2025 21 Global mean sea level will continue to rise and there is high confidence that the consequent increases in extreme levels will result in local sea levels in most locations that historically occurred once per century occurring at least annually by the end of the century under all RCP scenarios high 2 train +2260 AR6_WGI 2026 9 Recent analysis of station data showed spatial variations in the mostly decreasing but non-significant trends in annual and extreme rainfall over the western Pacific from 1961 to 2011 low 0 train +2261 AR6_WGI 2026 33 In summary, the ability of climate models to simulate the climate over the region has improved in many key respects with the application of increased model resolution and a better representation of the land surface processes of particular importance in these advances high 2 train +2262 AR6_WGI 2026 34 Regional climate models (RCMs) simulate realistically seasonal surface temperature and precipitation patterns including the bimodal rainfall in the precipitation annual cycle although with some timing biases in some regions high 2 train +2263 AR6_WGI 2026 35 The important regional circulation and precipitation features, the Caribbean low-level jet and the midsummer drought (MSD), are well represented over a variety of RCM domains covering the region high 2 train +2264 AR6_WGI 2028 22 There are fewer significant trends in precipitation in these regions though several locations in the Caribbean have detectable decreasing trends high 2 train +2265 AR6_WGI 2029 9 These modes show no sustained trend since the late 19th century high 2 train +2266 AR6_WGI 2029 12 The SROCC (IPCC, 2019a) finds very high confidence that global mean sea level rise has accelerated in recent decades which has exacerbated extreme sea level events and flooding high 2 train +2267 AR6_WGI 2029 13 It will continue to rise with consequent increases in extreme levels so that the historical one-in-a-century extreme local sea level will become an annual event by the end of the century under all RCP scenarios high 2 train +2268 AR6_WGI 2029 18 Some general observed climate trends include higher magnitude and frequency of temperatures including warm extremes (high confidence) (Section 12.4.7.1, Table 11.13 and Atlas.10.2), declines in high-intensity rainfall events (low to medium confidenc e) (Table 11.14), regional sea level rises with strong storm surges and waves resulting in increased coastal flood intensity (high confidence) (Section 12.4.7.4 and Atlas.10.2), and increased intensity and intensification rates of tropical cyclones at global scale medium 1 train +2269 AR6_WGI 2029 19 No significant long-term trends are observed for annual Caribbean rainfall over the 20th century low 0 train +2270 AR6_WGI 2029 20 Over the western Pacific, generally decreasing but non-significant trends are noted in annual total rainfall from 1961 to 2011 low 0 train +2271 AR6_WGI 2031 8 Information on future climate changes Small Islands will very likely continue to warm this century, though at a rate less than the global average (Figure Atlas.28), with consequent increased frequency of warm extremes for the Caribbean and western Pacific islands, and heatwave events for the Caribbean high 2 train +2272 AR6_WGI 2031 9 Annual and JJA rainfall declines are likely for some Indian and southern Pacific ocean regions with drying over southern French Polynesia (attributed partially to greenhouse gas increases) and farther east clearly evident in CMIP5 and CMIP6 projections high 2 train +2273 AR6_WGI 2031 12 This JJA drying has been linked to a future strengthening of the Caribbean low level jet (CLLJ) (Taylor et al., 2013a), a westward expansion and intensification of the NASH, stronger low-level easterlies over the region, a southwardly-placed eastern Pacific ITCZ (Rauscher et al., 2008), and changing dynamics due to increased greenhouse gas concentrations very high 3 train +2274 AR6_WGI 2033 18 In the AP, concomitant increase in temperature and foehn winds due to positive SAM caused increased surface melting over the Larsen ice shelves medium 1 train +2275 AR6_WGI 2033 19 Strong warming between the mid-1950s and the late 1990s led to the collapse of the Larsen B ice shelf in 2002, which had been intact for 11,000 years (medium confidence).Snowfall increased over the Antarctic Ice Sheet over AP and WAN, offsetting some of the 20th-century sea level rise medium 1 train +2276 AR6_WGI 2033 20 Longer records suggest either a decrease in snowfall over the Antarctic Ice Sheet over the last 1000 years or a statistically negligible change over the last 800 years low 0 train +2277 AR6_WGI 2033 23 There is medium agreement but limited evidence of an anthropogenic forcing effect on Antarctic ice-sheet mass balance low 0 train +2278 AR6_WGI 2034 3 The century-scale warming trend in the AP is very likely an emerging signal compared to natural variability, while the WAN warming trend falls in the high end of century-scale trends over the last 2000 years medium 1 train +2279 AR6_WGI 2034 21 The SMB of EAN increased during the 20th century which mitigated global mean sea level rise by 0.77 ± 0.40 mm per decade during 1901–2000 medium 1 train +2280 AR6_WGI 2035 4 More evidence has emerged showing the importance of the Pacific–South American pattern, ENSO and Pacific Ocean convection, and large-scale blocking causing warm-air intrusions and both extreme precipitation and melt events, responsible for large interannual SMB variability high 2 train +2281 AR6_WGI 2038 3 Significant warming trends are observed in other West Antarctic regions and at selected stations in East Antarctica medium 1 train +2282 AR6_WGI 2038 4 Antarctic precipitation and SMB showed a significant positive trend over the 20th century according to the ice cores, while large interannual variability masks any existing trend over the satellite period since the end of the 1970s medium 1 train +2283 AR6_WGI 2038 8 Under all assessed emissions scenarios, both West and East Antarctica are very likely to have higher annual mean surface air temperatures and more precipitation, which will have a dominant influence on determining future changes in the SMB high 2 train +2284 AR6_WGI 2038 17 Arctic surface air temperatures have increased from the mid-1950s, with feedbacks from loss of sea ice and snow cover contributing to the amplified warming (high confidence) (IPCC, 2018c), and have likely increased by more than double the global average over the last two decades high 2 train +2285 AR6_WGI 2038 18 Arctic snow cover in June has declined from 1967 to 2018 high 2 train +2286 AR6_WGI 2038 19 Arctic glaciers are losing mass (very high confidence) and this along with changes in high-mountain snowmelt have caused changes in hydrology, including river runoff, that are projected to continue in the near term high 2 train +2287 AR6_WGI 2038 20 The rate of ice loss from the Greenland Ice Sheet has increased; during 2006–2015 the loss was 278 ± 11 Gt yr–1 with the rate for 2012–2016 higher than for 2002–2011 and several times higher than during 1992–2001 high 2 train +2288 AR6_WGI 2038 21 The Arctic sea ice area is declining in all months of the year (very high confidence) with the September sea ice minimum very likely having reduced by 12.8 ± 2.3% per decade during the satellite era (1979–2018) to levels unprecedented for at least 1000 years medium 1 train +2289 AR6_WGI 2038 25 No further losses are projected under RCP2.6 whereas a further 15–25% reduction in snow cover duration is projected by the end of century under RCP8.5 high 2 train +2290 AR6_WGI 2039 16 AMAP reported Arctic precipitation increases of 1.5–2.0% per decade, with the strongest increase in the cold season (October–May) medium 1 train +2291 AR6_WGI 2039 22 Rainfall frequency is estimated to have increased over the Arctic by 2.7–5.4% over 2000–2016 (Boisvert et al., 2018) with more frequent rainfall events reported for NEU and ARO (Svalbard; Maturilli et al., 2015; AMAP , 2019), and winter rain totals and frequency have increased in Svalbard since 2000 medium 1 train +2292 AR6_WGI 2039 28 The CMIP5 models reproduce the observed Arctic warming over the past century medium 1 train +2293 AR6_WGI 2040 34 In the south-west, north-east and north-west, SMB turns negative or close to zero after 2000 and remains above zero in other regions medium 1 test +2294 AR6_WGI 2040 38 The Arctic warming may be as much as 4°C in the annual mean and 7°C in late autumn under 2°C global warming, regardless of which scenario is considered high 2 train +2295 AR6_WGI 2040 43 In projections from 30 CMIP5 models, winter warming over ARO varies regionally from 3°C to 5°C by mid-century and 5°C to 9°C by late-century under RCP4.5 high 2 train +2296 AR6_WGI 2041 5 End-of-the-century warming is approximately twice as large under RCP8.5 demonstrating the impact of the lower emissions under RCP4.5 high 2 train +2297 AR6_WGI 2041 22 The interannual variability of Arctic precipitation will likely increase markedly (up to 40% over the 21st century), especially in summer medium 1 train +2298 AR6_WGI 2041 26 Consistent with the generally higher warming in CMIP6, compared to CMIP5, the projected precipitation increase is also higher high 2 train +2299 AR6_WGI 2041 30 The pattern and amplitude of precipitation changes agree in CORDEX simulations with their driving CMIP5 models high 2 train +2300 AR6_WGI 2041 33 The CMIP5 models show a decrease in annual Arctic snowfall under both RCP4.5 and RCP8.5 high 2 train +2301 AR6_WGI 2041 36 With ongoing warming and polar amplification in the Arctic, the Greenland Ice Sheet SMB will inevitably continue to change high 2 train +2302 AR6_WGI 2042 1 The CMIP5 models reproduce the observed Arctic warming over the past century but overestimate the amplified Arctic warming in the recent decades medium 1 train +2303 AR6_WGI 2042 2 Arctic CORDEX simulations show adequate skill in capturing regional temperature and precipitation patterns and precipitation extremes high 2 train +2304 AR6_WGI 2042 3 SMB models have improved due to increased availability and quality of remotely sensed and in situ observations, and an ensemble mean of SMB model simulations provides the best estimate of the present- day SMB medium 1 train +2305 AR6_WGII 20 9 These include increased heat-related human mortality (medium confidence), warm-water coral bleaching and mortality (high confidence), and increased drought-related tree mortality high 2 train +2306 AR6_WGII 20 11 Adverse impacts from tropical cyclones, with related losses and damages19, have increased due to sea level rise and the increase in heavy precipitation medium 1 train +2307 AR6_WGII 20 12 Impacts in natural and human systems from slow-onset processes29 such as ocean acidification, sea level rise or regional decreases in precipitation have also been attributed to human induced climate change high 2 train +2308 AR6_WGII 20 14 The extent and magnitude of climate change impacts are larger than estimated in previous assessments high 2 train +2309 AR6_WGII 20 15 Widespread deterioration of ecosystem structure and function, resilience and natural adaptive capacity, as well as shifts in seasonal timing have occurred due to climate change (high confidence), with adverse socioeconomic consequences high 2 train +2310 AR6_WGII 20 16 Approximately half of the species assessed globally have shifted polewards or, on land, also to higher elevations very high 3 test +2311 AR6_WGII 20 17 Hundreds of local losses of species have been driven by increases in the magnitude of heat extremes (high confidence), as well as mass mortality events on land and in the ocean (very high confidence) and loss of kelp forests high 2 train +2312 AR6_WGII 20 18 Some losses are already irreversible, such as the first species extinctions driven by climate change medium 1 train +2313 AR6_WGII 20 19 Other impacts are approaching irreversibility such as the impacts of hydrological changes resulting from the retreat of glaciers, or the changes in some mountain (medium confidence) and Arctic ecosystems driven by permafrost thaw high 2 train +2314 AR6_WGII 20 22 Although overall agricultural productivity has increased, climate change has slowed this growth over the past 50 years globally (medium confidence), related negative impacts were mainly in mid- and low latitude regions but positive impacts occurred in some high latitude regions high 2 train +2315 AR6_WGII 20 23 Ocean warming and ocean acidification have adversely affected food production from shellfish aquaculture and fisheries in some oceanic regions high 2 train +2316 AR6_WGII 20 24 Increasing weather and climate extreme events have exposed millions of people to acute food insecurity30 and reduced water security, with the largest impacts observed in many locations and/or communities in Africa, Asia, Central and South America, Small Islands and the Arctic high 2 train +2317 AR6_WGII 20 25 Jointly, sudden losses of food production and access to food compounded by decreased diet diversity have increased malnutrition in many communities (high confidence), especially for Indigenous Peoples, small-scale food producers and low-income households (high confidence), with children, elderly people and pregnant women particularly impacted high 2 train +2318 AR6_WGII 20 26 Roughly half of the world’s population currently experience severe water scarcity for at least some part of the year due to climatic and non-climatic drivers medium 1 train +2319 AR6_WGII 22 1 Climate change impacts on health are mediated through natural and human systems, including economic and social conditions and disruptions high 2 train +2320 AR6_WGII 22 2 In all regions extreme heat events have resulted in human mortality and morbidity very high 3 train +2321 AR6_WGII 22 3 The occurrence of climate-related food-borne and water-borne diseases has increased very high 3 train +2322 AR6_WGII 22 4 The incidence of vector-borne diseases has increased from range expansion and/or increased reproduction of disease vectors high 2 train +2323 AR6_WGII 22 5 Animal and human diseases, including zoonoses, are emerging in new areas high 2 train +2324 AR6_WGII 22 8 Although diarrheal diseases have decreased globally, higher temperatures, increased rain and flooding have increased the occurrence of diarrheal diseases, including cholera (very high confidence) and other gastrointestinal infections high 2 train +2325 AR6_WGII 22 9 In assessed regions, some mental health challenges are associated with increasing temperatures (high confidence), trauma from weather and climate extreme events (very high confidence), and loss of livelihoods and culture high 2 train +2326 AR6_WGII 22 10 Increased exposure to wildfire smoke, atmospheric dust, and aeroallergens have been associated with climate-sensitive cardiovascular and respiratory distress high 2 train +2327 AR6_WGII 22 11 Health services have been disrupted by extreme events such as floods high 2 train +2328 AR6_WGII 22 13 Multiple climate and non-climate hazards impact cities, settlements and infrastructure and sometimes coincide, magnifying damage high 2 train +2329 AR6_WGII 22 14 Hot extremes including heatwaves have intensified in cities (high confidence), where they have also aggravated air pollution events (medium confidence) and limited functioning of key infrastructure high 2 train +2330 AR6_WGII 22 15 Observed impacts are concentrated amongst the economically and socially marginalized urban residents, e.g., in informal settlements high 2 train +2331 AR6_WGII 22 16 Infrastructure, including transportation, water, sanitation and energy systems have been compromised by extreme and slow-onset events, with resulting economic losses, disruptions of services and impacts to well-being high 2 train +2332 AR6_WGII 22 18 Some positive economic effects have been identified in regions that have benefited from lower energy demand as well as comparative advantages in agricultural markets and tourism high 2 train +2333 AR6_WGII 22 19 Economic damages from climate change have been detected in climate-exposed sectors, with regional effects to agriculture, forestry, fishery, energy, and tourism (high confidence), and through outdoor labour productivity high 2 train +2334 AR6_WGII 22 20 Some extreme weather events, such as tropical cyclones, have reduced economic growth in the short-term high 2 train +2335 AR6_WGII 22 21 Non-climatic factors including some patterns of settlement, and siting of infrastructure have contributed to the exposure of more assets to extreme climate hazards increasing the magnitude of the losses high 2 train +2336 AR6_WGII 22 22 Individual livelihoods have been affected through changes in agricultural productivity, impacts on human health and food security, destruction of homes and infrastructure, and loss of property and income, with adverse effects on gender and social equity high 2 train +2337 AR6_WGII 22 24 Climate and weather extremes are increasingly driving displacement in all regions (high confidence), with Small Island States disproportionately affected high 2 train +2338 AR6_WGII 22 25 Flood and drought-related acute food insecurity and malnutrition have increased in Africa (high confidence) and Central and South America high 2 train +2339 AR6_WGII 22 26 While non-climatic factors are the dominant drivers of existing intrastate violent conflicts, in some assessed regions extreme weather and climate events have had a small, adverse impact on their length, severity or frequency, but the statistical association is weak medium 1 train +2340 AR6_WGII 22 27 Through displacement and involuntary migration from extreme weather and climate events, climate change has generated and perpetuated vulnerability medium 1 train +2341 AR6_WGII 23 3 Approximately 3.3 to 3.6 billion people live in contexts that are highly vulnerable to climate change high 2 train +2342 AR6_WGII 23 4 A high proportion of species is vulnerable to climate change high 2 train +2343 AR6_WGII 23 5 Human and ecosystem vulnerability are interdependent high 2 train +2344 AR6_WGII 23 6 Current unsustainable development patterns are increasing exposure of ecosystems and people to climate hazards high 2 train +2345 AR6_WGII 23 8 Unsustainable land-use and land cover change, unsustainable use of natural resources, deforestation, loss of biodiversity, pollution, and their interactions, adversely affect the capacities of ecosystems, societies, communities and individuals to adapt to climate change high 2 train +2346 AR6_WGII 23 9 Loss of ecosystems and their services has cascading and long-term impacts on people globally, especially for Indigenous Peoples and local communities who are directly dependent on ecosystems, to meet basic needs high 2 train +2347 AR6_WGII 23 11 Globally, and even within protected areas, unsustainable use of natural resources, habitat fragmentation, and ecosystem damage by pollutants increase ecosystem vulnerability to climate change high 2 train +2348 AR6_WGII 23 13 In most protected areas, there is insufficient stewardship to contribute to reducing damage from, or increasing resilience to, climate change high 2 train +2349 AR6_WGII 23 15 Projected climate change, combined with non-climatic drivers, will cause loss and degradation of much of the world’s forests (high confidence), coral reefs and low-lying coastal wetlands very high 3 train +2350 AR6_WGII 23 16 While agricultural development contributes to food security, unsustainable agricultural expansion, driven in part by unbalanced diets32, increases ecosystem and human vulnerability and leads to competition for land and/or water resources high 2 train +2351 AR6_WGII 23 18 Global hotspots of high human vulnerability are found particularly in West-, Central- and East Africa, South Asia, Central and South America, Small Island Developing States and the Arctic high 2 train +2352 AR6_WGII 23 19 Vulnerability is higher in locations with poverty, governance challenges and limited access to basic services and resources, violent conflict and high levels of climate-sensitive livelihoods (e.g., smallholder farmers, pastoralists, fishing communities) high 2 train +2353 AR6_WGII 23 20 Between 2010–2020, human mortality from floods, droughts and storms was 15 times higher in highly vulnerable regions, compared to regions with very low vulnerability high 2 train +2354 AR6_WGII 23 21 Vulnerability at different spatial levels is exacerbated by inequity and marginalization linked to gender, ethnicity, low income or combinations thereof (high confidence), especially for many Indigenous Peoples and local communities high 2 train +2355 AR6_WGII 23 22 Present development challenges causing high vulnerability are influenced by historical and ongoing patterns of inequity such as colonialism, especially for many Indigenous Peoples and local communities high 2 train +2356 AR6_WGII 24 1 In rural areas vulnerability will be heightened by compounding processes including high emigration, reduced habitability and high reliance on climate-sensitive livelihoods high 2 train +2357 AR6_WGII 24 2 Key infrastructure systems including sanitation, water, health, transport, communications and energy will be increasingly vulnerable if design standards do not account for changing climate conditions high 2 train +2358 AR6_WGII 24 3 Vulnerability will also rapidly rise in low-lying Small Island Developing States and atolls in the context of sea level rise and in some mountain regions, already characterised by high vulnerability due to high dependence on climate-sensitive livelihoods, rising population displacement, the accelerating loss of ecosystem services and limited adaptive capacities high 2 train +2359 AR6_WGII 24 4 Future exposure to climatic hazards is also increasing globally due to socioeconomic development trends including migration, growing inequality and urbanization high 2 train +2360 AR6_WGII 24 6 The level of risk will depend on concurrent near- term trends in vulnerability, exposure, level of socioeconomic development and adaptation high 2 train +2361 AR6_WGII 24 7 Near-term actions that limit global warming to close to 1.5°C would substantially reduce projected losses and damages related to climate change in human systems and ecosystems, compared to higher warming levels, but cannot eliminate them all very high 3 train +2362 AR6_WGII 24 9 Near-term risks for biodiversity loss are moderate to high in forest ecosystems (medium confidence), kelp and seagrass ecosystems (high to very high confidence), and high to very high in Arctic sea-ice and terrestrial ecosystems (high confidence) and warm-water coral reefs very high 3 train +2363 AR6_WGII 24 10 Continued and accelerating sea level rise will encroach on coastal settlements and infrastructure (high confidence) and commit low-lying coastal ecosystems to submergence and loss medium 1 train +2364 AR6_WGII 24 11 If trends in urbanisation in exposed areas continue, this will exacerbate the impacts, with more challenges where energy, water and other services are constrained medium 1 train +2365 AR6_WGII 24 12 The number of people at risk from climate change and associated loss of biodiversity will progressively increase medium 1 train +2366 AR6_WGII 24 13 Violent conflict and, separately, migration patterns, in the near-term will be driven by socioeconomic conditions and governance more than by climate change medium 1 train +2367 AR6_WGII 24 15 Regional differences exist, and risks are highest where species and people exist close to their upper thermal limits, along coastlines, in close association with ice or seasonal rivers high 2 train +2368 AR6_WGII 24 16 Risks are also high where multiple non-climate drivers persist or where vulnerability is otherwise elevated high 2 train +2369 AR6_WGII 24 17 Many of these risks are unavoidable in the near-term, irrespective of emissions scenario high 2 train +2370 AR6_WGII 24 18 Several risks can be moderated with adaptation high 2 train +2371 AR6_WGII 24 20 Between 1.2°C and 4.5°C global warming level very high risks emerge in all five RFCs compared to just two RFCs in AR5 high 2 train +2372 AR6_WGII 24 21 Two of these transitions from high to very high risk are associated with near-term warming: risks to unique and threatened systems at a median value of 1.5 [1.2 to 2.0] °C (high confidence) and risks associated with extreme weather events at a median value of 2.0 [1.8 to 2.5] °C medium 1 train +2373 AR6_WGII 24 22 Some key risks contributing to the RFCs are projected to lead to widespread, pervasive, and potentially irreversible impacts at global warming levels of 1.5–2°C if exposure and vulnerability are high and adaptation is low medium 1 train +2374 AR6_WGII 24 23 Near-term actions that limit global warming to close to 1.5°C would substantially reduce projected losses and damages related to climate change in human systems and ecosystems, compared to higher warming levels, but cannot eliminate them all very high 3 train +2375 AR6_WGII 25 2 For 127 identified key risks, assessed mid- and long-term impacts are up to multiple times higher than currently observed high 2 train +2376 AR6_WGII 25 3 The magnitude and rate of climate change and associated risks depend strongly on near-term mitigation and adaptation actions, and projected adverse impacts and related losses and damages escalate with every increment of global warming very high 3 train +2377 AR6_WGII 25 6 In ocean and coastal ecosystems, risk of biodiversity loss ranges between moderate and very high by 1.5°C global warming level and is moderate to very high by 2°C but with more ecosystems at high and very high risk high 2 train +2378 AR6_WGII 25 7 Very high extinction risk for endemic species in biodiversity hotspots is projected to at least double from 2% between 1.5°C and 2°C global warming levels and to increase at least tenfold if warming rises from 1.5°C to 3°C medium 1 train +2379 AR6_WGII 25 9 At approximately 2°C global warming, snowmelt water availability for irrigation is projected to decline in some snowmelt dependent river basins by up to 20%, and global glacier mass loss of 18 ± 13% is projected to diminish water availability for agriculture, hydropower, and human settlements in the mid- to long-term, with these changes projected to double with 4°C global warming medium 1 train +2380 AR6_WGII 25 10 In Small Islands, groundwater availability is threatened by climate change high 2 train +2381 AR6_WGII 25 11 Changes to streamflow magnitude, timing and associated extremes are projected to adversely impact freshwater ecosystems in many watersheds by the mid- to long-term across all assessed scenarios medium 1 train +2382 AR6_WGII 25 12 Projected increases in direct flood damages are higher by 1.4 to 2 times at 2°C and 2.5 to 3.9 times at 3°C compared to 1.5°C global warming without adaptation medium 1 train +2383 AR6_WGII 25 13 At global warming of 4°C, approximately 10% of the global land area is projected to face increases in both extreme high and low river flows in the same location, with implications for planning for all water use sectors medium 1 train +2384 AR6_WGII 25 14 Challenges for water management will be exacerbated in the near, mid and long term, depending on the magnitude, rate and regional details of future climate change and will be particularly challenging for regions with constrained resources for water management high 2 train +2385 AR6_WGII 25 16 Increases in frequency, intensity and severity of droughts, floods and heatwaves, and continued sea level rise will increase risks to food security (high confidence) in vulnerable regions from moderate to high between 1.5°C and 2°C global warming level, with no or low levels of adaptation medium 1 train +2386 AR6_WGII 25 17 At 2°C or higher global warming level in the mid-term, food security risks due to climate change will be more severe, leading to malnutrition and micro-nutrient deficiencies, concentrated in Sub-Saharan Africa, South Asia, Central and South America and Small Islands high 2 train +2387 AR6_WGII 25 18 Global warming will progressively weaken soil health and ecosystem services such as pollination, increase pressure from pests and diseases, and reduce marine animal biomass, undermining food productivity in many regions on land and in the ocean medium 1 train +2388 AR6_WGII 25 19 At 3°C or higher global warming level in the long term, areas exposed to climate-related hazards will expand substantially compared with 2°C or lower global warming level (high confidence), exacerbating regional disparity in food security risks high 2 train +2389 AR6_WGII 26 1 Globally, population exposure to heatwaves will continue to increase with additional warming, with strong geographical differences in heat-related mortality without additional adaptation very high 3 train +2390 AR6_WGII 26 2 Climate-sensitive food-borne, water-borne, and vector-borne disease risks are projected to increase under all levels of warming without additional adaptation high 2 train +2391 AR6_WGII 26 3 In particular, dengue risk will increase with longer seasons and a wider geographic distribution in Asia, Europe, Central and South America and sub-Saharan Africa, potentially putting additional billions of people at risk by the end of the century high 2 train +2392 AR6_WGII 26 4 Mental health challenges, including anxiety and stress, are expected to increase under further global warming in all assessed regions, particularly for children, adolescents, elderly, and those with underlying health conditions very high 3 test +2393 AR6_WGII 26 6 Globally, population change in low-lying cities and settlements will lead to approximately a billion people projected to be at risk from coastal-specific climate hazards in the mid-term under all scenarios, including in Small Islands high 2 train +2394 AR6_WGII 26 7 The population potentially exposed to a 100-year coastal flood is projected to increase by about 20% if global mean sea level rises by 0.15 m relative to 2020 levels; this exposed population doubles at a 0.75 m rise in mean sea level and triples at 1.4 m without population change and additional adaptation medium 1 train +2395 AR6_WGII 26 8 Sea level rise poses an existential threat for some Small Islands and some low-lying coasts medium 1 train +2396 AR6_WGII 26 9 By 2100 the value of global assets within the future 1-in-100 year coastal floodplains is projected to be between US$7.9 and US$12.7 trillion (2011 value) under RCP4.5, rising to between US$8.8 and US$14.2 trillion under RCP8.5 medium 1 train +2397 AR6_WGII 26 10 Costs for maintenance and reconstruction of urban infrastructure, including building, transportation, and energy will increase with global warming level (medium confidence), the associated functional disruptions are projected to be substantial particularly for cities, settlements and infrastructure located on permafrost in cold regions and on coasts high 2 train +2398 AR6_WGII 26 12 The existence of higher estimates than assessed in AR5 indicates that global aggregate economic impacts could be higher than previous estimates (low confidence).36 Significant regional variation in aggregate economic damages from climate change is projected (high confidence) with estimated economic damages per capita for developing countries often higher as a fraction of income high 2 train +2399 AR6_WGII 26 13 Economic damages, including both those represented and those not represented in economic markets, are projected to be lower at 1.5°C than at 3°C or higher global warming levels high 2 train +2400 AR6_WGII 26 15 At progressive levels of warming, involuntary migration from regions with high exposure and low adaptive capacity would occur medium 1 train +2401 AR6_WGII 26 16 Compared to other socioeconomic factors the influence of climate on conflict is assessed as relatively weak high 2 train +2402 AR6_WGII 26 17 Along long-term socioeconomic pathways that reduce non-climatic drivers, risk of violent conflict would decline medium 1 train +2403 AR6_WGII 26 18 At higher global warming levels, impacts of weather and climate extremes, particularly drought, by increasing vulnerability will increasingly affect violent intrastate conflict medium 1 train +2404 AR6_WGII 26 20 There is evidence that some regions could benefit from low levels of warming high 2 train +2405 AR6_WGII 29 4 Multiple risks interact, generating new sources of vulnerability to climate hazards, and compounding overall risk high 2 train +2406 AR6_WGII 29 5 Increasing concurrence of heat and drought events are causing crop production losses and tree mortality high 2 train +2407 AR6_WGII 29 6 Above 1.5°C global warming increasing concurrent climate extremes will increase risk of simultaneous crop losses of maize in major food-producing regions, with this risk increasing further with higher global warming levels medium 1 train +2408 AR6_WGII 29 7 Future sea level rise combined with storm surge and heavy rainfall will increase compound flood risks high 2 train +2409 AR6_WGII 29 8 Risks to health and food production will be made more severe from the interaction of sudden food production losses from heat and drought, exacerbated by heat-induced labour productivity losses high 2 train +2410 AR6_WGII 29 9 These interacting impacts will increase food prices, reduce household incomes, and lead to health risks of malnutrition and climate-related mortality with no or low levels of adaptation, especially in tropical regions high 2 train +2411 AR6_WGII 29 10 Risks to food safety from climate change will further compound the risks to health by increasing food contamination of crops from mycotoxins and contamination of seafood from harmful algal blooms, mycotoxins, and chemical contaminants high 2 train +2412 AR6_WGII 29 12 These hazards and cascading risks also trigger tipping points in sensitive ecosystems and in significantly and rapidly changing social-ecological systems impacted by ice melt, permafrost thaw and changing hydrology in polar regions high 2 train +2413 AR6_WGII 30 1 In Amazonia, and in some mountain regions, cascading impacts from climatic (e.g., heat) and non-climatic stressors (e.g., land use change) will result in irreversible and severe losses of ecosystem services and biodiversity at 2°C global warming level and beyond medium 1 train +2414 AR6_WGII 30 2 Unavoidable sea level rise will bring cascading and compounding impacts resulting in losses of coastal ecosystems and ecosystem services, groundwater salinisation, flooding and damages to coastal infrastructure that cascade into risks to livelihoods, settlements, health, well-being, food and water security, and cultural values in the near to long-term high 2 train +2415 AR6_WGII 30 5 Climate change causes the redistribution of marine fish stocks, increasing risk of transboundary management conflicts among fisheries users, and negatively affecting equitable distribution of food provisioning services as fish stocks shift from lower to higher latitude regions, thereby increasing the need for climate-informed transboundary management and cooperation high 2 train +2416 AR6_WGII 30 6 Precipitation and water availability changes increases the risk of planned infrastructure projects, such as hydropower in some regions, having reduced productivity for food and energy sectors including across countries that share river basins medium 1 train +2417 AR6_WGII 30 8 Deployment of afforestation of naturally unforested land, or poorly implemented bioenergy, with or without carbon capture and storage, can compound climate-related risks to biodiversity, water and food security, and livelihoods, especially if implemented at large scales, especially in regions with insecure land tenure high 2 train +2418 AR6_WGII 30 10 Solar radiation modification approaches have potential to offset warming and ameliorate some climate hazards, but substantial residual climate change or overcompensating change would occur at regional scales and seasonal timescales high 2 train +2419 AR6_WGII 30 12 Solar radiation modification would not stop atmospheric CO 2 concentrations from increasing or reduce resulting ocean acidification under continued anthropogenic emissions high 2 train +2420 AR6_WGII 30 15 Depending on the mag- nitude and duration of overshoot, some impacts will cause release of additional greenhouse gases (medium confidence) and some will be irreversible, even if global warming is reduced high 2 train +2421 AR6_WGII 30 17 Additional warming, e.g., above 1.5°C during an overshoot period this century, will result in irreversible impacts on certain ecosystems with low resilience, such as polar, mountain, and coastal ecosystems, impacted by ice-sheet, glacier melt, or by accelerating and higher committed sea level rise (high confidence).38 Risks to human systems will increase, including those to infrastructure, low-lying coastal settlements, some ecosystem-based adaptation measures, and associated livelihoods (high confidence), cultural and spiritual values medium 1 train +2422 AR6_WGII 30 18 Projected impacts are less severe with shorter duration and lower levels of overshoot medium 1 train +2423 AR6_WGII 31 1 In high-carbon ecosystems (currently storing 3,000 to 4,000 GtC) such impacts are already observed and are projected to increase with every additional increment of global warming, such as increased wildfires, mass mortality of trees, drying of peatlands, and thawing of permafrost, weakening natural land carbon sinks and increasing releases of greenhouse gases medium 1 test +2424 AR6_WGII 31 2 The resulting contribution to a potential amplification of global warming indicates that a return to a given global warming level or below would be more challenging medium 1 train +2425 AR6_WGII 31 9 However, adaptation progress is unevenly distributed with observed adaptation gaps40 high 2 train +2426 AR6_WGII 31 10 Many initiatives prioritize immediate and near-term climate risk reduction which reduces the opportunity for transformational adaptation high 2 train +2427 AR6_WGII 31 12 Growing public and political awareness of climate impacts and risks has resulted in at least 170 countries and many cities including adaptation in their climate policies and planning processes high 2 train +2428 AR6_WGII 31 13 Decision support tools and climate services are increasingly being used very high 3 train +2429 AR6_WGII 31 14 Pilot projects and local experiments are being implemented in different sectors high 2 train +2430 AR6_WGII 31 15 Adaptation can generate multiple additional benefits such as improving agricultural productivity, innovation, health and well-being, food security, livelihood, and biodiversity conservation as well as reduction of risks and damages very high 3 test +2431 AR6_WGII 31 17 Most observed adaptation is fragmented, small in scale, incremental, sector-specific, designed to respond to current impacts or near-term risks, and focused more on planning rather than implementation high 2 train +2432 AR6_WGII 31 18 Observed adaptation is unequally distributed across regions (high confidence), and gaps are partially driven by widening disparities between the estimated costs of adaptation and documented finance allocated to adaptation high 2 train +2433 AR6_WGII 31 19 The largest adaptation gaps exist among lower income population groups high 2 train +2434 AR6_WGII 31 20 At current rates of adaptation planning and implementation the adaptation gap will continue to grow high 2 train +2435 AR6_WGII 31 21 As adaptation options often have long implementation times, long-term planning and accelerated implementation, particularly in the next decade, is important to close adaptation gaps, recognising that constraints remain for some regions high 2 train +2436 AR6_WGII 32 6 The feasibility of implementing adaptation options in the near-term differs across sectors and regions very high 3 train +2437 AR6_WGII 32 7 The effec- tiveness of adaptation to reduce climate risk is documented for specific contexts, sectors and regions (high confidence) and will decrease with increasing warming high 2 train +2438 AR6_WGII 32 8 Integrated, multi-sectoral solutions that address social in- equities, differentiate responses based on climate risk and cut across systems, increase the feasibility and effectiveness of adaptation in multiple sectors high 2 train +2439 AR6_WGII 32 10 For inland flooding, combinations of non-structural measures like early warning systems and structural measures like levees have reduced loss of lives medium 1 train +2440 AR6_WGII 32 11 Enhancing natural water retention such as by restoring wetlands and rivers, land use planning such as no build zones or upstream forest management, can further reduce flood risk medium 1 train +2441 AR6_WGII 32 12 On-farm water management, water storage, soil moisture conservation and irrigation are some of the most common adaptation responses and provide economic, institutional or ecological benefits and reduce vulnerability high 2 train +2442 AR6_WGII 32 13 Irrigation is effective in reducing drought risk and climate impacts in many regions and has several livelihood benefits, but needs appropriate management to avoid potential adverse outcomes, which can include accelerated depletion of groundwater and other water sources and increased soil salinization medium 1 train +2443 AR6_WGII 32 14 Large scale irrigation can also alter local to regional temperature and precipitation patterns (high confidence), including both alleviating and exacerbating temperature extremes medium 1 train +2444 AR6_WGII 32 15 The effectiveness of most water-related adaptation options to reduce projected risks declines with increasing warming high 2 train +2445 AR6_WGII 32 17 Effective options include cultivar improvements, agroforestry, community-based adaptation, farm and landscape diversification, and urban agriculture high 2 train +2446 AR6_WGII 32 19 Agroecological principles and practices, ecosystem-based management in fisheries and aquaculture, and other approaches that work with natural processes support food security, nutrition, health and well-being, livelihoods and biodiversity, sustainability and ecosystem services high 2 train +2447 AR6_WGII 32 20 These services include pest control, pollination, buffering of temperature extremes, and carbon sequestration and storage high 2 train +2448 AR6_WGII 32 21 Trade-offs and barriers associated with such approaches include costs of establishment, access to inputs and viable markets, new knowledge and management (high confidence) and their potential effectiveness varies by socioeconomic context, ecosystem zone, species combinations and institutional support medium 1 train +2449 AR6_WGII 32 22 Integrated, multi-sectoral solutions that address social inequities and differentiate responses based on climate risk and local situation will enhance food security and nutrition high 2 train +2450 AR6_WGII 32 23 Adaptation strategies which reduce food loss and waste or support balanced diets33 (as described in the IPCC Special Report on Climate Change and Land) contribute to nutrition, health, biodiversity and other environmental benefits high 2 train +2451 AR6_WGII 35 1 The resilience of species, biological communities and ecosystem processes increases with size of natural area, by restoration of degraded areas and by reducing non-climatic stressors high 2 train +2452 AR6_WGII 35 2 To be effective, conservation and restoration actions will increasingly need to be responsive, as appropriate, to ongoing changes at various scales, and plan for future changes in ecosystem structure, community composition and species’ distributions, especially as 1.5°C global warming is approached and even more so if it is exceeded high 2 train +2453 AR6_WGII 35 3 Adaptation options, where circumstances allow, include facilitating the movement of species to new ecologically appropriate locations, particularly through increasing connectivity between conserved or protected areas, targeted intensive management for vulnerable species and protecting refugial areas where species can survive locally medium 1 train +2454 AR6_WGII 35 5 Ecosystem-based Adaptation is vulnerable to climate change impacts, with effectiveness declining with increasing global warming high 2 train +2455 AR6_WGII 35 6 Urban greening using trees and other vegetation can provide local cooling very high 3 train +2456 AR6_WGII 35 7 Natural river systems, wetlands and upstream forest ecosystems reduce flood risk by storing water and slowing water flow, in most circumstances high 2 train +2457 AR6_WGII 35 8 Coastal wetlands protect against coastal erosion and flooding associated with storms and sea level rise where sufficient space and adequate habitats are available until rates of sea level rise exceeds natural adaptive capacity to build sediment very high 3 train +2458 AR6_WGII 35 10 The urgent provision of basic services, infrastructure, livelihood diversification and employment, strengthening of local and regional food systems and community-based adaptation enhance lives and livelihoods, particularly of low-income and marginalised groups high 2 train +2459 AR6_WGII 35 11 Inclusive, integrated and long-term planning at local, municipal, sub-national and national scales, together with effective regulation and monitoring systems and financial and technological resources and capabilities foster urban and rural system transition high 2 train +2460 AR6_WGII 35 14 Globally, more financing is directed at physical infrastructure than natural and social infrastructure medium 1 train +2461 AR6_WGII 35 15 Ecosystem-based adaptation (e.g., urban agriculture and forestry, river restoration) has increasingly been applied in urban areas high 2 train +2462 AR6_WGII 35 16 Combined ecosystem-based and structural adaptation responses are being developed, and there is growing evidence of their potential to reduce adaptation costs and contribute to flood control, sanitation, water resources management, landslide prevention and coastal protection medium 1 train +2463 AR6_WGII 36 1 Such adaptation challenges would occur much earlier under high rates of sea level rise, in particular if low-likelihood, high impact outcomes associated with collapsing ice sheets occur high 2 train +2464 AR6_WGII 36 2 Responses to ongoing sea level rise and land subsidence in low-lying coastal cities and settlements and small islands include protection, accommodation, advance and planned relocation high 2 train +2465 AR6_WGII 36 3 These responses are more effective if combined and/or sequenced, planned well ahead, aligned with sociocultural values and development priorities, and underpinned by inclusive community engagement processes high 2 train +2466 AR6_WGII 36 9 Energy generation diversification, including with renewable energy resources and generation that can be decentralised depending on context (e.g., wind, solar, small scale hydroelectric) and demand side management (e.g., storage, and energy efficiency improvements) can reduce vulnerabilities to climate change, especially in rural populations high 2 train +2467 AR6_WGII 36 10 Adaptations for hydropower and thermo-electric power generation are effective in most regions up to 1.5°C to 2°C, with decreasing effectiveness at higher levels of warming medium 1 train +2468 AR6_WGII 36 11 Climate responsive energy markets, updated design standards on energy assets according to current and projected climate change, smart-grid technologies, robust transmission systems and improved capacity to respond to supply deficits have high feasibility in the medium- to long-term, with mitigation co-benefits very high 3 train +2469 AR6_WGII 36 14 Heat Health Action Plans that include early warning and response systems are effective adaptation options for extreme heat high 2 train +2470 AR6_WGII 36 15 Effective adaptation options for water-borne and food-borne diseases include improving access to potable water, reducing exposure of water and sanitation systems to flooding and extreme weather events, and improved early warning systems very high 3 train +2471 AR6_WGII 36 16 For vector-borne diseases, effective adaptation options include surveillance, early warning systems, and vaccine development very high 3 train +2472 AR6_WGII 36 17 Effective adaptation options for reducing mental health risks under climate change include improving surveillance, access to mental health care, and monitoring of psychosocial impacts from extreme weather events high 2 train +2473 AR6_WGII 36 18 Health and well-being would benefit from integrated adaptation approaches that mainstream health into food, livelihoods, social protection, infrastructure, water and sanitation policies requiring collaboration and coordination at all scales of governance very high 3 test +2474 AR6_WGII 36 20 This improves the degree of choice under which migration decisions are made, ensuring safe and orderly movements of people within and between countries high 2 train +2475 AR6_WGII 36 21 Some development reduces underlying vulnerabilities associated with conflict, and adaptation contributes by reducing the impacts of climate change on climate sensitive drivers of conflict high 2 train +2476 AR6_WGII 36 22 Risks to peace are reduced, for example, by supporting people in climate-sensitive economic activities (medium confidence) and advancing women’s empowerment high 2 train +2477 AR6_WGII 37 1 For example, climate services that are inclusive of different users and providers can improve agricultural practices, inform better water use and efficiency, and enable resilient infrastructure planning high 2 train +2478 AR6_WGII 37 3 Hard limits to adaptation have been reached in some ecosystems high 2 train +2479 AR6_WGII 37 4 With increasing global warming, losses and damages will increase and additional human and natural systems will reach adaptation limits high 2 train +2480 AR6_WGII 37 6 For example, individuals and households in low-lying coastal areas in Australasia and Small Islands and smallholder farmers in Central and South America, Africa, Europe and Asia have reached soft limits medium 1 train +2481 AR6_WGII 37 7 Inequity and poverty also constrain adaptation, leading to soft limits and resulting in disproportionate exposure and impacts for most vulnerable groups high 2 train +2482 AR6_WGII 37 8 Lack of climate literacy46 at all levels and limited availability of information and data pose further constraints to adaptation planning and implementation medium 1 train +2483 AR6_WGII 37 10 Although global tracked climate finance has shown an upward trend since AR5, current global financial flows for adaptation, including from public and private finance sources, are insufficient for and constrain implementation of adaptation options especially in developing countries high 2 train +2484 AR6_WGII 37 11 The overwhelming majority of global tracked climate finance was targeted to mitigation while a small proportion was targeted to adaptation very high 3 train +2485 AR6_WGII 37 12 Adaptation finance has come predominantly from public sources very high 3 train +2486 AR6_WGII 37 13 Adverse climate impacts can reduce the availability of financial resources by incurring losses and damages and through impeding national economic growth, thereby further increasing financial constraints for adaptation, particularly for developing and least developed countries medium 1 train +2487 AR6_WGII 37 15 Ecosystems already reaching or surpassing hard adaptation limits include some warm- water coral reefs, some coastal wetlands, some rainforests, and some polar and mountain ecosystems high 2 train +2488 AR6_WGII 37 16 Above 1.5°C global warming level, some Ecosystem-based Adaptation measures will lose their effectiveness in providing benefits to people as these ecosystems will reach hard adaptation limits high 2 train +2489 AR6_WGII 37 18 Above 1.5°C global warming level, limited freshwater resources pose potential hard limits for Small Islands and for regions dependent on glacier and snow-melt medium 1 train +2490 AR6_WGII 37 19 By 2°C global warming level, soft limits are projected for multiple staple crops in many growing areas, particularly in tropical regions high 2 train +2491 AR6_WGII 37 20 By 3°C global warming level, soft limits are projected for some water management measures for many regions, with hard limits projected for parts of Europe medium 1 train +2492 AR6_WGII 37 21 Transitioning from incremental to transformational adaptation can help overcome soft adaptation limits high 2 train +2493 AR6_WGII 38 4 The implementation of these maladaptive actions can result in infrastructure and institutions that are inflexible and/or expensive to change high 2 train +2494 AR6_WGII 38 5 For example, seawalls effectively reduce impacts to people and assets in the short-term but can also result in lock-ins and increase exposure to climate risks in the long-term unless they are integrated into a long-term adaptive plan high 2 train +2495 AR6_WGII 38 6 Adaptation integrated with development reduces lock-ins and creates opportunities (e.g., infrastructure upgrading) medium 1 train +2496 AR6_WGII 38 15 Maladaptation is also minimized by planning that accounts for the time it takes to adapt (high confidence), the uncertainty about the rate and magnitude of climate risk (medium confidence) and a wide range of potentially adverse consequences of adaptation actions high 2 train +2497 AR6_WGII 38 19 Implementing actions can require large upfront investments of human, financial and technological resources (high confidence), whilst some benefits could only become visible in the next decade or beyond medium 1 train +2498 AR6_WGII 38 20 Accelerating commitment and follow-through is promoted by rising public awareness, building business cases for adaptation, accountability and transparency mechanisms, monitoring and evaluation of adaptation progress, social movements, and climate-related litigation in some regions medium 1 train +2499 AR6_WGII 39 1 Sustained adaptation actions are strengthened by mainstreaming adaptation into institutional budget and policy planning cycles, statutory planning, monitoring and evaluation frameworks and into recovery efforts from disaster events high 2 train +2500 AR6_WGII 39 2 Instruments that incorporate adaptation such as policy and legal frameworks, behavioural incentives, and economic instruments that address market failures, such as climate risk disclosure, inclusive and deliberative processes strengthen adaptation actions by public and private actors medium 1 train +2501 AR6_WGII 39 4 A wide range of top-down, bottom-up and co-produced processes and sources can deepen climate knowledge and sharing, including capacity building at all scales, educational and information programmes, using the arts, participatory modelling and climate services, Indigenous knowledge and local knowledge and citizen science high 2 train +2502 AR6_WGII 39 5 These measures can facilitate awareness, heighten risk perception and influence behaviours high 2 train +2503 AR6_WGII 39 7 Building capacity and removing some barriers to accessing finance is fundamental to accelerate adaptation, especially for vulnerable groups, regions and sectors high 2 train +2504 AR6_WGII 39 9 Public finance is an important enabler of adaptation high 2 train +2505 AR6_WGII 39 10 Public mechanisms and finance can leverage private sector finance for adaptation by addressing real and perceived regulatory, cost and market barriers, for example via public-private partnerships high 2 train +2506 AR6_WGII 39 11 Financial and technological resources enable effective and ongoing implementation of adaptation, especially when supported by institutions with a strong understanding of adaptation needs and capacity high 2 train +2507 AR6_WGII 39 13 M&E implementation is currently limited high 2 train +2508 AR6_WGII 39 14 Although most of the monitoring of adaptation is focused towards planning and implementation, the monitoring of outcomes is critical for tracking the effectiveness and progress of adaptation high 2 train +2509 AR6_WGII 39 16 M&E systems are most effective when supported by capacities and resources and embedded in enabling governance systems high 2 train +2510 AR6_WGII 39 18 Vulnerabilities and climate risks are often reduced through carefully designed and implemented laws, policies, processes, and interventions that address context specific inequities such as based on gender, ethnicity, disability, age, location and income high 2 train +2511 AR6_WGII 39 19 These approaches, which include multi-stakeholder co-learning platforms, transboundary collaborations, community-based adaptation and participatory scenario planning, focus on capacity-building, and meaningful participation of the most vulnerable and marginalised groups, and their access to key resources to adapt high 2 train +2512 AR6_WGII 40 7 This in turn undermines efforts to achieve sustainable development, particularly for vulnerable and marginalized communities very high 3 train +2513 AR6_WGII 40 9 This is especially challenging in localities with persistent development gaps and limited resources high 2 train +2514 AR6_WGII 40 11 Integrated and inclusive system-oriented solutions based on equity and social and climate justice reduce risks and enable climate resilient development high 2 train +2515 AR6_WGII 40 13 Climate resilient devel- opment is facilitated by international cooperation and by governments at all levels working with communities, civil society, educational bodies, scientific and other institutions, media, investors and businesses; and by developing partner - ships with traditionally marginalised groups, including women, youth, Indigenous Peoples, local communities and ethnic minorities high 2 train +2516 AR6_WGII 40 16 These practices build on diverse knowledges about climate risk and chosen development pathways account for local, regional and global climate impacts, risks, barriers and opportunities high 2 train +2517 AR6_WGII 40 17 Structural vulnerabilities to climate change can be reduced through carefully designed and implemented legal, policy, and process interventions from the local to global that address inequities based on gender, ethnicity, disability, age, location and income very high 3 train +2518 AR6_WGII 40 18 This includes rights-based approaches that focus on capacity-building, meaningful participation of the most vulnerable groups, and their access to key resources, including financing, to reduce risk and adapt high 2 test +2519 AR6_WGII 40 19 Evidence shows that climate resilient development processes link scientific, Indigenous, local, practitioner and other forms of knowledge, and are more effective and sustainable because they are locally appropriate and lead to more legitimate, relevant and effective actions high 2 train +2520 AR6_WGII 42 1 Planning processes and decision analysis tools can help identify ‘low regrets’ options47 that enable mitigation and adaptation in the face of change, complexity, deep uncertainty and divergent views medium 1 test +2521 AR6_WGII 42 3 Inclusive processes strengthen the ability of governments and other stakeholders to jointly consider factors such as the rate and magnitude of change and uncertainties, associated impacts, and timescales of different climate resilient development pathways given past development choices leading to past emissions and scenarios of future global warming high 2 train +2522 AR6_WGII 42 5 The quality and outcome of these interactions helps determine whether development pathways shift towards or away from climate resilient development medium 1 train +2523 AR6_WGII 42 13 However, the global trend of urbanisation also offers a critical opportunity in the near-term, to advance climate resilient development high 2 train +2524 AR6_WGII 42 15 Equitable outcomes contributes to multiple benefits for health and well-being and ecosystem services, including for Indigenous Peoples, marginalised and vulnerable communi- ties high 2 train +2525 AR6_WGII 42 16 Climate resilient development in urban areas also supports adaptive capacity in more rural places through maintaining peri-urban supply chains of goods and services and financial flows medium 1 test +2526 AR6_WGII 42 17 Coastal cities and settlements play an especially important role in advancing climate resilient development high 2 train +2527 AR6_WGII 43 5 Dominant models of energy intensive and market-led urbanisation, insufficient and misaligned finance and a predominant focus on grey infrastructure in the absence of integration with ecological and social approaches, risks missing opportunities for adaptation and locking in maladaptation high 2 train +2528 AR6_WGII 43 6 Poor land use planning and siloed approaches to health, ecological and social planning also exacerbates, vulnerability in already marginalised communities medium 1 train +2529 AR6_WGII 43 7 Urban climate resilient development is observed to be more effective if it is responsive to regional and local land use development and adaptation gaps, and addresses the underlying drivers of vulnerability high 2 train +2530 AR6_WGII 43 8 The greatest gains in well-being can be achieved by prioritizing finance to reduce climate risk for low-income and marginalized residents including people living in informal settlements high 2 train +2531 AR6_WGII 43 14 Recent analyses, drawing on a range of lines of evidence, suggest that maintaining the resilience of biodiversity and ecosystem services at a global scale depends on effective and equitable conservation of approximately 30% to 50% of Earth’s land, freshwater and ocean areas, including currently near-natural ecosystems high 2 train +2532 AR6_WGII 43 17 Degradation and loss of ecosystems is also a cause of greenhouse gas emissions and is at increasing risk of being exacerbated by climate change impacts, including droughts and wildfire high 2 train +2533 AR6_WGII 43 18 Climate resilient development avoids adaptation and mitigation measures that damage ecosystems high 2 train +2534 AR6_WGII 43 19 Documented examples of adverse impacts of land-based measures intended as mitigation, when poorly implemented, include afforestation of grasslands, savannas and peatlands, and risks from bioenergy crops at large scale to water supply, food security and biodiversity high 2 train +2535 AR6_WGII 44 1 Consequences of current and future global warming for climate resilient development include reduced effectiveness of Ecosystem-based Adaptation and approaches to climate change mitigation based on ecosystems and amplifying feedbacks to the climate system high 2 train +2536 AR6_WGII 44 3 Past and current development trends (past emissions, development and climate change) have not advanced global climate resilient development very high 3 train +2537 AR6_WGII 44 4 Societal choices and actions implemented in the next decade determine the extent to which medium- and long-term pathways will deliver higher or lower climate resilient development high 2 train +2538 AR6_WGII 44 5 Importantly climate resilient development prospects are increasingly limited if current greenhouse gas emissions do not rapidly decline, es- pecially if 1.5°C global warming is exceeded in the near-term high 2 train +2539 AR6_WGII 44 6 These prospects are constrained by past development, emissions and climate change, and enabled by inclusive governance, adequate and appropriate human and technological resources, information, capacities and finance high 2 train +2540 AR6_WGII 44 8 The prospects for climate resilient development will be further limited if global warming levels exceeds 1.5°C (high confidence) and not be possible in some regions and sub-regions if the global warming level exceeds 2°C medium 1 train +2541 AR6_WGII 44 9 Climate resilient development is most constrained in regions/subregions in which climate impacts and risks are already advanced, including low-lying coastal cities and settlements, small islands, deserts, mountains and polar regions high 2 train +2542 AR6_WGII 44 10 Regions and subregions with high levels of poverty, water, food and energy insecurity, vulnerable urban environments, degraded ecosystems and rural environments, and/or few enabling conditions, face many non-climate challenges that inhibit climate resilient development which are further exacerbated by climate change high 2 train +2543 AR6_WGII 54 2 Feasible, integrated mitigation and adaptation solutions can be tailored to specific locations and monitored for their effectiveness while avoiding conflict with sustainable development objectives and managing risks and trade- offs high 2 train +2544 AR6_WGII 56 1 Ecosystems and biodiversity TS.B.1 Climate change has altered marine, terrestrial and fresh - water ecosystems all around the world very high 3 train +2545 AR6_WGII 56 2 Effects were experienced earlier and are more widespread with more far-reaching consequences than anticipated medium 1 train +2546 AR6_WGII 56 3 Biological responses, including changes in physi - ology, growth, abundance, geographic placement and shifting seasonal timing, are often not sufficient to cope with recent climate change very high 3 train +2547 AR6_WGII 56 4 Climate change has caused local species losses, increases in disease (high confidence) and mass mortality events of plants and animals (very high confidence), resulting in the first climate-driven extinctions (medium confidence), ecosystem restructuring, increases in areas burned by wildfire (high confidence) and declines in key ecosystem services high 2 train +2548 AR6_WGII 56 5 Climate- driven impacts on ecosystems have caused measurable eco - nomic and livelihood losses and altered cultural practices and recreational activities around the world high 2 train +2549 AR6_WGII 56 8 Consistent with expectations, species in all ecosystems have shifted their geographic ranges and altered the timing of seasonal events very high 3 train +2550 AR6_WGII 56 9 Among thousands of species spread across terrestrial, freshwater and marine systems, half to two-thirds have shifted their ranges to higher latitudes (very high confidence), and approximately two-thirds have shifted towards earlier spring life events very high 3 train +2551 AR6_WGII 56 10 The move of diseases and their vectors has brought new diseases into the high Arctic and at higher elevations in mountain regions to which local wildlife and humans are not resistant high 2 train +2552 AR6_WGII 56 11 These processes have led to emerging hybridisation, competition, temporal or spatial mismatches in predator–prey, insect– plant and host–parasite relationships and invasion of alien plant pests or pathogens medium 1 train +2553 AR6_WGII 56 14 Range shifts reduce biodiversity in the warmest regions and locations as adaptation limits are exceeded high 2 train +2554 AR6_WGII 56 15 Simultaneously, these shifts homogenise biodiversity (medium confidence) in regions receiving climate-migrant species, alter food webs and eliminate the distinctiveness of communities medium 1 train +2555 AR6_WGII 56 16 Increasing losses of habitat-forming species such as trees, corals, kelp and seagrass have caused irreversible shifts in some ecosystems and threaten associated biodiversity in marine systems high 2 train +2556 AR6_WGII 56 17 Human-introduced invasive (non-native) species can reduce or replace native species and alter ecosystem characteristics if they fare better than endemic species in new climate-altered ecological niches high 2 train +2557 AR6_WGII 56 18 Such invasive species effects are most prominent in geographically constrained areas, including islands, semi-enclosed seas and mountains, and they increase vulnerability in these systems high 2 train +2558 AR6_WGII 56 19 Phenological shifts increase the risks of temporal mismatches between trophic levels within ecosystems (medium confidence), which can lead to reduced food availability and population abundances medium 1 train +2559 AR6_WGII 56 21 Box 3.2, Box 3.4, 3.5.2, 3.5.3, 4.3.5, 9.6.1, 10.4.2, 11.3.1, 11.3.2, 11.3.11, 13.3.1, 13.4.1, 13.10.2, 14.5.1, 15.3.3, 15.3.4, 15.8, Box CCP1.1, CCP1.2.2, CCP1.2.1, CCP3.2.1, CCP5.2.1, CCB EXTREMES} TS.B.1.3 At the warm (equatorward and lower) edges of distribu tions, adaptation limits to human-induced warming have led to widespread local population losses (extirpations) that result in range contractions very high 3 train +2560 AR6_WGII 56 22 Among land plants and animals, local population loss was detected in around 50% of studied species and is often attributable to extreme events high 2 train +2561 AR6_WGII 56 24 Many mountain-top species have suffered population losses along lower elevations, leaving them increasingly restricted to a smaller area and at higher risk of extinction medium 1 train +2562 AR6_WGII 56 25 Global extinctions due to climate change are already being observed, with two extinctions currently attributed to anthropogenic climate change medium 1 train +2563 AR6_WGII 56 26 Climate-induced extinctions, including mass extinctions, are common in the palaeo record, underlining the potential of climate change to have catastrophic impacts on species and ecosystems high 2 train +2564 AR6_WGII 56 29 The Arctic is showing increased arrival of species from warmer areas on land and in the sea, with a declining extent of tundra and ice-dependent species, such as the polar bear high 2 train +2565 AR6_WGII 58 1 Coral reefs are suffering global declines, with abrupt shifts in community composition persisting for years very high 3 train +2566 AR6_WGII 58 2 Deserts and tropical systems are decreasing in diversity due to heat stress and extreme events high 2 train +2567 AR6_WGII 58 3 In contrast, arid lands are displaying varied responses around the globe in response to regional changes in the hydrological cycle high 2 train +2568 AR6_WGII 58 5 In terrestrial ecosystems, carbon uptake services linked to CO 2 fertilisation effects are being increasingly limited by drought and warming and exacerbated by non-climatic anthropogenic impacts high 2 train +2569 AR6_WGII 58 6 Deforestation, draining and burning of peatlands and tropical forests and thawing of Arctic permafrost have already shifted some areas from being carbon sinks to carbon sources high 2 train +2570 AR6_WGII 58 7 The severity and outbreak extent of forest insect pests increased in several regions high 2 train +2571 AR6_WGII 58 8 Woody plant expansion into grasslands and savannahs, linked to increased CO 2, has reduced grazing land, while invasive grasses in semiarid lands increased the risk of fire high 2 train +2572 AR6_WGII 58 9 Coastal ‘blue carbon’ systems are already impacted by multiple climate and non- climate drivers very high 3 test +2573 AR6_WGII 58 10 Warming and CO 2 fertilisation have altered coastal ecosystem biodiversity, making carbon storage or release regionally variable high 2 train +2574 AR6_WGII 58 12 Indigenous knowledge contains unique information sources about past changes and potential solutions to present issues medium 1 train +2575 AR6_WGII 58 13 Tangible heritage, such as traditional harvesting sites or species and archaeological and cultural heritage sites, and intangible heritage, such as festivals and rites associated with nature-based activities, endemic knowledge and unique insights about plants and animals, are being lost high 2 train +2576 AR6_WGII 58 15 Cultural losses threaten adaptive capacity and may accumulate into intergenerational trauma and irrevocable losses of sense of belonging, valued cultural practices, identity and home medium 1 train +2577 AR6_WGII 58 17 Extremes are surpassing the resil- ience of some ecological and human systems and challenging the adaptation capacities of others, including impacts with irre- versible consequences high 2 train +2578 AR6_WGII 58 18 Vulnerable people and human systems and climate-sensitive species and ecosystems are most at risk very high 3 train +2579 AR6_WGII 58 20 The most severe impacts are occurring in the most climate-sensitive species and ecosystems, characterised by traits that limit their abilities to regenerate between events or to adapt, and those most exposed to climate hazards high 2 train +2580 AR6_WGII 58 21 Losses of local plant and animal populations have been widespread, many associated with large increases in hottest yearly temperatures and heatwave events very high 3 train +2581 AR6_WGII 58 22 Marine heatwave events have led to widespread, abrupt and extensive mortality of key habitat-forming species among tropical corals, kelps, seagrasses and mangroves, as well as mass mortality of wildlife species, including benthic sessile species high 2 train +2582 AR6_WGII 59 7 Droughts, floods, wildfires and marine heatwaves contribute to reduced food availability and increased food prices, threatening food security, nutrition and livelihoods of millions of people across regions high 2 train +2583 AR6_WGII 59 8 Extreme events caused economic losses in forest productivity and crops and livestock farming, including losses in wheat production in 2012, 2016 and 2018, with the severity of impacts from extreme heat and drought tripling over the last 50 years in Europe high 2 train +2584 AR6_WGII 59 9 Forests were impacted by extreme heat and drought impacting timber sales, for example, in Europe high 2 train +2585 AR6_WGII 59 10 Marine heatwaves, including well-documented events along the west coast of North America (2013–2016) and east coast of Australia (2015–2016, 2016–2017 and 2020), have caused the collapse of regional fisheries and aquaculture high 2 train +2586 AR6_WGII 59 11 Human populations exposed to extreme weather and climate events are at risk of food insecurity with lower diversity in diets, leading to malnutrition and increased risk of disease high 2 train +2587 AR6_WGII 59 13 Since AR5, the impacts of climate change and extreme weather events such as wildfires, extreme heat, cyclones, storms and floods have adversely affected or caused loss and damage to human health, shelter, displacement, incomes and livelihoods, security and inequality high 2 train +2588 AR6_WGII 59 14 Over 20 million people have been internally displaced annually by weather-related extreme events since 2008, with storms and floods the most common drivers high 2 train +2589 AR6_WGII 59 15 Climate-related extreme events are followed by negative impacts on mental health, well-being, life satisfaction, happiness, cognitive performance and aggression in exposed populations very high 3 test +2590 AR6_WGII 59 19 Human-induced global warming has slowed the growth of agricultural productivity over the past 50 years in mid and low latitudes medium 1 train +2591 AR6_WGII 59 20 Crop yields are compromised by surface ozone high 2 train +2592 AR6_WGII 59 21 Methane emissions have negatively impacted crop yields by increasing temperatures and surface ozone concentrations medium 1 train +2593 AR6_WGII 59 22 Warming is negatively affecting crop and grassland quality and harvest stability high 2 train +2594 AR6_WGII 59 23 Warmer and drier conditions have increased tree mortality and forest disturbances in many temperate and boreal biomes (high confidence), negatively impacting provisioning services medium 1 train +2595 AR6_WGII 59 24 Ocean warming has decreased sustainable yields of some wild fish populations high 2 train +2596 AR6_WGII 59 25 Ocean acidification and warming have already affected farmed aquatic species high 2 train +2597 AR6_WGII 59 28 At higher latitudes, warming has expanded the available area but has also altered phenology (high confidence), potentially causing plant–pollinator and pest mismatches medium 1 train +2598 AR6_WGII 59 29 At low latitudes, temperatures have crossed upper tolerance thresholds, more frequently leading to heat stress and/ or shifts in distribution and losses for crops, livestock, fisheries and aquaculture high 2 train +2599 AR6_WGII 60 2 The impacts of climate-related extremes on food security, nutrition and livelihoods are particularly acute and severe for people living in sub-Saharan Africa, Asia, small islands, Central and South America and the Arctic and small-scale food producers globally high 2 test +2600 AR6_WGII 60 3 Droughts induced by the 2015–2016 El Niño, partially attributable to human influences (medium confidence), caused acute food insecurity in various regions, including eastern and southern Africa and the Dry Corridor of Central America high 2 train +2601 AR6_WGII 60 4 In the northeast Pacific, a 5-year warm period (2013 to 2017) impacted the migration, distribution and abundance of key fish resources high 2 train +2602 AR6_WGII 60 5 Increasing variability in grazing systems has negatively affected animal fertility, mortality and herd recovery rates, reducing livestock keepers’ resilience medium 1 train +2603 AR6_WGII 60 7 Higher temperatures and humidity caused by climate change increases toxigenic fungi on many food crops very high 3 train +2604 AR6_WGII 60 8 Harmful algal blooms and water-borne diseases threaten food security and the economy and livelihoods of many coastal communities high 2 train +2605 AR6_WGII 60 9 Increasing ocean warming and acidification are enhancing movement and bioaccumulation of toxins and contaminants into marine food webs (medium confidence) and with bio-magnification of persistent organic pollutants and methyl mercury already affecting fisheries medium 1 test +2606 AR6_WGII 60 10 Indigenous Peoples and local communities, especially where food safety monitoring is underdeveloped, are among the most vulnerable to these risks, in particular in the Arctic high 2 train +2607 AR6_WGII 60 12 Women, the elderly and children in low-income households, Indigenous Peoples, minority groups, small-scale producers and fishing communities and people in high-risk regions more often experience malnutrition, livelihood loss and rising costs high 2 train +2608 AR6_WGII 60 13 Increasing competition for critical resources, such as land, energy and water, can exacerbate the impacts of climate change on food security high 2 train +2609 AR6_WGII 60 16 Water insecurity is manifested through climate-induced water scar -city and hazards and is further exacerbated by inadequate water governance high 2 train +2610 AR6_WGII 60 18 Economic and societal impacts of water insecurity are more pronounced in low-income countries than in middle- and high-income ones high 2 train +2611 AR6_WGII 60 21 Human-induced climate change has affected physical aspects of water security through increasing water scarcity and exposing more people to water-related extreme events like floods and droughts, thereby exacerbating existing water-related vulnerabilities caused by other socioeconomic factors high 2 train +2612 AR6_WGII 60 22 Many of these changes in water availability and water-related hazards can be directly attributed to anthropogenic climate change high 2 train +2613 AR6_WGII 60 23 Water insecurity disproportionately impacts the poor, women, children, Indigenous Peoples and the elderly in low-income countries high 2 train +2614 AR6_WGII 60 24 Water insecurity can contribute to social unrest in regions where inequality is high and water governance and institutions are weak medium 1 train +2615 AR6_WGII 60 26 Nearly half a billion people now live in areas where the long-term average precipitation is now as high as was previously seen in only about 1 in 6 years medium 1 train +2616 AR6_WGII 60 27 Approximately 163 million people now live in unfamiliarly dry areas medium 1 test +2617 AR6_WGII 60 29 Substantially more people (around 709 million) live in regions where annual maximum 1-d precipitation has increased than in re- gions where it has decreased (around 86 million) medium 1 train +2618 AR6_WGII 61 2 Over the last two decades, the global glacier mass loss rate has been the highest since the glacier mass balance measurements began a century ago high 2 train +2619 AR6_WGII 61 4 Cryosphere changes have impacted cultural uses of water among vulnerable mountain and Arctic communities and Indigenous Peoples high 2 train +2620 AR6_WGII 61 5 Cryosphere change has affected ecosystems, water resources, livelihoods and cultural uses of water in all cryosphere- dependent regions across the world very high 3 train +2621 AR6_WGII 61 7 Anthropogenic climate change has led to increased likelihood, severity and societal impacts of droughts (primarily agricultural and hydrological droughts) in many regions high 2 train +2622 AR6_WGII 61 8 Between 1970 and 2019, drought-related disaster events worldwide caused billions of dollars in economic damages medium 1 train +2623 AR6_WGII 61 9 Drylands are particularly exposed to climate change related droughts high 2 train +2624 AR6_WGII 61 10 Recent heavy rainfall events that have led to catastrophic flooding were made more likely by anthropogenic climate change high 2 train +2625 AR6_WGII 61 11 Observed mortality and losses due to floods and droughts are much greater in regions with high vulnerability and vulnerable populations such as the poor, women, children, Indigenous Peoples and the elderly due to historical, political and socioeconomic inequities high 2 train +2626 AR6_WGII 61 13 Climate change and changes in land use and water pollution are key drivers of ecosystem loss and degradation (high confidence), with negative impacts observed on culturally significant terrestrial and freshwater species and ecosystems in the Arctic, mountain regions and other biodiversity hotspots high 2 train +2627 AR6_WGII 61 14 Climate trends and extreme events have had major impacts on many natural systems high 2 train +2628 AR6_WGII 61 15 For example, periodic droughts in parts of the Amazon since the 1990s, partly attributed to climate change, resulted in high tree mortality rates and basin-wide reductions in forest productivity, momentarily turning Amazon forests from a carbon sink into a net carbon source high 2 train +2629 AR6_WGII 61 16 Fire risks have increased due to heat and drought conditions in many parts of the world medium 1 train +2630 AR6_WGII 61 17 Increased precipitation has resulted in range shifts of species in some regions high 2 train +2631 AR6_WGII 61 19 Climate-induced trends and extremes in the water cycle have impacted agricultural production positively and negatively, with negative impacts outweighing the positive ones high 2 train +2632 AR6_WGII 61 20 Droughts, floods and rainfall variability have contributed to reduced food availability and increased food prices, threatening food and nutrition security, and the livelihoods of millions globally (high confidence), with the poor in parts of Asia, Africa and South and Central America being disproportionately affected high 2 train +2633 AR6_WGII 61 21 Drought years have reduced thermoelectric and hydropower production by around 4–5% compared to long-term average production since the 1980s (medium confidence), reducing economic growth in Africa and with billions in US dollars of existing and planned hydropower infrastructure assets in mountain regions worldwide and in Africa exposed to increasing hazards high 2 train +2634 AR6_WGII 61 22 Changes in temperature, precipitation and water-related disasters are linked to increased incidences of water- borne diseases such as cholera, especially in regions with limited access to safe water, sanitation and hygiene infrastructure high 2 train +2635 AR6_WGII 61 25 Women, children, the elderly, Indigenous People, low-income households and socially marginalised groups within cities, set- tlements, regions and countries are the most vulnerable high 2 train +2636 AR6_WGII 62 3 Distress sufficient to impair mental health has been caused by climate-related ecological grief associated with environmental change (e.g., solastalgia) or extreme weather and climate events (very high confidence), vicarious experience or anticipation of climate events (medium confidence) and climate-related loss of livelihoods and food insecurity very high 3 train +2637 AR6_WGII 62 4 Vulnerability to mental health effects of climate change varies by region and population, with evidence that Indigenous Peoples, agricultural communities, first responders, women and members of minority groups experience greater impacts high 2 train +2638 AR6_WGII 62 6 A significant proportion of warm-season heat-related mortality in temperate regions is attributed to observed anthropogenic climate change (medium confidence), with fewer data available for tropical regions in Africa high 2 train +2639 AR6_WGII 62 7 For some heatwave events over the last two decades, associated health impacts have been partially attributed to observed climate change high 2 train +2640 AR6_WGII 62 9 Potential hours of work lost due to heat have increased significantly over the past two decades high 2 train +2641 AR6_WGII 62 10 Some regions are already experiencing heat stress conditions at or approaching the upper limits of labour productivity high 2 train +2642 AR6_WGII 62 12 Extreme climate events have been key drivers in rising undernutrition of millions of people, primarily in Africa and Central America high 2 train +2643 AR6_WGII 62 13 For example, anthropogenic warming contributed to climate extremes induced by the 2015–2016 El Niño, which resulted in severe droughts, leading to an additional 5.9 million children in 51 countries becoming underweight high 2 train +2644 AR6_WGII 62 14 Undernutrition can in turn increase susceptibility to other health problems, including mental health problems, and impair cognitive and work performance, with resulting economic impacts very high 3 train +2645 AR6_WGII 62 15 Children and pregnant women experience disproportionate adverse health and nutrition impacts high 2 train +2646 AR6_WGII 62 17 These risks include Salmonella, Campylobacter and Cryptosporidium infections (medium confidence) mycotoxins associated with cancer and stunting in children (high confidence) and seafood contamination with marine toxins and pathogens high 2 train +2647 AR6_WGII 62 18 Climate-related food-borne disease risks vary temporally and are influenced, in part, by food availability, accessibility, preparation and preferences (medium confidence), as well as adequate food safety monitoring high 2 train +2648 AR6_WGII 62 20 More extreme weather events have contributed to vector-borne disease outbreaks in humans through direct effects on pathogens and vectors and indirect effects on human behaviour and emergency response destabilisation medium 1 train +2649 AR6_WGII 62 21 Climate change and variability are facilitating the spread of chikungunya virus in North, Central and South America, Europe and Asia (medium to high confidence); tick-borne encephalitis in Europe (medium confidence); Rift Valley fever in Africa; West Nile fever in southeastern Europe, western Asia, the Canadian prairies and parts of the USA (medium confidence); Lyme disease vectors in North America (high confidence) and Europe (medium confidence); malaria in eastern and southern Africa (high confidence); and dengue globally high 2 train +2650 AR6_WGII 62 22 For example, in Central and South America, the reproduction potential for the transmission of dengue increased between 17% and 80% for the period 1950–1954 to 2016–2021, depending on the sub-region, as a result of changes in temperature and precipitation high 2 train +2651 AR6_WGII 62 24 Water insecurity and inadequate water, sanitation and hygiene increase disease risk high 2 train +2652 AR6_WGII 63 2 Exposure to wildfires and associated smoke has increased in several regions very high 3 train +2653 AR6_WGII 63 4 Spring pollen season start dates in northern mid-latitudes are occurring earlier due to climate change, increasing the risks of allergic respiratory diseases high 2 train +2654 AR6_WGII 63 6 Most climate-related displacement and migration occur within national boundaries, with interna- tional movements occurring primarily between countries with contiguous borders high 2 train +2655 AR6_WGII 63 7 Since 2008, an annual average of over 20 million people have been internally dis- placed annually by weather-related extreme events, with storms and floods being the most common high 2 train +2656 AR6_WGII 63 9 Extreme climate events act as both direct drivers (e.g., destruction of homes by tropical cyclones) and indirect drivers (e.g., rural income losses during prolonged droughts) of involuntary migration and displacement very high 3 train +2657 AR6_WGII 63 10 The largest absolute number of people displaced by extreme weather each year occurs in Asia (South, Southeast and East), followed by sub-Saharan Africa, but small island states in the Caribbean and South Pacific are disproportionately affected relative to their small population size high 2 train +2658 AR6_WGII 63 12 Specific climate events and conditions cause migration to increase, decrease or flow in new directions high 2 train +2659 AR6_WGII 63 13 One of the main pathways for climate- induced migration is through deteriorating economic conditions and livelihoods high 2 train +2660 AR6_WGII 63 14 Climate change has influenced changes in temporary, seasonal or permanent migration, often rural to urban or rural to rural, that is associated with labour diversification as a risk- reduction strategy in Central America, Africa, South Asia and Mexico high 2 train +2661 AR6_WGII 63 15 This movement is often followed by remittances medium 1 train +2662 AR6_WGII 63 16 However, the same economic losses can also undermine household resources and savings, limiting mobility and compounding people’s exposure and vulnerability high 2 train +2663 AR6_WGII 63 20 Involuntary migration occurs when adaptation alternatives are exhausted or not viable and reflects non-climatic factors that constrain adaptive capacity and create high levels of exposure and vulnerability high 2 train +2664 AR6_WGII 63 21 These outcomes are also shaped by policy and planning decisions at regional, national and local scales that relate to housing, infrastructure, water provisioning, schools and healthcare to support the integration of migrants into receiving communities high 2 train +2665 AR6_WGII 63 25 Existing examples of relocations of Indigenous Peoples in coastal Alaska and villages in the Solomon Islands and Fiji suggest that relocated people can experience significant financial and emotional distress as cultural and spiritual bonds to place and livelihoods are disrupted high 2 train +2666 AR6_WGII 63 27 Vulnerability to climate change is a multi-dimension- al, dynamic phenomenon shaped by intersecting historical and contemporary political, economic and cultural processes of marginalisation high 2 train +2667 AR6_WGII 63 28 Societies with high levels of inequity are less resilient to climate change high 2 train +2668 AR6_WGII 64 2 Complex human vulnerability patterns are shaped by past developments, such as colonialism and its ongoing legacy (high confidence), are worsened by compounding and cascading risks high 2 train +2669 AR6_WGII 64 3 For example, low-income, young, poor and female- headed households face greater livelihood risks from climate hazards high 2 train +2670 AR6_WGII 64 6 Supporting Indigenous self-determination, recognising Indigenous Peoples’ rights and supporting Indigenous knowledge-based adaptation are critical to reducing climate change risks and effective adaptation very high 3 train +2671 AR6_WGII 64 8 There is evidence that present adaptation strategies do not sufficiently include poverty reduction and the underlying social determinants of human vulnerability such as gender, ethnicity and governance high 2 train +2672 AR6_WGII 64 10 Compared to other socioeconomic factors, the influence of climate on conflict is exacerbated by insecure land tenure, weather-sensitive economic activities, weak institutions and fragile governance, poverty and inequality medium 1 test +2673 AR6_WGII 64 11 The literature also suggests a larger climate-related influence on the dynamics of conflict than on the likelihood of initial conflict outbreak low 0 train +2674 AR6_WGII 64 14 More people and key assets are exposed to climate-induced impacts, and loss and damage in cities, set - tlements and key infrastructure since AR5 high 2 train +2675 AR6_WGII 64 15 Sea level rise, heatwaves, droughts, changes in runoff, floods, wildfires and permafrost thaw cause disruptions of key infra - structure and services such as energy supply and transmission, communications, food and water supply and transport systems in and between urban and peri-urban areas high 2 train +2676 AR6_WGII 64 16 The most rapid growth in urban vulnerability and exposure has been in cities and settlements where adaptive capacity is limited, including informal settlements in low- and middle-in - come communities and in smaller and medium-sized urban communities high 2 train +2677 AR6_WGII 64 18 The most rapid growth in urban vulnerability has been in unplanned and informal settlements and in smaller to medium urban centres in low- and middle-income nations where adaptive capacity is limited high 2 train +2678 AR6_WGII 64 19 Since AR5, observed impacts of climate change on cities, peri-urban areas and settlements have extended from direct, climate-driven impacts to compound, cascading and systemic impacts high 2 train +2679 AR6_WGII 64 22 The high degree of informality limits adaptation and increases differential vulnerability to climate change high 2 train +2680 AR6_WGII 64 23 Globally, exposure to climate-driven impacts such as heatwaves, extreme precipitation and storms in combination with rapid urbanisation and lack of climate-sensitive planning, along with continuing threats from urban heat islands, is increasing the vulnerability of marginalised urban populations and key infrastructure to climate change, for example, more frequent and/ or extreme rainfall and drought stress existing design and capacity of current urban water systems and heighten urban and peri-urban water insecurity high 2 train +2681 AR6_WGII 64 24 COVID-19 has had a substantial urban impact and generated new climate-vulnerable populations high 2 train +2682 AR6_WGII 65 1 Coastal cities are disproportionately affected by interacting, cascading and climate-compounding climate- and ocean-driven impacts, in part because of the exposure of multiple assets, economic activities and large populations concentrated in narrow coastal zones high 2 train +2683 AR6_WGII 65 2 Early impacts of accelerating sea level rise have been detected at sheltered or subsiding coasts, manifesting as nuisance and chronic flooding at high tides, water-table salinisation, ecosystem and agricultural transitions, increased erosion and coastal flood damage medium 1 train +2684 AR6_WGII 65 3 Coastal settlements with high inequality, for example a high proportion of informal settlements, as well as deltaic cities prone to land subsidence (e.g., Bangkok, Jakarta, Lagos, New Orleans, Mississippi, Nile, Ganges-Brahmaputra deltas) and small island states are highly vulnerable and have experienced impacts from severe storms and floods in addition to, or in combination with, those from accelerating sea level rise high 2 train +2685 AR6_WGII 65 4 Currently, coastal cities already dependent on extensive protective works face the prospects of significantly increasing costs to maintain current protection levels, especially if the local sea level rises to the point that financial and technical limits are reached; systemic changes, such as relocation of millions of people, will be necessary medium 1 test +2686 AR6_WGII 65 6 Vulnerabilities vary by location and are shaped by intersecting processes of marginalization, including gender, class, race, income, ethnic origin, age, level of ability, sexuality and non- conforming gender orientation high 2 train +2687 AR6_WGII 65 8 Due to the connectivity of infrastructure systems, climate impacts, such as with thawing permafrost or severe storms affecting energy and transport networks, can propagate outside the reach of the hazard footprint and cause larger impacts and widespread regional disruption high 2 train +2688 AR6_WGII 65 9 Interdependencies between infrastructure systems have created new pathways for compounding climate risk, which has been accelerated by trends in information and communication technologies, increased reliance on energy, and complex (often global) supply chains high 2 train +2689 AR6_WGII 65 11 Recent extreme weather and climate-induced events have been associated with large costs through damaged property, infrastructure and supply chain disruptions, although development patterns have driven much of these increases high 2 train +2690 AR6_WGII 65 12 Adverse impacts on economic growth have been identified from extreme weather events (high confidence) with large effects in developing countries high 2 train +2691 AR6_WGII 65 13 Widespread climate impacts have undermined economic livelihoods, especially among vulnerable populations high 2 train +2692 AR6_WGII 65 14 Climate impacts and projected risks have been insufficiently internalised into private- and public-sector planning and budgeting practices and adaptation finance medium 1 train +2693 AR6_WGII 65 16 Greater economic losses are observed for sectors with high direct climate exposure, including regional losses to agriculture, forestry, fisheries, energy and tourism high 2 train +2694 AR6_WGII 65 17 Many industrial and service sectors are indirectly affected through supply disruptions, especially during and following extreme events high 2 train +2695 AR6_WGII 65 18 Costs are also incurred from adaptation, disaster spending, recovery and rebuilding of infrastructure high 2 train +2696 AR6_WGII 65 19 Estimates of the global effects of climate change on aggregate measures of economic performance and gross domestic product (GDP) range from negative to positive, in part due to uncertainty in how weather variability and climate impacts manifest in GDP high 2 train +2697 AR6_WGII 65 20 Climate change is estimated to have slowed trends of decreasing economic inequality between developed and developing countries (low confidence), with particularly negative effects for Africa medium 1 train +2698 AR6_WGII 65 22 Extreme weather events, such as tropical cyclones, droughts and severe fluvial floods, have reduced economic growth in the short term (high confidence) and will continue to reduce it in the coming decades medium 1 train +2699 AR6_WGII 65 23 Patterns of development have augmented the exposure of more assets to extreme hazards, increasing the magnitude of the losses high 2 train +2700 AR6_WGII 65 24 Small Island Developing States have reported economic losses and a wide range of damage from tropical cyclones and increases in sea level rise high 2 train +2701 AR6_WGII 66 2 Climate-sensitive livelihoods are more concentrated in regions that have higher socioeconomic vulnerabilities and lower adaptive capacities, exacerbating existing inequalities medium 1 train +2702 AR6_WGII 66 3 Extreme events have also had more pronounced adverse effects in poorer regions and on more vulnerable populations medium 1 train +2703 AR6_WGII 66 4 These greater economic effects have further reduced the ability of these populations to adapt to existing impacts medium 1 train +2704 AR6_WGII 66 5 Within populations, the poor, women, children, elderly and Indigenous populations have been especially vulnerable due to a combination of factors, including gendered divisions of paid and/ or unpaid labour high 2 train +2705 AR6_WGII 66 7 Existing adaptation has prevented greater economic losses (medium confidence), yet adaptation gaps remain due to limited financial resources, including gaps in international adaptation finance and competing priorities in budget allocations medium 1 train +2706 AR6_WGII 66 8 Insufficient consideration of these impacts, however, has placed more assets in areas that are highly exposed to climate hazards medium 1 train +2707 AR6_WGII 66 15 Threats to species and ecosystems in oceans, coastal regions and on land, particularly in biodiversity hotspots, present a global risk that will increase with every additional tenth of a degree of warming high 2 train +2708 AR6_WGII 66 16 The transformation of terrestrial and ocean/ coastal ecosystems and loss of biodiversity, exacerbated by pollution, habitat fragmentation and land use changes, will threaten livelihoods and food security high 2 train +2709 AR6_WGII 66 18 Risks escalate with additional near-term warming in all regions and domains high 2 train +2710 AR6_WGII 66 20 Unique and threatened ecosystems are expected to be at high risk in the very near term at 1.2°C global warming levels very high 3 train +2711 AR6_WGII 66 21 Even for less vulnerable species and systems, projected climate change risks surpass hard limits to natural adaptation, increasing species at high risk of population declines (medium confidence) and loss of critical habitats (medium to high confidence) and compromising ecosystem structure, functioning and resilience medium 1 train +2712 AR6_WGII 66 22 At a global warming of 2°C with associated changes in precipitation global land area burned by wildfire is projected to increase by 35% medium 1 train +2713 AR6_WGII 66 24 Beginning at 1.5°C warming, natural adaptation faces hard limits, driving high risks of biodiversity decline, mortality, species extinction and loss of related livelihoods high 2 train +2714 AR6_WGII 66 25 At 1.6°C (median estimate), >10% of species are projected to become endangered, increasing to >20% at 2.1°C, representing severe biodiversity risk medium 1 train +2715 AR6_WGII 66 26 These risks escalate with warming, most rapidly and severely in areas at both extremes of temperature and precipitation high 2 train +2716 AR6_WGII 66 27 With warming of 3°C, >80% of marine species across large parts of the tropical Indian and Pacific Ocean will experience potentially dangerous climate conditions medium 1 train +2717 AR6_WGII 67 2 Space for nature is shrinking as large areas of forest are lost to deforestation (high confidence), peat draining and agricultural expansion, land reclamation and protection structures in urban and coastal settlements high 2 train +2718 AR6_WGII 67 3 Currently less than 15% of the land and 8% of the ocean are under some form of protection, and enforcement of protection is often weak high 2 train +2719 AR6_WGII 67 4 Future ecosystem vulnerability will strongly depend on developments in society, including demographic and economic change high 2 train +2720 AR6_WGII 67 6 Coordinated and well-monitored habitat restoration, protection and management, combined with consumer pressure and incentives, can reduce non- climatic impacts and increase resilience high 2 train +2721 AR6_WGII 67 7 Adaptation and mitigation options, such as afforestation, dam construction and coastal infrastructure placements, can increase vulnerability, compete for land and water and generate risks for the integrity and functioning of ecosystems high 2 train +2722 AR6_WGII 67 9 In terrestrial ecosystems, the fertilisation effects of high atmospheric CO 2 concentrations on carbon uptake will be increasingly saturated and limited by warming and drought medium 1 train +2723 AR6_WGII 67 10 Increases in wildfires, tree mortality, insect pest outbreaks, peatland drying and permafrost thaw (high confidence) all exacerbate self-reinforcing feedbacks between emissions from high-carbon ecosystems and warming with the potential to turn many ecosystems that are currently net carbon sinks into sources medium 1 train +2724 AR6_WGII 67 11 In coastal areas beyond 1.5°C warming, blue carbon storage by mangroves, marshes and seagrass habitats are increasingly threatened by rising sea levels and the intensity, duration and extent of marine heatwaves, as well as adaptation options (including coastal development) high 2 train +2725 AR6_WGII 67 12 Changes in ocean stratification are projected to reduce nutrient supply and alter the magnitude and efficiency of the biological carbon pump medium 1 train +2726 AR6_WGII 67 14 The median values for percentage of species at very high risk of extinction are 9% at 1.5°C, 10% at 2°C, 12% at 3°C, 13% at 4°C and 15% at 5°C high 2 train +2727 AR6_WGII 67 15 Extinction risks are higher for species in biodiversity hotspots (medium confidence), reaching 24% of species at very high extinction risk above 1.5°C, with yet higher proportions for endemic species of 84% in mountains (medium confidence) and 100% on islands medium 1 train +2728 AR6_WGII 67 16 Thousands of individual populations are projected to be locally lost, which will reduce species diversity in some areas where there are no species moving in to replace them, for example, in tropical systems high 2 train +2729 AR6_WGII 67 17 Novel species interactions at the cold edge of species’ distribution may also lead to extirpations and extinctions of newly encountered species low 0 train +2730 AR6_WGII 67 18 Palaeo records indicate that at extreme warming levels (>5°C), mass extinctions of species occur medium 1 train +2731 AR6_WGII 67 21 These processes will exacerbate both stress on systems already at risk from climate impacts and non-climate impacts like habitat fragmentation and pollution high 2 train +2732 AR6_WGII 67 22 The increasing frequency and severity of extreme events will decrease the recovery time available for ecosystems high 2 train +2733 AR6_WGII 67 23 Irreversible changes will occur from the interaction of stressors and the occurrence of extreme events very high 3 train +2734 AR6_WGII 67 25 Ecosystem integrity is threatened by the positive feedback between direct human impacts (land use change, pollution, overexploitation, fragmentation and destruction) and climate change high 2 train +2735 AR6_WGII 67 26 In the case of the Amazon forest, this could lead to large-scale ecological transformations and shifts from a closed, wet forest into a drier and lower-biomass vegetation medium 1 train +2736 AR6_WGII 67 27 If these pressures are not successfully addressed, the combined and interactive effects between climate change, deforestation and forest degradation, and forest fires are projected to lead to a reduction of over 60% of the area covered by forest in response to 2.5°C global warming level medium 1 train +2737 AR6_WGII 67 28 Some habitat-forming coastal ecosystems, including many coral reefs, kelp forests and seagrass meadows, will undergo irreversible phase shifts due to marine heatwaves with global warming levels >1.5°C and are at high risk this century even in <1.5°C scenarios that include periods of temperature overshoot beyond 1.5°C high 2 train +2738 AR6_WGII 67 29 Under SSP1–2.6, coral reefs are at risk of widespread decline, loss of structural integrity and transitioning to net erosion by mid-century due to the increasing intensity and frequency of marine heatwaves very high 3 train +2739 AR6_WGII 68 1 In response to heatwaves, bleaching of the Great Barrier Reef is projected to occur annually if warming increases above 2.0°C, resulting in widespread decline and loss of structural integrity very high 3 train +2740 AR6_WGII 68 2 Global warming of 3.0°C–3.5°C increases the likelihood of extreme and lethal heat events in western and northern Africa medium 1 train +2741 AR6_WGII 68 3 Drought risks are projected to increase in many regions over the 21st century very high 3 train +2742 AR6_WGII 68 5 Invasive plant species are predicted to expand both in latitude and altitude high 2 train +2743 AR6_WGII 68 8 Depending on location and human–wildlife interactions, climate-driven shifts in distributions of wild animals increase the risk of emergence of novel human infectious diseases, as has occurred with SARS, MERS and SARS-CoV-2 medium 1 train +2744 AR6_WGII 68 11 These control measures will become costlier under climate change medium 1 train +2745 AR6_WGII 68 13 Limiting the global temperature increase to 1.5°C, compared to 2°C, could reduce projected permafrost CO 2 losses by 2100 by 24.2 GtC low 0 train +2746 AR6_WGII 68 18 For example, by 2100, 18.8% ± 19.0% to 38.9% ± 9.4% of the ocean will very likely undergo a change of more than 20 days (advances and delays) in the start of the phytoplankton growth period under SSP1-2.6 and SSP5-8.5 respectively low 0 train +2747 AR6_WGII 68 19 This altered timing increases the risk of temporal mismatches between plankton blooms and fish spawning seasons (medium to high confidence) and increases the risk of fish recruitment failure for species with restricted spawning locations, especially in mid- to high latitudes of the northern hemisphere (low confidence) but provide short-term opportunities to countries benefiting from shifting fish stocks medium 1 train +2748 AR6_WGII 68 21 Major risks include loss of coastal ecosystems such as wetlands and marshlands from committed sea level rise associated with overshoot warming (medium confidence), coral reefs and kelps from heat-related mortality and associated ecosystem transitions high 2 train +2749 AR6_WGII 68 23 With every increment of warming, exposure to climate hazards will grow substantially (high confidence), and adverse impacts on all food sectors will become prevalent, further stressing food security high 2 train +2750 AR6_WGII 68 24 Regional disparity in risks to food security will grow with warming levels, increasing poverty traps, particularly in regions characterised by a high level of human vulnerability high 2 train +2751 AR6_WGII 68 27 Compared to 1.5°C global warming level, 2°C global warming level will even further negatively impact food production where current temperatures are already high as in lower latitudes high 2 train +2752 AR6_WGII 68 28 Increased and potentially concurrent climate extremes will increase simultaneous losses in major food-producing regions medium 1 train +2753 AR6_WGII 68 29 The adverse effects of climate change on food production will become more severe when global temperatures rise by more than 2°C high 2 train +2754 AR6_WGII 68 30 At 3°C or higher global warming levels, exposure to climate hazards will grow substantially (high confidence), further stressing food production, notably in sub-Saharan Africa and South and South East Asia high 2 train +2755 AR6_WGII 71 1 Catch composition will change regionally, and the vulnerability of fishers will partially depend on their ability to move, diversify and leverage technology medium 1 train +2756 AR6_WGII 71 2 Global marine aquaculture will decline under increasing temperature and acidification conditions by 2100, with potential short-term gains for finfish aquaculture in some temperate regions and overall negative impacts on bivalve aquaculture due to habitat reduction medium 1 train +2757 AR6_WGII 71 3 Changes in precipitation, sea level rise, temperature and extreme events will negatively affect food provisioning from inland aquatic systems medium 1 train +2758 AR6_WGII 71 5 Extreme weather events will increase risks of food insecurity via spikes in food prices, reduced food diversity and reduced income for agricultural and fishery livelihoods (high confidence), preventing achievement of the UN SDG 2 (‘Zero Hunger’) by 2030 in regions with limited adaptive capacities, including Africa, small island states and South Asia high 2 train +2759 AR6_WGII 71 6 With about 2°C warming, climate-related changes in food availability and diet quality are estimated to increase nutrition-related diseases and the number of undernourished people by 2050, affecting tens (under low vulnerability and low warming) to hundreds of millions of people (under high vulnerability and high warming, i.e., SSP-3-RCP6.0), particularly among low-income households in low- and middle-income countries in sub- Saharan Africa, South Asia and Central America high 2 train +2760 AR6_WGII 71 7 At 3°C or higher global warming levels, adverse impacts on all food sectors will become prevalent, further stressing food availability (high confidence), agricultural labour productivity and food access medium 1 train +2761 AR6_WGII 71 8 Regional disparity in risks to food security will grow at these higher warming levels, increasing poverty traps, particularly in regions characterised by a high level of human vulnerability high 2 train +2762 AR6_WGII 71 10 Increased CO 2 concentrations promote crop growth and yield but reduce the density of important nutrients in some crops (high confidence) with projected increases in undernutrition and micronutrient deficiency, particularly in countries that currently have high levels of nutrient deficiency (high confidence) and regions with low access to diverse foods medium 1 train +2763 AR6_WGII 71 11 Marine-dependent communities, including Indigenous Peoples and local peoples, will be at increased risk of malnutrition due to losses of seafood-sourced nutrients medium 1 train +2764 AR6_WGII 72 1 Climate change will reduce the effectiveness of pollination as species are lost from certain areas, or the coordination of pollinator activity and flower receptiveness will be disrupted in some regions high 2 train +2765 AR6_WGII 72 2 Greenhouse gas emissions will negatively impact air, soil and water quality, exacerbating direct climatic impacts on yields high 2 train +2766 AR6_WGII 72 4 Higher temperatures and humidity will expand the risk of aflatoxin contamination into higher- latitude regions high 2 train +2767 AR6_WGII 72 5 More frequent and intense flood events and increased melting of snow and ice will increase food contamination high 2 train +2768 AR6_WGII 72 6 Aquatic food safety will decrease through increased detrimental impacts from harmful algal blooms high 2 train +2769 AR6_WGII 72 7 These negative food safety impacts will be greater without adaptation and fall disproportionately on low-income countries and communities with high consumption of seafood, including coastal Indigenous communities medium 1 train +2770 AR6_WGII 72 9 Regions and populations with higher exposure and vulnerability are pro- jected to face greater risks than others medium 1 train +2771 AR6_WGII 72 10 Projected changes in the water cycle, water quality, cryosphere changes, drought and flood will negatively impact natural and human systems high 2 train +2772 AR6_WGII 72 12 About 800 million to 3 billion people at 2°C and about 4 billion at 4°C warming are projected to experience different levels of water scarcity medium 1 train +2773 AR6_WGII 72 13 At 4°C global warming by the end of the century, approximately 10% of the global land area is projected to face simultaneously increasing high extreme streamflow and decreasing low extreme streamflow, affecting over 2.1 billion people medium 1 train +2774 AR6_WGII 72 14 Globally, the greatest risks to attaining global sustainability goals come from risks to water security high 2 train +2775 AR6_WGII 72 16 Glacier mass loss, permafrost thaw and decline in snow cover are projected to continue beyond the 21st century high 2 train +2776 AR6_WGII 72 17 Many low- elevation and small glaciers around the world will lose most of their total mass at 1.5°C warming high 2 train +2777 AR6_WGII 72 19 Glacier lake outburst flood will threaten the security of local and downstream communities in High Mountain Asia high 2 train +2778 AR6_WGII 72 20 By 2100, annual runoff in one-third of the 56 large-scale glacierised catchments are projected to decline by over 10%, with the most significant reductions in Central Asia and the Andes medium 1 train +2779 AR6_WGII 72 21 Cryosphere related changes in floods, landslides and water availability have the potential to lead to severe consequences for people, infrastructure and the economy in most mountain regions high 2 train +2780 AR6_WGII 72 23 By 2050, environmentally critical streamflow is projected to be affected in 42% to 79% of the world’s watersheds, causing negative impacts on freshwater ecosystems medium 1 train +2781 AR6_WGII 72 24 Increased wildfire, combined with soil erosion due to deforestation, could degrade water supplies medium 1 train +2782 AR6_WGII 72 25 Projected climate-driven water cycle changes, including increases in evapotranspiration, altered spatial patterns and amount of precipitation, and associated changes in groundwater recharge, runoff and streamflow, will impact terrestrial, freshwater, estuarine and coastal ecosystems and the transport of materials through the biogeochemical cycles, impacting humans and societal well-being medium 1 train +2783 AR6_WGII 72 27 In Central and South America, disruption in water flows will significantly degrade ecosystems such as high-elevation wetlands high 2 train +2784 AR6_WGII 72 29 Under RCP6.0 and SSP2, the population that is projected to be exposed to extreme to exceptional low total water storage will reach up to 7% over the 21st century medium 1 train +2785 AR6_WGII 72 31 In southern Europe, more than a third of the population will be exposed to water scarcity at 2°C, and the risk doubles at 3°C, with significant economic losses medium 1 train +2786 AR6_WGII 73 1 Above 2°C, the frequency and duration of meteorological drought are projected to double over North Africa, the western Sahel and southern Africa medium 1 train +2787 AR6_WGII 73 2 More droughts and extreme fire weather are projected in southern and eastern Australia (high confidence) and over most of New Zealand medium 1 train +2788 AR6_WGII 73 4 The projected increase in precipitation intensity (high confidence) will increase rain-generated local flooding medium 1 train +2789 AR6_WGII 73 5 Direct flood damage is projected to increase by four to five times at 4°C compared to 1.5°C medium 1 train +2790 AR6_WGII 73 6 A higher sea level with storm surge further inland may create more severe coastal flooding high 2 train +2791 AR6_WGII 73 7 Projected intensifications of the hydrological cycle pose increasing risks, including potential doubling of flood risk and 1.2- to 1.8-fold increase in GDP loss due to flooding between 1.5°C and 3°C medium 1 train +2792 AR6_WGII 73 8 Projected increase in heavy rainfall events at all levels of warming in many regions in Africa will cause increasing exposure to pluvial and riverine flooding high 2 train +2793 AR6_WGII 73 9 A 1.5°C increase would result in an increase of 100–200% in the population affected by floods in Colombia, Brazil and Argentina, 300% in Ecuador and 400% in Peru medium 1 train +2794 AR6_WGII 73 12 Agricultural water use will increase globally as a consequence of population increase and dietary changes, as well as increased water requirements due to climate change high 2 train +2795 AR6_WGII 73 14 Increased floods and droughts, together with heat stress, will have an adverse impact on food availability and prices, resulting in increased undernourishment in South and Southeast Asia high 2 train +2796 AR6_WGII 73 19 Historically rare extreme sea level events will occur annu- ally by 2100, compounding these risks high 2 train +2797 AR6_WGII 73 21 These risks will be compounded where coastal development prevents upshore migration of habitats or where terrestrial sediment inputs are limited and tidal ranges are small high 2 train +2798 AR6_WGII 73 22 Loss of these habitats disrupts associated ecosystem services, including wave-energy attenuation, habitat provision for biodiversity, climate mitigation and food and fuel resources high 2 train +2799 AR6_WGII 73 23 Near- to mid-term sea level rise will also exacerbate coastal erosion and submersion and the salinisation of coastal groundwater, expanding the loss of many different coastal habitats, ecosystems and ecosystem services medium 1 train +2800 AR6_WGII 73 25 High population growth and urbanisation in low-lying coastal zones will be the major driver of increasing exposure to sea level rise in the coming decades high 2 train +2801 AR6_WGII 73 26 By 2030, 108– 116 million people will be exposed to sea level rise in Africa (compared to 54 million in 2000), increasing to 190–245 million by 2060 medium 1 train +2802 AR6_WGII 73 27 By 2050, more than a billion people located in low-lying cities and settlements will be at risk from coast-specific climate hazards, influenced by coastal geomorphology, geographical location and adaptation action high 2 train +2803 AR6_WGII 73 31 By 2100, compound and cascading risks will result in the submergence of some low-lying island states and damage to coastal heritage, livelihoods and infrastructure very high 3 train +2804 AR6_WGII 74 4 For example, in Europe, coastal flood damage is projected to increase at least 10-fold by the end of the 21st century, and even more or earlier with current adaptation and mitigation high 2 test +2805 AR6_WGII 74 5 By 2100, 158–510 million people and USD7,919–12,739 billion in assets are projected to be exposed to the 1-in-100-year coastal floodplain under RCP4.5, and 176–880 million people and USD8,813–14,178 billion assets under RCP8.5 high 2 train +2806 AR6_WGII 74 6 Projected impacts reach far beyond coastal cities and settlements, with damage to ports potentially severely compromising global supply chains and maritime trade, with local to global geopolitical and economic ramifications medium 1 train +2807 AR6_WGII 74 7 Compounded and cascading climate risks, such as tropical cyclone storm surge damage to coastal infrastructure and supply chain networks, are expected to increase medium 1 train +2808 AR6_WGII 74 9 Changes in wave climate superimposed on sea level rise will significantly increase coastal flooding high 2 train +2809 AR6_WGII 74 10 The frequency, extent and duration of coastal flooding will significantly increase from 2050 (high confidence), unless coastal and marine ecosystems are able to naturally adapt to sea level rise through vertical growth and landward migration low 0 train +2810 AR6_WGII 74 11 Permafrost thaw, sea level rise, and reduced sea ice protection is projected to damage or cause loss to many cultural heritage sites, settlements and livelihoods across the Arctic very high 3 train +2811 AR6_WGII 74 12 Deltaic cities and settlements characterised by high inequality and informal settlements are especially vulnerable high 2 train +2812 AR6_WGII 74 13 Although risks are distributed across cities and settlements at all levels of economic development, wealthier and more urbanised coastal cities and settlements are more likely to be able to limit impacts and risk in the near- to mid-term through infrastructure resilience and coastal protection interventions, with highly uncertain prospects in many of these locations beyond 2100 high 2 train +2813 AR6_WGII 74 14 Prospects for enabling and contributing to climate resilient development thus vary markedly within and between coastal cities and settlements high 2 train +2814 AR6_WGII 74 17 Health risks will be differentiated by gender, age, income, social status and region high 2 train +2815 AR6_WGII 74 19 Projections under mid-range emissions scenarios show an additional 250,000 deaths per year by 2050 (compared to 1961–1990) due to malaria, heat, childhood undernutrition and diarrhoea high 2 train +2816 AR6_WGII 74 21 Mortality and morbidity will continue to escalate as exposures become more frequent and intense, putting additional strain on health and economic systems high 2 train +2817 AR6_WGII 74 22 Vulnerable groups include young children (<5 years old), the elderly (>65 years old), pregnant women, Indigenous Peoples, those with pre-existing diseases, physical labourers and those in low socioeconomic conditions high 2 train +2818 AR6_WGII 74 24 Children and adolescents, particularly girls, as well as people with existing mental, physical and medical challenges, are particularly at risk high 2 train +2819 AR6_WGII 74 25 Mental health impacts are expected to arise from exposure to extreme weather events, displacement, migration, famine, malnutrition, degradation or destruction of health and social care systems, climate-related economic and social losses and anxiety and distress associated with worry about climate change very high 3 train +2820 AR6_WGII 74 27 Globally, temperature- related mortality is projected to increase under RCP4.5 to RCP8.5, even with adaptation very high 3 train +2821 AR6_WGII 74 29 In Melbourne, Sydney and Brisbane, urban heat-related excess deaths are projected to increase by about 300 yr-1 (low emission pathway) to 600 yr-1 (high emission pathway) during 2031–2080 relative to 142 yr-1 during 1971–2020 high 2 train +2822 AR6_WGII 75 2 Reduced marine and freshwater fisheries catch potential is projected to increase malnutrition in East, West and Central Africa (medium to high confidence) and in subsistence- dependent communities across North America high 2 train +2823 AR6_WGII 75 4 These projected changes will increase diet-related risk factors and related non-communicable diseases globally and increase undernutrition, stunting and related childhood mortality, particularly in Africa and Asia high 2 train +2824 AR6_WGII 75 6 Climate change is projected to put 8 million (SSP1-6.0) to 80 million people (SSP3-6.0) at risk of hunger in mid-century, concentrated in sub-Saharan Africa, South Asia and Central America high 2 train +2825 AR6_WGII 75 7 These climate change impacts on nutrition could undermine progress towards the eradication of child undernutrition high 2 train +2826 AR6_WGII 75 9 Dengue vector ranges will increase in North America, Asia, Europe and sub-Saharan Africa under RCP6 and RCP8.5, potentially putting another 2.25 billion people at risk high 2 train +2827 AR6_WGII 75 10 Higher incidence rates of Lyme disease are projected for the Northern Hemisphere high 2 train +2828 AR6_WGII 75 11 Climate change is projected to increase malaria’s geographic distribution in endemic areas of sub-Saharan and southern Africa, Asia and South America (high confidence), exposing tens of millions more people to malaria, predominately in east and southern Africa, and up to hundreds of millions more exposed under RCP8.5 high 2 train +2829 AR6_WGII 75 13 At 2.1°C, thousands to tens of thousands of additional cases of diarrhoeal disease are projected, mainly in central and east Africa medium 1 train +2830 AR6_WGII 75 14 Morbidity from cholera will increase in central and east Africa (medium confidence), and increased schistosomiasis risk is projected for eastern Africa high 2 train +2831 AR6_WGII 75 15 In Asia and Africa, 1°C warming can cause a 7% increase in diarrhoea, an 8% increase in E. coli and a 3% to 11% increase in deaths medium 1 train +2832 AR6_WGII 75 16 Warming increases the risk of food-borne disease outbreaks, including Salmonella and Campylobacter infections medium 1 train +2833 AR6_WGII 75 17 Warming supports the growth and geographical expansion of toxigenic fungi in crops (medium confidence) and potentially toxic marine and freshwater algae medium 1 train +2834 AR6_WGII 75 18 Food safety risks in fisheries and aquaculture are projected through harmful algal blooms (high confidence), pathogens (e.g., Vibrio) (high confidence), and human exposure to elevated bioaccumulation of persistent organic pollutants and mercury medium 1 train +2835 AR6_WGII 75 20 Cardiovascular disease mortality could increase by 18.4%, 47.8% and 69.0% in the 2020s, 2050s and 2080s respectively under RCP4.5, and by 16.6%, 73.8% and 134% under RCP8.5 compared to the 1980s high 2 train +2836 AR6_WGII 75 21 Future risks of respiratory disease associated with aeroallergens and ozone exposure are expected to increase high 2 train +2837 AR6_WGII 75 23 In many regions, the frequency and/or severity of floods, extreme storms and droughts is projected to increase in coming decades, es- pecially under high emissions scenarios, raising future risk of displacement in the most exposed areas high 2 train +2838 AR6_WGII 75 24 Under all global warming levels, some regions that are pres- ently densely populated will become unsafe or uninhabitable, with movement from these regions occurring autonomously or through planned relocation high 2 train +2839 AR6_WGII 75 26 Future migration and displacement patterns in a changing climate will depend not only on the physical impacts of climate change, but also on future policies and planning at all scales of governance high 2 train +2840 AR6_WGII 75 27 Projecting the number of people migrating due to slow onset events is difficult due to the multi- causal nature of migration and the dominant role that socioeconomic factors have in determining migration responses high 2 train +2841 AR6_WGII 75 29 High emissions/low development scenarios raise the potential for both increased rates of migration and displacement and larger involuntary immobile populations that are highly exposed to climatic risks but lack the means of moving to other locations medium 1 train +2842 AR6_WGII 76 1 Uncertainties about socioeconomic development are reflected in the wide range of projected population displacements by 2050 in Central and South America, sub-Saharan Africa and South Asia due to climate change, ranging from 31 million to 143 million people high 2 train +2843 AR6_WGII 76 2 Projections of the number of people at risk of future displacement by sea level rise range from tens of millions to hundreds of millions by the end of this century, depending on the level of warmings and assumptions about exposure high 2 train +2844 AR6_WGII 76 4 Planned relocation will be increasingly required as climate change undermines livelihoods, safety and overall habitability, especially for coastal areas and small islands medium 1 train +2845 AR6_WGII 76 5 This will have implications for traditional livelihood practices, social cohesion and knowledge systems that have inherent value as intangible culture as well as introduce new risks for communities by amplifying existing and generating new vulnerabilities high 2 train +2846 AR6_WGII 76 7 Future climate change may increase involuntary displacement, but severe impacts also undermine the capacity of households to use mobility as a coping strategy, causing high exposure to climate risks, with consequences for basic survival, health and well- being high 2 train +2847 AR6_WGII 76 8 The COVID-19 pandemic is expected to increase the adverse consequences of climate change since the financial consequences have led to a shift in priorities and constrain vulnerability reduction medium 1 train +2848 AR6_WGII 76 10 Under higher emissions scenarios and increasing climate hazards, the potential for societal risks also increases medium 1 train +2849 AR6_WGII 76 13 Cultural heritage is already being impacted by climate change and variability, for example in Africa, Small Island Developing States and the Arctic, where heritage sites are exposed to future climate change risk high 2 train +2850 AR6_WGII 76 18 Climate change may produce severe risks to peace within this century through climate variability and extremes, especially in contexts marked by low economic development, high economic dependence on climate-sensitive activities, high or increasing social marginalisation and fragile governance medium 1 train +2851 AR6_WGII 76 19 The largest impacts are expected in weather-sensitive communities with low resilience to climate extremes and high prevalence of underlying risk factors medium 1 train +2852 AR6_WGII 76 20 Trajectories that prioritise economic growth, political rights and sustainability are associated with lower conflict risk medium 1 train +2853 AR6_WGII 76 22 In all cities and urban areas, projected risks faced by people from climate-driven impacts has increased high 2 train +2854 AR6_WGII 76 24 Communities in informal settlements will have higher exposure and lower capacity to adapt high 2 train +2855 AR6_WGII 76 25 Most at risk are women and children who make up the majority populations of these settlements high 2 train +2856 AR6_WGII 76 26 Risks to critical physical infrastructure in cities can be severe and pervasive under higher warming levels, potentially resulting in compound and cascading risks, and can disrupt livelihoods both within and across cities high 2 train +2857 AR6_WGII 76 27 In coastal cities and settlements, risks to people and infrastructure will get progressively worse in a changing climate, sea level rise and with ongoing coastal development very high 3 train +2858 AR6_WGII 76 30 Growth is most pronounced in smaller and medium- sized urban settlements of up to one million people high 2 train +2859 AR6_WGII 77 1 These could amplify pre-existing stresses related to poverty, informality, exclusion and governance, such as in African cities high 2 train +2860 AR6_WGII 77 2 Climate change increases heat stress risks in cities (high confidence) and amplifies the urban heat island across Asian cities at 1.5°C and 2°C warming levels, both substantially larger than under present climates medium 1 train +2861 AR6_WGII 77 3 Urban population exposure to extreme heat in Africa is projected to increase from 2 billion person-days per year in 1985–2005 to 45 billion person-days by the 2060s (1.7°C global warming with low population growth) and to 95 billion person- days (2.8°C global warming with medium-high population growth) medium 1 train +2862 AR6_WGII 77 4 Risks driven by flooding and droughts will also increase in cities high 2 train +2863 AR6_WGII 77 5 Urban populations exposed to severe droughts in West Africa will increase (65.3±34.1 million) at 1.5°C warming and increase further at 2°C medium 1 train +2864 AR6_WGII 77 7 Higher risks from temperature and precipitation extremes are projected for almost all Asian cities under RCP8.5 medium 1 train +2865 AR6_WGII 77 9 Unplanned rapid urbanisation is a major driver of risk, particularly where increasing climate-driven risks affect key infrastructure and potentially result in compounding and cascading risks as cities expand into coastal and mountain regions prone to flooding or landslides that disrupt transportation networks, or where water and energy resources are inadequate to meet the needs of growing settlements high 2 train +2866 AR6_WGII 77 10 These infrastructure risks expand beyond city boundaries; climate-related transport and energy infrastructure damage is projected to be a significant financial burden for African countries, reaching tens to hundreds of billions of US dollars under moderate and high emissions scenarios high 2 train +2867 AR6_WGII 77 11 Projected changes in both the hydrological cycle and the cryosphere will threaten urban water infrastructure and resource management in most regions very high 3 train +2868 AR6_WGII 77 12 South and Southeast Asian coastal cities can experience significant increases in average annual economic losses between 2005 and 2050 due to flooding, with very high losses in east Asian cities under RCP8.5 high 2 train +2869 AR6_WGII 77 14 In small islands, degraded terrestrial ecosystems decrease resource provision (e.g., potable water) and amplify the vulnerability of island inhabitants high 2 train +2870 AR6_WGII 77 15 Projections suggest that 350 million (± 158.8 million) more people in urban areas will be exposed to water scarcity from severe droughts at 1.5°C warming and 410.7 million (± 213.5) at 2°C warming low 0 train +2871 AR6_WGII 77 18 Climate change risks, including sea level rise, interact in intricate ways with non-climatic drivers of coastal change, such as land subsidence, continued infrastructure development in coastal floodplains, the rise of asset values and landward development adversely impacting coastal ecosystems, to shape future risk in coastal settlements high 2 train +2872 AR6_WGII 77 20 Some recent estimates of projected global economic damage from climate impacts are higher than previous estimates and generally increase with global average temperature high 2 train +2873 AR6_WGII 77 21 However, the spread in the estimates of the magnitude of this damage is substantial and does not allow for robust range to be established high 2 train +2874 AR6_WGII 77 22 Non-market, non-economic damage and adverse impacts on livelihoods will be concentrated in regions and populations that are already more vulnerable high 2 train +2875 AR6_WGII 77 23 Socioeconomic drivers and more inclusive development will largely determine the extent of this damage high 2 train +2876 AR6_WGII 77 26 In addition to market damage and disaster management costs, substantial costs of climate inaction are projected for human health high 2 train +2877 AR6_WGII 77 27 At higher levels of warming, climate impacts will pose risks to financial and insurance markets, especially if climate risks are incompletely internalised (medium confidence), with adverse implications for the stability of markets low 0 train +2878 AR6_WGII 78 2 Much smaller effects are estimated for less warming, lower vulnerability and more adaptation medium 1 train +2879 AR6_WGII 78 3 Regional estimates of GDP damage vary high 2 train +2880 AR6_WGII 78 4 Severe risks are more likely in (typically hotter) developing countries medium 1 train +2881 AR6_WGII 78 5 For Africa, GDP damage is projected to be negative across models and approaches high 2 train +2882 AR6_WGII 78 7 If future climate change under high emissions scenarios continues and increases risks, without strong adaptation measures, losses and damage will likely be concentrated among the poorest vulnerable populations high 2 train +2883 AR6_WGII 78 9 Higher growth scenarios along higher warming levels increase exposure to hazards and assets at risk, such as sea level rise for coastal regions, which will have large implications for economic activities, including shipping and ports high 2 train +2884 AR6_WGII 78 10 The high sensitivity of developing economies to climate impacts will pose increasing challenges to economic growth and performance, although projections depend as much or more on future socioeconomic development pathways and mitigation policies as on warming levels medium 1 train +2885 AR6_WGII 78 12 This wide range of effects underscore the impact of climate change on welfare and the adverse effects on vulnerable populations medium 1 train +2886 AR6_WGII 78 15 They exacerbate existing stressors and constrain adaptation options medium 1 train +2887 AR6_WGII 78 17 Some compound and cascading impacts occur locally, some spread across sectors and socioeconomic and natural systems, while others can be driven by events in other regions, for instance through trade and flows of commodities and goods through supply chain linkages high 2 train +2888 AR6_WGII 78 19 For example, cascading effects on food webs have been reported in the Baltic due to detrimental oxygen levels high 2 train +2889 AR6_WGII 78 21 Compound risks to health and food systems (especially in tropical regions) are projected from simultaneous reductions in food production across crops, livestock and fisheries (high confidence), heat-related loss of labour productivity in agriculture (high confidence), increased heat-related mortality (high confidence), contamination of seafood (high confidence), malnutrition (high confidence) and flooding from sea level rise high 2 train +2890 AR6_WGII 78 22 Malnourished populations will increase through direct impacts on food production with cascading impacts on food prices and household incomes, reducing access to safe and nutritious food high 2 train +2891 AR6_WGII 78 23 Food safety will be undermined from increased food contamination for seafood with marine toxins from harmful algal blooms and chemical contaminants, worsening health risks high 2 train +2892 AR6_WGII 78 25 Extreme weather events result in cascading and compounding risks that affect health and are expected to increase with warming very high 3 train +2893 AR6_WGII 78 26 Compound climate hazards can overwhelm adaptive capacity and substantially increase damage high 2 train +2894 AR6_WGII 79 2 Fewer habitats, less biodiversity, lower coastal protection (medium confidence) and decreased food and water security will result (medium confidence), reducing the habitability of some small islands high 2 train +2895 AR6_WGII 79 6 Losses become systemic when they affect entire systems and can even jump from one system to another (e.g., drought impacting rural food production contributing to urban food insecurity) medium 1 train +2896 AR6_WGII 79 8 Flows of commodities and goods, as well as people, finance and innovation, can be driven or disrupted by distant climate change impacts on rural populations, transport networks and commodity speculation high 2 train +2897 AR6_WGII 79 9 For example, Europe faces climate risks from outside the area due to global supply chain positioning and shared resources high 2 train +2898 AR6_WGII 79 10 Climate risks in Europe also impact finance, food production and marine resources beyond Europe medium 1 train +2899 AR6_WGII 79 12 Impacts and risks include reduced access to and productivity of future fisheries, regional and global food and nutritional security (high confidence), local livelihoods, health and well-being (high confidence) and loss to sociocultural assets, including heritage sites in all Arctic regions very high 3 train +2900 AR6_WGII 79 14 Deforestation, fires and urbanisation have increased the exposure of Indigenous Peoples to respiratory problems, air pollution and diseases high 2 train +2901 AR6_WGII 79 17 Regions characterised by compound challenges of high levels of poverty, a significant number of people without access to basic services, such as water and sanitation and wealth and gender inequalities, and governance challenges are among the most vulnerable regions and are particularly located in East, Central and West Africa, South Asia, Micronesia and Melanesia and in Central America high 2 train +2902 AR6_WGII 79 19 Solar radiation modification (SRM) approaches attempt to offset warming and ameliorate some climate risks but introduce a range of new risks to people and ecosystems, which are not well understood high 2 train +2903 AR6_WGII 79 22 Transitions from high to very high risk emerge in all five RFCs, compared to just two RFCs in AR5 high 2 train +2904 AR6_WGII 80 1 Remaining below 2°C warming (but above 1.5°C) would imply that risk for RFC3 through RFC5 would be transitioning to high, and risk for RFC1 and RFC2 would be transitioning to very high high 2 train +2905 AR6_WGII 80 2 By 2.5°C warming, RFC1 will be at very high risk high 2 train +2906 AR6_WGII 80 5 Once such risks materialise, the impacts would persist even if global temperatures subsequently declined to levels associated with lower levels of risk in an ‘overshooting’ scenario, for example where temperatures increase over ‘well below 2°C above pre-industrial’ for multi-decadal time spans before decreasing high 2 train +2907 AR6_WGII 80 8 Even if the Paris temperature goal is still reached by 2100, this ‘overshoot’ entails severe risks and irreversible impacts on many natural and human systems (e.g., glacier melt, loss of coral reefs, loss of human life due to heat) high 2 train +2908 AR6_WGII 80 12 The exact timing and magnitude of climate–biosphere feedbacks and potential tipping points of carbon loss are characterised by large uncertainty, but studies of feedbacks indicate increased ecosystem carbon losses can cause large future temperature increases medium 1 train +2909 AR6_WGII 80 16 Even the lowest estimates of species extinctions (9% lost) are 1000 times the natural background rates medium 1 train +2910 AR6_WGII 80 19 SRM effects on climate hazards are highly dependent on deployment scenarios, and substantial residual climate change or overcompensating change would occur at regional scales and seasonal time scales high 2 train +2911 AR6_WGII 80 22 SRM would not stop CO 2 from increasing in the atmosphere or reduce resulting ocean acidification under continued anthropogenic emissions high 2 train +2912 AR6_WGII 80 26 Examples include coral reefs, the Arctic and its Indigenous Peoples, mountain glaciers and biodiversity hotspots.Coral bleaching, mass tree and animal mortalities, species extinction; decline in sea-ice dependent species, range shifts in multiple ecosystemsIn transition from moderate to high1.1°C very high 3 train +2913 AR6_WGII 82 7 Adaptation progress and gaps TS.D.1 Increasing adaptation is being observed in natural and human systems (very high confidence), yet the majority of climate risk management and adaptation currently being planned and implemented are incremental high 2 train +2914 AR6_WGII 82 8 There are gaps between current adaptation and the adaptation needed to avoid the increase of climate impacts that can be observed across sectors and regions, especially under medium and high warming levels high 2 test +2915 AR6_WGII 82 10 Growing adaptation knowledge in public and private sectors, increasing numbers of policy and legal frameworks and dedicated spending on adaptation are all clear indications that the availability of response options has expanded high 2 train +2916 AR6_WGII 82 11 However, observed adaptation in human systems across all sectors and regions is dominated by small incremental, reactive changes to usual practices often after extreme weather events, while evidence of transformative adaptation in human systems is limited high 2 train +2917 AR6_WGII 82 12 Droughts, pluvial, fluvial and coastal flooding are the most common hazards for which adaptation is being implemented, and many of these have physical, affordability and social limits high 2 train +2918 AR6_WGII 82 15 These measures can increase the resilience, productivity and sustainability of both natural and food systems under climate change high 2 train +2919 AR6_WGII 82 17 Investment in climate service provision has benefited the agricultural sector in many regions, with limited uptake of climate service information into decision- making frameworks medium 1 train +2920 AR6_WGII 82 19 There are large gaps in risk management and risk transfer in low- income contexts, and even larger gaps in conflict-affected contexts high 2 train +2921 AR6_WGII 82 20 Adaptive capacity is highly uneven across and within regions high 2 train +2922 AR6_WGII 82 21 Current adaptation efforts are not expected to meet existing goals high 2 train +2923 AR6_WGII 82 23 Many plans focus on climate risk reduction, missing opportunities to advance co- benefits of climate mitigation and sustainable development and risking compounding inequality and reduced well-being medium 1 train +2924 AR6_WGII 82 24 The largest adaptation gaps exist in projects that manage complex risks, for example in the food–energy–water–health nexus or the inter- relationships of air quality and climate risk high 2 train +2925 AR6_WGII 82 25 Most innovation in adaptation has occurred through advances in social and ecological infrastructures, including disaster risk management, social safety nets and green/blue infrastructure medium 1 train +2926 AR6_WGII 82 26 However, most financial investment continues to be directed narrowly at large-scale hard engineering projects after climate events have caused harm medium 1 train +2927 AR6_WGII 95 2 Most of the adaptation options to the key risks depend on limited water and land resources high 2 train +2928 AR6_WGII 95 3 Governance capacity, financial support and the legacy of past urban infrastructure investment constrain how cities and settlements are able to adapt high 2 train +2929 AR6_WGII 95 4 Critical urban capacity gaps include limited ability to identify social vulnerability and community strengths, the absence of integrated planning to protect communities, the lack of access to innovative funding arrangements and a limited capability to manage finance and commercial insurance medium 1 train +2930 AR6_WGII 95 6 For example, Africa faces severe climate data constraints and inequities in research funding and leadership that reduce adaptive capacity very high 3 test +2931 AR6_WGII 95 8 Annual finance flows targeting adaptation for Africa, for example, are billions of US dollars less than the lowest adaptation cost estimates for near-term climate change high 2 train +2932 AR6_WGII 95 11 Tracked private-sector finance for climate change action has grown substantially since 2015, but the proportion directed towards adaptation has remained small high 2 train +2933 AR6_WGII 95 12 Globally, private-sector financing of adaptation has been limited, especially in developing countries high 2 train +2934 AR6_WGII 95 16 The success of adaptation will depend on our understanding of which adaptation options are feasible and effective in their local context high 2 train +2935 AR6_WGII 95 18 To close the adaptation gap, political commitment, persistent and consistent action across scales of government and upfront mobilisation of human and financial capital are key high 2 train +2936 AR6_WGII 95 20 In some natural systems, hard limits have been reached (high confidence) and more will be reached beyond 1.5°C medium 1 train +2937 AR6_WGII 95 21 Surpassing such hard, evolutionary limits causes local species extinctions and displacements if suitable habitats exist high 2 train +2938 AR6_WGII 95 22 Otherwise, species’ existence is at very high risk high 2 train +2939 AR6_WGII 95 23 In human, managed and natural systems, soft limits are already being experienced high 2 train +2940 AR6_WGII 95 24 Financial constraints are key determinants of adaptation limits in human and managed systems, particularly in low-income settings (high confidence), while in natural systems key determinants for limits are inherent traits of the species or ecosystem very high 3 train +2941 AR6_WGII 95 28 Hard limits will increasingly emerge at higher levels of warming high 2 train +2942 AR6_WGII 95 30 Evidence and signals of the thresholds at which constraints result in limits is still sparse and, in human systems, are expected to remain contested even with increasing knowledge high 2 train +2943 AR6_WGII 95 32 Beginning at below 1.5°C, autonomous and evolutionary adaptation responses by more terrestrial and aquatic species and ecosystems will face hard limits, resulting in species extinctions, loss of ecosystem integrity and a resulting loss of livelihoods high 2 train +2944 AR6_WGII 95 35 Soft limits are currently being experienced in particular by individuals, households, cities and settlements along the coast and by small-scale farmers medium 1 train +2945 AR6_WGII 96 2 Hard limits beginning at 1.5°C are also projected for coastal communities reliant on nature-based coastal protection medium 1 train +2946 AR6_WGII 96 3 Adaptation to address the risks of heat stress, heat mortality and reduced capacities for outdoor work for humans face soft and hard limits across regions that become significantly more severe at 1.5°C and are particularly relevant for regions with warm climates high 2 train +2947 AR6_WGII 96 4 Beginning at 3°C, hard limits are projected for water management measures, leading to decreased water quality and availability, negative impacts on health and well-being, economic losses in water and energy-dependent sectors and potential migration of communities medium 1 train +2948 AR6_WGII 96 5 Soft and hard limits for agricultural production are related to water availability and the uptake and effectiveness of climate resilient crops, which are constrained by socioeconomic and political challenges medium 1 train +2949 AR6_WGII 96 6 In terms of settlements, limits to adaptation are often most pronounced in smaller and rapidly growing towns and cities, including those without dedicated local government medium 1 train +2950 AR6_WGII 96 7 At the same time, legacy infrastructure in large and mega cities, designed without taking climate change risk into account, constrains innovation, leading to stranded assets and with increasing numbers of people unable to avoid harm, including heat stress and flooding, without transformative adaptation medium 1 train +2951 AR6_WGII 96 11 Information, awareness and technological constraints are also high in multiple regions high 2 train +2952 AR6_WGII 96 12 For example, awareness of anthropogenic climate change ranges between 23% and 66% of people across 33 African countries, with low climate literacy limiting potential for transformative adaptation medium 1 train +2953 AR6_WGII 96 15 The ability of actors to overcome socioeconomic constraints determines whether additional adaptation can be implemented and prevent soft limits from becoming hard limits medium 1 train +2954 AR6_WGII 96 16 Above 1.5°C of warming, limits to adaptation are reported for human and natural systems, including coral reefs high 2 train +2955 AR6_WGII 96 18 Decreasing maladaptation requires attention to justice and a shift in enabling conditions towards those that enable timely adjustments for avoiding or minimising damage and for seizing opportunities high 2 train +2956 AR6_WGII 96 20 Policy decisions that ignore the risks of adverse effects can be maladaptive by worsening the impacts of and vulnerabilities to climate change high 2 train +2957 AR6_WGII 96 21 Examples include in coastal systems (e.g., sea walls that enable further exposure through intensification of developments in low-lying coastal areas), urban areas (e.g., inflexible infrastructure in cities and settlements that cannot be adjusted easily or affordably for increased heavy rainfall), agriculture (e.g., the use of high cost irrigation in areas that are projected to have more intense drought conditions), forestry (e.g., planting of unsuitable trees species which displace Indigenous Peoples and other forest- dependent communities) and human settlements (e.g., stranded assets and stranded vulnerable communities that cannot afford to shift away or adapt and require an increase in social safety nets) high 2 train +2958 AR6_WGII 97 1 Rights-based approaches to adaptation, participatory methodologies and inclusion of local and Indigenous knowledge, combined with informed consent, deliver mechanisms to avoid these pitfalls medium 1 train +2959 AR6_WGII 97 2 Adaptation solutions benefit from engagement with Indigenous and marginalised groups, solve past equity and justice issues and offer novel approaches medium 1 train +2960 AR6_WGII 97 3 Indigenous knowledge is a powerful tool to assess interlinked ecosystem functions across terrestrial, marine and freshwater systems, bypassing siloed approaches and sectoral problems high 2 train +2961 AR6_WGII 97 4 Lastly, engagement with Indigenous knowledge and marginalised groups often offers an intergenerational context for adaptation solutions needed to avoid maladaptation high 2 train +2962 AR6_WGII 97 6 Avoiding maladaptive responses to sea level rise depends on immediate mitigation and application of adaptive planning that sets out near-term, low-regret actions while keeping open options to account for ongoing committed sea level rise very high 3 train +2963 AR6_WGII 97 7 Such forward-looking adaptive pathway planning and iterative risk management can address the current path dependencies that lead to maladaptation and can enable timely adaptation alignment with long implementation lead times, as well as addressing uncertainty about rate and magnitude of local sea level rise, and ensuring that adaptation will be more effective medium 1 train +2964 AR6_WGII 97 8 As sea level rise advances, only avoidance and relocation will eliminate coastal risks high 2 test +2965 AR6_WGII 97 9 Other measures only delay impacts for a time, increasing residual risk, perpetuating risk and creating ongoing legacy effects and inevitable property and ecosystem losses high 2 train +2966 AR6_WGII 97 10 While relocation may in the near term appear socially unacceptable, economically inefficient or technically infeasible, it may become the only feasible option as protection costs become unaffordable and technical limits are reached medium 1 train +2967 AR6_WGII 97 13 Integrated approaches, such as the water–energy–food nexus and inter-regional considerations of risks can reduce the risk of maladaptation, building on existing adaptation strategies, increasing community participation and consultation, integration of Indigenous knowledge and local 7 Ecosystem-based adaptation is defined as the use of ecosystem management activities to increase the resilience and reduce the vulnerability of people and ecosystems to climate change.knowledge, focusing on the most vulnerable small-scale producers, anticipating risks of maladaptation in decision-making for long-lived activities, including infrastructure decisions, and the impact of trade- offs and co-benefits high 2 train +2968 AR6_WGII 97 15 Better ecosystem protection and management is key to reduce the risks that climate change poses to biodiversity and ecosystem services and build resilience; it is also essential that climate change adaptation be integrated into the planning and implementation of conservation and environmental management if it is to be fully effective in future high 2 train +2969 AR6_WGII 97 16 Risks to ecosystems from climate change can be reduced by protection and restoration and also by a range of targeted actions to adapt conservation practice to climate change high 2 train +2970 AR6_WGII 98 7 Ambitious and swift global mitigation offers more adaptation options and pathways to sustain ecosystems and their services high 2 train +2971 AR6_WGII 98 8 Even under current climate change, it is necessary to take account of climate change impacts, which are already occurring or are inevitable, in environmental management to maintain biodiversity and ecosystem services high 2 train +2972 AR6_WGII 98 10 Ecosystem-based adaptation approaches are increasingly being used as part of strategies to manage flood risk, at the coast in the face of rising sea levels and inland in the context of more extreme rainfall events high 2 train +2973 AR6_WGII 98 11 Flood-risk measures that work with nature by allowing flooding within coastal and wetland ecosystems and support sediment accretion can reduce costs and bring substantial co-benefits to ecosystems, liveability and livelihoods high 2 train +2974 AR6_WGII 98 12 In urban areas, trees and natural areas can lower temperatures by providing shade and cooling from evapotranspiration high 2 train +2975 AR6_WGII 98 13 Restoration of ecosystems in catchments can also support water supplies during periods of variable 8 Actions to protect, sustainably manage and restore natural or modified ecosystems that address societal challenges effectively and adaptively, simultaneously providing human well-being and biodiversity benefits.rainfall and maintain water quality and, combined with inclusive water regimes that overcome social inequalities, provide disaster risk reduction and sustainable development high 2 train +2976 AR6_WGII 98 15 Restoration of wetlands could support livelihoods and help sequester carbon medium 1 train +2977 AR6_WGII 98 22 Taking account of interdisciplinary scientific information, Indigenous knowledge and local knowledge and practical expertise is essential to effective ecosystem-based adaptation high 2 train +2978 AR6_WGII 98 23 There is a large risk of maladaptation where this does not happen medium 1 train +2979 AR6_WGII 98 27 Adaptation responses reduce future climate risks at 1.5°C warming, but effectiveness decreases above 2°C high 2 train +2980 AR6_WGII 98 28 Resilience is strengthened by eco- system-based adaptation (high confidence) and sustainable resource management of terrestrial and aquatic species medium 1 train +2981 AR6_WGII 101 1 Competition, trade-offs and conflict between mitigation and adaptation priorities will in- crease with climate change impacts high 2 train +2982 AR6_WGII 101 2 Integrated, multi-sectoral, inclusive and systems-oriented solutions reinforce long-term resilience (high confidence), along with supportive public policies medium 1 train +2983 AR6_WGII 101 4 Frequently documented options include rainwater harvesting, soil moisture conservation, cultivar improvements, community-based adaptation, agricultural diversification, climate services and adaptive eco-management in fisheries high 2 train +2984 AR6_WGII 101 5 Roughly 25% of assessed water-related adaptations have co-benefits, while 33% of the assessed reported current or future maladaptive outcomes high 2 train +2985 AR6_WGII 101 7 Integration of Indigenous knowledge and local knowledge increase their effectiveness high 2 train +2986 AR6_WGII 101 9 Currently known adaptation responses generally perform more effectively at 1.5°C than at 2°C or more, with increasing risks remaining after adaptation at higher warming levels high 2 train +2987 AR6_WGII 101 10 Irrigation expansion will face increasing limits due to water availability beyond 1.5°C (medium confidence), with a potential doubling of regional risks to irrigation water availability between 2°C and 4°C medium 1 train +2988 AR6_WGII 101 11 Negative risks even with adaptation will become greater beyond 2°C warming in an increasing number of regions high 2 train +2989 AR6_WGII 101 13 Options such as ecosystem approaches to fisheries, agricultural diversification, agroforestry and other ecological practices support long-term productivity and ecosystem services such as pest control, soil health, pollination and buffering of temperature extremes (high confidence), but potential and trade-offs vary by socioeconomic context, ecosystem zone, species combinations and institutional support medium 1 train +2990 AR6_WGII 101 14 Ecosystem-based approaches support food security, nutrition and livelihoods when inclusive equitable governance processes are used high 2 train +2991 AR6_WGII 101 16 Adaptation options exist to reduce the vulnerability of fisheries through better management, governance and socioeconomic dimensions (medium confidence) to eliminate overexploitation and pollution high 2 train +2992 AR6_WGII 101 17 Indigenous knowledge and local knowledge can facilitate adaptation in small-scale fisheries, especially when combined with scientific knowledge and utilised in management regimes medium 1 train +2993 AR6_WGII 101 18 Adaptive transboundary governance and ecosystem-based management, livelihood diversification, capacity development and improved knowledge-sharing will reduce conflict and promote the fair distribution of sustainably harvested wild products and revenues medium 1 train +2994 AR6_WGII 101 20 Agricultural intensification addresses short-term food security and livelihood goals but has trade-offs in equity, biodiversity and ecosystem services high 2 train +2995 AR6_WGII 101 21 Irrigation is widely used and effective for yield stability, but with several negative outcomes, including water demand (high confidence), groundwater depletion (high confidence), alteration of local to regional climates (high confidence), increasing soil salinity (medium confidence), widening inequalities and loss of rural smallholder livelihoods with weak governance medium 1 train +2996 AR6_WGII 102 1 Genetic improvements through modern biotechnology have the potential to increase climate resilience in food production systems (high confidence), but with biophysical ceilings, and technical, agroecosystem, socioeconomic and political variables strongly influence and limit the uptake of climate resilient crops, particularly for smallholders medium 1 train +2997 AR6_WGII 102 3 Large-scale land deals for climate mitigation have trade-offs with livelihoods, water and food security high 2 train +2998 AR6_WGII 102 4 Afforestation programmes without adequate safeguards adversely affect Indigenous Peoples’ rights, land tenure and adaptive capacity high 2 train +2999 AR6_WGII 102 5 Some mitigation measures, such as carbon capture and storage, bio-energy and afforestation, have a high water footprint high 2 train +3000 AR6_WGII 102 6 Increased demand for aquaculture, animal and marine foods and energy products will intensify competition and potential conflict over land and water resources, particularly in low- and medium-income countries (high confidence), with negative impacts on food security and deforestation medium 1 train +3001 AR6_WGII 102 7 Integrated, systems-oriented solutions reduce competition and trade-offs and include inclusive governance, behavioural (e.g., healthier diets with lower carbon and water footprints) and technical (e.g., novel feeds) responses high 2 train +3002 AR6_WGII 102 11 Differentiated responses based on water and food security level and climate risk increase effectiveness, such as social protection programmes for extreme events, medium-term responses such as local food procurement for school meals, community seed banks or well construction to build adaptive capacity medium 1 train +3003 AR6_WGII 102 12 Longer- term responses include strengthening ecosystem services, local and regional markets, enhanced capacity and reducing systemic gender, land tenure and other social inequalities as part of a rights-based approach medium 1 train +3004 AR6_WGII 102 15 Collective efforts across sectors, with the involvement of food producers and water users and including Indigenous knowledge and local knowledge, are a pre-condition to reaching sustainable water and food systems high 2 train +3005 AR6_WGII 102 16 Policies that support system transitions include shifting subsidies, certification, green public procurement, capacity building, payments for ecosystem services and social protection medium 1 train +3006 AR6_WGII 102 18 The concentration and interconnection of people, infrastructure and assets within and across cities and into rural areas drives the creation of risks and solutions at a global scale high 2 train +3007 AR6_WGII 102 19 Concentrated inequalities in risk are broken through prioritising affordable housing and upgrading of informal and precarious settlements, paying special attention to including marginalised groups and women high 2 train +3008 AR6_WGII 102 20 Such actions are most effective when deployed across grey/ physical infrastructure, nature-based solutions and social policy and between local and city-wide or national actions medium 1 train +3009 AR6_WGII 102 26 Moreover, an additional 2.5 billion people are projected to be living in urban areas by 2050, with up to 90% of this increase concentrated in the regions of Asia and Africa high 2 train +3010 AR6_WGII 104 2 Governance capacity, financial support and the legacy of past urban infrastructure investment constrain how cities and settlements can adapt to key climate risks medium 1 train +3011 AR6_WGII 104 4 The adaptation gap is also geographically uneven; it is highest in Africa medium 1 train +3012 AR6_WGII 104 6 At the same time, legacy infrastructure in large and mega cities, designed without taking climate change risk into account, and past adaptation decisions constrain innovation, leading to stranded assets and with increasing numbers of people unable to avoid harm, including heat stress and flooding, without transformative adaptation medium 1 train +3013 AR6_WGII 104 9 As ecosystems provide important additional benefits to human well-being and coastal livelihoods, urban adaptation strategies can be developed for settlements and nearby ecosystems; combining these with engineering solutions can extend their lifetime under high rates of sea level rise medium 1 train +3014 AR6_WGII 104 16 Where inclusive approaches to adaptation policy and action are supported, this can enable wider gains of more equitable urbanisation medium 1 train +3015 AR6_WGII 104 19 Urban adaptation measures have many opportunities to contribute to climate resilient development pathways medium 1 test +3016 AR6_WGII 104 23 Targeted development planning across the range of innovation and investment in social policy, nature-based solutions and grey/physical infrastructure can significantly increase the adaptive capacity of urban settlements and cities and their contribution to climate resilient development high 2 train +3017 AR6_WGII 105 1 City and local action can complement—and at times go further than—national and international interventions high 2 train +3018 AR6_WGII 105 2 Adaptation policy that focuses on informality and sub-serviced or inadequately serviced neighbourhoods and supports inclusive urbanisation by considering the social and economic root causes of unequal vulnerability and exposure can contribute to the broader goals of the 2030 Sustainable Development Agenda and reduce vulnerability to non-climatic risks, including pandemic risk high 2 train +3019 AR6_WGII 105 3 More comprehensive and clearly articulated global ambitions for city and community adaptation will contribute to inclusive urbanisation by addressing the root causes of social and economic inequalities that drive social exclusion and marginalisation, so that adaptation can directly support the 2030 Agenda for Sustainable Development high 2 train +3020 AR6_WGII 105 5 Adaptation pathways break adaptation planning into manageable steps based on near- term, low-regret actions and aligning adaptation choices with societal goals that account for changing risk, interests and values, uncertain futures and the long-term commitment to adapting to sea level rise high 2 train +3021 AR6_WGII 105 6 In charting adaptation pathways, reconciling divergent interests and values is a priority high 2 train +3022 AR6_WGII 105 8 Nature-based interventions, for example wetlands and salt marshes, can reduce impacts and costs while supporting biodiversity and livelihoods but have limits under high warming levels and rapid sea level rise high 2 train +3023 AR6_WGII 105 9 Ecological limits and socioeconomic, financial and governance barriers will be reached first and are determined by the type of coastline and city or settlement medium 1 train +3024 AR6_WGII 105 13 With higher warming, faster sea level rise and increasing human pressures due to coastal development, the ability to adapt decreases high 2 train +3025 AR6_WGII 105 14 Adaptation options, such as providing sufficient space for a coastal system to migrate inland, when combined with ambitious and urgent mitigation measures, can reduce impacts, but they depend on the type of coastline and patterns of coastal development high 2 train +3026 AR6_WGII 105 15 With rapid sea level rise, these options will become insufficient to limit risks for marine ecosystems and their services such as food provision, coastal protection and carbon sequestration high 2 train +3027 AR6_WGII 105 18 The options include vulnerability-reducing measures, avoidance (e.g., disincentivising developments in high-risk areas and addressing existing social vulnerabilities), hard and soft protection (e.g., sea walls, coastal wetlands), accommodation (e.g., elevating houses), advance (e.g., building up and out to sea) and staged, managed retreat (e.g., landward movement of people and development) interventions very high 3 train +3028 AR6_WGII 105 20 Local government barriers to coastal adaptation could lead to courts’ becoming de facto decision makers for local adaptation, and this could be compounded by legislative shortcomings and fragmentation, insufficient leadership, lack of coordination between governance levels and disagreement about financial responsibility high 2 train +3029 AR6_WGII 105 22 Protection has a high benefit-cost ratio during the 21st century but can become unaffordable and insufficient to reduce coastal risk (e.g., due to salinisation, drainage of rivers and excess water), reaching technical limits high 2 train +3030 AR6_WGII 105 23 Hard protection sets up lock-in of assets and people to risks and reaches limits by the end of the century or sooner, depending on the scenario, local sea level rise effects and community tolerance thresholds high 2 train +3031 AR6_WGII 105 24 Considering coastal retreat as part of the solution space could lower global adaptation costs but would result in large land losses and high levels of migration for South and Southeast Asia in particular and in relative terms, small island nations would suffer most high 2 train +3032 AR6_WGII 105 25 Solutions include disincentivising developments in high-risk areas and addressing existing social vulnerabilities now high 2 train +3033 AR6_WGII 105 27 Many drivers and root causes of coastal risk are historically and institutionally embedded very high 3 train +3034 AR6_WGII 106 2 Reconciling divergent worldviews, values and interests can unlock the productive potential of conflict for transitioning towards pathways that foster climate resilient development, generate equitable adaptation outcomes and remove governance constraints high 2 train +3035 AR6_WGII 106 3 Shared understanding and locally appropriate responses are enabled by deliberate experimentation, innovation and social learning medium 1 train +3036 AR6_WGII 106 4 External assistance and government support can enhance community capabilities to reduce coastal hazard risk high 2 train +3037 AR6_WGII 106 7 Flexible options enable responses to be adjusted as climate risk escalates and circumstances change, which may increase exposure medium 1 train +3038 AR6_WGII 106 8 Legal and financial provisions can enable managed retreat from the most at-risk locations (medium confidence) but require coordination, trust and legitimate decisions by and across policy domains and sectors (high confidence) that prioritise vulnerability, justice and equity medium 1 train +3039 AR6_WGII 106 9 Inclusive, informed and meaningful deliberation and collaborative problem- solving depend on safe arenas for engagement by all stakeholders high 2 train +3040 AR6_WGII 106 11 Building adaptive capacity through sustainable development and encouraging safe and orderly movements of people within and between states represent key adaptation responses to prevent climate-related involuntary migration high 2 train +3041 AR6_WGII 106 12 Reducing poverty, inequity and food and water insecurity and strengthening institutions in particular reduce the risk of conflict and supports climate resilient peace high 2 train +3042 AR6_WGII 106 14 The COVID-19 pandemic demon- strated the value of coordinated planning across sectors, safety nets and other capacities in societies to cope with a range of shocks and stresses and to alleviate system-wide risks to health high 2 train +3043 AR6_WGII 106 15 A significant adaptation gap exists for human health and well-being and for responses to disaster risks very high 3 train +3044 AR6_WGII 106 16 Most Nation- ally Determined Contributions to the Paris Agreement from low- and middle-income countries identify health as a priority concern very high 3 train +3045 AR6_WGII 106 17 Effective governance institutions, arrangements, funding and mandates are key for adaptation to climate-related health risks high 2 train +3046 AR6_WGII 106 19 Although some mortality and morbidity from climate change are already unavoidable, targeted adaptation and mitigation actions can reduce risks and vulnerabilities high 2 train +3047 AR6_WGII 106 20 The burden of diseases could be reduced and resilience increased through health systems, generating awareness of climate change impacts on health (medium confidence), strengthening access to water and sanitation (high confidence), integrating vector control management approaches (very high confidence), expanding existing early-warning monitoring systems (high confidence), increasing vaccine development and coverage (medium confidence), improving the heat resistance of the built environment (medium confidence) and building financial safety nets medium 1 train +3048 AR6_WGII 106 22 Such cross-sectoral solutions include improved air quality through renewable energy sources (very high confidence), active transport (e.g., walking and cycling) (high confidence) and sustainable food systems that lead to healthier diets high 2 train +3049 AR6_WGII 106 23 Heat Action Plans have strong potential to prevent mortality from extreme heat events and elevated temperature high 2 train +3050 AR6_WGII 106 24 Nature- based solutions reduce a variety of risks to both physical and mental health and well-being high 2 train +3051 AR6_WGII 106 25 For example, integrated agroecological food systems offer opportunities to improve dietary diversity while building climate-related local resilience to food insecurity high 2 train +3052 AR6_WGII 106 27 The greatest gaps between policy and action are in failures to manage adaptation of social infrastructure (e.g., community facilities, services and networks) and failure to address complex interconnected risks for example in the food–energy–water–health nexus or the inter- relationships of air quality and climate risk medium 1 train +3053 AR6_WGII 106 29 Building climate resilient health systems will require multi-sectoral, multi-system and collaborative efforts at all governance scales very high 3 train +3054 AR6_WGII 107 1 The health sectors in some countries have focused on implementing incremental changes to policies and measures to respond to impacts very high 3 train +3055 AR6_WGII 107 2 As the likelihood of dangerous risks to human health continues to increase, there is a greater need for transformational changes to health and other systems very high 3 train +3056 AR6_WGII 107 3 This highlights an urgent and immediate need to address the wider interactions between environmental change, socioeconomic development and human health and well-being high 2 train +3057 AR6_WGII 107 5 Financial support for health adaptation is currently less than 0.5% of overall dispersed multilateral climate finance projects high 2 train +3058 AR6_WGII 107 6 This level of investment is insufficient to protect human health and health systems from most climate-sensitive health risks very high 3 train +3059 AR6_WGII 107 7 Adaptation financing often does not reach places where the climate sensitivity of the health sector is greatest high 2 train +3060 AR6_WGII 107 9 Properly support- ed and where levels of agency and assets are high, migration as an adaptation to climate change can reduce exposure and socioeconomic vulnerability medium 1 train +3061 AR6_WGII 107 10 However, migration becomes a risk when climate hazards cause an individual, household or community to move involuntarily or with low agency high 2 train +3062 AR6_WGII 107 11 Inability to migrate (i.e., involuntary immobility) in the face of climate hazards is also a potential risk to exposed populations medium 1 train +3063 AR6_WGII 107 12 Broad-based institutional and cross-sectoral efforts to build adaptive capacity, including meeting the SDGs, reduce future risks of climate- related involuntary displacement and immobility (medium confidence), while policies such as the Global Compact on Safe, Orderly and Reg- ular Migration medium 1 train +3064 AR6_WGII 107 14 Residents of small island states do not view relocation as an appropriate or desirable means of adapting to the impacts of climate change high 2 train +3065 AR6_WGII 107 15 Previous disaster- and development- related relocation has been expensive and contentious, posed multiple challenges for governments and amplified existing ones and generated new vulnerabilities for the people involved high 2 train +3066 AR6_WGII 107 16 In locations where permanent, government-assisted relocation becomes unavoidable, active involvement of local populations in planning and decision-making may lead to more successful outcomes medium 1 train +3067 AR6_WGII 107 18 By addressing vulner - ability, improving livelihoods and strengthening institutions, meeting the SDGs reduces the risks of armed conflict and violence medium 1 train +3068 AR6_WGII 107 19 Formal institutional arrangements for natural resource management and environmental peacebuilding, conflict-sensitive adaptation and climate-sensitive peacebuilding and gender-sensitive approaches offer potential new avenues to build peace in conflict- prone regions vulnerable to climate change medium 1 train +3069 AR6_WGII 107 22 Concepts of justice, consent and rights-based deci- sion-making, together with societal measures of well-being, are increasingly used to legitimate adaptation actions and evaluate the impacts on individuals and ecosystems, diverse communities and across generations medium 1 train +3070 AR6_WGII 107 23 Applying these principles as part of monitoring and evaluating the outcomes of adaptation, particularly during system transitions, provide a basis for ensuring that the distribution of benefits and costs are identified medium 1 train +3071 AR6_WGII 107 25 Adaptation and mitigation approaches that exacerbate inequitable access to resources and fail to address injustice increase suffering, including water and food insecurity and malnutrition rates for vulnerable groups that rely directly or indirectly on natural resources for their livelihoods high 2 test +3072 AR6_WGII 107 28 Insurance solutions are difficult for low- income groups to access medium 1 train +3073 AR6_WGII 107 29 Formal insurance policies come with risks when implemented in a stand-alone manner, including risks of maladaptation medium 1 train +3074 AR6_WGII 108 3 Understanding the positive and negative links of adaptation actions with gender equality goals (i.e., SDG 5) is important to ensure that adaptive actions do not exacerbate existing gender-based and other social inequalities high 2 train +3075 AR6_WGII 108 4 Climate literacy varies across diverse communities, compounding vulnerability {2.6.3, 2.6.7, 4.3, 4.6, 4.6.9, 5.12.5, 5.14, 6.4.4, Box 6.1, 9.4.5, Box 9.1, 12.5.8, 16.1.4, CCB GENDER} TS.D.9.4 Empowering marginalised communities in the co-pro- duction of policy at all scales of decision-making advances equi- table adaptation efforts and reduces the risks of maladaptation high 2 train +3076 AR6_WGII 108 5 Recognising Indigenous rights and local knowledge in the design and implementation of climate change responses contrib- utes to equitable adaptation outcomes high 2 train +3077 AR6_WGII 108 6 Indigenous knowledge and local knowledge play an important role in finding solu- tions and often creates critical linkages between cultures, policy frame- works, economic systems and natural resource management medium 1 train +3078 AR6_WGII 108 7 Intergenerational approaches to future climate planning and policy will become increasingly important in relation to the manage- ment, use and valuation of social-ecological systems high 2 train +3079 AR6_WGII 108 8 Many regions benefit from the significant diversity of local knowledge and systems of production, informed by long-standing experience with natural variability, providing a rich foundation for adaptation actions ef- fective at local scales high 2 train +3080 AR6_WGII 108 10 The greatest gains are achieved by prioritising investment to reduce climate risk for low-income and marginalised residents, particularly in informal settlements and rural communities high 2 train +3081 AR6_WGII 108 12 Legislative frameworks will assist business and insurance sector investment in key infrastructure to drive adaptive action at scale for equitable outcomes medium 1 train +3082 AR6_WGII 108 15 There are gender differences in climate literacy in many regions exacerbating vulnerability in agricultural contexts in access to resources and opportunities for climate resilient crops (high confidence) {3.6.4, 4.6.5, 4.8.5, 5.4.4, 5.13.4, Table 5.6, 6.3.6, 9.4.2, 9.4.5, Box 9.2, CCB FEASIB, CCB MOVING PLATE}TS.D.9.7 Local leadership, especially among women and youth, can advance equity within and between generations medium 1 train +3083 AR6_WGII 108 18 Climate justice initiatives that explicitly address multi- dimensional inequalities as part of a climate change adaptation strategy can reduce inequities in access to resources, assets and services as well as participation in decision-making and leadership, and are essential to achieving gender and climate justice high 2 train +3084 AR6_WGII 108 20 Various tools, measures and processes are available that can enable, accelerate and sustain adaptation implementation (high confidence), in particular when anticipating climate change impacts, and empower inclusive decision-making and action when they are supported by adaptation finance and leadership across all sectors and groups in society high 2 train +3085 AR6_WGII 108 22 Breaking adaptation down into manageable steps over time, while acknowledging potential long-term adaptation needs and options, can increase the prospect that effective adaptation plans will be actioned in timely and effective ways by stakeholders, sectors and institutions high 2 train +3086 AR6_WGII 108 25 Opportunities exist to integrate adaptation into institutionalised decision cycles (e.g., budget reforms, statutory monitoring and evaluation, election cycles) and during windows of opportunity (e.g., recovery after disastrous events, designing new or replacing existing critical infrastructure or developing COVID recovery projects) high 2 train +3087 AR6_WGII 108 26 Appraisal of adaptation options for policy and implementation that considers the risks of adverse effects can help prevent maladaptive adaptation and take advantage of possible co-benefits medium 1 train +3088 AR6_WGII 108 27 Instruments such as behavioural nudges, re-directing subsidies and taxes and the regulation of marketing and insurance schemes have proven useful to strengthening societal responses beyond governmental actors medium 1 train +3089 AR6_WGII 109 8 Integrated adaptation frameworks and decision-support tools that anticipate multi-dimensional risks and accommodate community values are more effective than those with a narrow focus on single risks medium 1 train +3090 AR6_WGII 109 9 Approaches that integrate the adaptation needs of multiple sectors such as disaster management, account for different risk perceptions and integrate multiple knowledge systems are better suited to addressing key risks medium 1 train +3091 AR6_WGII 109 10 Reliable climate services, monitoring and early warning systems are the most commonly used strategies for managing the key risks, complementing long-term investments in risk reduction high 2 train +3092 AR6_WGII 109 13 Integrated pathways for managing climate risks will be most suitable when so- called ‘low-regret’ anticipatory options are established jointly across sectors in a timely manner and are feasible and effective in their local context, when path dependencies are avoided so as not to limit future options for climate resilient development and when maladaptations across sectors are avoided high 2 train +3093 AR6_WGII 109 14 Integration of risks across sectors can be assisted by mainstreaming climate considerations across institutions and decision-making processes high 2 train +3094 AR6_WGII 109 19 Approaches that break down adaptation into manageable steps over time and use pathway analyses to determine low-regret actions for the near-term and long- term options are a useful starting point for adaptation medium 1 train +3095 AR6_WGII 109 21 Considering socioeconomic developments and climatic changes beyond 2100 is particularly relevant for long-lived investment decisions such as new harbours, airports, urban expansions and flood defences to avoid lock-ins medium 1 train +3096 AR6_WGII 109 22 Monitoring climate change, socioeconomic developments and progress on implementation is critical for learning about adaptation success and maladaptation and to assess whether, when and what further actions are needed for informing iterative risk management high 2 train +3097 AR6_WGII 109 26 Narratives can effectively communicate climate information and link this to societal goals and the actions needed to achieve them high 2 train +3098 AR6_WGII 109 28 Implementing actions often requires large upfront investments of human and financial resources and political capital by public, private and societal actors, while the benefits of these actions may only become visible in the mid to long term medium 1 train +3099 AR6_WGII 110 5 These options, such as disaster risk management, climate services and risk sharing, increase the feasibility and effectiveness of other options by expanding the solution space available high 2 train +3100 AR6_WGII 110 6 For example, carefully designed and implemented disaster risk management and climate services can increase the feasibility and effectiveness of adaptation responses to improve agricultural practices, income diversification, urban and critical services and infrastructure planning very high 3 train +3101 AR6_WGII 110 7 Risk insurance can be a feasible tool to adapt to transfer climate risks and support sustainable development high 2 train +3102 AR6_WGII 110 12 Distributed generation utilities, such as microgrids, are increasingly being considered, with growing evidence of their role in reducing vulnerability, especially within underserved populations high 2 train +3103 AR6_WGII 110 13 Infrastructure resilience and reliable power are particularly important in reducing risk in peri-urban and rural areas when they are supported by distributed generation of renewable energy by isolated systems high 2 train +3104 AR6_WGII 110 15 Efficient water use and water management especially in hydropower and combined cycle power plants in drought-prone areas have a high feasibility (high confidence) with multiple co-benefits medium 1 train +3105 AR6_WGII 110 16 Water-related adaptation in the energy sector is highly effective up to 1.5°C but declines with increasing warming medium 1 train +3106 AR6_WGII 110 19 Providing critical infrastructure, including through distributed generation power systems through renewable energy, has provided many co-benefits high 2 train +3107 AR6_WGII 110 21 Strengthening local and regional food systems through strategies such as collective trademarks, participatory guarantee systems and city–rural links build rural livelihoods, resilience and self- reliance medium 1 train +3108 AR6_WGII 110 24 Key barriers to livelihood diversification include sociocultural and institutional barriers as well as inadequate resources and livelihood opportunities that hinder the full adaptive possibilities of existing livelihood diversification practices high 2 train +3109 AR6_WGII 110 26 Indigenous knowledge and local knowledge, ecosystem-based adaptation and community-based adaptation are often found together in effective adaptation strategies and actions and together can generate transformative sustainable changes, but they need the resources, legal basis and an inclusive decision process to be most effective medium 1 train +3110 AR6_WGII 110 29 Risk responsibilities across the globe are unclear and unevenly defined high 2 train +3111 AR6_WGII 110 31 There are at least two contrasting approaches for pursuing deliberate transformation: one seeking rapid, system-wide change and the other a collection of incremental actions that together catalyse desired system changes medium 1 train +3112 AR6_WGII 111 2 There is only limited opportunity to widen the remaining solution space and take advantage of many potentially effective, yet unimplemented, options for reducing society and ecosystem vulnerability high 2 train +3113 AR6_WGII 111 7 With progressive climate change, enabling conditions will diminish, and opportunities for successfully transitioning systems for both mitigation and adaptation will become more limited high 2 train +3114 AR6_WGII 111 8 Investments in economic recovery from COVID-19 offer opportunities to promote climate resilient development high 2 train +3115 AR6_WGII 111 12 Collectively, these system transitions can widen the solution space and accelerate and deepen the implementation of sustainable development, adaptation and mitigation actions by equipping actors and decision makers with more effective options high 2 train +3116 AR6_WGII 111 13 For example, urban ecological infrastructure linked to an appropriate land use mix, street connectivity, open and green spaces and job-housing proximity provides adaptation and mitigation benefits that can aid urban transformation medium 1 train +3117 AR6_WGII 111 16 Such transitions can generate benefits across different sectors and regions, provided they are facilitated by appropriate enabling conditions, including ef- fective governance, policy implementation, innovation and climate and development finance, which are currently insufficient high 2 train +3118 AR6_WGII 111 28 Improving health systems through enhancing access to medical services and developing or strengthening surveillance systems can have high feasibility when there is a robust institutional and regulatory framework high 2 train +3119 AR6_WGII 111 30 Moving towards different pathways involves confronting complex synergies and trade-offs between development pathways and the options, contested values and interests that underpin climate mitigation and adaptation choices very high 3 train +3120 AR6_WGII 112 2 Economic sectors and global regions are exposed to different opportunities and challenges in facilitating climate resilient development, suggesting adaptation and mitigation options should be aligned to local and regional context and development pathways very high 3 train +3121 AR6_WGII 112 13 The current decade is critical to charting climate resilient development pathways that catalyse the transformation of prevailing development practices and offer the greatest promise and potential for human well-being and planetary health very high 3 train +3122 AR6_WGII 112 19 They often require rights-based approaches to protect the livelihoods, priorities and survival of marginalised groups including Indigenous Peoples, women, ethnic minorities and children high 2 train +3123 AR6_WGII 112 21 People who have experienced climate shocks are more likely to implement risk management measures high 2 train +3124 AR6_WGII 112 22 Autonomous adaptation is very common in locations where people are more exposed to extreme events and have the resources and the temporal capacity to act on their own, for example in remote communities high 2 train +3125 AR6_WGII 112 24 Diverse actors including youth, women, Indigenous communities and business leaders are the agents of societal changes and transformations that enable climate resilient development high 2 train +3126 AR6_WGII 112 30 Efforts are needed to change unequal power dynamics and to foster inclusive decision-making for climate adaptation to have a positive impact for gender equality high 2 train +3127 AR6_WGII 112 31 There are very few examples of successful integration of gender and other social inequities in climate policies to address climate change vulnerabilities and questions of social justice very high 3 test +3128 AR6_WGII 118 16 Integrating adaptation into social protection programmes can build long-term resilience to climate change high 2 train +3129 AR6_WGII 118 17 Nevertheless, social protection programmes can increase resilience to climate related shocks, even if they do not specifically address climate risks high 2 train +3130 AR6_WGII 118 18 Climate adaptation actions are grounded in local realities so understanding links with SDGs is important to ensure that adaptive actions do not worsen existing gender and other inequities within society, leading to maladaptation practices high 2 train +3131 AR6_WGII 118 21 Overcoming institutional and financial constraints (govern- ance, institutions, policies), including path dependency, is among the most important requirements enabling effective adaptation in the water sector high 2 train +3132 AR6_WGII 118 22 Water-related challenges, despite reported adaptation efforts, indicate limits of adaptation in the absence of water neutral mitigation action medium 1 train +3133 AR6_WGII 118 25 Reorienting existing institutions to become more flexible (e.g., through capacity building and institutional reform) and inclusive is key to building adaptive governance systems that are equipped to take long-term decisions medium 1 train +3134 AR6_WGII 118 28 Changes in lifestyles, human behaviour and preferences can have a significant impact on adaptation implementation, demand and hence emissions and decision-making around climate action high 2 train +3135 AR6_WGII 118 29 Additionally, the use of customary and traditional justice systems, such as those of Indigenous peoples, can enhance the equity of adaptation policy processes high 2 train +3136 AR6_WGII 118 31 Enabling environments share common governance characteristics, including the meaningful involvement of multiple actors and assets, alongside multiple centres of power at different levels that are well integrated, vertically and horizontally high 2 train +3137 AR6_WGII 118 32 Enabling conditions harness synergies, address moral and ethical choices and divergent values and interests and support just approaches to livelihood transitions that do not undermine human well-being medium 1 train +3138 AR6_WGII 119 1 To address regionally specific adaptation and developmental needs, five key dimensions of climate resilient development are identified for Africa: climate finance, governance, cross-sectoral and transboundary solutions, adaptation law and climate services and climate change literacy high 2 train +3139 AR6_WGII 119 5 The urgency of climate action is a potential enabler of climate decision-making medium 1 train +3140 AR6_WGII 119 6 Perceptions of urgency encourage communities, businesses and leaders to undertake climate adaptation and mitigation measures more quickly and to prioritise climate action high 2 train +3141 AR6_WGII 119 8 The integration of consideration of non-climatic drivers into adaptation pathways can reduce climate impacts across food systems, human settlements, health, water, economies and livelihoods high 2 train +3142 AR6_WGII 119 9 Strengthened health, education and basic social services are vital for improving population well-being and supporting climate resilient development high 2 train +3143 AR6_WGII 119 10 The use of climate-smart agriculture technologies that strengthen synergies among productivity and mitigation is growing as an important adaptation strategy high 2 train +3144 AR6_WGII 119 11 Pertinent information for farmers provided by climate information services is helping them to understand the role of climate compared with other drivers in perceived productivity changes medium 1 train +3145 AR6_WGII 119 12 Index insurance builds resilience and contributes to adaptation both by protecting farmers’ assets in the face of major climate shocks, by promoting access to credit and by adopting improved farm technologies and practices high 2 train +3146 AR6_WGII 119 14 Societal resilience is founded on strengthening local democracy, empowering citizens to shape societal choices to support gender and equity inclusive climate resilient development very high 3 train +3147 AR6_WGII 119 17 Adaptation responses to climate- driven impacts in mountain regions vary significantly in terms of goals and priorities, scope, depth and speed of implementation, governance and modes of decision-making and the extent of financial and other resources to implement them high 2 train +3148 AR6_WGII 119 18 Adaptation in Africa has multiple benefits, and most assessed adaptation options have medium effectiveness at reducing risks for present-day global warming, but their efficacy at future warming levels is largely unknown high 2 train +3149 AR6_WGII 119 19 In Australia and New Zealand, a range of incremental and transformative adaptation options and pathways is available as long as enablers are in place to implement them high 2 train +3150 AR6_WGII 119 20 Several enablers can be used to improve adaptation outcomes and to build resilience (high confidence), including better governance and legal reforms; improving justice, equity and gender considerations; building human resource capacity; increased finance and risk transfer mechanisms; education and awareness programmes; increased access to climate information; adequately downscaled climate data; inclusion of Indigenous knowledge; and integrating cultural resources into decision-making high 2 train +3151 AR6_WGII 119 22 One area of sustained effort is community-based adaptation planning actions that have potential to be better integrated to enhance well-being and create synergies with the SDG ambitions of leaving no one behind high 2 train +3152 AR6_WGII 119 23 Complex trade-offs and gaps in alignment between mitigation and adaptation over scale and across policy areas where sustainable development is hindered or reversed also remain medium 1 train +3153 AR6_WGII 119 24 Globally, decisions about key infrastructure systems and urban expansion drive risk creation and potential action on climate change high 2 train +3154 AR6_WGII 119 26 Indigenous Peoples have been faced with adaptation challenges for centuries and have developed strategies for resilience in changing environments that can enrich and strengthen other adaptation efforts high 2 train +3155 AR6_WGII 120 1 Indigenous knowledge underpins successful understanding of, responses to and governance of climate change risks high 2 train +3156 AR6_WGII 120 2 For example, Indigenous knowledge contains resource-use practices and ecosystem stewardship strategies that conserve and enhance both wild and domestic biodiversity, resulting in terrestrial and aquatic ecosystems and species that are often less degraded in Indigenous managed lands in other lands medium 1 test +3157 AR6_WGII 120 3 Valuing Indigenous knowledge systems is a key component of climate justice high 2 train +3158 AR6_WGII 120 6 Inclusion of interdisciplinary scientific information, Indigenous knowledge and practical expertise is essential to effective ecosystem-based adaptation (high confidence), and there is a large risk of maladaptation where this does not happen high 2 train +3159 AR6_WGII 120 9 Effective ecosystem conservation on approximately 30% to 50% of Earth’s land, freshwater and ocean areas, including all remaining areas with a high degree of naturalness and ecosystem integrity, will help protect biodiversity, build ecosystem resilience and ensure essential ecosystem services high 2 train +3160 AR6_WGII 120 12 Ecosystem services that are under threat from a combination of climate change and other anthropogenic pressures include climate change mitigation, flood-risk management and water supply high 2 train +3161 AR6_WGII 120 14 The loss of species also lowers the resilience of the ecosystem as a whole, including its capacity to persist through climate change and recover from extreme events high 2 train +3162 AR6_WGII 120 15 Species extinction levels that are more than 1000 times natural background rates as a result of anthropogenic pressures, and climate change will increasingly exacerbate this high 2 train +3163 AR6_WGII 120 16 Conservation efforts are more effective when integrated into local spatial plans inclusive of adaptation responses, alongside sustainable food and fiber production systems high 2 train +3164 AR6_WGII 120 17 Strong inclusive governance systems and participatory planning processes that support equitable and effective adaptation outcomes, are gender sensitive and reduce intergroup conflict are required for enhanced ecosystem protection and restoration high 2 train +3165 AR6_WGII 120 19 Limiting warming to 2°C and protecting 30% of high-biodiversity regions in Africa, Asia and Latin America is estimated to reduce the risk of species extinctions by half high 2 train +3166 AR6_WGII 120 20 Meeting the increasing needs of the human population for food and fibre production requires transformation in management regimes to recognise dependencies on local healthy ecosystems, with greater sustainability, including through increased use of agroecological farming approaches and adaptation to the changing climate high 2 train +3167 AR6_WGII 120 21 People with higher levels of contact with nature have been found to be significantly happier, healthier and more satisfied with their lives high 2 train +3168 AR6_WGII 120 22 Participatory, inclusive governance approaches such as adaptive co-management or community-based planning, which integrate those groups who rely on these ecosystems (e.g., Indigenous Peoples, local communities), support equitable and effective adaptation outcomes high 2 train +3169 AR6_WGII 120 24 Degradation and loss of ecosystems is a major cause of greenhouse gas emissions, which is increasingly exacerbated by climate change very high 3 train +3170 AR6_WGII 120 25 Globally, there is a 38% overlap between areas of high carbon storage and high intact biodiversity, but only 12% of that is protected high 2 train +3171 AR6_WGII 120 28 Climate resilient development will require strategies for land-based climate change mitigation to be integrated with adaptation, biodiversity and sustainable development objectives; there is good potential for positive synergies, but also the potential for conflict, including with afforestation and bioenergy crops, when these objectives are pursued in isolation high 2 train +3172 AR6_WGII 121 1 Feedback from monitoring and assessments of the changing state of planetary conditions and local ecosystems enables proactive adaptation to manage risks and minimise impacts medium 1 train +3173 AR6_WGII 121 2 Integrated sectoral approaches promoting climate resilience, particularly for addressing the impacts of extreme events, are key to effective climate resilient development medium 1 train +3174 AR6_WGII 121 5 Conservation and restoration will alone be insufficient to protect coral reefs beyond 2030 (high confidence) and to protect mangroves beyond the 2040s high 2 train +3175 AR6_WGII 121 6 Deep cuts in emissions will be necessary to minimise irreversible loss and damage high 2 train +3176 AR6_WGII 121 8 Governance for climate resilient development involves diverse societal actors, including the most vulnerable, who can work collectively, drawing upon local and Indigenous knowledges and science, and are supported by strong political will and climate change leadership medium 1 train +3177 AR6_WGII 121 9 Governance practices will work best when they are coordinated within and between multiple scales and levels (institutional, geographical and temporal) and sectors, with supporting financial resources, are tailored for local conditions, are gender-responsive and gender-inclusive and are founded upon enduring institutional and social learning capabilities to address the complexity, dynamism, uncertainty and contestation that characterise escalating climate risk medium 1 train +3178 AR6_WGII 121 11 Institutional fragmentation, under- resourcing of services, inadequate adaptation funding, uneven capability to manage uncertainties and conflicting values and reactive governance across competing policy domains collectively lock in existing exposures and vulnerabilities, creating barriers and limits to adaptation, and undermine climate resilient development prospects high 2 train +3179 AR6_WGII 121 12 This is amplified by inequity, poverty, population growth and high population density, land use change, especially deforestation, soil degradation, biodiversity loss, high dependence of national and local economies on natural resources for production of commodities, weak governance, unequal access to safe water and sanitation services and a lack of infrastructure and financing, which reduce adaptation capacity and deepen vulnerability high 2 train +3180 AR6_WGII 121 14 Collective action and strengthened networked collaboration, more inclusive governance, spatial planning and risk-sensitive infrastructure delivery will contribute to reducing risks medium 1 train +3181 AR6_WGII 121 15 Enablers for climate governance include better practices and legal reforms, improving justice, equity and gender considerations, building human resource capacity, increased finance and risk transfer mechanisms, education and climate change literacy programmes, increased access to climate information, adequately downscaled climate data and embedding Indigenous knowledge and local knowledge as well as integrating cultural resources into decision-making high 2 train +3182 AR6_WGII 121 18 Many forms of adaptation are more effective, more cost-efficient and more equitable when organised inclusively high 2 train +3183 AR6_WGII 121 19 Greater coordination and engagement across levels of government, business and community serves to move from planning to action and from reactive to proactive adaptation high 2 train +3184 AR6_WGII 121 22 City and local governments remain key actors facilitating climate change adaptation in cities and settlements medium 1 test +3185 AR6_WGII 121 23 Private and business investment in key infrastructure, housing construction and insurance can drive adaptive action at scale but can exclude the priorities of the poor medium 1 train +3186 AR6_WGII 121 24 Networked community actions can address neighbourhood-scale improvements and vulnerability at scale very high 3 test +3187 AR6_WGII 122 1 Dedicated climate change acts can play a foundational and distinctive role in supporting effective climate governance, and are drivers of subsequent activity in both developing and developed countries high 2 train +3188 AR6_WGII 122 2 The transboundary nature of many climate change risks and species responses will require transboundary solutions through multi-national or regional governance processes on land (medium confidence) and at sea high 2 train +3189 AR6_WGII 122 5 Formal institutional arrangements for natural resource management can contribute to wider cooperation and peacebuilding high 2 train +3190 AR6_WGII 122 7 Strong governance and gender-sensitive approaches to natural resource management reduce the risk of intergroup conflict in climate-disrupted areas medium 1 train +3191 AR6_WGII 122 9 National guidance and laws, policies and regulations, decision tools that can be tailored to local circumstances, innovative engagement processes and collaborative governance can motivate better understanding of climate risk and build climate resilient development high 2 train +3192 AR6_WGII 122 10 Collaborative networks and institutions, including among local communities and their governing authorities, can help resolve conflicts high 2 train +3193 AR6_WGII 122 11 A combination of robust climate information, adaptive decision-making under uncertainty, land use planning, public engagement and conflict resolution approaches can help to address governance constraints to prepare for climate risks and build adaptive capacity high 2 train +3194 AR6_WGII 122 12 New modelling, monitoring and evaluation approaches, alongside disruptive technologies, can help understand the societal implications of trade-offs and build integrated pathways of low-regret anticipatory options, established jointly across sectors in a timely manner, to avoid locked-in development pathways high 2 train +3195 AR6_WGII 122 14 Transformation towards climate resilient development is advanced most effectively when actors work in inclusive and enabling ways to reconcile divergent interests, values and worldviews, building on information and knowledge on climate risk and adaptation options derived from different knowledge systems high 2 train +3196 AR6_WGII 122 18 A deliberate shift from primarily technological adaptation strategies to those that additionally incorporate behavioural and institutional changes, adaptation finance, equity and environmental justice and that align policy with global sustainability goals will facilitate transformational adaptation high 2 train +3197 AR6_WGII 122 19 Application and efficacy testing of climate resilient development, or adaptation pathways, show promise for implementing transformational approaches medium 1 train +3198 AR6_WGII 122 20 Climate information services that are demand driven and context specific, combined with climate change literacy, have the potential to improve adaptation responses high 2 train +3199 AR6_WGII 122 24 This implies the need for wider arenas of engagement for diverse actors to collectively solve problems and to unlock the synergies between adaptation and mitigation and sustainable development high 2 train +3200 AR6_WGII 123 3 Opportunities exist to promote synergies between sustainable development, adaptation and mitigation, but trade-offs are likely unavoidable, and managing trade-offs and synergies will be important high 2 train +3201 AR6_WGII 123 4 Climate resilient development risks and opportunities vary by location with uncertainty about global mitigation effort and future climates relevant to local planning high 2 train +3202 AR6_WGII 123 9 Civic engagement is an important element of building societal consensus and reducing barriers to action on adaptation, mitigation and sustainable development very high 3 train +3203 AR6_WGII 124 9 Key risks cover scales from the local to the global, are especially prominent in particular regions or systems and are particularly large for vulnerable sub-groups, especially low-income populations, and already at-risk ecosystems high 2 train +3204 AR6_WGII 124 17 For most RKRs, potentially global and systemically pervasive risks become severe in the case of high levels of warming, combined with high exposure/vulnerability, low adaptation or both high 2 train +3205 AR6_WGII 125 3 Under these conditions there would be severe and pervasive risks associated with water scarcity and water- related disasters high 2 train +3206 AR6_WGII 125 7 Tropical and polar low-lying coastal human communities are experiencing severe impacts today high 2 train +3207 AR6_WGII 125 8 Some systems will experience severe risks before the end of the century (medium confidence), for example critical infrastructure affected by extreme events medium 1 train +3208 AR6_WGII 125 9 Food security for millions of people, particularly low-income populations, also faces significant risks with moderate to high warming or high vulnerability, with a growing challenge by 2050 in terms of providing nutritious and affordable diets high 2 train +3209 AR6_WGII 125 15 Priority areas for regions are indicated by the intersection of hazards, risks and challenges, where, in the near term, challenges to SDGs indicate probable systemic vulnerabilities and issues in responding to climatic hazards high 2 train +3210 AR6_WGII 134 2 Since IPCC AR5, human influence on the Earth’s climate has become unequivocal, increasingly apparent and widespread, reflected in both the growing scientific literature and in the perception and experiences of people worldwide high 2 train +3211 AR6_WGII 134 25 The total risk in any location may thus differ from the sum of individual risks if these interactions, as well as risks from responses themselves, are not considered high 2 train +3212 AR6_WGII 135 1 Many recent impacts are not detected, due to a shortage of monitoring and robust attribution analysis high 2 train +3213 AR6_WGII 136 1 Since AR5, a growing literature provides initial inventories of adaptation plans and implementation worldwide, but information on effectiveness remains scare high 2 train +3214 AR6_WGII 136 14 Adaptation is urgent to the extent that soft adaptation limits are currently being approached or exceeded and that achieving levels of adaptation adequate to address these soft limits requires action at a speed and scale faster than that represented by current trends high 2 train +3215 AR6_WGII 137 13 As described in this report, however, current climate policies and actions alone are not sufficient to meet stated policy goals (Section 1.1.3) high 2 train +3216 AR6_WGII 141 23 While understanding regarding the extent of adaptation gaps remains limited, the available evidence suggests significant adaptation gaps exist high 2 train +3217 AR6_WGII 153 11 The other half of the 10-year difference arises because, for central estimates of climate sensitivity, most scenarios show stronger warming over the near term than was assessed as ‘current’ in SR1.5 medium 1 train +3218 AR6_WGII 153 25 However, uncertainties in regional climate responses at a given GWL are large (Cross-Chapter Box CLIMATE in Chapter 1, Table CLIMATE.3a) and natural climate variability occurs in parallel with ongoing warming, so the potential for impacts higher than central estimates could be a more urgent consideration for risk assessments and adaptation planning than the earlier projected timing of reaching 1.5°C high 2 train +3219 AR6_WGII 163 4 Climate has always changed, often with severe effects on nature, including species loss Observations provided by the historical, archaeological, and palaeontological records, together with paleoclimatic data, demonstrate that climatic variability has high potential to affect biodiversity and human society high 2 train +3220 AR6_WGII 163 5 The evolution of the Earth’s biota has been punctuated by global biodiversity crises often triggered by rapid warming high 2 train +3221 AR6_WGII 163 9 Mass extinctions, each with greater than 70% marine species extinctions, occurred when the magnitude of temperature change exceeded 5.2°C (Song et al., 2021), albeit species extinctions occurred at lower magnitudes of warming medium 1 train +3222 AR6_WGII 163 10 Responses of biota to rapid climate change have included range shifts (very high confidence), phenotypic plasticity (high confidence), evolutionary adaptation (medium confidence), and species extinctions, including mass extinctions very high 3 test +3223 AR6_WGII 163 11 While knowledge about the relative roles of these processes in promoting survival during times of climate change is still limited (Nogués-Bravo et al., 2018), they have influenced the evolutionary trajectories of species and entire ecosystems (high confidence), and also the course of human history medium 1 train +3224 AR6_WGII 163 13 Temperature and deoxygenation were key drivers of past biotic responses in the oceans (Gibbs et al., 2016; Penn et al., 2018; Section 3.3) (high confidence), whereas on land the interplay between temperature and precipitation is less well established in ancient hyperthermals (Frank et al., 2021) medium 1 train +3225 AR6_WGII 164 1 Marine invertebrates and fishes are at greater extinction risk in response to warming than terrestrial ones because of reduced availability of thermal refugia in the sea (Pinsky et al., 2019) high 2 train +3226 AR6_WGII 164 2 Terrestrial plants showed reduced extinction during past rapid warming compared to animals high 2 train +3227 AR6_WGII 164 3 Population range shifts including migrations are common adaptations to climate changes across multiple time scales and ecological systems in the past and in response to current warming high 2 train +3228 AR6_WGII 164 5 During warming periods, diversity loss was common near the equator medium 1 train +3229 AR6_WGII 164 8 This variability may have favoured key hominin adaptations such as bipedality, increased brain size, complex sociality, and more diverse tools (Potts, 1998; Potts et al., 2020) (medium confidence), but extinctions of five species of Homo have also been attributed partly to climate change (Raia et al., 2020) low 0 train +3230 AR6_WGII 164 10 Most late Pleistocene megafaunal extinctions are attributed to direct and indirect human impacts (Sandom et al., 2014), although some were likely accelerated by climate change (Wan and Zhang, 2017; Westaway et al., 2017; Carotenuto et al., 2018; Saltré et al., 2019) low 0 train +3231 AR6_WGII 164 12 Variability in resource availability and agricultural production, entrained by climatic variability, is implicated in the disruption and decline of numerous past human societies medium 1 train +3232 AR6_WGII 164 19 This will overturn the long-lasting stability of interactions between humans and domesticated plants and animals as well as challenge the habitability for humans in several world regions (Horton et al., 2021) medium 1 train +3233 AR6_WGII 168 16 Medium confidenc e In many regions, changing preciptiation or melting snow and ice are altering hydrological systems, affecting water resources in terms of quantity and quality medium 1 train +3234 AR6_WGII 171 7 Articulating the goals of adaptation at the international, national and local levels thus requires engaging with the concepts of equity, justice and effectiveness high 2 train +3235 AR6_WGII 171 18 Without recognition, actors may not benefit from the two other aspects of justice medium 1 train +3236 AR6_WGII 181 5 Soft limits are usually associated with human systems whereas hard limits are more proximate for natural systems due to inability to adapt to biophysical changes (Chapter 16) medium 1 train +3237 AR6_WGII 182 28 Technology-led, market-led or state-led transitions aimed at meeting Paris Agreement and SDGs may fail without integrating dimensions of social justice and addressing the social and political exclusion that prevent the disadvantaged from accessing such improvements and increasing their incomes (Burkett et al., 2014; Scoones et al., 2015) medium 1 train +3238 AR6_WGII 186 1 However, incremental strategies can fail to move fast enough, can succumb to path-dependency that locks in initially helpful but long-term adverse responses (such as the well-known levee effect) (Sadoff 2015; Haasnoot 2019) or can result in a transition that meets some goals (e.g., environmental) but not others (e.g., equity) high 2 train +3239 AR6_WGII 186 7 A focus on single or overly aggregated measures (Section 1.4.1.2) and single scenarios can privilege some actors’ views over others, reduce transparency and make it more difficult to identify resilient and equitable solutions to complex, deeply uncertain, non-linear and contested problems (Schoen and and Rein, 1994; Renn, 2008; Jones et al., 2014; Lempert and Turner, 2020) medium 1 train +3240 AR6_WGII 187 8 The norms, institutions and power relationships that mediate such choices determine the extent to which the process unfolds consistent with principles of equity and social justice high 2 train +3241 AR6_WGII 211 4 The most severe impacts are occurring in the most vulnerable species and ecosystems, characterised by inherent physiological, ecological or behavioural traits that limit their abilities to adapt, as well as those most exposed to climatic hazards high 2 train +3242 AR6_WGII 211 6 Where attribution was assessed (>4,000 species globally), approximately half of the species had shifted their ranges to higher latitudes or elevations and two-thirds of spring phenological events had advanced, driven by regional climate changes very high 3 test +3243 AR6_WGII 211 7 Shifts in species ranges are altering community make-up, with exotic species exhibiting a greater ability to adapt to climate change than natives, especially in more northern latitudes, potentially leading to new invasive species medium 1 train +3244 AR6_WGII 211 8 New analyses demonstrate that prior reports underestimated impacts due to the complexity of biological responses to climate change high 2 train +3245 AR6_WGII 211 20 Changes in flow have led to reduced connectivity in rivers high 2 train +3246 AR6_WGII 211 21 Indirect changes include alterations in river morphology, substrate composition, oxygen concentrations and thermal regime in lakes very high 3 test +3247 AR6_WGII 211 23 Warming and browning (increase in organic matter) have occurred in boreal freshwaters, with both positive and negative repercussions on water temperature profiles (lower vs. upper water) (high confidence) and primary productivity (medium confidence) as well as reduced water quality high 2 train +3248 AR6_WGII 211 24 Climate change has increased wildlife diseases high 2 train +3249 AR6_WGII 211 26 Many vector-borne diseases and those caused by ticks, helminth worms and the chytrid fungus (Batrachochytrium dendrobatidis, Bd) have shifted polewards and upwards and are emerging in new regions high 2 train +3250 AR6_WGII 211 29 Local population extinctions caused by climate change have been widespread among plants and animals, detected in 47% of 976 species examined and associated with increases in the hottest yearly temperatures very high 3 test +3251 AR6_WGII 212 1 The white sub-species of the lemuroid ringtail possum (Hemibelideus lemuroides) in Queensland, Australia, disappeared after heat waves in 2005 high 2 train +3252 AR6_WGII 212 2 The Bramble Cay melomys (BC melomys, Melomys rubicola) was not seen after 2009 and was declared extinct in 2016, with sea-level rise (SLR) and increased storm surge associated with climate change being the most probable drivers high 2 train +3253 AR6_WGII 212 3 Additionally, the interaction of climate change and chytrid fungus (Bd) has driven many of the observed global declines in amphibian populations and the extinction of many species high 2 train +3254 AR6_WGII 212 4 A growing number of studies have documented genetic evolution within populations in response to recent climate change very high 3 train +3255 AR6_WGII 212 5 To date, genetic changes remain within the limits of known variation for species high 2 train +3256 AR6_WGII 212 6 Controlled selection experiments and field observations indicate that evolution would not prevent a species becoming extinct if its climate space disappears globally high 2 train +3257 AR6_WGII 212 7 Climate hazards outside of those to which species have adapted are occurring on all continents high 2 train +3258 AR6_WGII 212 8 More frequent and intense extreme events, superimposed on longer-term climate trends, have pushed sensitive species and ecosystems towards tipping points that are beyond the ecological and evolutionary capacity to adapt, causing abrupt and possibly irreversible changes medium 1 train +3259 AR6_WGII 212 10 New studies are documenting the changes that were projected in prior IPCC reports have now been observed, including upward shifts in the forest/alpine tundra ecotone, northward shifts in the deciduous/boreal forest ecotones, increased woody vegetation in the sub-Arctic tundra and shifts in the thermal habitat in lakes high 2 train +3260 AR6_WGII 212 11 A combination of changes in grazing, browsing, fire, climate and atmospheric CO 2 is leading to observed woody encroachment into grasslands and savannah, consistent with projections from process-based models driven by precipitation, atmospheric CO 2 and wildfires high 2 train +3261 AR6_WGII 212 13 Observed changes impact the structure, functioning and resilience of ecosystems as well as ecosystem services, such as climate regulation high 2 train +3262 AR6_WGII 212 14 Regional increases in the area burned by wildfire (up to double natural levels), tree mortality of up to 20%, and biome shifts of up to 20 km latitudinally and 300 m up-slope have been at- tributed to anthropogenic climate change in tropical, temper -ate and boreal ecosystems around the world high 2 train +3263 AR6_WGII 212 15 This degrades the survival of vegetation, habitat for biodiversity, water supplies, carbon sequestration, and other key aspects of the integrity of ecosystems and their ability to provide services for people high 2 train +3264 AR6_WGII 212 17 Field evidence shows that anthropogenic climate change increased area burned by wildfire above natural levels in western North America in the period 1984–2017: a doubling above natural for the western USA and 11 times higher than natural in one extreme year in British Columbia high 2 train +3265 AR6_WGII 212 19 Wildfires generate up to one-third of ecosystem carbon emissions globally, a feedback that exacerbates climate change high 2 train +3266 AR6_WGII 212 21 Increase in wildfire from the levels to which ecosystems are adapted degrades vegetation, habitat for biodiversity, water supplies and other key aspects of the integrity of ecosystems and their ability to provide services for people high 2 train +3267 AR6_WGII 212 23 It has also potentially contributed to over 100 other cases of drought-induced tree mortality across Africa, Asia, Australia, Europe, and North and South America high 2 train +3268 AR6_WGII 212 24 Field observations have documented post-mortality vegetation shifts high 2 train +3269 AR6_WGII 212 26 Increases in forest insect pests driven by climate change have contributed to tree mortality and shifts in carbon dynamics in many temperate and boreal forest areas very high 3 train +3270 AR6_WGII 212 27 The direction of changes in carbon balance and wildfires following insect outbreaks depends on the local forest insect communities medium 1 train +3271 AR6_WGII 212 29 Intact tropical rainforests, Arctic permafrost, peatlands and other healthy high-carbon ecosystems provide a vital global ecosystem service of preventing the release of stored carbon high 2 train +3272 AR6_WGII 212 30 Terrestrial ecosystems contain stocks of ~3500 GtC in vegetation, permafrost, and soils, three to five times the amount of carbon in unextracted fossil fuels (high confidence) and >4 times the carbon currently in the atmosphere high 2 train +3273 AR6_WGII 213 1 Deforestation, draining, burning or drying of peatlands, and thawing of Arctic permafrost, due to climate change, has already shifted some areas of these ecosystems from carbon sinks to carbon sources high 2 train +3274 AR6_WGII 213 8 The median values for percentage of species at very high risk of extinction (categorized as “critically endangered” by IUCN Red List categories)(IUCN, 2001) are 9% at 1.5°C rise in GSAT, 10% at 2°C, 12% at 3.0°C, 13% at 4°C and 15% at 5°C high 2 train +3275 AR6_WGII 213 10 All groups fare substantially better at lower warming of 2°C, with extinction projections reducing to <3% for all groups, except salamanders that reduced to 7% medium 1 train +3276 AR6_WGII 213 14 As species become rare, their role in the functioning of the ecosystem diminishes high 2 train +3277 AR6_WGII 213 15 Loss of species locally reduces the ability of an ecosystem to provide services and lowers its resilience to climate change high 2 train +3278 AR6_WGII 213 16 At 1.58°C GSAT warming, >10% of species are projected to become endangered (median estimate, with “endangered” equating to a high risk of extinction, sensu IUCN), and at 2.07°C this rises to >20% of species, representing a high and very high risk of biodiversity loss, respectively medium 1 train +3279 AR6_WGII 213 17 Biodiversity loss is projected for more regions with increasing warming, and will be worst in northern South America, southern Africa, most of Australia and at northern high latitudes medium 1 train +3280 AR6_WGII 213 18 Climate change increases risks of biome shifts on up to 35% of global land at ≥4°C GSAT warming, that emission reductions could limit to <15% for <2°C warming medium 1 train +3281 AR6_WGII 213 19 Under high-warming scenarios, models indicate shifts of extensive parts of the Amazon rainforest to drier and lower- biomass vegetation (medium confidence), poleward shifts of boreal forest into treeless tundra across the Arctic, and upslope shifts of montane forests into alpine grassland high 2 train +3282 AR6_WGII 213 20 Area at high risk of biome shifts from changes in climate and land use combined can double or triple compared to climate change alone medium 1 test +3283 AR6_WGII 213 21 Novel ecosystems, with no historical analogue, are expected to become increasingly common in the future medium 1 train +3284 AR6_WGII 213 23 With 4°C GSAT warming by 2100, wildfire frequency is projected to have a net increase of ~30% medium 1 train +3285 AR6_WGII 213 24 Increased wildfire, combined with soil erosion due to deforestation, could degrade water supplies medium 1 train +3286 AR6_WGII 213 25 For ecosystems with an historically low frequency of fires, a projected 4°C global temperature rise increases the risk of fires, with potential increases in tree mortality and the conversion of extensive parts of the Amazon rainforest to drier and lower-biomass vegetation medium 1 train +3287 AR6_WGII 214 3 This includes minimising additional stresses or disturbances; reducing fragmentation; increasing natural habitat extent, connectivity and heterogeneity; maintaining taxonomic, phylogenetic, and functional diversity and redundancy; and protecting small-scale refugia where micro-climate conditions can allow species to persist high 2 train +3288 AR6_WGII 214 8 There is also new evidence that species can persist locally because of plasticity including changes in phenology or behavioural changes that move an individual into cooler micro-climates, and genetic adaptation may allow species to persist for longer than might be expected from local climatic changes high 2 train +3289 AR6_WGII 214 9 There is no evidence to indicate that these mechanisms will prevent global extinctions of rare, very localised species already near their climatic limits or species inhabiting climate/habitat zones that are disappearing high 2 train +3290 AR6_WGII 214 12 Many proposed adaptation measures have not been implemented low 0 test +3291 AR6_WGII 214 15 Adaptation can reduce risks but cannot prevent all damaging impacts so is not a substitute for reductions in greenhouse gas (GHG) emissions high 2 train +3292 AR6_WGII 214 17 An increasing body of evidence demonstrates that climatic risks to people including floods, drought, fire and overheating, can be lowered by a range of EbA techniques in urban and rural areas medium 1 train +3293 AR6_WGII 214 21 Interdisciplinary scientific information and practical expertise, including Indigenous and local knowledge (IKLK), are essential to effectiveness high 2 train +3294 AR6_WGII 214 22 There is a large risk of maladaptation where this does not happen high 2 train +3295 AR6_WGII 214 27 Ecosystem services that are under threat from a combination of climate change and other anthropogenic pressures include climate change mitigation, flood risk management, food provisioning and water supply high 2 train +3296 AR6_WGII 216 13 Land use and land cover change (LULCC) as well as the unsustainable exploitation of resources in terrestrial and freshwater systems continue to be major factors contributing to the loss of natural ecosystems and biodiversity high 2 train +3297 AR6_WGII 216 15 Likewise, for biodiversity, invasive alien species have been identified as a major threat, especially in freshwater systems, on islands and in coastal regions high 2 train +3298 AR6_WGII 219 6 Non-climatic hazards such as LUC, habitat fragmentation, pollution and invasive species have been the primary drivers of change in terrestrial and freshwater ecosystems in the past high 2 train +3299 AR6_WGII 219 10 Increased temperatures and changes to rainfall and runoff patterns; greater variability in temperature, rainfall, river flow and water levels; and rising sea levels and the increased frequency of extreme events means that greater areas of the world are being exposed to climate hazards outside of those to which they are adapted high 2 train +3300 AR6_WGII 219 11 Extreme events are a natural and important part of many ecosystems, and many organisms have adapted to cope with long-term and short-term climate variability within the disturbance regime experienced during their evolutionary history high 2 train +3301 AR6_WGII 219 12 However, climate changes, disturbance regime changes and the magnitude and frequency of extreme events such as floods, droughts, cyclones, heat waves and fire have increased in many regions high 2 train +3302 AR6_WGII 219 13 These disturbances affect ecosystem functioning, biodiversity and ecosystem services high 2 train +3303 AR6_WGII 219 17 Increases in the frequency and severity of heat waves, droughts and aridity, floods, fires and extreme storms have been observed in many regions (Seneviratne et al., 2012; Ummenhofer and Meehl, 2017), and these trends are projected to continue high 2 train +3304 AR6_WGII 219 24 High climate velocity (Loarie et al., 2009) is expected to be associated with distribution shifts, incomplete range filling and species extinctions high 2 train +3305 AR6_WGII 219 25 It is generally assumed that the more rapid the rate of change, the greater the impact on species and ecosystems, but responses are taxonomically and geographically variable high 2 test +3306 AR6_WGII 219 28 The ability to track suitable climates is substantially reduced by habitat fragmentation and human modifications of the landscape such as dams on rivers and urbanisation high 2 train +3307 AR6_WGII 220 2 Many species, both terrestrial and freshwater, are not expected to be able to disperse fast enough to track suitable climates under mid- and high-emission scenarios medium 1 train +3308 AR6_WGII 225 7 Rivers already under stress from human activities such as urban development and farming on floodplains are prone to reduced resilience to future extreme events medium 1 train +3309 AR6_WGII 225 10 If instream habitat is strongly affected, recovery, if it occurs, takes much longer, resulting in a decline of biodiversity medium 1 train +3310 AR6_WGII 225 14 In summary, extreme events (heat waves, storms and loss of ice) affect lakes in terms of water temperature, water level, light, oxygen concentrations and nutrient dynamics, which, in turn, affect primary production, fish communities and GHG emissions high 2 train +3311 AR6_WGII 225 16 Droughts have a negative impact on water quality in streams and lakes by increasing water temperature, salinity, the frequency of algal blooms and contaminant concentrations, and reducing concentrations of nutrients and dissolved oxygen medium 1 train +3312 AR6_WGII 225 22 For RCP8.5, warming increases to 5.4°C and duration increases dramatically to 95.5 days medium 1 train +3313 AR6_WGII 225 24 Most prominently, monomictic lakes—undergoing one mixing event in most years—will become permanently stratified, while lakes that are currently dimictic—mix- ing twice per year—will become monomictic by 2080–2100 medium 1 train +3314 AR6_WGII 225 29 Moreover, melting of ice decreases the ratio of sensible to latent heat flux, thus channelling more energy into evaporation medium 1 train +3315 AR6_WGII 226 1 Introduction Increases in the frequency and magnitudes of extreme events, attributed to anthropogenic climate change by WGI (IPCC, 2021a), are now causing profound negative effects across all realms of the world (marine, terrestrial, freshwater and polar) medium 1 train +3316 AR6_WGII 226 2 Changes to population abundance, species distributions, local extirpations, and global extinctions are leading to long-term, potentially irreversible shifts in the composition, structure and function of natural systems medium 1 train +3317 AR6_WGII 226 5 This has collapsed the timeline that organisms and natural communities have to acclimate or adapt to climate change medium 1 train +3318 AR6_WGII 226 8 Marine Heat Waves Consensus is emerging that anthropogenic climate change has significantly increased the likelihood of recent marine heat waves (MHWs) medium 1 train +3319 AR6_WGII 226 12 Modelling suggests rapid shifts in the geographic distributions of important fish species in response to MHWs (Cheung and Frolicher, 2020), with projected decreased biomass and distributional shifts of fish at least four times faster and larger than the effects of decadal-scale mean changes throughout the 21st century under RCP8.5 high 2 train +3320 AR6_WGII 226 14 The Arctic region is warming more than twice as fast as the global mean, and polar organisms and ecosystems are likely to be particularly vulnerable to heat waves due to their specific thermal niches and physiological thresholds and also the lack of poleward ‘refugia’ high 2 train +3321 AR6_WGII 227 6 Terrestrial Heat Waves Heat waves are now regularly occurring that exceed the physiological thresholds of some species, including birds and other small endotherms such as flying foxes high 2 train +3322 AR6_WGII 230 17 Attribution is strong for species and species-interactions for which there is a robust mechanistic understanding of the role of climate on biological processes high 2 train +3323 AR6_WGII 231 9 Changes of temperature extremes are often more important to these local extinction rates than changes of mean annual temperature high 2 train +3324 AR6_WGII 231 27 Major declines have been recorded for several species, population extinctions have occured at lower elevations since the early 2000s, and the white sub-species of the lemuroid ringtail possum (Hemibelideus lemuroides) in Queensland, Australia, disappeared after heat waves in 2005 high 2 train +3325 AR6_WGII 232 1 The interaction between expansion of chytrid fungus globally and local climate change is implicated in the extinction of a wide range of tropical amphibians high 2 train +3326 AR6_WGII 232 8 In summary, local population extinctions caused by climate-change-driven increases in extreme weather and climate events have been widespread among plants and animals very high 3 train +3327 AR6_WGII 235 7 As novel climate conditions develop, novel communities made up of new combinations of species are emerging as populations and species adapt and shift their ranges differentially, not always with negative consequences high 2 train +3328 AR6_WGII 235 10 However, observations, experimental mesocosms (Bastazini et al., 2021), and theoretical models (Lurgi et al., 2012; Sentis et al., 2021) provide support that novel communities will continue to emerge with climate change medium 1 train +3329 AR6_WGII 236 21 The 15% of species that did both (shifting northward by 113.1 km and advancing their flight period by 2.7 days per decade, on average, over a 20-year period) had the largest population increases, and the 40% of species that showed no response had the largest population declines.2.4.2.6 Observed Changes to Physiology and Morphology Driven by Climate Change Impacts on species physiology in terrestrial and freshwater systems have been observed, and attributed to climate change medium 1 train +3330 AR6_WGII 236 25 Behavioural plasticity (flexibility) such as nest-site selection can provide a partial buffer from the effects of increasing temperature by placing the individual in a slightly cooler microclimate, but there are environmental and physical limits to this plasticity medium 1 train +3331 AR6_WGII 236 26 Plasticity in heat tolerance (e.g., due to reversible acclimation or acclimatisation) can also potentially compensate for rising temperatures (Angilletta Jr, 2009), but ectotherms have relatively low acclimation in thermal tolerance and acclimation is expected to only slightly reduce the risk of overheating in even the most plastic taxa low 0 train +3332 AR6_WGII 236 28 In many ectotherms, plasticity in thermal tolerance increases polewards, as thermal seasonality increases (Chown et al., 2004), contributing to higher vulnerability to warming in tropical organisms low 0 train +3333 AR6_WGII 236 30 The most heat-tolerant species, such as those from extreme environments, may therefore be at a greater risk of warming because of an inability to physiologically adjust to thermal change low 0 test +3334 AR6_WGII 236 31 Physiological changes have observable impacts on morphology, such as changes in body size (and length of appendages), and colour changes in butterflies, dragonflies and birds medium 1 train +3335 AR6_WGII 241 2 However, the lack of systematic empirical evidence in fresh waters, and confounding effects such as interactions between temperature, nutrient availability and predation, limit generalisations in attributing observed body size changes to climate change low 0 train +3336 AR6_WGII 241 3 Evidence is weak for a consistent reduction in body size across taxonomic groups in terrestrial animals low 0 train +3337 AR6_WGII 241 6 Several lines of evidence suggest the evolution of melanism in response to climate change low 0 train +3338 AR6_WGII 241 7 Such changes may represent decreased phenotypic diversity and, potentially, genetic diversity low 0 train +3339 AR6_WGII 241 10 Interactions between morphological changes and changes in phenology may facilitate or constrain adaptation to climate change medium 1 train +3340 AR6_WGII 242 3 There is increasing evidence of the role of extreme events in disease outbreaks very high 3 train +3341 AR6_WGII 242 20 These documented changes in climate, hosts and pathogens have been linked to a higher incidence and more frequent outbreaks of disease in livestock across Europe very high 3 train +3342 AR6_WGII 243 9 At least six major VBDs affected by climate drivers have recently emerged in Nepal and are now considered endemic, with climate change implicated as a primary driver as LULCC has been assessed to have a minimal influence on these diseases high 2 train +3343 AR6_WGII 243 16 In concert with these poleward shifts of hosts and vectors, pathogens, particularly tick-borne pathogens and helminth infections, have increased dramatically in incidence and severity from once-rare occurrences and have appeared in new regions very high 3 train +3344 AR6_WGII 244 5 These increases in introduction risk compounded with climate change have already begun to harm Indigenous Peoples dependent on hunting and herding livestock (horses and reindeer) that are suffering increased pathogen infection high 2 train +3345 AR6_WGII 248 10 In summary, with our present knowledge, evolution is not expected to be sufficient to prevent the extinction of whole species if a species’ climate space disappears within the region they inhabit high 2 train +3346 AR6_WGII 251 4 The global extent of grasslands is declining significantly because of climate change medium 1 train +3347 AR6_WGII 254 16 In summary, new studies since AR5 have explicitly estimated the effects of warming and browning on freshwaters in boreal areas, with complex positive and negative repercussions on water temperature profiles (lower vs. upper water) (high confidence) and primary production medium 1 train +3348 AR6_WGII 255 4 In summary, field evidence shows that anthropogenic climate change has increased the area burned by wildfire above natural levels across western North America in the period 1984–2017, at GMST increases of 0.6°C–0.9°C, increasing burned area up to 11 times in one extreme year and doubling it (over natural levels) in a 32-year period high 2 train +3349 AR6_WGII 255 18 Overall, human land use exerts an influence on wildfire trends for global terrestrial area as a whole that can be stronger than climate change medium 1 train +3350 AR6_WGII 257 2 Deforestation, peat draining, agricultural expansion or abandonment, fire suppression and inter-decadal cycles such as the ENSO exert a stronger influence than climate change on wildfire trends in numerous regions outside of North America high 2 train +3351 AR6_WGII 259 5 Nevertheless, the evidence from the forests of western North America shows that human-caused climate change has, at least on one continent, clearly driven increases in wildfire.Box FAQ 2.3 (continued) In summary, anthropogenic climate change caused drought-induced tree mortality of up to 20% in the period 1945–2007 in western North America, the African Sahel and North Africa, via global temperature increases of 0.3°C–0.9°C above the pre-industrial period and increases in aridity, and it contributed to over 100 other cases of drought-induced tree mortality in Africa, Asia, Australia, Europe and North and South America high 2 train +3352 AR6_WGII 259 6 Field observations document accelerating mortality rates, rising background mortality and post-mortality vegetation shifts high 2 train +3353 AR6_WGII 260 12 In summary, climate change has contributed to tree mortality in Europe high 2 train +3354 AR6_WGII 260 17 Mammals that use trees as refugia showed declines with tree mortality high 2 train +3355 AR6_WGII 260 19 Ground-nesting, ground-foraging, tree-hole nesting and bark-foraging birds increased most, but nectar-feeding and foliage-gleaning birds declined high 2 train +3356 AR6_WGII 260 20 Within invertebrates, declines were strongest in ground-foraging predators and detritivores medium 1 train +3357 AR6_WGII 260 31 Thus, tropical rainforests, Arctic permafrost and other ecosystems provide the global ecosystem service of naturally preventing carbon from contributing to climate change high 2 train +3358 AR6_WGII 261 10 These ecosystems form natural sinks that prevent the emission to the atmosphere of 1400– 1800 GtC that would otherwise increase the magnitude of climate change high 2 train +3359 AR6_WGII 261 21 Conservation of high biodiversity areas, particularly in protected areas, protects ecosystem carbon, prevents emissions to the atmosphere and reduces the magnitude of climate change high 2 train +3360 AR6_WGII 262 7 These factors combined—recent impacts of climate change on undisturbed forest, coupled with deforestation and agricultural expansion, along with associated intentional burning—have caused Amazon rainforest to become an overall net carbon emitter medium 1 train +3361 AR6_WGII 262 15 Tropical deforestation, the draining and burning of peatlands and the thawing of Arctic permafrost due to climate change have caused these ecosystems to emit more carbon to the atmosphere than they naturally remove through vegetation growth high 2 train +3362 AR6_WGII 263 5 In large, nutrient-poor lakes, warming-induced prolonged thermal stratification can reduce primary production medium 1 train +3363 AR6_WGII 263 9 However, increases or declines of algae cannot entirely be attributed to climate change; they are lake-specific and modulated through weather conditions, lake morphology, salinity, land use and restoration and biotic interactions medium 1 train +3364 AR6_WGII 263 11 Global and regional meta-analyses of diverse systems, habitats and taxonomic groupings document that approximately half of all species with long-term records have shifted their ranges poleward and/or upward in elevation and ~2/3 have advanced their timing of spring events (phenology) very high 3 train +3365 AR6_WGII 263 12 Changes in abundance tend to match predictions from climate warming, with warm-adapted species significantly outperforming cold-adapted species in warming habitats (Feeley et al., 2020) and the composition of local communities becoming more ‘thermophilised’, that is, experiencing an ‘increase in relative abundance of heat-loving or heat-tolerant species’ high 2 train +3366 AR6_WGII 263 13 New studies since AR5, with more sophisticated analyses designed to capture complex responses, indicate that past estimates of the proportion of species impacted by recent climate change were underestimates due to unspoken assumptions that local or regional warming should lead solely to poleward/upward range shifts and advancements of spring timing high 2 train +3367 AR6_WGII 263 14 More complex analyses have documented cases of winter warming driving delayed spring timing of northern temperate species due to chilling requirements, and increased precipitation driving species’ range shifts downslope in elevation, and eastward and westward in arid regions high 2 train +3368 AR6_WGII 263 15 Further new studies have shown that phenological changes have, in some cases, successfully compensated for local climate change and reduced the extent of range shifts medium 1 train +3369 AR6_WGII 263 17 Responses in freshwater species are consistent with responses in terrestrial species, including poleward and upward range shifts, earlier timing of spring plankton development, earlier spawning by fish and the extension of the growing season high 2 train +3370 AR6_WGII 263 18 Observed changes in freshwater species are strongly related to anthropogenic climate change-driven changes in the physical environment (e.g., increased water temperature, reduced ice cover, reduced mixing in lakes, loss of oxygen and reduced river connectivity) high 2 train +3371 AR6_WGII 263 19 While evidence is robust for an increase in primary production in nutrient rich lakes along with warming trends high 2 train +3372 AR6_WGII 263 21 This has caused waters to become brown, resulting in an acceleration of upper-water warming and an overall cooling of deep water high 2 train +3373 AR6_WGII 263 22 Browning may accelerate primary production through the input of nutrients associated with DOM in nutrient-poor lakes and increases the growth of cyanobacteria, which cope better with low light intensity medium 1 train +3374 AR6_WGII 263 23 Field research since the AR5 has detected biome shifts at numerous sites, poleward and upslope, that are consistent with increased temperatures and altered precipitation patterns driven by climate change, and support prior studies that attributed such shifts to anthropogenic climate change high 2 train +3375 AR6_WGII 263 25 Globally, woody encroachment into open areas (grasslands, arid regions and tundra) is likely being driven by climate change and increased CO 2, in concert with changes in grazing and fire regimes medium 1 train +3376 AR6_WGII 264 1 Analyses of causal factors have attributed increasing tree mortality at sites in Africa and North America to anthropogenic climate change, and field evidence has detected tree mortality due to drought, wildfires and insect pests in temperate and tropical forests around the world high 2 train +3377 AR6_WGII 264 2 Water stress, leading to plant hydraulic failure, is a principal mechanism of drought-induced tree mortality, along with the indirect effects of climate change mediated by community interactions high 2 train +3378 AR6_WGII 264 3 Terrestrial ecosystems sequester and store globally critical stocks of carbon, but these stocks are at risk from deforestation and climate change high 2 train +3379 AR6_WGII 264 5 In the Arctic, increased temperatures have thawed permafrost at numerous sites, dried some areas and increased fires, causing net emissions of carbon from soils high 2 train +3380 AR6_WGII 264 6 Globally, increases in temperature, aridity and drought have increased the length of fire seasons and doubled the potentially burnable land area medium 1 train +3381 AR6_WGII 264 7 Increases in the area burned have been attributed to anthropogenic climate change in North America high 2 test +3382 AR6_WGII 264 11 The changes in biodiversity and ecosystem health that we have observed, and project will continue, pose a risk of declines in human health and well-being (e.g., tourism, recreation, food, livelihoods and quality of life) medium 1 train +3383 AR6_WGII 268 28 AR5 stated: ‘a large fraction of terrestrial and freshwater species face increased extinction risk under projected climate change during and beyond the 21st century, especially as climate change interacts with other pressures high 2 train +3384 AR6_WGII 270 5 Species’ losses are projected to be worst in northern South America, southern Africa, most of Australia and at northern high latitudes medium 1 train +3385 AR6_WGII 270 8 Ten-twenty percent losses represent high and very high risk of biodiversity losses, respectively, substantial enough to reduce ecosystem integrity and functioning medium 1 train +3386 AR6_WGII 270 14 The roughly equivalent estimate of this risk as expressed in AR4 (Fischlin et al., 2007) is indicated by the dotted block indicating the medium confidence statement ‘Approximately 20–30% of plant and animal species assessed so far (in an unbiased sample) are likely to be at increasingly high risk of extinction as global mean temperatures exceed a warming of 2–3°C above pre-industrial levels medium 1 train +3387 AR6_WGII 270 28 All groups fare substantially better at 2°C, with extinction projections reducing to <3% for all groups, except salamanders at 7% medium 1 train +3388 AR6_WGII 272 19 In sum, climate change is expected to expand and redistribute the burden of vector-borne and other environmentally transmitted diseases of wild animals, domesticated animals and humans, by shifting many regions toward the thermal optima of VBD transmission for multiple parasites, thereby increasing risk of transmission, while pushing temperatures above optimal and towards upper thermal limits for other vectors and pathogens, thus decreasing their transmission high 2 train +3389 AR6_WGII 274 2 Large discrepancies exist between models and between scenarios regarding the region and the speed of change (Gonzalez et al., 2010; Pereira et al., 2010; Pecl et al., 2017), but robust understanding is emerging in that the degree of impact increases in high-emission and high-warming scenarios high 2 train +3390 AR6_WGII 274 6 In particular, in cold (boreal and tundra) regions, as well as in dry regions high 2 train +3391 AR6_WGII 274 11 However, ‘novel ecosystems’, that is, communities with no current or historical equivalent because of the novel combinations of abiotic conditions under climate change, are expected to be increasingly common in the future medium 1 train +3392 AR6_WGII 274 12 The possibility of these novel ecosystems and the communities that live within them are a challenge for current modelling of ecosystem shifts, and new approaches to conservation will be required that are designed to adapt to rapid changes in species composition and the ensuing challenges.2.5.2.3 Risk to Arid Regions Shifts in arid system structure and functioning that have been observed to date (Section 2.4.3.3) are projected to continue medium 1 train +3393 AR6_WGII 275 16 Similarly, much of the MTEs are open shrublands and grasslands and proposed extensive tree-planting to sequester atmospheric CO 2 could result in a loss of biodiversity and threaten water security (Doblas-Miranda et al., 2017; Bond et al., 2019).2.5.2.5 Risk to Grasslands and Savannas Worldwide, woody cover is increasing in savannas (Buitenwerf et al., 2012; Donohue et al., 2013; Stevens et al., 2017), as a result of interactions of elevated CO 2 and altered fire and herbivory impacts, some of which stems from LULCC high 2 train +3394 AR6_WGII 275 19 Increases in woody vegetation in grassy ecosystems could provide some carbon increase (medium confidence) (Zhou et al., 2017; Mureva et al., 2018), but is expected to decrease biodiversity (Smit and Prins, 2015; Abreu et al., 2017; Andersen and Steidl, 2019) and water availability (Honda and Durigan, 2016; Stafford et al., 2017) and alter ecosystem services like grazing and wood provision high 2 train +3395 AR6_WGII 275 21 It has been shown that simulation studies that do not account for CO 2 interactions but only consider climate change impacts do not realistically capture the future distribution of savannas high 2 train +3396 AR6_WGII 275 25 Future fire-spread is expected to be reduced with increased woody dominance (Scheiter et al., 2015; Knorr et al., 2016b; Scheiter et al., 2020), feeding back to further increase tree-to-grass ratios high 2 train +3397 AR6_WGII 276 27 As a consequence, boreal tree species are expected to move northwards (or in mountainous regions, upwards) into regions dominated by tundra, unless constrained by edaphic features, and temperate species are projected to grow in regions currently occupied by southern boreal forest high 2 train +3398 AR6_WGII 278 2 SR1.5 classified tundra and boreal forests as particularly vulnerable to degradation and encroachment by woody shrubs high 2 train +3399 AR6_WGII 278 3 The SROCC projected climate-related changes to arctic hydrology, wildfires and abrupt thaw (high confidence) and the broad disappearance of arctic near-surface permafrost this century, with important consequences for global climate very high 3 train +3400 AR6_WGII 278 11 Models of vegetation response to climate project acceleration in the coming decades of observed increases in shrub dominance and boreal forest encroachment that have been driven by recent warming (Settele et al., 2014), leading to a shrinking of the area of tundra globally medium 1 train +3401 AR6_WGII 278 21 Therefore, the trends of changing vegetation cover identified in simulations of transient warming continue to show up in simulations that hold climate change at low levels of warming medium 1 train +3402 AR6_WGII 281 24 Projected climate change could expose an extensive part of the global protected area to disappearing and novel climate conditions high 2 train +3403 AR6_WGII 282 2 Projected disappearance of suitable climate conditions in protected areas increase risks to the survival of species and vegetation types of conservation concern in tropical, temperate and boreal ecosystems high 2 train +3404 AR6_WGII 282 11 Protected areas conserve refugia from climate change under a temperature increase of 4°C, which is important for biodiversity conservation but is limited to <10% of the current protected area medium 1 train +3405 AR6_WGII 283 17 In summary, under a high-emission scenario that increases global temperature 4°C by 2100, climate change could increase the global burned area by 50–70% and the global mean fire frequency by ~30%, with increases on one- to two-thirds and decreases on one-fifth of global land medium 1 train +3406 AR6_WGII 283 18 Lower emissions that would limit the global temperature increase to <2°C would reduce projected increases of burned area to ~35% and projected increases of fire frequency to ~20% medium 1 train +3407 AR6_WGII 283 19 Increased wildfire, combined with erosion due to deforestation, could degrade water supplies high 2 train +3408 AR6_WGII 283 20 For ecosystems with an historically low fire frequency, a projected 4°C rise in global temperature increases risks of fire, contributing to potential tree mortality and conversion of over half the Amazon rainforest to grassland and thawing of the Arctic permafrost that could release 11–200 GtC that could substantially exacerbate climate change medium 1 train +3409 AR6_WGII 285 15 Similar to tropical peatlands, given projected human population growth and socioeconomic changes, the continued conversion of forests and savannas into agricultural or pasture systems very likely poses a significant risk of rapid carbon loss which will amplify the climate change-induced risks substantially high 2 train +3410 AR6_WGII 285 17 Cascading trophic effects triggered by top predators or the largest herbivores propagate through food webs and reverberate through to the functioning of whole ecosystems, notably altering productivity, carbon and nutrient turnover and net carbon storage medium 1 train +3411 AR6_WGII 285 18 Across different field experiments, the ecosystem consequences of the presence or absence of herbivores and carnivores have been found to be quantitatively as large as the effects of other environmental change drivers such as warming, enhanced CO 2, fire and variable nitrogen deposition medium 1 train +3412 AR6_WGII 285 23 It is virtually certain that land cover changes affect regional and global climate through changes to albedo, evapotranspiration and roughness very high 3 train +3413 AR6_WGII 285 24 There is growing evidence that biosphere-related climate processes are being affected by climate change in combination with disturbance and LULCC high 2 train +3414 AR6_WGII 285 25 It is virtually certain that land surface change caused by disturbances such as forest fires, hurricanes, phenological changes, insect outbreaks and deforestation affect carbon, water and energy exchanges, thereby influencing weather and climate very high 3 train +3415 AR6_WGII 285 27 Due to the positive impacts of CO 2 on vegetation growth and ecosystem carbon storage high 2 train +3416 AR6_WGII 287 10 For instance, the impacts of climate-induced altered animal composition and trophic cascades on ecosystem carbon turnover (see Sections 2.4.4.4, 2.5.3.4) could be a substantive contribu- tion to carbon–climate feedbacks low 0 train +3417 AR6_WGII 287 12 Climate-induced shifts towards forests in what is currently tundra would be expected to reduce regional albedo especially in spring, but also during parts of winter when trees are snow-free (whereas tundra vegetation would be covered in snow), which amplifies warming regionally high 2 train +3418 AR6_WGII 288 2 While this has not yet been systematically explored, similar feedbacks might also emerge from a CO2-induced woody cover increase in savannas low 0 test +3419 AR6_WGII 288 5 Locally, both directly human-mediated and climate change-mediated changes in vegetation cover can therefore notably affect annual average freshwater availability to human societies, especially if negative feedbacks amplify the reduction of vegetation cover, evapotranspiration and precipitation medium 1 test +3420 AR6_WGII 290 16 As species become rare, their roles in the functioning of the ecosystem diminishes and disappears altogether if they become locally extinct high 2 train +3421 AR6_WGII 290 17 Loss of species and functional groups reduces the ability of an ecosystem to provide services, and lowers its resilience to climate change high 2 train +3422 AR6_WGII 290 23 Continued climate change substantially increases the risk of carbon losses due to wildfires, tree mortality from drought and insect pest outbreaks, peatland drying, permafrost thaw and changes in the structure of ecosystems; these could exacerbate self-reinforcing feedbacks between emissions from high-carbon ecosystems and increasing global temperatures medium 1 train +3423 AR6_WGII 290 24 Thawing of Arctic permafrost alone could release 11–200 GtC medium 1 train +3424 AR6_WGII 290 26 The exact timing and magnitude of climate–biosphere feedbacks and the potential tipping points of carbon loss are characterised by broad ranges of the estimates, but studies indicate that increased ecosystem carbon losses could cause extreme future temperature increases medium 1 train +3425 AR6_WGII 302 3 Cross-Chapter Box ILLNESS | Infectious Diseases, Biodiversity and Climate: Serious Risks Posed by Vector- and Water-Borne Diseases Authors: Marie-Fanny Racault (UK/France, Chapter 3), Stavana E. Strutz (USA, Chapter 2), Camille Parmesan (France/UK/USA, Chapter 2), Rita Adrian (Germany, Chapter 2), Guéladio Cissé (Mauritania/Switzerland/France, Chapter 7), Sarah Cooley (USA, Chapter 3), Meghnath Dhimal (Nepal), Luis E. Escobar (Guatemala/USA), Adugna Gemeda (Ethiopia, Chapter 9), Nathalie Jeanne Marie Hilmi (Monaco/France, Chapter 18), Salvador E. Lluch-Cota (Mexico, Chapter 5), Erin Mordecai (USA), Gretta Pecl (Australia, Chapter 11), A. Townsend Peterson (USA), Joacim Rocklöv (Germany/Sweden), Marina Romanello (UK/Argentina/Italy), David Schoeman (Australia, Chapter 3), Jan C. Semenza (Italy, Chapter 7), Maria Cristina Tirado (USA/Spain, Chapter 7), Gautam Hirak Talukdar (India, Chapter 2), Yongyut Trisurat (Thailand, Chapter 2) Climate change is altering the life cycles of many pathogenic organisms and changing the risk of transmission of vector- and water-borne infectious diseases to humans high 2 train +3426 AR6_WGII 302 5 There are substantial non-climatic drivers (LUC, wildlife exploitation, habitat degradation, public health and socioeconomic conditions) to the geographic and seasonal range suitability of pathogens and vectors that affect the attribution of the overall impacts on the prevalence or severity of some vector- and water-borne infectious diseases over recent decades high 2 test +3427 AR6_WGII 302 6 Adaptation options that involve sustained and rapid surveillance systems as well as the preservation and restoration of natural habitats with their associated higher levels of biodiversity, both marine and terrestrial, will be key to reducing the risk of epidemics and the large-scale transmission of diseases medium 1 train +3428 AR6_WGII 303 16 The area of coastline suitable for Cholera, Dengue or Malaria outbreak is increasing in North and West Africa, not changing in Central and East Africa, decreasing but potentially expanding in South Africa low 0 test +3429 AR6_WGII 304 2 South America Endemicity EpidemicEndemic in all regions except southern South AmericaEndemic Climate driversAbundance of coastal V. cholerae: northwestern South America: SST, Plankton (low confidence)Temperature, precipitation, droughtNorthern South America: temperature (low confidence) northern and southeastern South America: Tmax, Tmin, humidity (low confidence) Direction of ChangeArea of coastline suitable for outbreak: no change (low confidence)Increasing due to urbanisation and decreased vector control programmes, not strongly linked to climateHigher elevation regions: Increase (low confidence) Europe Endemicity Not endemic Southern Europe: focal outbreaks Not endemic Climate driversNo evidence for disease incidence Abundance of coastal V. cholerae: northern Europe: SST, Plankton (medium confidence) Direction of ChangeArea of coastline suitable for outbreak: increase (low confidence)Mediterranean regions of southern Europe: outbreaks (low confidence)No change North America Endemicity Not endemicPartially endemic in southern North AmericaNot endemic Climate driversNo evidence for disease incidence Abundance of coastal V. cholerae: eastern North America: SST (low confidence due to limited evidence)Winter Tmin (low confidence) Direction of ChangeArea of coastline suitable for outbreak: increase (low confidence)Declining No change Small Islands Endemicity EpidemicEndemic on many small islands in the TropicsEndemic on many small islands in the Tropics Climate driversDisease incidence: Caribbean: SST, LST, rainfall (low to medium confidence)Caribbean: SPI, Tmin (low confidence) Direction of ChangeArea of coastline suitable for outbreak: Caribbean and Pacific small islands: Decrease (low confidence)Increasing low 0 train +3430 AR6_WGII 304 3 Warming, acidification, hypoxia, SLR and increases in extreme weather and climate events (e.g., MHWs, storm surges, flooding and drought), which are projected to intensify in the 21st century (high confidence) (IPCC, 2021b), are driving species’ geographic range shifts and global rearrangements in the location and extent of areas with suitable conditions for many harmful pathogens, including viruses, bacteria, algae, protozoa and helminths high 2 train +3431 AR6_WGII 304 5 Our understanding of the impacts of climate-change drivers on the dynamics of Vibrio pathogens and related infections has been strengthened through improved observations from long-term monitoring programmes (Vezzulli et al., 2016) and statistical modelling supported by large-scale and high-resolution satellite observations high 2 train +3432 AR6_WGII 305 3 Already, studies have noted greater numbers of Vibrio-related human infections and, most notably, disease outbreaks linked to extreme weather events such as heat waves in temperate regions such as Northern Europe (Baker-Austin et al., 2013; Baker- Austin et al., 2017; Baker-Austin et al., 2018) high 2 train +3433 AR6_WGII 305 10 Climate-driven increase in temperature, the frequency and intensity of extreme events as well as changes in precipitation and relative humidity have provided opportunities for rearrangements of disease geography and seasonality, and emergence into new areas high 2 train +3434 AR6_WGII 305 17 In addition, the effective management and treatment of domestic and waste-water effluent, through better infrastructure and preservation of aquatic systems acting as natural water purifiers, have been key to securing the integrity of the surrounding water bodies, such as groundwater, reservoirs and lakes, and agricultural watersheds as well as protecting public health high 2 train +3435 AR6_WGII 305 18 The preservation and restoration of natural ecosystems, with their associated higher levels of biodiversity, have been reported as significant buffers against epidemics and large-scale pathogen transmission medium 1 train +3436 AR6_WGII 305 19 Furthermore, the timely allocation of financial resources and sufficient political will in support of a ‘One Health’ scientific research approach, recognising the health of humans, animals and ecosystems as interconnected (Rubin et al., 2014; Whitmee et al., 2015; Zinsstag et al., 2018), holds potential for improving surveillance and prevention strategies that may help to reduce the risks of further spread and new emergence of pathogens and vectors medium 1 train +3437 AR6_WGII 313 8 A large body of evidence has demonstrated the extent to which human life, well-being and economies are dependent on healthy ecosystems and also the range of threats that these are faced with high 2 train +3438 AR6_WGII 313 10 The health of ecosystems is, in turn, reliant upon the maintenance of natural levels of species’ richness and functional diversity high 2 train +3439 AR6_WGII 313 14 Analyses suggest that 30% or even 50% of land and sea needs to be protected or restored to confer adequate protection for species and ecosystem services high 2 train +3440 AR6_WGII 313 21 There is also increasing evidence, reported in this chapter, that the loss and degradation of natural and semi-natural habitats exacerbates the impacts of climate change and climatic extreme events on biodiversity and ecosystem services high 2 train +3441 AR6_WGII 313 25 Globally, there is a 38% overlap between areas of high carbon storage and high intact biodiversity (mainly in the peatland tropical forests of Asia, the western Amazon and the high Arctic), but only 12% of this is protected high 2 train +3442 AR6_WGII 314 17 Cross-Chapter Box NATURAL | Nature-Based Solutions for Climate Change Mitigation and Adaptation Authors: Camille Parmesan (France/USA/UK, Chapter 2), Gusti Anshari (Indonesia, Chapter 2, CCP7), Polly Buotte (USA, Chapter 4), Donovan Campbell (Jamaica, Chapter 15), Edwin Castellanos (Guatemala, Chapter 12), Annette Cowie (Australia, WGIII Chapter 12), Marta Rivera Ferre (Spain, Chapter 8), Patrick Gonzalez (USA, Chapter 2, CCP3), Elena López Gunn (Spain, Chapter 4), Rebecca Harris (Australia, Chapter 2, CCP3), Jeff Hicke (USA, Chapter 14), Rachel Bezner Kerr (USA/Canada, Chapter 5), Rodel Lasco (Philippines, Chapter 5), Robert Lempert (USA, Chapter 1), Brendan Mackey (Australia, Chapter 11), Paulina Martinetto (Argentina, Chapter 3), Robert Matthews (UK, WGIII, Chapter 3), Timon McPhearson (USA, Chapter 6), Mike Morecroft (UK, Chapter 2, CCP5), Aditi Mukherji (India, Chapter 4), Gert-Jan Nabuurs (the Netherlands, WGIII Chapter 7), Henry Neufeldt (Denmark/Germany, Chapter 5), Roque Pedace (Argentina, WGIII Chapter 3), Julio Postigo (USA/Peru, Chapter 12), Jeff Price (UK, Chapter 2, CCP1), Juan Pulhin (Philippines, Chapter 10), Joeri Rogelj (UK/Belgium, WGI Chapter 5), Daniela Schmidt (UK/Germany, Chapter 13), Dave Schoeman (Australia, Chapter 3), Pramod Kumar Singh (India, Chapter 18), Pete Smith (UK, WGIII Chapter 12), Nicola Stevens (South Africa, Chapter 2, CCP3), Stavana E. Strutz (USA, Chapter 2), Raman Sukumar (India, Chapter 1), Gautam Hirak Talukdar (India, Chapter 2, CCP1), Maria Cristina Tirado (USA/Spain, Chapter 7), Christopher Trisos (South Africa, Chapter 9) Nature-based solutions provide adaptation and mitigation benefits for climate change as well as contributing to other sustainable development goals high 2 train +3443 AR6_WGII 314 20 Poorly conceived and poorly designed nature-based mitigation efforts have the potential for multiple negative impacts, including competing for land and water with other sectors, reducing human well-being and failing to provide mitigation that is sustainable in the long term high 2 train +3444 AR6_WGII 315 2 Agro-ecological practices mitigate and adapt to climate change and can promote native biodiversity high 2 train +3445 AR6_WGII 315 10 Supporting local livelihoods and providing benefits to indigenous local communities and millions of private landowners, together with their active engagement in decision-making, are critical to ensuring support for NbS and their successful delivery high 2 train +3446 AR6_WGII 315 11 Forests Intact natural forest ecosystems are major stores of carbon and support large numbers of species that cannot survive in degraded habitats very high 3 train +3447 AR6_WGII 315 13 Deforestation and land degradation continue to be a source of global GHG emissions very high 3 train +3448 AR6_WGII 315 14 Protection of existing natural forests and sustainable management of semi-natural forests that continue to provide goods and services are highly effective NbS (Bauhus et al., 2009) high 2 train +3449 AR6_WGII 315 15 Natural forests and sustainably managed biodiverse forests play important roles in climate change mitigation and adaptation while providing many other ecosystem goods and services very high 3 train +3450 AR6_WGII 315 20 Reforestation of previously forested land can help to protect and recover biodiversity and is one of the most practical and cost-effective ways of sequestering and storing carbon high 2 train +3451 AR6_WGII 316 1 It can also restore hydrological processes, thereby improving water supply and quality (Ellison et al., 2017) and reducing the risk of soil erosion and floods high 2 train +3452 AR6_WGII 316 4 Adaptation measures, such as increasing the diversity of forest stands through ecological restoration rather than monoculture plantations can help to reduce these risks high 2 train +3453 AR6_WGII 316 5 When plantations are established without effective landscape planning and meaningful engagement including free prior and informed consent, they can present risks to biodiversity and the rights, well-being and livelihoods of indigenous and local communities as well as being less climate-resilient than natural forests very high 3 train +3454 AR6_WGII 316 6 Afforesting areas such as savannas and temperate peatlands, which would not naturally be forested, damages biodiversity and increases vulnerability to climate change high 2 train +3455 AR6_WGII 316 9 Draining, cutting and burning peat lead to oxidation and the release of CO2 very high 3 test +3456 AR6_WGII 316 10 Re-wetting by blocking drainage and preventing cutting and burning can reverse this process on temperate peatlands medium 1 test +3457 AR6_WGII 316 14 Naturally treeless temperate and boreal peatlands have, in some cases, been drained to enable trees to be planted, which then leads to CO 2 emissions, and restoration requires the removal of trees as well as re-blocking drainage high 2 train +3458 AR6_WGII 316 17 Therefore, blue carbon strategies, referring to climate change mitigation and adaptation actions based on the conservation and restoration of blue carbon ecosystems, can be effective NbS, with evidence of the recovery of carbon stocks following restoration, although their global or regional carbon sequestration potential and net mitigation potential may be limited medium 1 train +3459 AR6_WGII 316 18 They can also significantly attenuate wave energy, raise the seafloor (thereby counteracting the effects of SLR) and buffer storm surges and erosion from flooding high 2 train +3460 AR6_WGII 316 19 Additionally, they provide a suite of cultural (e.g., tourism and the livelihoods and well-being of native and local communities), provision (e.g., mangrove wood, edible fish and shellfish) and regulation (e.g., nutrient cycling) services high 2 train +3461 AR6_WGII 317 9 Agro-forestry, cover crops and other practices that increase vegetation cover and enhance soil organic matter, carefully managed and varying by agro-ecosystem, mitigate climate change high 2 train +3462 AR6_WGII 317 19 The adoption of agro-ecology principles and practices will therefore be highly beneficial to maintaining healthy, productive food systems under climate change high 2 train +3463 AR6_WGII 317 20 AF practices such as hedgerows and poly-cultures maintain habitat and connectivity for biodiversity, thus aiding the ability of wild species to respond to climate change via range shifts, and support ecosystem functioning under climate stress compared to conventional agriculture high 2 train +3464 AR6_WGII 317 22 Biodiverse agro-forestry systems increase ecosystem services and biodiversity benefits compared to simple agro-forestry and conventional agriculture high 2 train +3465 AR6_WGII 317 25 AF significantly improves food security and nutrition by increasing access to healthy, diverse diets and raising incomes for food producers, due to the increased biodiversity of crops, animals and landscapes high 2 train +3466 AR6_WGII 318 4 Yields of agro- forestry and organic farming can be lower than high-input agricultural systems but, conversely, AF can boost productivity and profit, varying according to the time frame and the socioeconomic, political or ecosystem context medium 1 train +3467 AR6_WGII 318 9 Conclusions NbS provide adaptation and mitigation benefits for climate change as well as contributing to achieving other sustainable development goals high 2 train +3468 AR6_WGII 319 1 Conversely, well- designed and implemented mitigation efforts have the potential to provide co-benefits in terms of climate change adaptation as well as providing multiple goods and services, including the conservation of biodiversity, clean and abundant water resources, flood mitigation, sustainable livelihoods, food and fibre security and human health and well-being high 2 train +3469 AR6_WGII 323 1 Meta-analyses of 162 studies involving 51,738 people documented that individuals with high levels of contact with nature throughout their lives felt significantly happier, healthier and more satisfied with their lives, and engaged in more pro-nature behaviours than those with little or no contact with nature high 2 train +3470 AR6_WGII 323 2 Meta-analyses of manipulative human trials across 65 studies documented a significant increase in positive feelings and attitudes and a decline in negative feelings after experimental treatments involving nature medium 1 train +3471 AR6_WGII 392 4 Observations: vulnerabilities and impacts Anthropogenic climate change has exposed ocean and coastal ecosystems to conditions that are unprecedented over millennia (high confidence2), and this has greatly impacted life in the ocean and along its coasts very high 3 train +3472 AR6_WGII 392 5 Fundamental changes in the physical and chemical characteristics of the ocean acting individually and together are changing the timing of seasonal activities (very high confidence), distribution (very high confidence) and abundance very high 3 train +3473 AR6_WGII 392 7 Geographic range shifts of marine species generally follow the pace and direction of climate warming (high confidence): surface warming since the 1950s has shifted marine taxa and communities poleward at an average (mean ± very likely3 range) of 59.2 ± 15.5 km per decade high 2 train +3474 AR6_WGII 392 8 Seasonal events occur 4.3 ± 1.8 d to 7.5 ± 1.5 d earlier per decade among planktonic organisms (very high confidence) and on average 3 ± 2.1 d earlier per decade for fish very high 3 train +3475 AR6_WGII 392 9 Warming, acidification and deoxygenation are altering ecological communities by increasing the spread of physiologically suboptimal conditions for many marine fish and invertebrates medium 1 train +3476 AR6_WGII 392 10 These and other responses have subsequently driven habitat loss (very high confidence), population declines (high confidence), increased risks of species extirpations and extinctions medium 1 train +3477 AR6_WGII 392 18 This Report also uses the term ‘likely range’ to indicate that the assessed likelihood of an outcome lies within the 17–83% probability range.Marine heatwaves lasting weeks to several months are exposing species and ecosystems to environmental conditions beyond their tolerance and acclimation limits very high 3 train +3478 AR6_WGII 392 19 WGI AR6 concluded that marine heatwaves are more frequent (high confidence), more intense and longer medium 1 train +3479 AR6_WGII 392 20 Open-ocean, coastal and shelf-sea ecosystems, including coral reefs, rocky shores, kelp forests, seagrasses, mangroves, the Arctic Ocean and semi-enclosed seas, have recently undergone mass mortalities from marine heatwaves very high 3 train +3480 AR6_WGII 392 21 Marine heatwaves, including well-documented events along the west coast of North America (2013–2016) and east coast of Australia (2015–2016, 2016–2017 and 2020), drive abrupt shifts in community composition that may persist for years (very high confidence), with associated biodiversity loss (very high confidence), collapse of regional fisheries and aquaculture (high confidence) and reduced capacity of habitat-forming species to protect shorelines high 2 train +3481 AR6_WGII 392 23 Although impacts of multiple climate and non-climate drivers can be beneficial or neutral to marine life, most are detrimental high 2 train +3482 AR6_WGII 392 24 Warming exacerbates coastal eutrophication and associated hypoxia, causing ‘dead zones’ (very high confidence), which drive severe impacts on coastal and shelf-sea ecosystems (very high confidence), including mass mortalities, habitat reduction and fisheries disruptions medium 1 train +3483 AR6_WGII 392 25 Overfishing exacerbates effects of multiple climate-induced drivers on predators at the top of the marine food chain medium 1 train +3484 AR6_WGII 392 26 Urbanisation and associated changes in freshwater and sediment dynamics increase the vulnerability of coastal ecosystems like sandy beaches, salt marshes and mangrove forests to sea level rise and changes in wave energy very high 3 train +3485 AR6_WGII 392 27 Although these non-climate drivers confound attribution of impacts to climate change, adaptive, inclusive and evidence-based management reduces the cumulative pressure on ocean and coastal ecosystems, which will decrease their vulnerability to climate change high 2 train +3486 AR6_WGII 393 3 Interacting climate- induced drivers and non-climate drivers are enhancing movement and bioaccumulation of toxins and contaminants into marine food webs (medium evidence, high agreement), and increasing salinity of coastal waters, aquifers and soils (very high confidence), which endangers human health very high 3 train +3487 AR6_WGII 393 4 Combined climate- induced drivers and non-climate drivers decrease physical protection of people, property and culturally important sites from flooding very high 3 test +3488 AR6_WGII 393 6 Marine species richness near the equator and in the Arctic is projected to continue declining, even with less than 2°C warming by the end of the century medium 1 train +3489 AR6_WGII 393 7 In the deep ocean, all global warming levels will cause faster movements of temperature niches by 2100 than those that have driven extensive reorganisation of marine biodiversity at the ocean surface over the past 50 years medium 1 train +3490 AR6_WGII 393 8 At warming levels beyond 2°C by 2100, risks of extirpation, extinction and ecosystem collapse escalate rapidly high 2 train +3491 AR6_WGII 393 9 Paleorecords indicate that at extreme global warming levels (>5.2°C), mass extinction of marine species may occur medium 1 train +3492 AR6_WGII 393 11 Some habitat- forming coastal ecosystems including many coral reefs, kelp forests and seagrass meadows, will undergo irreversible phase shifts due to marine heatwaves with global warming levels >1.5°C and are at high risk this century even in <1.5°C scenarios that include periods of temperature overshoot beyond 1.5°C high 2 train +3493 AR6_WGII 393 12 Under SSP1-2.6, coral reefs are at risk of widespread decline, loss of structural integrity and transitioning to net erosion by mid-century due to increasing intensity and frequency of marine heatwaves very high 3 train +3494 AR6_WGII 393 14 Other coastal ecosystems, including kelp forests, mangroves and seagrasses, are vulnerable to phase shifts towards alternate states as marine heatwaves intensify high 2 train +3495 AR6_WGII 393 15 Loss of kelp forests are expected to be greatest at the low-latitude warm edge of species’ ranges high 2 train +3496 AR6_WGII 393 18 Modest projected declines in global phytoplankton biomass translate into larger declines of total animal biomass (by 2080–2099 relative to 1995–2014) ranging from (mean ± very likely range) −5.7 ± 4.1% to −15.5 ± 8.5% under SSP1-2.6 and SSP5-8.5, respectively medium 1 train +3497 AR6_WGII 393 19 Projected declines in upper-ocean nutrient concentrations, likely associated with increases in stratification, will reduce carbon export flux to the mesopelagic and deep-sea ecosystems medium 1 train +3498 AR6_WGII 393 21 By 2100, 18.8 ± 19.0% to 38.9 ± 9.4% of the ocean will very likely undergo a change of more than 20 d (advances and delays) in the start of the phytoplankton growth period under SSP1-2.6 and SSP5- 8.5, respectively low 0 train +3499 AR6_WGII 393 22 This altered timing increases the risk of temporal mismatches between plankton blooms and fish spawning seasons (medium to high confidence) and increases the risk of fish- recruitment failure for species with restricted spawning locations, especially in mid-to-high latitudes of the Northern Hemisphere low 0 train +3500 AR6_WGII 393 23 Projected range shifts among marine species medium 1 train +3501 AR6_WGII 393 27 Sea level rise under emission scenarios that do not limit warming to 1.5°C will increase the risk of coastal erosion and submergence of coastal land (high confidence), loss of coastal habitat and ecosystems (high confidence) and worsen salinisation of groundwater (high confidence), compromising coastal ecosystems and livelihoods high 2 train +3502 AR6_WGII 393 28 Under SSP1-2.6, most coral reefs (very high confidence), mangroves (likely, medium confidence) and salt marshes (likely, medium confidence) will be unable to keep up with sea level rise by 2050, with ecological impacts escalating rapidly beyond 2050, especially for scenarios coupling high emissions with aggressive coastal development very high 3 train +3503 AR6_WGII 393 29 Resultant decreases in natural shoreline protection will place increasing numbers of people at risk very high 3 train +3504 AR6_WGII 393 30 The ability to adapt to current coastal impacts, cope with future coastal risks and prevent further acceleration of sea level rise beyond 2050 depends on immediate implementation of mitigation and adaptation actions very high 3 train +3505 AR6_WGII 394 1 Catch composition and diversity of regional fisheries will change (high confidence), and fishers who are able to move, diversify and leverage technology to sustain harvests decrease their own vulnerability medium 1 train +3506 AR6_WGII 394 2 Management that eliminates overfishing facilitates successful future adaptation of fisheries to climate change very high 3 train +3507 AR6_WGII 394 3 Marine-dependent communities, including Indigenous Peoples and local peoples, will be at increased risk of losing cultural heritage and traditional seafood- sourced nutrition medium 1 train +3508 AR6_WGII 394 4 Without adaptation, seafood- dependent people face increased risk of exposure to toxins, pathogens and contaminants (high confidence), and coastal communities face increasing risk from salinisation of groundwater and soil high 2 train +3509 AR6_WGII 394 5 Early-warning systems and public education about environmental change, developed and implemented within the local and cultural context, can decrease those risks high 2 train +3510 AR6_WGII 394 6 Coastal development and management informed by sea level rise projections will reduce the number of people and amount of property at risk (high confidence), but historical coastal development and policies impede change high 2 train +3511 AR6_WGII 394 7 Current financial flows are globally uneven and overall insufficient to meet the projected costs of climate impacts on coastal and marine social–ecological systems very high 3 train +3512 AR6_WGII 394 8 Inclusive governance that (a) accommodates geographically shifting marine life, (b) financially supports needed human transformations, (c) provides effective public education and (d) incorporates scientific evidence, Indigenous knowledge and local knowledge to manage resources sustainably shows greatest promise for decreasing human vulnerability to all of these projected changes in ocean and coastal ecosystem services very high 3 train +3513 AR6_WGII 394 10 Low-emission scenarios permit a wider array of feasible, effective and low-risk nature-based adaptation options (e.g., restoration, revegetation, conservation, early-warning systems for extreme events and public education) high 2 train +3514 AR6_WGII 394 11 Under high-emission scenarios, adaptation options (e.g., hard infrastructure for coastal protection, assisted migration or evolution, livelihood diversification, migration and relocation of people) are more uncertain and require transformative governance changes high 2 train +3515 AR6_WGII 394 12 Transformative climate adaptation will reinvent institutions to overcome obstacles arising from historical precedents, reducing current barriers to climate adaptation in cultural, financial and governance sectors high 2 train +3516 AR6_WGII 394 13 Without transformation, global inequities will likely increase between regions high 2 train +3517 AR6_WGII 394 15 Adaptation solutions implemented at appropriate scales, when combined with ambitious and urgent mitigation measures, can meaningfully reduce impacts high 2 train +3518 AR6_WGII 394 16 Increasing evidence from implemented adaptations indicates that multi-level governance, early-warning systems for climate-associated marine hazards, seasonal and dynamic forecasts, habitat restoration, ecosystem-based management, climate-adaptive management and sustainable harvesting tend to be both feasible and effective high 2 train +3519 AR6_WGII 394 17 Marine protected areas (MPAs), as currently implemented, do not confer resilience against warming and heatwaves (medium confidence) and are not expected to provide substantial protection against climate impacts past 2050 high 2 train +3520 AR6_WGII 394 18 However, MPAs can contribute substantially to adaptation and mitigation if they are designed to address climate change, strategically implemented and well governed high 2 train +3521 AR6_WGII 394 19 Habitat restoration limits climate-change-related loss of ecosystem services, including biodiversity, coastal protection, recreational use and tourism (medium confidence), provides mitigation benefits on local to regional scales (e.g., via carbon-storing ‘blue carbon’ ecosystems) high 2 train +3522 AR6_WGII 394 20 Ambitious and swift global mitigation offers more adaptation options and pathways to sustain ecosystems and their services high 2 train +3523 AR6_WGII 394 22 Nature-based solutions, such as ecosystem-based management, climate-smart conservation approaches (i.e., climate- adaptive fisheries and conservation) and coastal habitat restoration, can be cost-effective and generate social, economic and cultural co- benefits while contributing to the conservation of marine biodiversity and reducing cumulative anthropogenic drivers high 2 train +3524 AR6_WGII 394 23 The effectiveness of nature-based solutions declines with warming; conservation and restoration alone will be insufficient to protect coral reefs beyond 2030 (high confidence) and to protect mangroves beyond the 2040s high 2 train +3525 AR6_WGII 394 24 The multidimensionality of climate-change impacts and their interactions with other anthropogenic stressors calls for integrated approaches that identify trade-offs and synergies across sectors and scales in space and time to build resilience of ocean and coastal ecosystems and the services they deliver high 2 train +3526 AR6_WGII 395 1 Furthermore, existing inequalities and entrenched practices limit effective and just responses to climate change in coastal communities high 2 train +3527 AR6_WGII 397 22 Previous IPCC assessments have established that many marine species ‘have shifted their geographic ranges, seasonal activities, migration patterns, abundances and species interactions in response to climate change’ (high confidence) (IPCC, 2014b; IPCC, 2014c), which has had global impacts on species composition, abundance and biomass, and on ecosystem structure and function medium 1 train +3528 AR6_WGII 397 23 Warming and acidification have affected coastal ecosystems in concert with non-climate drivers (high confidence), which have affected habitat area, biodiversity, ecosystem function and services high 2 train +3529 AR6_WGII 397 25 AR5 and SROCC assessed how physiological sensitivity to climate-induced drivers is the underlying cause of most marine organisms’ vulnerability to climate high 2 train +3530 AR6_WGII 403 9 MHWs became more frequent over the 20th century (high confidence) and into the beginning of the 21st century, approximately doubling in frequency (high confidence) and becoming more intense and longer since the 1980s medium 1 train +3531 AR6_WGII 404 11 RSL rise is driving a global increase in the frequency of extreme sea levels high 2 train +3532 AR6_WGII 406 1 The expected frequency of the current 1-in-100-year extreme sea level is projected to increase by a median of 20–30 times across tide-gauge sites by 2050, regardless of emission scenario medium 1 train +3533 AR6_WGII 406 2 In addition, extreme-sea-level frequency may be affected by changes in tropical cyclone climatology (low confidence), wave climatology (low confidence) and tides high 2 train +3534 AR6_WGII 406 8 Recent evidence has strengthened estimates of the rate of change (Yamaguchi and Suga, 2019; Li et al., 2020a; Sallée et al., 2021), with an estimated increase of 1.0 ± 0.3% (very likely range) per decade over the period 1970–2018 high 2 train +3535 AR6_WGII 406 12 WGI AR6 assessed that only the California Current system has undergone large-scale upwelling-favourable wind intensification since the 1980s medium 1 train +3536 AR6_WGII 406 13 While no consistent pattern of contemporary changes in upwelling- favourable winds emerges from observation-based studies, numerical and theoretical work projects that summertime winds near poleward boundaries of upwelling zones will intensify, while winds near equatorward boundaries will weaken high 2 train +3537 AR6_WGII 406 14 Nevertheless, projected future annual cumulative upwelling wind changes at most locations and seasons remain within ±10–20% of present-day values medium 1 train +3538 AR6_WGII 406 16 Direct observational records since the mid-2000s remain too short to determine the relative contributions of internal variability, natural forcing and anthropogenic forcing to AMOC change high 2 train +3539 AR6_WGII 407 4 Ocean acidification is also developing in the ocean interior very high 3 train +3540 AR6_WGII 407 5 There, it leads to the shoaling of saturation horizons of aragonite and calcite high 2 train +3541 AR6_WGII 407 26 SROCC concluded that a loss of oxygen had occurred in the upper 1000 m of the ocean medium 1 train +3542 AR6_WGII 408 1 New findings since SROCC also confirm that the volume of oxygen minimum zones (OMZs) are expanding at many locations high 2 train +3543 AR6_WGII 408 3 Based on these CMIP6 projections, WGI AR6 concludes that the oxygen content of the subsurface ocean is projected to decline to historically unprecedented conditions over the 21st century medium 1 train +3544 AR6_WGII 408 8 Nitrogen availability tends to limit phytoplankton productivity throughout most of the low-latitude ocean, whereas dissolved iron availability limits productivity in high-nutrient, low- chlorophyll regions, such as in the main upwelling region of the Southern Ocean and the Eastern Equatorial Pacific high 2 train +3545 AR6_WGII 408 14 It is concluded that the surface ocean will encounter reduced nitrate concentrations in the 21st century medium 1 train +3546 AR6_WGII 408 16 The rates and magnitudes of these changes largely depend on the extent of future emissions very high 3 train +3547 AR6_WGII 408 22 The Arctic Ocean is characterised by the highest rates of acidification and warming, strong nutrient depletion, and it will likely become practically sea ice free in the September mean for the first time before the year 2050 in all SSP scenarios high 2 train +3548 AR6_WGII 408 23 In general, the projected changes in climate-induced drivers are less in absolute terms in the deep-sea (mesopelagic and bathypelagic domains and deep-sea habitats) than in the surface ocean and in shallow-water habitats (e.g., kelp ecosystems, warm-water corals) very high 3 train +3549 AR6_WGII 408 25 Significant differences in projected trends between the SSPs show that mitigation strategies will limit exposure of deep-sea ecosystems to potential warming, acidification and deoxygenation during the 21st century very high 3 train +3550 AR6_WGII 411 7 Ancient intervals of rapid climate warming that occurred between 300 and 50 million years ago (Ma) were triggered by the release of greenhouse gases high 2 train +3551 AR6_WGII 411 10 Warming and deoxygenation in the oceans were closely associated in hyperthermal events (high confidence), with anoxia reaching the photic zone and abyssal depths (Kaiho et al., 2014; Müller et al., 2017; Penn et al., 2018; Weissert, 2019), whereas ocean acidification has not been demonstrated consistently medium 1 train +3552 AR6_WGII 411 12 There is little evidence for ocean acidification in the past 2.6 Ma (low confidence) (Hönisch et al., 2012), but ocean ventilation was highly sensitive to even modest warming such as observed in the past 10,000 years medium 1 train +3553 AR6_WGII 411 17 Temperature affects the movement and transport of molecules and, thereby, the rates of all biochemical reactions; thus, ongoing and projected warming that remains below an organism’s physiological optimum will generally raise metabolic rates very high 3 test +3554 AR6_WGII 411 20 For example, organisms adapted to thermally stable environments (e.g., tropical, polar, deep sea) are often more sensitive to warming than those from thermally variable environments (e.g., estuaries) very high 3 train +3555 AR6_WGII 411 21 Heat tolerance also decreases with increasing organisational complexity (Storch et al., 2014; Pörtner and Gutt, 2016) and is lower in eggs, embryos and spawning fish than for their larval stages or adults outside the spawning season high 2 train +3556 AR6_WGII 411 22 By altering physiological responses, projected changes in ocean warming (Section 3.2.2.1) will modify growth, migration, distribution, competition, survival and reproduction very high 3 train +3557 AR6_WGII 412 2 Detrimental impacts of acidification include decreased growth and survival, and altered development, especially in early life stages high 2 train +3558 AR6_WGII 412 4 Calcifiers are generally more sensitive to acidification (e.g., for growth and survival) than non-calcifying groups high 2 train +3559 AR6_WGII 412 5 For calcifying primary producers, including phytoplankton and coralline algae, ocean acidification has different, often opposing effects, for example, decreasing calcification while photosynthetic rates increase high 2 train +3560 AR6_WGII 412 11 Under hypoxia (oxygen concentrations ≤2 mg l–1; Limburg et al., 2020), physiological and ecological processes are impaired and communities undergo species migration, replacement and loss, transforming community composition very high 3 train +3561 AR6_WGII 412 12 Hypoxia can lead to expanding OMZs, which will favour specialised microbes and hypoxia-tolerant organisms medium 1 train +3562 AR6_WGII 412 13 As respiration consumes oxygen and produces CO 2, lowered oxygen levels are often interlinked with acidification in coastal and tropical habitats (Rosa et al., 2013; Gobler and Baumann, 2016; Feely et al., 2018) and is an example of a compound hazard (Sections 3.2.4.1, 3.4.2.4).Increased density stratification and mixed-layer shallowing, caused by warming, freshening and sea ice decline, can alter light climate and nutrient availability within the surface mixed layer high 2 train +3563 AR6_WGII 412 15 Decreased upward nutrient supply is expected to decrease primary production in the low-latitude ocean medium 1 train +3564 AR6_WGII 412 16 Alternatively, higher mean underwater light levels resulting from changes in sea ice and/or mixed layer shallowing can increase primary production in high-latitude offshore regions, provided nutrient levels remain sufficiently high medium 1 train +3565 AR6_WGII 412 17 In some parts of the open Southern Ocean, where iron limitation largely controls primary productivity (Tagliabue et al., 2017), changes in wind fields will deepen the summer mixed-layer depth (Panassa et al., 2018), entrain more nutrients, and raise primary productivity in the future medium 1 train +3566 AR6_WGII 412 20 Marine heatwaves exacerbate the impacts of rising mean temperatures, with major ecological consequences very high 3 train +3567 AR6_WGII 412 25 The amplitude of diel and seasonal pH and CO 2 changes are projected to increase in the future due to lowered CO 2 seawater buffering capacity very high 3 train +3568 AR6_WGII 415 2 Non-climate drivers (Section 3.1) can dominate outcomes or amplify vulnerability to climate- induced drivers, with mostly detrimental effects such as extirpation very high 3 train +3569 AR6_WGII 415 7 Co-occurring environmental drivers often cause complex organismal responses high 2 train +3570 AR6_WGII 415 17 Ocean acidification poses a large risk for coralline algae that is further amplified by warming medium 1 train +3571 AR6_WGII 415 19 For seagrass, warming beyond a species’ thermal tolerance will limit growth and impact germination, but ocean acidification appears to increase thermal tolerance of some eelgrass species by increasing the photosynthesis-to-respiration ratio medium 1 train +3572 AR6_WGII 415 20 Thermal sensitivity of pelagic primary producers changes with nutrient supply high 2 train +3573 AR6_WGII 415 22 This trend may hold for open-ocean phytoplankton, which are often iron- limited medium 1 train +3574 AR6_WGII 415 26 Rising metabolic rates due to warming will be restricted to primary producers in high- nutrient regions medium 1 test +3575 AR6_WGII 415 28 The effects of ocean acidification on growth, metabolic rates or elemental composition of primary producers changes with nutrient availability and light conditions high 2 train +3576 AR6_WGII 417 2 Given the expected mixed-layer shallowing in some regions (Section 3.2.2.3), the exposure to overall higher mean irradiances could shift the effects of acidification from beneficial to detrimental for some primary producers, depending on species and organismal traits medium 1 train +3577 AR6_WGII 417 4 The few experimental studies that have addressed three or more drivers (Xu et al., 2014; Boyd et al., 2015b; Brennan and Collins, 2015; Brennan et al., 2017; Hoppe et al., 2018b; Moreno-Marín et al., 2018) indicate that one or two drivers generally dominate the cumulative outcome, with others playing a subordinate role medium 1 train +3578 AR6_WGII 417 7 Higher ocean CO 2 influences the thermal tolerance of species adapted to extreme but stable habitats in tropical and polar regions, more than that of thermally tolerant generalists high 2 train +3579 AR6_WGII 417 10 As with ocean acidification, reduced oxygen availability further alters the influence of warming on metabolic rates high 2 train +3580 AR6_WGII 417 17 In consequence, expansion of OMZs and other regions where warming, hypoxia and acidification combine will further reduce habitat for many fish and invertebrates high 2 train +3581 AR6_WGII 418 3 It is difficult to generalise to what extent co-occurring ocean warming ameliorates or exacerbates effects of acidification on behaviour (Laubenstein et al., 2019); outcomes depend upon species and life stage (Faleiro et al., 2015; Chan et al., 2016; Tills et al., 2016; Wang et al., 2018b; Jarrold et al., 2020), interactions between species (e.g., Paula et al., 2019) along with confounding factors including food availability and salinity medium 1 test +3582 AR6_WGII 418 5 Other influential drivers include ocean acidification, salinity (high confidence) (Lefevre, 2016; Whiteley et al., 2018; Reddin et al., 2020) or food availability/quality medium 1 train +3583 AR6_WGII 418 11 In highly fluctuating environments (e.g., upwelling regions, coastal zones), multiple drivers can change and interact across temporal and spatial scales, generating geographic mosaics of tolerances and sensitivities to environmental and climate change in marine organisms medium 1 train +3584 AR6_WGII 418 14 Some studies have documented higher phenotypic plasticity and tolerance to ocean warming and acidification in marine invertebrates (Dam, 2013; Kelly et al., 2013; Pespeni et al., 2013; Gaitán-Espitia et al., 2017a; Vargas et al., 2017; Li et al., 2018a), seaweeds (Noisette et al., 2013; Padilla-Gamiño et al., 2016; Machado Monteiro et al., 2019) and fish medium 1 train +3585 AR6_WGII 418 17 For instance, transgenerational effects and/or developmental acclimation, both ‘carry-over effects’ (where the early- life environment affects the expression of traits in later life stages or generations), can influence within- and cross-generational changes in the tolerances of marine organisms medium 1 train +3586 AR6_WGII 418 21 Although plasticity provides an adaptive mechanism, it is unlikely to provide a long-term solution for species undergoing sustained directional environmental change (e.g., global warming) medium 1 train +3587 AR6_WGII 419 5 Experimental evolution suggests that microbial populations can rapidly adapt (i.e., over 1–2 years) to environmental changes mimicking projected effects of climate change medium 1 train +3588 AR6_WGII 419 7 The evolutionary responses of microbes are conditioned by the number and characteristics of interacting drivers low 0 train +3589 AR6_WGII 420 9 Specifically, associations between vulnerabilities and traits of marine ectotherms in laboratory experiments correspond with organismal responses to ancient hyperthermal events medium 1 train +3590 AR6_WGII 420 14 On a global scale, ecosystem models project a −5.7 ± 4.1% (very likely range) to −15.5 ± 8.5% decline in marine animal biomass with warming under SSP1-2.6 and SSP5-8.5, respectively, by 2080–2099 relative to 1995–2014, albeit with significant regional variation in both trends and uncertainties medium 1 train +3591 AR6_WGII 420 16 For instance, trophic amplification (strengthening of responses to climate-induced drivers at higher trophic levels) may result from combined direct and indirect food-web-mediated effects medium 1 train +3592 AR6_WGII 420 17 Alternatively, compensatory species interactions can dampen strong impacts on species from ocean acidification, resulting in weaker responses at functional-group or community level than at species level medium 1 train +3593 AR6_WGII 420 18 Globally, the projected reduction of biomass due to climate-induced drivers is relatively unaffected by fishing pressure, indicating additive responses of fisheries and climate change low 0 train +3594 AR6_WGII 420 19 Regionally, projected interactions of climate-induced drivers, fisheries and other regional non-climate drivers can be both synergistic and antagonistic, varying across regions, functional groups and species, and can cause nonlinear dynamics with counterintuitive outcomes, underlining the importance of adaptations and associated trade-offs high 2 train +3595 AR6_WGII 423 3 Heat stress and mass bleaching events caused decreases in live coral cover (virtually certain) (Graham et al., 2014; Hughes et al., 2018b), loss of sensitive species (extremely likely) (Donner and Carilli, 2019; Lange and Perry, 2019; Toth et al., 2019; Courtney et al., 2020), vulnerability to disease (extremely likely) (van Woesik and Randall, 2017; Hadaidi et al., 2018; Brodnicke et al., 2019; Howells et al., 2020) and declines in coral recruitment in the tropics medium 1 train +3596 AR6_WGII 423 5 Changes in coral community structure due to bleaching have caused declines in reef carbonate production high 2 train +3597 AR6_WGII 423 7 Bleaching and other drivers promote phase shifts to ecosystems dominated by macroalgae or other stress-tolerant species (very high confidence) (Graham et al., 2015; Stuart-Smith et al., 2018), leading to changes in reef-fish species assemblages high 2 train +3598 AR6_WGII 423 8 Ocean acidification and associated declines in aragonite saturation state (Ω aragonite) decrease rates of calcification by corals and other calcifying reef organisms (very high confidence), reduce coral settlement (medium confidence) and increase bioerosion and dissolution of reef substrates high 2 train +3599 AR6_WGII 423 11 However, experimental evidence suggests that coral responses to ocean acidification are species specific medium 1 train +3600 AR6_WGII 424 6 The increased water depth due to coral loss and reef erosion, as well as reduced structural complexity, will limit wave attenuation and exacerbate the risk of flooding from SLR on reef- Table 3.3 | Summary of previous IPCC assessments of coral reefs Observations Projections AR5 (Hoegh-Guldberg et al., 2014; Wong et al., 2014) Coral reefs are one of the most vulnerable marine ecosystems high 2 train +3601 AR6_WGII 424 7 Mass coral bleaching and mortality, triggered by positive temperature anomalies high 2 train +3602 AR6_WGII 424 8 Ocean acidification reduces biodiversity and the calcification rate of corals (high confidence) while at the same time increasing the rate of dissolution of the reef framework medium 1 train +3603 AR6_WGII 424 12 SROCC (Bindoff et al., 2019a) ‘New evidence since AR5 and SR15 confirms the impacts of ocean warming and acidification on coral reefs (high confidence), enhancing reef dissolution and bioerosion (high confidence), affecting coral species distribution and leading to community changes high 2 train +3604 AR6_WGII 424 13 The rate of SLR (primarily noticed in small reef islands) may outpace the growth of reefs to keep up, although there is low agreement in the literature (low confidence).’ ‘Reefs are further exposed to other increased impacts, such as enhanced storm intensity, turbidity and increased runoff from the land high 2 train +3605 AR6_WGII 424 14 Recovery of coral reefs resulting from repeated disturbance events is slow high 2 train +3606 AR6_WGII 424 15 Only few coral reef areas show some resilience to global change drivers (low confidence).’‘Coral reefs will face very high risk at temperatures 1.5°C of global sea surface warming (very high confidence).’ ‘Almost all coral reefs will degrade from their current state, even if global warming remains below 2°C (very high confidence), and the remaining shallow coral reef communities will differ in species composition and diversity from present reefs very high 3 train +3607 AR6_WGII 424 16 This will greatly diminish the services they provide to society, such as food provision (high confidence), coastal protection (high confidence) and tourism (medium confidence).’ ‘The very high vulnerability of coral reefs to warming, ocean acidification, increasing storm intensity and SLR under climate change, including enhanced bioerosion high 2 train +3608 AR6_WGII 425 1 Local coral reef fish species richness is projected to decline due to the impacts of warming on coral cover and diversity high 2 train +3609 AR6_WGII 425 3 Major reef crises in the past 300 million years were governed by hyperthermal events (medium confidence) (Section 3.2.4.4; Cross-Chapter Box PALEO in Chapter 1) longer in time scale than anthropogenic climate change, during which net coral reef accretion was more strongly affected than biodiversity medium 1 test +3610 AR6_WGII 425 10 Recovery and restoration efforts that target heat- resistant coral populations and culture heat-tolerant algal symbionts have the greatest potential of effectiveness under future warming high 2 train +3611 AR6_WGII 425 12 In summary, additional evidence since SROCC and SR15 (Table 3.3) finds that living coral and reef growth are declining due to warming and MHWs very high 3 train +3612 AR6_WGII 425 13 Coral reefs are under threat of transitioning to net erosion with >1.5°C of global warming high 2 train +3613 AR6_WGII 425 14 The effectiveness of conservation efforts to sustain living coral area, coral diversity and reef growth is limited for the majority of the world’s reefs with >1.5°C of global warming high 2 train +3614 AR6_WGII 425 20 For example, the collapse of sea star populations in the Northeast Pacific due to a MHW-related disease outbreak (Hewson et al., 2014; Menge et al., 2016; Miner et al., 2018; Schiebelhut et al., 2018), including 80–100% loss of the common predatory sunflower star, Pycnopodia helianthoides very high 3 train +3615 AR6_WGII 425 21 Multiple lines of evidence find that foundational calcifying organisms such as mussels are at high risk of decline due to both the individual and synergistic effects of warming, acidification and hypoxia high 2 train +3616 AR6_WGII 425 23 Experiments show that ocean acidification negatively impacts mussel physiology very high 3 train +3617 AR6_WGII 425 24 Net calcification and abundance of mussels and other foundational species, including oysters, are expected to decline due to ocean acidification (very high confidence) (Kwiatkowski et al., 2016; Sunday et al., 2016; McCoy et al., 2018; Meng et al., 2018), causing the reorganisation of communities high 2 train +3618 AR6_WGII 426 10 Experiments indicate that warming reduces calcification by coralline algae high 2 train +3619 AR6_WGII 429 1 The upper vertical limits of some species will also be constrained by climate change high 2 train +3620 AR6_WGII 429 2 Experimental evidence since previous assessments further indicates that acidification decreases abundance and richness of calcifying species high 2 train +3621 AR6_WGII 429 3 Synergistic effects of warming and acidification will promote shifts towards macroalgal dominance in some ecosystems (medium confidence) and lead to reorganisation of communities medium 1 train +3622 AR6_WGII 429 8 Recent research (Straub et al., 2019; Butler et al., 2020; Filbee-Dexter et al., 2020b; Tait et al., 2021) supports the findings of previous assessments (Table 3.5) that kelp and other seaweeds in most regions are undergoing mass mortalities from high temperature extremes and range shifts from warming very high 3 train +3623 AR6_WGII 429 12 The most prominent effects are range shifts of species in response to ocean warming (high confidence) and changes in species distribution and abundance (high confidence) mostly in relation to ocean warming and acidification.’ ‘The dramatic decline of biodiversity in mussel beds of the Californian coast has been attributed to large-scale processes associated with climate-related drivers [...] (high confidence).’‘The abundance and distribution of rocky shore species will continue to change in a warming world high 2 train +3624 AR6_WGII 429 13 For example, the long-term consequences of ocean warming on mussel beds of the northeast Pacific are both positive (increased growth) and negative (increased susceptibility to stress and of exposure to predation) (medium confidence).’ ‘Observations performed near natural CO 2 vents in the Mediterranean Sea show that diversity, biomass and trophic complexity of rocky shore communities will decrease at future pH levels (high confidence).’ SR15 (Hoegh-Guldberg et al., 2018a) ‘Changes in ocean circulation can have profound impacts on [temperate] marine ecosystems by connecting regions and facilitating the entry and establishment of species in areas where they were unknown before (‘tropicalization’ ...) as well as the arrival of novel disease agents (medium agreement, limited evidence).’‘In the transition to 1.5°C, changes to water temperatures are expected to drive some species (e.g., plankton, fish) to relocate to higher latitudes and cause novel ecosystems to assemble high 2 train +3625 AR6_WGII 429 14 Other ecosystems (e.g., kelp forests, coral reefs) are relatively less able to move, however, and are projected to experience high rates of mortality and loss very high 3 train +3626 AR6_WGII 429 15 SROCC (Bindoff et al., 2019a) Intertidal rocky shores ecosystems are highly sensitive to ocean warming, acidification and extreme heat exposure during low tide emersion high 2 train +3627 AR6_WGII 429 17 These ecosystems have low to moderate adaptive capacity, as they are highly sensitive to ocean temperatures and acidification.’ ‘Benthic species will continue to relocate in the intertidal zones and experience mass mortality events due to warming high 2 train +3628 AR6_WGII 429 18 Interactive effects between acidification and warming will exacerbate the negative impacts on rocky shore communities, causing a shift towards a less diverse ecosystem in terms of species richness and complexity, increasingly dominated by macroalgae high 2 train +3629 AR6_WGII 430 1 Warming is driving range contraction and extirpation at the warm edge of species’ ranges and expansions at the cold range edge very high 3 train +3630 AR6_WGII 430 2 Local declines in populations of kelp and other canopy-forming seaweeds driven by MHWs and other stressors have caused irreversible shifts to turf- or urchin-dominated ecosystems, with lower productivity and biodiversity (high confidence) (Filbee-Dexter and Scheibling, 2014; Filbee-Dexter and Wernberg, 2018; Rogers-Bennett and Catton, 2019; Beas-Luna et al., 2020; Stuart-Smith et al., 2021), ecosystems dominated by warm- affinity seaweeds or coral high 2 train +3631 AR6_WGII 430 9 While reducing non-climate drivers can help prevent kelp loss from warming and MHWs, there is limited potential for restoration of kelp ecosystems after transition to urchin-dominant ecosystems high 2 train +3632 AR6_WGII 430 12 Active reseeding of wild kelp populations through transplantation and propagation of warm-tolerant genotypes (Coleman et al., 2020b; Alsuwaiyan et al., 2021) can overcome low dispersal rates of many kelp species and facilitate effective restoration medium 1 train +3633 AR6_WGII 430 13 Building on the conclusions of SROCC, this assessment finds that kelp ecosystems are expected to decline and undergo changes in community structure in the future due to warming and increasing frequency and intensity of MHWs high 2 train +3634 AR6_WGII 430 14 Risk of loss of kelp ecosystems Table 3.5 | Summary of previous IPCC assessments of kelp ecosystems Observations Projections AR5 (Wong et al., 2014) ‘Kelp forests have been reported to decline in temperate areas in both hemispheres, a loss involving climate change high 2 train +3635 AR6_WGII 430 15 Decline in kelp populations attributed to ocean warming has been reported in southern Australia and the north coast of Spain.’‘Kelp ecosystems will decline with the increased frequency of heatwaves and sea temperature extremes as well as through the impact of invasive subtropical species (high confidence).’ ‘Climate change will contribute to the continued decline in the extent of [...] kelps in the temperate zone (medium confidence) and the range of [...] kelps in the Northern Hemisphere will expand poleward (high confidence).’ SR15 (Hoegh-Guldberg et al., 2018a) Observed movement of kelp ecosystems not assessed.‘In the transition to 1.5°C of warming, changes to water temperatures will drive some species (e.g., plankton, fish) to relocate to higher latitudes and cause novel ecosystems to assemble high 2 train +3636 AR6_WGII 430 16 Other ecosystems (e.g., kelp forests, coral reefs) are relatively less able to move, however, and are projected to experience high rates of mortality and loss (very high confidence).’ SROCC (Bindoff et al., 2019a) ‘Kelp forests have experienced large-scale habitat loss and degradation of ecosystem structure and functioning over the past half century, implying a moderate to high level of risk at present conditions of global warming (high confidence).’ ‘The abundance of kelp forests has decreased at a rate of ~2% per year over the past half century, mainly due to ocean warming and marine heat waves [...], as well as from other human stressors (high confidence).’ ‘Changes in ocean currents have facilitated the entry of tropical herbivorous fish into temperate kelp forests decreasing their distribution and abundance (medium confidence).’ ‘The loss of kelp forests is followed by the colonisation of turfs, which contributes to the reduction in habitat complexity, carbon storage and diversity (high confidence).’Kelp forests will face moderate to high risk at temperatures above 1.5°C global sea surface warming high 2 train +3637 AR6_WGII 431 2 Although these coastal ecosystems have historically been sensitive to erosion-accretion cycles driven by sea level, drought and storms (high confidence) (Peteet et al., 2018; Wang et al., 2018c; Jones et al., 2019b; Urrego et al., 2019; Hapsari et al., 2020; Zhao et al., 2020b), they were impacted for much of the 20th century primarily by non-climate drivers very high 3 train +3638 AR6_WGII 431 3 Nevertheless, the influence of climate-induced drivers has become more apparent over recent decades medium 1 train +3639 AR6_WGII 431 4 Estuarine biota are sensitive to warming high 2 train +3640 AR6_WGII 431 5 MHWs can be more severe in estuaries than in adjacent coastal seas (Lonhart et al., 2019), causing conspicuous impacts very high 3 train +3641 AR6_WGII 431 6 Relative SLR extends the upstream limit of saline waters (high confidence) (Harvey et al., 2020; Jiang et al., 2020) and alters tidal ranges high 2 train +3642 AR6_WGII 431 7 Elevated water levels also alter submergence patterns for intertidal habitat (high confidence) (Andres et al., 2019), moving high-water levels inland (high confidence) (Peteet et al., 2018; Appeaning Addo et al., 2020; Liu et al., 2020e) and increasing the salinity of coastal water tables and soils high 2 train +3643 AR6_WGII 431 8 These processes favour inland and/or upstream migration of intertidal habitat, where it is unconstrained by infrastructure, topography or other environmental features high 2 train +3644 AR6_WGII 431 11 Overall, changing salinity and submergence patterns decrease the ability of shoreline vegetation to trap sediment (Xue et al., 2018), reducing accretion rates and increasing the vulnerability of estuarine shorelines to submergence by SLR and erosion by wave action medium 1 train +3645 AR6_WGII 431 14 The same phenomena alter salinity gradients, which are the primary drivers of estuarine species distributions high 2 train +3646 AR6_WGII 431 16 Acidification of estuarine water is a growing hazard (medium confidence) (Doney et al., 2020; Scanes et al., 2020a; Cai et al., 2021), and resident organisms display sensitivity to altered pH in laboratory settings medium 1 train +3647 AR6_WGII 431 20 Warming (including MHWs) and eutrophication interact to decrease estuarine oxygen content and pH, increasing the vulnerability of animals to MHWs (Brauko et al., 2020) and exacerbating the incidence and impact of dead zones medium 1 train +3648 AR6_WGII 431 22 All these impacts are projected to escalate under future climate change, but their magnitude depends on the amount of warming, the socioeconomic development pathway and implementation of adaptation strategies medium 1 train +3649 AR6_WGII 431 23 Modelling studies (Lopes et al., 2019; Rodrigues et al., 2019; White et al., 2019; Zhang and Li, 2019; Hong et al., 2020; Krvavica and Ružić, 2020; Liu et al., 2020e; Shalby et al., 2020) suggest that responses of estuaries to SLR will be complex and context dependent (Khojasteh et al., 2021), but project that salinity, tidal range, storm-surge amplitude, depth and stratification will increase with SLR (medium confidence), and that marine-dominated waters will penetrate farther upstream high 2 train +3650 AR6_WGII 431 24 Without careful management of freshwater inputs, sediment augmentation and/or the restoration of shorelines to more natural states, transformation and loss of intertidal areas and wetland vegetation will increase with SLR (high confidence) (Doughty et al., 2019; Leuven et al., 2019; Yu et al., 2019; Raw et al., 2020; Shih, 2020; Stein et al., 2020), with small, shallow microtidal estuaries being more vulnerable to impacts than deeper estuaries with well-developed sediments medium 1 train +3651 AR6_WGII 431 25 Warming and MHWs will enhance stratification and deoxygenation in shallow lagoons medium 1 train +3652 AR6_WGII 432 3 Since AR5 and SROCC, syntheses have emphasised that the vulnerability of rooted wetland ecosystems to climate-induced drivers is exacerbated by non-climate drivers (high confidence) (Elliott et al., 2019; Ostrowski et al., 2021; Williamson and Guinder, 2021) and climate Table 3.6 | Summary of previous IPCC assessments of estuaries, deltas and coastal lagoons Observations Projections AR5 (Wong et al., 2014) Humans have impacted lagoons, estuaries and deltas (high to very high confidence), but non-climate drivers have been the primary agents of change very high 3 train +3653 AR6_WGII 432 4 In estuaries and lagoons, nutrient inputs have driven eutrophication, which has modified food-web structures (high confidence) and caused more-intense and longer-lasting hypoxia, more-frequent occurrence of harmful algal blooms and enhanced emissions of nitrous oxide high 2 train +3654 AR6_WGII 432 5 In deltas, land-use changes and associated disruption of sediment dynamics and land subsidence have driven changes that have been exacerbated by relative SLR and episodic events, including river floods and oceanic storm surges very high 3 train +3655 AR6_WGII 432 6 Increased coastal flooding, erosion and saltwater intrusions have led to degradation of ecosystems (very high confidence).Future changes in climate impact-drivers such as warming, acidification, waves, storms, sea level rise (SLR) and runoff will have consequences for ecosystem function and services in lagoons and estuaries high 2 train +3656 AR6_WGII 432 7 Warming, changes in precipitation and changes in wind strength can interact to alter water-column salinity and stratification (medium confidence), which could impact water column oxygen content medium 1 train +3657 AR6_WGII 432 8 Land-use change, SLR and intensifying storms will alter deposition-erosion dynamics, impacting shoreline vegetation and altering turbidity medium 1 train +3658 AR6_WGII 432 10 The projected impacts of climate change on deltas are associated mainly with pluvial floods and SLR, which will amplify observed impacts of interacting climate and non-climate drivers high 2 train +3659 AR6_WGII 432 11 Estuaries, deltas and lagoons were not assessed in this report.Under both a 1.5°C and 2°C of warming, relative to the pre-industrial era, deltas are expected to be highly threatened by SLR and localised subsidence high 2 test +3660 AR6_WGII 432 12 The slower rate of SLR associated with 1.5°C of warming poses smaller risks of flooding and salinisation (high confidence), and facilitates greater opportunities for adaptation, including managing and restoring natural coastal ecosystems and infrastructure reinforcement medium 1 train +3661 AR6_WGII 432 14 Other feedbacks, such as landward migration of wetlands and the adaptation of infrastructure, remain important (medium confidence).’ SROCC (Bindoff et al., 2019a) Increased seawater intrusion caused by SLR has driven upstream redistribution of marine biotic communities in estuaries (medium confidence) where physical barriers, such as the availability of benthic substrates, do not limit availability of suitable habitats medium 1 train +3662 AR6_WGII 432 15 Warming has driven poleward range shifts in species’ distributions among estuaries medium 1 train +3663 AR6_WGII 432 16 Interactions between warming, eutrophication and hypoxia have increased the incidence of harmful algal blooms (high confidence), pathogenic bacteria, such as Vibrio species, (low confidence) and mortalities of invertebrates and fish communities (medium confidence).‘Salinisation and expansion of hypoxic conditions will intensify in eutrophic estuaries, especially in mid and high latitudes with microtidal regimes (high confidence).’ ‘The effects of warming will be more pronounced in high-latitude and temperate shallow estuaries with limited exchange with the open ocean [...] and seasonality that already leads to dead zone development [...] (medium confidence).’ Interaction between SLR and changes in precipitation will have greater impacts on shallow than deep estuaries medium 1 train +3664 AR6_WGII 432 17 Estuaries characterised by large tidal exchanges and associated well-developed sediments will be more resilient to projected SLR and changes in river flow medium 1 train +3665 AR6_WGII 432 18 Human activities that inhibit sediment dynamics in coastal deltas increase their vulnerability to SLR medium 1 train +3666 AR6_WGII 433 1 Global rates of mangrove loss have been extensive but are slowing high 2 train +3667 AR6_WGII 433 2 From 2000 to 2010 mangrove loss averaged 0.16% yr–1, globally, but with greatest loss in Southeast Asia high 2 train +3668 AR6_WGII 433 3 Salt-marsh ecosystems have also suffered extensive losses (up to 60% in places since the 1980s), especially in developed and rapidly developing countries medium 1 train +3669 AR6_WGII 433 4 Similarly, 29% of seagrass meadows were lost from 1879–to 2006 due primarily to coastal development and degradation of water quality, with climate-change impacts escalating since 1990 medium 1 train +3670 AR6_WGII 433 5 Local examples of habitat stability or growth (e.g., de los Santos et al., 2019; Laengner et al., 2019; Sousa et al., 2019; Suyadi et al., 2019; Derolez et al., 2020; Goldberg et al., 2020; McKenzie and Yoshida, 2020) indicate some resilience to climate change in the absence of non-climate drivers high 2 train +3671 AR6_WGII 433 6 Nevertheless, previous declines have left wetland ecosystems more vulnerable to impacts from climate-induced drivers and non-climate drivers high 2 train +3672 AR6_WGII 433 8 Warming is allowing some, but not all (Rogers and Krauss, 2018; Saintilan et al., 2018), mangrove Table 3.7 | Summary of previous IPCC assessments of mangroves, salt marshes and seagrass beds Observations Projections AR5 (Wong et al., 2014) Seagrasses occurring close to their upper thermal limits are already stressed by climate change high 2 train +3673 AR6_WGII 433 11 As a result, interactions between climate change and non-climate drivers will continue to cause declines in estuarine vegetated systems very high 3 train +3674 AR6_WGII 433 12 SR15 (Hoegh-Guldberg et al., 2018a) Vegetated blue carbon systems were not assessed in this report.Intact wetland ecosystems can reduce the adverse impacts of rising sea levels and intensifying storms by protecting shorelines medium 1 train +3675 AR6_WGII 433 13 Under 1.5°C of warming, natural sedimentation rates are projected to outpace SLR (medium confidence), but ‘other feedbacks, such as landward migration of wetlands and the adaptation of infrastructure, remain important (medium confidence).’ SROCC (Bindoff et al., 2019a; Oppenheimer et al., 2019) Coastal ecosystems, including salt marshes, mangroves, vegetated dunes and sandy beaches, can build vertically and expand laterally in response to SLR, though this capacity varies across sites high 2 train +3676 AR6_WGII 433 15 However, as a consequence of human actions that fragment wetland habitats and restrict landward migration, coastal ecosystems progressively lose their ability to adapt to climate-induced changes and provide ecosystem services, including acting as protective barriers high 2 train +3677 AR6_WGII 433 16 Examples include mangrove encroachment into subtropical salt marshes (high confidence) and contraction in extent of low-latitude seagrass meadows high 2 train +3678 AR6_WGII 433 17 Plants with low tolerance to flooding and extreme temperatures are particularly vulnerable, increasing the risk of extirpation medium 1 train +3679 AR6_WGII 433 18 Extreme-weather events, including heatwaves, droughts and storms, are causing mass mortalities and changes in community composition in coastal wetlands high 2 train +3680 AR6_WGII 433 19 Severe disturbance of wetlands or transitions among wetland community types can favour invasive species medium 1 train +3681 AR6_WGII 433 20 The degradation or loss of vegetated coastal ecosystems reduces carbon storage, with positive feedbacks to the climate system (high confidence).‘Seagrass meadows (high confidence) [...] will face moderate to high risk at temperature above 1.5°C global sea surface warming.’ ‘The transition from undetectable to moderate risk in salt marshes [...] takes place between 0.7°C–1.2°C of global sea surface warming (medium/high confidence), and between 0.9°C–1.8°C (medium confidence) in sandy beaches, estuaries and mangrove forests.’ ‘The ecosystems at moderate to high risk under future emission scenarios are mangrove forests (transition from moderate to high risk at 2.5°C–2.7°C of global sea surface warming), estuaries and sandy beaches (2.3°C–3.0°C) and salt marshes (transition from moderate to high risk at 1.8°C–2.7°C and from high to very high risk at 3.0°C–3.4°C) (medium confidence).’ ‘Global coastal wetlands will lose between 20–90% of their area depending on emissions scenario with impacts on their contributions to carbon sequestration and coastal protection high 2 train +3682 AR6_WGII 433 21 But SLR and warming are projected to drive global loss of up to 90% of vegetated wetlands by the end of the century under the RCP8.5 (medium confidence), especially if landward migration and sediment supply are limited by human modification of shorelines and river flows medium 1 train +3683 AR6_WGII 434 1 This expansion can affect species interactions (Guo et al., 2017; Friess et al., 2019), and enhance sediment accretion and carbon storage rates in some instances medium 1 train +3684 AR6_WGII 434 2 Drought, low sea levels and MHWs can cause significant die-offs among mangroves medium 1 train +3685 AR6_WGII 434 3 Seagrasses are similarly vulnerable to warming (high confidence) (Repolho et al., 2017; Duarte et al., 2018; Jayathilake and Costello, 2018; Savva et al., 2018), which has been attributed as one cause of observed changes in distribution and community structure medium 1 train +3686 AR6_WGII 434 4 MHWs, together with storm-driven turbidity and structural damage, can cause seagrass die-offs (high confidence) (Arias-Ortiz et al., 2018; Kendrick et al., 2019; Smale et al., 2019; Strydom et al., 2020), shifts to small, fast-growing species high 2 train +3687 AR6_WGII 434 5 The sensitivity of salt marshes and mangroves to RSLR depends on whether they accrete inorganic sediment and/or organic material at rates equivalent to rising water levels very high 3 train +3688 AR6_WGII 434 6 Otherwise, wetland ecosystems must migrate either inland or upstream, or face gradual submergence in deeper, increasingly saline water very high 3 train +3689 AR6_WGII 434 8 Submergence drives changes in community structure (high confidence) (Jones et al., 2019b; Yu et al., 2019; Douglass et al., 2020; Langston et al., 2020) and functioning (high confidence) (Charles et al., 2019; Buffington et al., 2020; Stein et al., 2020), and will eventually lead to extirpation of the most sensitive vegetation (medium confidence) (Schepers et al., 2017; Scalpone et al., 2020) and associated animals low 0 train +3690 AR6_WGII 434 11 On the basis of paleorecords (Table 3.8), we assess that mangroves and salt marshes are likely at high risk from future SLR, even under SSP1-1.9, with impacts manifesting in the mid- term medium 1 train +3691 AR6_WGII 434 12 Under SSP5-8.5, wetlands are very likely at high risk from SLR, with larger impacts manifesting before 2040 medium 1 train +3692 AR6_WGII 434 13 By 2100, these ecosystems are at high risk of impacts under all scenarios except SSP1-1.9 (high confidence), with impacts most severe along coastlines with gently sloping shorelines, limited sediment inputs, small tidal ranges and limited space for inland migration very high 3 train +3693 AR6_WGII 434 16 Other species, such as Posidonia oceanica in the Mediterranean, might lose as much as 75% of their habitat by 2050 under RCP8.5 and become functionally extinct low 0 test +3694 AR6_WGII 434 17 Observed impacts of MHWs (Kendrick et al., 2019; Strydom et al., 2020; Serrano et al., 2021) indicate that increasing intensity and frequency of MHWs (Section 3.2.2.1) will have escalating impacts on seagrass ecosystems high 2 train +3695 AR6_WGII 434 18 Habitat suitability can also be reduced by moderate RSLR, due to its impact on light attenuation medium 1 train +3696 AR6_WGII 434 19 Overall, warming will drive range shifts in wetland species (medium to high confidence), but SLR poses the greatest risk for mangroves and salt marshes, with significant losses projected under all future scenarios by mid-century medium 1 train +3697 AR6_WGII 435 1 Observations Projections AR5 (Wong et al., 2014) ‘Globally, beaches and dunes have in general undergone net erosion over the past century or longer.’ ‘Attributing shoreline changes to climate change is still difficult owing to the multiple natural and anthropogenic drivers contributing to coastal erosion.’‘In the absence of adaptation, beaches, sand dunes and cliffs currently eroding will continue to do so under increasing sea level high 2 train +3698 AR6_WGII 435 2 In many locations, finding sufficient sand to rebuild beaches and dunes artificially will become increasingly difficult and expensive as present supplies near project sites are depleted (high confidence).’ ‘In the absence of adaptation measures, beaches and sand dunes currently affected by erosion will continue to be affected under increasing sea levels (high confidence).’ SROCC (Bindoff et al., 2019a) Coastal ecosystems are already impacted by the combination of SLR, other climate-related ocean changes and adverse effects from human activities on ocean and land high 2 train +3699 AR6_WGII 435 3 Attributing such impacts to SLR, however, remains challenging due to the influence of other climate-related and non-climate drivers such as infrastructure development and human-induced habitat degradation high 2 train +3700 AR6_WGII 435 4 Coastal ecosystems, including salt marshes, mangroves, vegetated dunes and sandy beaches, can build vertically and expand laterally in response to SLR, though this capacity varies across sites high 2 train +3701 AR6_WGII 435 5 Coastal ecosystems also progressively lose their ability to adapt to climate-induced changes and provide ecosystem services, including acting as protective barriers high 2 train +3702 AR6_WGII 435 7 This will be especially important in low-lying coastal areas with high population and building densities (medium confidence).’ ‘Assuming that the physiological underpinning of the relationship between body size and temperature can be applied to warming (medium confidence), the body size of sandy beach crustaceans is expected to decrease under warming (low evidence, medium agreement).’ Sandy beaches transition from undetectable to moderate risk between 0.9°C and 1.8°C (medium confidence) of global sea surface warming and from moderate to high risk at 2.3°C–3.0°C of global sea surface warming medium 1 train +3703 AR6_WGII 435 9 However, pervasive coastal urbanisation lowers the buffering capacity and recovery potential of sandy beach ecosystems to impacts from SLR and warming, and thus is expected to limit their resilience to climate change (high confidence).’ ‘Coastal squeeze and human-driven habitat deterioration will reduce the natural capacity of these ecosystems to adapt to climate impacts (high confidence).’ by 2100 under all scenarios except SSP1-1.9 high 2 train +3704 AR6_WGII 435 10 MHWs pose the greatest risk to seagrasses high 2 train +3705 AR6_WGII 435 11 In all cases, losses will be greatest where accommodation space is constrained or where other non-climate drivers exacerbate risk from climate-induced drivers very high 3 train +3706 AR6_WGII 435 17 Nevertheless, RSLR, increases in wave energy and/or changes in wave direction, disruptions to sediment supplies (including sand mining) and other anthropogenic modifications of the coast have driven localised beach erosion very high 3 train +3707 AR6_WGII 435 18 Corresponding analyses of coarse-scale (30-m resolution) global data estimate that 15% of tidal flats (including beaches) have been lost since 1984 (medium confidence) (Mentaschi et al., 2018; Murray et al., 2019) but with a corresponding number of the world’s beaches accreting (28%) as eroding (24%) medium 1 train +3708 AR6_WGII 436 2 None of these local studies, however, considered high-energy storm events, which are known to also impact sandy coasts high 2 train +3709 AR6_WGII 436 3 Nevertheless, the most- advanced available models, which incorporate multiple coastal processes, including SLR, project that without anthropogenic barriers to erosion, 13.6–15.2% and 35.7–49.5% of the world’s beaches likely risk undergoing at least 100 m of shoreline retreat (relative to 2010) by 2050 and 2100, respectively low 0 train +3710 AR6_WGII 436 4 Aggregating these trends regionally suggests that relative rates of shoreline change under RCP4.5 and RCP8.5 diverge strongly after mid-century, with trends towards erosion escalating under RCP8.5 by 2100 medium 1 train +3711 AR6_WGII 436 5 This trend supports the WGI AR6 assessment that projected SLR will contribute to erosion of sandy beaches, especially under high-emissions futures high 2 train +3712 AR6_WGII 436 7 But even amongst the best-studied taxa, such as turtles, vulnerability to warming (high confidence) and SLR medium 1 train +3713 AR6_WGII 436 8 Moreover, although established vulnerabilities imply high projected future risk for turtles high 2 train +3714 AR6_WGII 437 2 For example, although feminisation poses a significant long-term risk to turtle populations (high confidence), it might contribute to population growth in the near to mid-term medium 1 train +3715 AR6_WGII 437 3 Resilience to climate change might be further enhanced by range extensions, alterations in nesting phenology and fine-scale nest-site selection medium 1 train +3716 AR6_WGII 437 6 Nevertheless, theoretical sensitivity to warming (Section 3.3.2), together with the projected loss of habitat under future climate scenarios, suggest substantial impacts for populations and communities of beach fauna into the future high 2 train +3717 AR6_WGII 437 7 These impacts will be exacerbated by coastal squeeze along urbanised coastlines high 2 train +3718 AR6_WGII 437 12 Since AR5, there is evidence for increasing frequency and duration of MHWs, extreme-weather events and a diversity of threats across depth strata causing mass-mortality events, local extirpations and coral reef decline high 2 train +3719 AR6_WGII 437 13 In most SES, non-climate drivers, including pollution, habitat destruction and especially overfishing, are decreasing the local adaptive capacity of organisms and the ability of ecosystems to cope with climate-change impacts high 2 train +3720 AR6_WGII 437 14 The SLR is accelerating faster than expected high 2 train +3721 AR6_WGII 437 15 The size and number of OMZs are increasing worldwide and in most SES high 2 train +3722 AR6_WGII 437 16 In the Persian Gulf and Red Sea, increasing nutrient loads associated with coastal activities and warming has increased the size of OMZs high 2 train +3723 AR6_WGII 437 19 Consequently, SES will respond faster than most other parts of the ocean high 2 train +3724 AR6_WGII 437 20 Due to a mixture of global and local human stressors, key fisheries have undergone fundamental changes in their abundance and distribution over the past 50 years (medium confidence).’‘Projected warming increases the risk of greater thermal stratification in some regions, which can lead to reduced O 2 ventilation [of underlying waters] and the formation of additional hypoxic zones, especially in the Baltic and Black seas medium 1 train +3725 AR6_WGII 437 21 These changes are likely to increase the risk of reduced bottom-water O2 levels to Baltic and Black Sea ecosystems (due to reduced solubility, increased stratification, and microbial respiration), which is very likely to affect fisheries.’ Persian Gulf, Red Sea: ‘Extreme temperature events, such as heat waves, are projected to increase (high confidence) [... and] temperatures are very likely to increase above established thresholds for mass coral bleaching and mortality very high 3 train +3726 AR6_WGII 437 22 In addition, geographic barriers, such as land boundaries [...] or lower oxygen water in deeper waters, are projected to limit species range shifts in SES, resulting in a larger relative decrease in species richness medium 1 train +3727 AR6_WGII 438 1 In the Baltic Sea, OMZs are affecting the extent of suitable spawning areas of cod, Gadus morhua (high confidence) (Hinrichsen et al., 2016), while in the Black Sea, the combined effect of OMZs and warming is influencing the distribution and physiology of fish species, and their migration and schooling behaviour in their overwintering grounds medium 1 train +3728 AR6_WGII 438 2 Cascading effects on food webs have been reported in the Baltic, where detrimental effects of changing oxygen levels on zooplankton production, pelagic and piscivorous fish are influencing seasonal succession and species composition of phytoplankton high 2 train +3729 AR6_WGII 438 3 In the Mediterranean Sea (Cross-Chapter Paper 4), the increase in climate extremes and mass-mortality events reported in AR5 has continued very high 3 train +3730 AR6_WGII 438 4 Extreme- weather events (including deep convection; González-Alemán et al., 2019) and MHWs have become more frequent (Darmaraki et al., 2019) and are associated with mass mortality of benthic sessile species across the basin high 2 train +3731 AR6_WGII 438 5 Since AR5, in the Persian Gulf and Red Sea, extreme temperatures, together with disease and predation, have continued to cause bleaching-induced mortality of corals, along with declines in the average coral-colony size high 2 train +3732 AR6_WGII 438 6 Poleward migration and tropicalisation of species (Section 3.4.2.3) has also continued in the Mediterranean, and these phenomena have also become an issue in the Black Sea high 2 train +3733 AR6_WGII 438 7 Climate impacts on phytoplankton production and phenology show high spatial heterogeneity across the Mediterranean Sea (medium evidence) (Marbà et al., 2015b; Salgado- Hernanz et al., 2019), with consequent effects on the diversity and abundance of zooplankton and fish species medium 1 train +3734 AR6_WGII 438 9 Evidence of impacts from ocean acidification is increasing, with the rates of coral calcification showing major decline in the Red Sea medium 1 train +3735 AR6_WGII 438 10 In the Mediterranean Sea, evidence of acidification events have been reported at the local scale (Hassoun et al., 2015), with impacts on bivalves and coralligenous species medium 1 train +3736 AR6_WGII 438 11 Climate models project increasing frequency and intensity of MHWs (high confidence) (Section 3.2.2.1), which will exacerbate warming- driven impacts in the Red Sea and Persian Gulf regions, and erode the resilience of Red Sea coral reefs high 2 train +3737 AR6_WGII 438 12 In the Persian Gulf region, extreme temperatures, >35°C (Pal and Eltahir, 2016), have been linked with high rates of extirpation and a decrease in fisheries catch potential medium 1 train +3738 AR6_WGII 438 22 Seasonal bottom-water hypoxia occurs in some shelf seas (e.g., northern Gulf of Mexico, Bohai Sea, East China Sea) due to riverine inputs of freshwater and nutrients, promoting stratification, enhanced primary production and organic carbon export to bottom waters high 2 train +3739 AR6_WGII 439 1 Eutrophication and seasonal bottom- water hypoxia in some shelf seas have been linked to warming (high confidence) (Wei et al., 2019; Del Giudice et al., 2020) and increased riverine nutrient loading high 2 train +3740 AR6_WGII 439 2 Since SROCC, some severe HABs have been attributed to extreme events, such as MHWs (Section 14.4.2; Roberts et al., 2019; Trainer et al., 2019); however, a recent worldwide assessment of HABs attributed the increase in observed HABs to intensified monitoring associated with increased aquaculture production high 2 train +3741 AR6_WGII 439 7 Changes in abundance, species composition and size of zooplankton have been detected in some shelf seas (Yellow Sea, North Sea, Celtic Sea and Tasman Sea), including a decline in stocks of larger copepods, increased abundances of gelatinous and meroplankton, and a shift to smaller species due to warming, increased river discharge, circulation change and/or extreme events (high confidence) (Wang et al., 2018a; Bedford et al., 2020; Evans et al., 2020; Shi et al., 2020; Edwards et al., 2021).Ocean warming has shifted distributions of fish (Free et al., 2019; Franco et al., 2020; Pinsky et al., 2020b; Fredston et al., 2021) and marine mammal species (Salvadeo et al., 2010; García-Aguilar et al., 2018; Davis et al., 2020) poleward high 2 train +3742 AR6_WGII 439 8 Warming has also tropicalised the pelagic and demersal fish assemblages of mid- and high-latitude shelves high 2 train +3743 AR6_WGII 439 9 Fisheries catch composition in many shelf-sea ecosystems has become increasingly dominated by warm- water species since the 1970s high 2 train +3744 AR6_WGII 439 12 For example, although, most species’ range edges are tracking temperature change on the northeast shelf of the USA medium 1 train +3745 AR6_WGII 439 15 For example, fisheries productivity Table 3.11 | Summary of past IPCC assessments of shelf seas Observations Projections AR5 (Hoegh-Guldberg et al., 2014) ‘Primary productivity, biomass yields and fish capture rates have undergone large changes within the ECS [East China Sea] over the past decades (limited evidence, medium agreement, low confidence).’ ‘Changing sea temperatures have influenced the abundance of phytoplankton, benthic biomass, cephalopod fisheries and the size of demersal trawl catches in the northern SCS [South China Sea] observed over the period 1976–2004 (limited evidence, medium agreement).’ ‘Concurrent with the retreat of the ‘cold pool’ [...] on the northern Bering Sea shelf, [...] bottom trawl surveys of fish and invertebrates show a significant community-wide northward distributional shift and a colonisation of the former cold pool areas by sub-Arctic fauna (high confidence).’ ‘Observed changes in the phenology of plankton groups in the North Sea over the past 50 years are driven by climate forcing, in particular regional warming (high confidence).’‘Global warming will result in more frequent extreme events and greater associated risks to ocean ecosystems high 2 train +3746 AR6_WGII 439 18 SROCC (Bindoff et al., 2019a) ‘Species composition of fisheries catches since the 1970s in many shelf seas ecosystems of the world is increasingly dominated by warm-water species medium 1 train +3747 AR6_WGII 439 19 Estuaries, shelf seas and a wide range of other intertidal and shallow-water habitats play an important role in the global carbon cycle through their primary production by rooted plants, seaweeds (macroalgae) and phytoplankton, and also by processing riverine organic carbon. The natural carbon dynamics of these systems have been greatly changed by human activities high 2 test +3748 AR6_WGII 440 5 Similar to the open ocean, large shifts in the phenology of phytoplankton blooms have been projected for shelf seas throughout subpolar and polar waters medium 1 train +3749 AR6_WGII 440 7 Trends towards tropicalisation will continue in the future high 2 train +3750 AR6_WGII 440 9 Under future climate change marked by more frequent and intense extreme events and the influences of multiple drivers, more flexible and adaptive management approaches could reduce climate impacts on species while also supporting industry adaptation high 2 train +3751 AR6_WGII 440 16 For example, trends in outputs from high-resolution, downscaled models in the California EBUS generally reflect those from underlying coarser-scale ESMs, but projections for physical variables are more convergent among modelling approaches than are those for biogeochemical variables high 2 train +3752 AR6_WGII 440 17 Models agree on general warming in the California EBUS, with concomitant declines in oxygen content medium 1 train +3753 AR6_WGII 441 1 More generally, changes in upwelling intensity are observed to affect organismal metabolism, population productivity and recruitment, and food-web structure medium 1 train +3754 AR6_WGII 441 4 Finally, although MHWs are an important emerging hazard in the global ocean, with intensity, Table 3.13 | Summary of previous IPCC assessments of eastern boundary upwelling systems (EBUS) Observations Projections AR5 (Hoegh-Guldberg et al., 2014; Lluch-Cota et al., 2014) ‘[EBUS] are vulnerable to changes that influence the intensity of currents, upwelling and mixing (and hence changes in sea surface temperature, wind strength and direction), as well as O 2 content, carbonate chemistry, nutrient content and the supply of organic carbon to deep offshore locations (high confidence).’ Climate-change-induced intensification of ocean upwelling in some EBUS, as observed in past decades, may lead to regional cooling, rather than warming, of surface waters and cause enhanced productivity medium 1 train +3755 AR6_WGII 441 5 Owing to contradictory observations, there is currently uncertainty about the future trends of major upwelling systems and how their drivers will shape ecosystem characteristics low 0 train +3756 AR6_WGII 441 7 These risks and uncertainties are likely to involve significant challenges for fisheries and associated livelihoods along the west coasts of South America, Africa and North America (low to medium confidence).’ ‘There is robust evidence and medium agreement that the California Current has experienced [...] an increase of the overall magnitude of upwelling events from 1967 to 2010 high 2 train +3757 AR6_WGII 441 8 This is consistent with changes expected under climate change yet remains complicated by the influence of decadal-scale variability low 0 test +3758 AR6_WGII 441 12 Ocean acidification and decrease in oxygen level in the California Current upwelling system have altered ecosystem structure, with direct negative impacts on biomass production and species composition (medium confidence).’ ‘Three out of the four major Eastern Boundary Upwelling Systems (EBUS) have shown large-scale wind intensification in the past 60 years high 2 train +3759 AR6_WGII 441 13 However, the interaction of coastal warming and local winds may have affected upwelling strength, with the direction of changes [varying] between and within EBUS low 0 test +3760 AR6_WGII 441 14 Increasing trends in ocean acidification in the California Current EBUS and deoxygenation in California Current and Humboldt Current EBUS are observed in the last few decades high 2 train +3761 AR6_WGII 441 15 The expanding California EBUS OMZ [oxygen minimum zone] has altered ecosystem structure and fisheries catches (medium confidence).’ ‘Overall, EBUS have been changing with intensification of winds that drives the upwelling, leading to changes in water temperature and other ocean biogeochemistry medium 1 train +3762 AR6_WGII 441 16 Moreover, the high natural variability of EBUS and their insufficient representation by global ESMs [Earth system models] gives low confidence that these observed changes can be attributed to anthropogenic causes.’‘Anthropogenic changes in EBUS will emerge primarily in the second half of the 21st century medium 1 train +3763 AR6_WGII 441 17 EBUS will be impacted by climate change in different ways, with strong regional variability with consequences for fisheries, recreation and climate regulation medium 1 train +3764 AR6_WGII 441 18 The Pacific EBUS are projected to have calcium carbonate undersaturation in surface waters within a few decades RCP8.5 (high confidence); combined with warming and decreasing oxygen levels, this will increase the impacts on shellfish larvae, benthic invertebrates, and demersal fishes (high confidence) and related fisheries and aquaculture (medium confidence).’ ‘The inherent natural variability of EBUS, together with uncertainties in present and future trends in the intensity and seasonality of upwelling, coastal warming and stratification, primary production and biogeochemistry of source waters poses large challenges in projecting the response of EBUS to climate change and to the adaptation of governance of biodiversity conservation and living marine resources in EBUS (high confidence).’ ‘Given the high sensitivity of the coupled human–natural EBUS to oceanographic changes, the future sustainable delivery of key ecosystem services from EBUS is at risk under climate change; those that are most at risk in the 21st century include fisheries (high confidence), aquaculture (medium confidence), coastal tourism (low confidence) and climate regulation (low confidence).’ ‘For vulnerable human communities with a strong dependence on EBUS services and low adaptive capacity, such as those along the Canary Current system, unmitigated climate-change effects on EBUS (complicated by other non-climatic stresses such as social unrest) have a high risk of altering their development pathways high 2 train +3765 AR6_WGII 442 1 Notwithstanding these trends, EBUS remain vulnerable both to MHWs (high confidence) (Sen Gupta et al., 2020) and to their long-lasting impacts high 2 train +3766 AR6_WGII 442 3 Despite low confidence in detailed projections for ecological changes in EBUS, the WGI assessment (WGI AR6 Chapter 9; Fox-Kemper et al., 2021) that upwelling-favourable winds will weaken (or be present for shorter durations) at low latitude but intensify at high latitude (high confidence), albeit by no more than 20% in either case medium 1 train +3767 AR6_WGII 442 11 Over the past two decades, Arctic Ocean surface temperature has increased in line with the global average, while there has been no uniform warming across the Antarctic high 2 train +3768 AR6_WGII 442 12 Thus, the rate of change due to warming, and associated sea ice loss, is greater in the Arctic than in the Antarctic high 2 train +3769 AR6_WGII 442 15 Previous assessments of polar seas (Table 3.14) concluded that climate change has already profoundly influenced polar ecosystems, through changing species distributions and abundances from primary producers to top predators, including both ecologically and economically important species (high confidence), and that it will continue to do so (Table 3.14).Since SROCC, evidence demonstrates that warmer oceans, less sea ice and increased advection results in increasing primary production in the Arctic, albeit with regional variation (high confidence), while trends remain spatially heterogeneous and less clear in the Antarctic medium 1 train +3770 AR6_WGII 442 17 Major community shifts, both gradual and abrupt, are observed in polar oceans in response to warming trends and MHWs (Arctic only) high 2 train +3771 AR6_WGII 442 18 In general, abundances and ranges of Arctic fish species are declining and contracting, while ranges of boreal fish species are expanding, both geographically and in terms of feeding interactions and ecological roles high 2 train +3772 AR6_WGII 442 21 Phenological, behavioural, physiological and distributional changes are observed in marine mammals and birds in response to altered ecological interactions and habitat degradation, especially to loss of sea ice high 2 train +3773 AR6_WGII 442 22 Reproductive failures and declining abundances attributed to warmer polar oceans and less sea ice cover are observed in populations of polar bears, Ursus maritimus, seals, whales and marine birds high 2 train +3774 AR6_WGII 442 24 Calcareous polar organisms are among the groups most sensitive to ocean acidification high 2 train +3775 AR6_WGII 443 23 Impacts of shifts in ocean conditions affect fish and shellfish abundances in the Arctic high 2 train +3776 AR6_WGII 443 24 Changes in sea ice and the physical environment to the west of the Antarctic Peninsula are altering phytoplankton stocks and productivity, and krill (high confidence).Some marine species will shift their ranges in response to changing ocean and sea ice conditions in the polar regions medium 1 train +3777 AR6_WGII 443 25 Loss of sea ice in summer and increased ocean temperatures are expected to impact secondary pelagic production in some regions of the Arctic Ocean, with associated changes in the energy pathways within the marine ecosystem medium 1 train +3778 AR6_WGII 443 26 Ocean acidification has the potential to inhibit embryo development and shell formation of some zooplankton and krill in the polar regions, with potentially far-reaching consequences to food webs in these regions medium 1 train +3779 AR6_WGII 443 27 Shifts in the timing and magnitude of seasonal biomass production could disrupt coupled phenologies in the food webs, leading to decreased survival of dependent species medium 1 train +3780 AR6_WGII 443 28 SR15 (Hoegh-Guldberg et al., 2018a) ‘A fundamental transformation is occurring in polar organisms and ecosystems, driven by climate change high 2 train +3781 AR6_WGII 443 29 There is high agreement and robust evidence that phytoplankton species will change because of sea ice retreat and related changes in temperature and radiation, and this is very likely to benefit fisheries productivity [in the Arctic spring bloom system].’ ‘‘Unique and threatened systems’ (RFC1), [including Arctic and coral reefs], display a transition from high to very high risk of transition at temperatures between 1.5°C and 2°C of global warming, as opposed to at 2.6°C of global warming in AR5 (high confidence).’ SROCC (Bindoff et al., 2019a) Climate-induced changes in seasonal sea ice extent and thickness as well as ocean stratification are altering marine primary production (high confidence), with impacts on ecosystems medium 1 train +3782 AR6_WGII 443 30 Changes in the timing, duration and magnitude of primary production have occurred in both polar oceans, with marked regional or local variability high 2 train +3783 AR6_WGII 443 31 In both polar regions, climate-induced changes in ocean and sea ice conditions have expanded the range of temperate species and contracted the range of polar fish and ice-associated species high 2 train +3784 AR6_WGII 443 32 Ocean acidification will affect several key Arctic species (medium confidence).Future climate-induced changes in the polar oceans, sea ice, snow and permafrost will drive habitat and biome shifts, with associated changes in the ranges and abundance of ecologically important species medium 1 train +3785 AR6_WGII 443 33 Projected range expansion of sub-Arctic marine species will increase pressure for high-Arctic species medium 1 train +3786 AR6_WGII 443 34 Both polar oceans will be increasingly affected by CO 2 uptake, causing corrosive conditions for calcium carbonate shell-producing organisms (high confidence), with associated impacts on marine organisms and ecosystems medium 1 train +3787 AR6_WGII 443 35 The projected effects of climate-induced stressors on polar marine ecosystems present risks for commercial and subsistence fisheries, with implications for regional economies, cultures and the global supply of fish, shellfish, and Antarctic krill high 2 train +3788 AR6_WGII 444 4 While levels of pollutants in biota (e.g., persistent organic pollutants, mercury) have generally declined over the past decades, recent increasing levels are associated with release from reservoirs in ice, snow and permafrost, and through changing food webs and pathways for trophic amplification medium 1 test +3789 AR6_WGII 444 5 Also, a warmer climate, altered ocean currents and increased human activities elevate the risk of invasive species in the Arctic (medium confidence), potentially changing ecosystems in this region high 2 train +3790 AR6_WGII 444 7 Fisheries are largely sustainably managed yet are expanding polewards following sea ice melt in the Arctic high 2 train +3791 AR6_WGII 444 8 Tourism is increasing and expanding in both polar regions, while shipping and hydrocarbon exploration are growing in the Arctic, increasing the risks of compound effects on vulnerable and already stressed populations and ecosystems high 2 train +3792 AR6_WGII 444 14 Nevertheless, increasing evidence supports that sustainable and adaptive ecosystem-based fisheries practices can reduce detrimental impacts of climate change on harvested populations medium 1 train +3793 AR6_WGII 445 1 On average, species’ distributions have shifted poleward by 72.0 ± 0.35 km per decade (high confidence).Spatial shifts of marine species due to projected warming will cause high-latitude invasions and high local-extinction rates in the tropics and semi-enclosed seas medium 1 train +3794 AR6_WGII 445 2 SROCC (Bindoff et al., 2019a) ‘Ocean warming has contributed to observed changes in biogeography of organisms ranging from phytoplankton to marine mammals (high confidence).’ ‘The direction of the majority of the shifts of epipelagic organisms are consistent with a response to warming (high confidence)’ but are also shaped by oxygen concentrations and ocean currents across depth, latitudinal and longitudinal gradients high 2 train +3795 AR6_WGII 445 3 Geographic ranges have shifted since the 1950s by 51.5 ± 33.3 km per decade (mean and very likely range) and 29.0 ± 15.5 km per decade for organisms in the epipelagic and seafloor ecosystems, respectively.‘Recent model projections since AR5 and SR15 continue to support global-scale range shifts of marine fishes at rates of tens to hundreds of km per decade in the 21st century, with rate of shifts being substantially higher under RCP8.5 than RCP2.6.’ range shifts consistent with climate change) estimates that marine species are moving poleward at a rate of 59.2 km per decade (very likely range: 43.7–74.7 km per decade), closely matching the local climate velocity high 2 train +3796 AR6_WGII 445 5 Biogeographic shifts lead to novel communities and biotic interactions (high confidence) (Zarco-Perello et al., 2017; Pecuchet et al., 2020b), with concomitant changes in ecosystem functioning and servicing high 2 train +3797 AR6_WGII 447 1 These changes subsequently affect biogenic carbon cycling through alteration of microbial remineralisation and carbon sequestration in deep water medium 1 train +3798 AR6_WGII 447 9 Although vertical redistribution of thermal niches is three to four orders of magnitude slower than horizontal displacement, maximum depth limits imposed by the seafloor and photic layer (both of which are projected to be reached in this century) will likely vertically compress suitable habitat for most marine organisms medium 1 train +3799 AR6_WGII 447 11 The volume of OMZs have been expanding at many locations (high confidence), and the oxygen content of the subsurface ocean is projected to decline to historically unprecedented conditions over the 21st century medium 1 train +3800 AR6_WGII 447 14 Projections also suggest that warming-related increases in trophic efficiency lead to a 17% increase in the biomass of the deep-scattering layer (zooplankton and fish in the mesopelagic) by 2100 low 0 test +3801 AR6_WGII 447 15 Observational studies appear to show that mesopelagic fishes adapted to warm water increased in abundance and distribution in the California Current associated with warming and the expansion of OMZ (Koslow et al., 2019), suggesting that some mesopelagic fish stocks might be resilient to a changing climate medium 1 train +3802 AR6_WGII 447 19 Moreover, different trophic levels within epipelagic food webs are responding at different rates very high 3 train +3803 AR6_WGII 448 4 There is high agreement in model projections that the start of the phytoplankton Table 3.18 | Summary of previous IPCC assessments of phenological shifts and trophic mismatches Observations Projections AR5 WGII (Hoegh-Guldberg et al., 2014; Larsen et al., 2014) ‘Changes to sea temperature have altered the phenology, or timing of key life-history events such as plankton blooms, and migratory patterns, and spawning in fish and invertebrates, over recent decades medium 1 train +3804 AR6_WGII 448 5 There is medium to high agreement that these changes pose significant uncertainties and risks to fisheries, aquaculture and other coastal activities.’ The highly productive high-latitude spring bloom systems in the northeast Atlantic are responding to warming (medium evidence, high agreement), with the greatest changes being observed since the late 1970s in the phenology, distribution and abundance of plankton assemblages, and the reorganisation of fish assemblages, with a range of consequences for fisheries high 2 train +3805 AR6_WGII 448 8 This impact would be exacerbated if shifts in timing occur rapidly medium 1 train +3806 AR6_WGII 448 9 SROCC (Bindoff et al., 2019a) ‘Phenology of marine ectotherms in the epipelagic systems related to ocean warming (high confidence) and the timing of biological events has shifted earlier high 2 train +3807 AR6_WGII 448 10 WGI AR6 Chapter 2 (Gulev et al., 2021) ‘Phenological metrics for many species of marine organisms have changed in the last half century high 2 train +3808 AR6_WGII 448 11 The changes vary with location and with species high 2 train +3809 AR6_WGII 448 12 There is a strong dependence of survival in higher trophic-level organisms (fish, exploited invertebrates, birds) on the availability of food at various stages in their life cycle, which in turn depends on the phenologies of both high 2 train +3810 AR6_WGII 448 16 Overall, the regional patterns are qualitatively similar under SSP1-2.6 and SSP5-8.5 but with greater magnitude and larger areas under SSP5-8.5 low 0 train +3811 AR6_WGII 448 18 Furthermore, under RCP8.5, trophic mismatch events exceeding ±30 days (Asch et al., 2019) leading to fish-recruitment failure are expected to increase tenfold for geographic spawners across much of the North Atlantic, North Pacific and Arctic Ocean basins low 0 train +3812 AR6_WGII 449 3 Temperatures during the last Interglacial (~125 ka), which were warmer than today, led to poleward range shifts of reef corals medium 1 train +3813 AR6_WGII 449 4 Temperature has also driven marine range shifts over multi-million-year time scales medium 1 train +3814 AR6_WGII 449 5 Warming climates, even with low ocean- warming rates, inevitably decreased tropical marine biodiversity compared with middle latitudes high 2 train +3815 AR6_WGII 449 6 The paleorecord confirms that marine biodiversity has been vulnerable to climate warming both globally and regionally very high 3 train +3816 AR6_WGII 449 7 In extreme cases of warming (e.g., >5.2°C), marine mass extinctions occurred in the geological past, and there may be a relationship between warming magnitude and extinction toll medium 1 train +3817 AR6_WGII 449 8 A combination of warming and spreading anoxia caused marine extinctions in ancient episodes of rapid climate warming high 2 train +3818 AR6_WGII 449 9 The role of ocean acidification in ancient extinctions is yet to be resolved low 0 train +3819 AR6_WGII 451 4 At the community level, the magnitude and shape of projected future biodiversity changes differ depending on which groups are considered medium 1 train +3820 AR6_WGII 452 5 On longer time scales, alteration of energy flow through marine food webs may lead to ecological tipping points (Wernberg et al., 2016; Harley et al., 2017) after which the food web collapses into shorter, bottom-heavy trophic pyramids medium 1 train +3821 AR6_WGII 452 7 Table 3.20 | Summary of previous IPCC assessments of community composition and biodiversity Observations Projections AR5 (Hoegh-Guldberg et al., 2014; Pörtner et al., 2014) The paleoecological record shows that global climate changes comparable in magnitudes to those projected for the 21st century under all scenarios resulted in large-scale biome shifts and changes in community composition, and that for rates projected under RCP6 and 8.5 those changes were associated with species extinctions in some groups high 2 train +3822 AR6_WGII 452 8 Loss of corals due to bleaching has a potentially critical influence on the maintenance of marine biodiversity in the tropics (high confidence).Spatial shifts of marine species due to projected warming will cause high-latitude invasions and high local-extinction rates in the tropics and semi-enclosed seas medium 1 train +3823 AR6_WGII 452 9 Species richness and fisheries catch potential are projected to increase, on average, at mid and high latitudes (high confidence) and decrease at tropical latitudes medium 1 train +3824 AR6_WGII 452 11 Thereby, climate change will reassemble communities and affect biodiversity, with differences over time and between biomes and latitudes high 2 train +3825 AR6_WGII 452 12 However, specific quantitative projections by these models remain imprecise (low confidence).’ SROCC (Bindoff et al., 2019a) ‘Ocean warming has contributed to observed changes in biogeography of organisms ranging from phytoplankton to marine mammals (high confidence), consequently changing community composition (high confidence), and in some cases altering interactions between organisms and ecosystem function medium 1 train +3826 AR6_WGII 452 14 In addition, geographic barriers, such as land, [bounding the] poleward species range edge in semi-enclosed seas or low-oxygen water in deeper waters are projected to limit range shifts, resulting in a larger relative decrease in species richness (medium confidence).’ ‘The large variation in sensitivity between different zooplankton taxa to future conditions of warming and ocean acidification suggests elevated risk on community structure and inter-specific interactions of zooplankton in the 21st century medium 1 train +3827 AR6_WGII 452 15 Under continuing climate change, the projected loss of biodiversity may ultimately threaten marine ecosystem stability (medium confidence) (Albouy et al., 2020; Nagelkerken et al., 2020; Henson et al., 2021), altering both the functioning and structure of marine ecosystems and thus affecting service provisioning medium 1 train +3828 AR6_WGII 453 3 Abrupt ecosystem shifts have been observed in both large open-ocean ecosystems and coastal ecosystems (Section 3.4.2), with dramatic social consequences through significant loss of diverse ecosystem services high 2 train +3829 AR6_WGII 453 5 Abrupt ecosystem shifts are associated with large-scale patterns of climate variability (Alheit et al., 2019; Beaugrand et al., 2019; Lehodey et al., 2020), some of which are projected to intensify with climate change medium 1 train +3830 AR6_WGII 453 6 Over the past 60 years, abrupt ecosystem shifts have generally followed El Niño/ Southern Oscillation events of any strength, but some periods had geographically limited ecological shifts (~0.25% of the global ocean in 1984–1987) and others more extensive shifts (14% of the global ocean in 2012–2015) medium 1 train +3831 AR6_WGII 453 7 Typically, interacting drivers, such as eutrophication and overharvest, reduce ecosystem resilience to climate extremes (e.g., MHWs, cyclones) or gradual warming, and hence promote ecosystem shifts high 2 train +3832 AR6_WGII 453 8 Shifts in different ecosystems may be connected through common drivers or through cascading effects medium 1 test +3833 AR6_WGII 453 9 Recent MHWs (Section 3.2.2.1) have caused major ecosystem shifts and mass mortality in oceanic and coastal ecosystems, including corals, kelp forests and seagrass meadows (Sections 3.4.2.1, 3.4.2.3, 3.4.2.5, 3.4.2.6, 3.4.2.10; Cross-Chapter Box MOVING SPECIES in Chapter 5; Cross-Chapter Box EXTREMES in Chapter 2), with dramatic declines in species foundational for habitat formation or trophic flow, biodiversity declines, and biogeographic shifts in fish stocks very high 3 train +3834 AR6_WGII 454 10 Marine birds and mammals are vulnerable to climate-induced loss of breeding and foraging habitats such as sea ice (Section 3.4.2.12), sandy beaches (Section 3.4.2.6), salt marshes (Section 3.4.2.5) and seagrass beds high 2 train +3835 AR6_WGII 454 12 Marine mammals dependent on sea ice habitat are particularly vulnerable to warming medium 1 train +3836 AR6_WGII 455 1 Nevertheless, even under an intermediate emission scenario RCP4.5, increasing ice-free periods will likely reduce both recruitment and adult survival across most polar bear populations over the next four decades, threatening their existence medium 1 train +3837 AR6_WGII 455 2 Climate change is affecting marine food-web dynamics high 2 train +3838 AR6_WGII 455 3 Higher-vulnerability species include central-place foragers (confined to, for example, breeding colonies fixed in space), diet and habitat specialists, and species with restricted distributions such as marine mammal populations in SES medium 1 train +3839 AR6_WGII 455 4 Surface-feeding and piscivorous marine birds appear to be more vulnerable to food-web changes than diving seabirds and planktivorous seabirds medium 1 train +3840 AR6_WGII 455 6 Marine birds are vulnerable to phenological shifts in food-web dynamics, as they have limited phenotypic plasticity of reproductive timing, with potentially little scope for evolutionary adaptation medium 1 train +3841 AR6_WGII 456 2 Also, climate-change driven distributional shifts have strengthened interactions with other anthropogenic impacts, through, for example, increasing risks of ship strikes and bycatch (medium confidence) (e.g., Hauser et al., 2018; Krüger et al., 2018; Record et al., 2019; Santora et al., 2020).Box 3.2 (continued) the risk of abrupt ecosystem shifts high 2 train +3842 AR6_WGII 456 4 However, where climate change is a dominant driver, ecosystem collapses increasingly cause permanent transitions high 2 train +3843 AR6_WGII 456 5 Over the coming decades, MHWs are projected to very likely become more frequent under all emission scenarios (Section 3.2; WGI AR6 Chapter 9; Fox-Kemper et al., 2021), with intensities and rates too high for recovery of degraded foundational species, habitats or biodiversity medium 1 train +3844 AR6_WGII 456 9 In contrast, the anthropogenic signal in phytoplankton community structure, which has a lower natural variability, will emerge under RCP8.5 across 63% of the ocean by 2100 when two standard deviations are used (limited evidence) (Dutkiewicz et al., 2019).Table 3.21 | Summary of previous IPCC assessments of observed and projected abrupt ecosystem shifts and extreme events Observations Projections AR5 (Wong et al., 2014) Observations of abrupt ecosystem shifts and extreme events were not assessed in this report.‘Warming and acidification will lead to coral bleaching, mortality, and decreased constructional ability high 2 train +3845 AR6_WGII 456 10 Temperate seagrass and kelp ecosystems will decline with the increased frequency of heatwaves and sea temperature extremes as well as through the impact of invasive subtropical species (high confidence).’ SROCC (Collins et al., 2019a) ‘Marine heatwaves (MHWs), periods of extremely high ocean temperatures, have negatively impacted marine organisms and ecosystems in all ocean basins over the last two decades, including critical foundation species such as corals, seagrasses and kelps (very high confidence).’‘Marine heatwaves are projected to further increase in frequency, duration, spatial extent and intensity (maximum temperature) very high 3 train +3846 AR6_WGII 456 11 Climate models project increases in the frequency of marine heatwaves by 2081–2100, relative to 1850–1900, by approximately 50 times under RCP8.5 and 20 times under RCP2.6 medium 1 train +3847 AR6_WGII 456 12 Extreme El Niño events are projected to occur about twice as often under both RCP2.6 and RCP8.5 in the 21st century when compared to the 20th century (medium confidence).’ ‘Limiting global warming would reduce the risk of impacts of MHWs, but critical thresholds for some ecosystems (e.g., kelp forests, coral reefs) will be reached at relatively low levels of future global warming high 2 train +3848 AR6_WGII 457 2 Better accounting for multiple interacting factors in ESMs (see Box 3.1) will provide insight into how marine ecosystems will respond to future climate high 2 train +3849 AR6_WGII 457 14 Using an ensemble of global-scale marine ecosystem and fisheries models (Fish-MIP) (Tittensor et al., 2018) with the CMIP5 ESM ensemble, SROCC concludes that projected ocean warming and decreased phytoplankton production and biomass will reduce global marine animal biomass during the 21st century medium 1 train +3850 AR6_WGII 460 1 These declines result from combined warming and decreased primary production (with low confidence in future changes in primary production; Section 3.4.3.5) and are amplified at each trophic level within all ESM and marine ecosystem model projections across all scenarios medium 1 train +3851 AR6_WGII 461 3 Owing to contradictory observations there is currently uncertainty about the future trends of major upwelling systems and how their drivers (enhanced productivity, acidification and hypoxia) will shape ecosystem characteristics low 0 test +3852 AR6_WGII 461 4 Animal biomass Observed changes in animal biomass were not assessed in this report.‘The climate-change-induced intensification of ocean upwelling in some eastern boundary systems, as observed in the last decades, may lead to regional cooling, rather than warming, of surface waters and cause enhanced productivity medium 1 train +3853 AR6_WGII 461 8 The strong dependence of the projected declines on phytoplankton production low 0 train +3854 AR6_WGII 461 11 Some increases are projected in the polar regions, due to enhanced stratification in the surface ocean, reduced primary production and shifts towards small phytoplankton medium 1 train +3855 AR6_WGII 462 4 Overall, ocean warming and decreased phytoplankton production and biomass will drive a global decline in biomass for zooplankton (low confidence), marine animals (medium confidence) and seafloor benthos (low confidence), with regional differences in the direction and magnitude of changes high 2 train +3856 AR6_WGII 462 5 There is increasing evidence that responses will amplify throughout the food web and at ocean depths, with relatively modest changes in surface primary producers translating into substantial changes at higher trophic levels and for deep-water benthic communities medium 1 train +3857 AR6_WGII 462 9 This is consistent with previous assessments that identified ocean warming and increased stratification as the main drivers high 2 train +3858 AR6_WGII 462 16 Furthermore, accurate simulation of many of the biogeochemical tracers upon which NPP depends (e.g., the distribution of iron; Tagliabue et al., 2016; Bindoff et al., 2019a) remains a significant and ongoing challenge to ESMs high 2 train +3859 AR6_WGII 464 1 Increasing model complexity with more widespread representation of ocean biogeochemical processes between CMIP5 and CMIP6, and inclusion of more than one or two classes of phyto- and zooplankton, will provide opportunities to improve assessments of how climate-induced drivers affect different components of biological carbon pump in the epipelagic ocean, as well as changes in the efficiency and magnitude of carbon export in the deep ocean high 2 train +3860 AR6_WGII 464 5 Model results indicate that sea surface temperatures (high confidence), Arctic sea ice (high confidence), surface ocean acidification (very high confidence) and surface ocean deoxygenation very high 3 train +3861 AR6_WGII 464 7 In an overshoot scenario in which CO 2 returns to 2040 levels by 2100 (SSP5-3.4-OS; O’Neill et al., 2016), SST and Arctic sea ice do not fully return by 2100 to levels prior to the CO 2 peak medium 1 train +3862 AR6_WGII 464 8 Models also indicate that global sea level rise, as well as warming, ocean acidification and deoxygenation at depth, are irreversible for centuries or longer very high 3 train +3863 AR6_WGII 464 11 A small decrease in productivity is evident globally for the period 1998–2015, but regional changes are larger and of opposing signs low 0 train +3864 AR6_WGII 465 1 The deep sea covers >63% of Earth’s surface (Costello and Cheung, 2010) and is exposed to climate-driven changes in abyssal, intermediate and surface waters that influence sinking fluxes of particulate organic matter high 2 train +3865 AR6_WGII 465 7 Acute mortality of some reef-forming cold-water corals to laboratory-simulated warming (Lunden et al., 2014) suggests that both long-term warming and the increase of MHWs in intermediate and deep waters (Elzahaby and Schaeffer, 2019) could pose significant risk to associated ecosystems high 2 test +3866 AR6_WGII 465 9 The extension and intensification of deep-water acidification (Section 3.2.3.1) has been identified as a further key risk to deep-water coral ecosystems medium 1 train +3867 AR6_WGII 465 12 Desmophyllum pertusum7and Madrepora oculata maintain calcification in moderately low pH (7.75) and near-saturation of aragonite (Hennige et al., 2014; Maier et al., 2016; Büscher et al., 2017), but lower pH (7.6) and corrosive conditions lead to net dissolution of D. pertusum skeletons high 2 train +3868 AR6_WGII 465 15 In OMZ regions (Section 3.2.3.2), benthic species distributions (Sperling et al., 2016; Levin, 2018; Gallo et al., 2020), abundance and composition of demersal fishes in canyons (De Leo et al., 2012) and deep-pelagic zooplankton (Wishner et al., 2018) follow oxygen gradients, indicating that deep-sea biodiversity and ecosystem structure will be impacted by extension of hypoxic areas medium 1 train +3869 AR6_WGII 465 18 Despite mortality and functional impacts from low oxygen concentrations observed in aquaria (Lunden et al., 2014), recent observations of the deep-water coral D. pertusum suggest adaptive capacity to hypoxia among specimens from OMZ regions that are highly productive low 0 train +3870 AR6_WGII 466 3 Climate-driven impacts further limit the resilience of deep-sea ecosystems to impacts from human activities high 2 train +3871 AR6_WGII 466 6 The spatial resolution of CMIP5 models is too coarse to robustly project changes in mesoscale circulation at the seafloor (Sulpis et al., 2019), on which deep-sea ecosystems depend for organic material supplies and dispersal of planktonic and planktotrophic larvae high 2 train +3872 AR6_WGII 467 5 Biodiversity has changed in association with ocean warming and loss of sea ice, sea level rise, coral bleaching, marine heat waves and upwelling changes high 2 test +3873 AR6_WGII 467 6 Overlapping non-climate drivers (Section 3.1) also decrease ocean and coastal ecosystem biodiversity very high 3 train +3874 AR6_WGII 467 9 Projected changes in biodiversity due to climate change (Section 3.4.3.3.3) are expected to alter the flow and array of ocean and coastal ecosystem services high 2 train +3875 AR6_WGII 467 10 Non-indigenous marine species are major agents of ocean and coastal biodiversity change, and climate and non-climate drivers interact to support their movement and success high 2 train +3876 AR6_WGII 467 11 At times, non-indigenous species act invasively and outcompete indigenous species, causing regional biodiversity shifts and altering ecosystem function, as seen in the Mediterranean region high 2 train +3877 AR6_WGII 467 14 Non-climate drivers, especially marine shipping in newly ice-free locations (Chan et al., 2019), fishing pressure (Last et al., 2011), aquaculture of non-indigenous species (Mach et al., 2017; Ruby and Ahilan, 2018) and marine pollution and debris (Gall and Thompson, 2015; Carlton et al., 2018; Carlton and Fowler, 2018; Lasut et al., 2018; Miralles et al., 2018; Rech et al., 2018; Therriault et al., 2018), promote range shifts and movement of non-indigenous species high 2 train +3878 AR6_WGII 467 16 Invasive marine species can alter species behaviour, reduce indigenous species abundance, reduce water clarity, bioaccumulate more heavy metals than indigenous species and inhibit ecosystem resilience in the face of extreme events medium 1 train +3879 AR6_WGII 468 23 Catch composition is changing in many locations fished by smaller-scale, less-mobile commercial, artisanal and recreational fisheries high 2 train +3880 AR6_WGII 468 25 Where possible, fishers are maintaining harvests by broadening catch diversity, traveling poleward and changing gear and strategies high 2 train +3881 AR6_WGII 469 1 Both positive and negative impacts result for food security through fisheries (medium confidence), local cultures and livelihoods (medium confidence), and tourism and recreation medium 1 train +3882 AR6_WGII 469 2 The impacts on ecosystem services have negative consequences for health and well-being (medium confidence), and for Indigenous Peoples and local communities dependent on fisheries (high confidence) (1.1, 1.5, 3.2.1, 5.4.1, 5.4.2, Figure SPM.2)’ (SROCC SPM A.8; IPCC, 2019c).‘Long-term loss and degradation of marine ecosystems compromises the ocean’s role in cultural, recreational, and intrinsic values important for human identity and well-being medium 1 train +3883 AR6_WGII 469 3 Biodiversity (Section 3.5.2)‘[Climate] Impacts are already observed on [coastal ecosystem] habitat area and biodiversity, as well as ecosystem functioning and services high 2 train +3884 AR6_WGII 469 4 Food provision (Section 3.5.3)‘Warming-induced changes in the spatial distribution and abundance of some fish and shellfish stocks have had positive and negative impacts on catches, economic benefits, livelihoods, and local culture high 2 train +3885 AR6_WGII 469 5 There are negative consequences for Indigenous Peoples and local communities that are dependent on fisheries high 2 train +3886 AR6_WGII 469 6 Shifts in species distributions and abundance has challenged international and national ocean and fisheries governance, including in the Arctic, North Atlantic and Pacific, in terms of regulating fishing to secure ecosystem integrity and sharing of resources between fishing entities (high confidence) (3.2.4, 3.5.3, 5.4.2, 5.5.2, Figure SPM.2)’ (SROCC SPM A.8.1; IPCC, 2019c).‘Future shifts in fish distribution and decreases in their abundance and fisheries catch potential due to climate change are projected to affect income, livelihoods, and food security of marine resource-dependent communities medium 1 train +3887 AR6_WGII 469 7 Long-term loss and degradation of marine ecosystems compromises the ocean’s role in cultural, recreational, and intrinsic values important for human identity and well-being medium 1 test +3888 AR6_WGII 469 9 The emerging demand for alternative energy sources is expected to generate economic opportunities for the ocean renewable energy sector (high confidence), although their potential may also be affected by climate change low 0 train +3889 AR6_WGII 469 10 Habitat creation and maintenance (Section 3.5.5.1)‘[Climate] Impacts are already observed on [coastal ecosystem] habitat area and biodiversity, as well as ecosystem functioning and services high 2 train +3890 AR6_WGII 469 13 Climate regulation and air quality (Section 3.5.5.2)‘Global ocean heat content continued to increase throughout [the 1951 to present] period, indicating continuous warming of the entire climate system very high 3 train +3891 AR6_WGII 469 15 This is projected to result in a higher proportion of emitted CO 2 remaining in the atmosphere high 2 train +3892 AR6_WGII 470 2 Technology-based adaptations (Section 3.6.3) have minimised aquaculture losses from ocean acidification, including early-warning systems to guide hatchery operations and culturing resilient shellfish Ecosystem service and chapter subsectionObserved impacts Projected impacts Observed impacts on marine organisms’ contribution to climate regulation not previously assessed.‘The effect of climate change on marine biota will alter their contribution to climate regulation, that is, the maintenance of the chemical composition and physical processes in the atmosphere and oceans high 2 train +3893 AR6_WGII 470 3 Provision of freshwater, maintenance of water quality, regulation of pathogens (Section 3.5.5.3)Observed climate impacts on salinisation of coastal soil and groundwater not previously assessed.‘In the absence of more ambitious adaptation efforts compared to today, and under current trends of increasing exposure and vulnerability of coastal communities, risks, such as erosion and land loss, flooding, salinisation, and cascading impacts due to mean sea level rise and extreme events are projected to significantly increase throughout this century under all greenhouse gas emissions scenarios very high 3 train +3894 AR6_WGII 470 8 Regulation of physical hazards (Section 3.5.5.4)‘Coastal ecosystems are already impacted by the combination of sea level rise, other climate-related ocean changes, and adverse effects from human activities on ocean and land (high confidence)... Coastal and near-shore ecosystems including saltmarshes, mangroves, and vegetated dunes in sandy beaches,...provide important services including coastal protection...(high confidence)’ (SROCC Chapter 4 Executive Summary; Oppenheimer et al., 2019).‘The decline in warm water coral reefs is projected to greatly compromise the services they provide to society, such as...coastal protection high 2 train +3895 AR6_WGII 470 10 However, the effect of these changes is not yet reflected in a weakening trend of the contemporary (1960–2019) ocean sink high 2 train +3896 AR6_WGII 470 14 Cultural services (Section 3.5.6)‘Climate change impacts on marine ecosystems and their services put key cultural dimensions of lives and livelihoods at risk medium 1 train +3897 AR6_WGII 470 15 This includes potentially rapid and irreversible loss of culture and local knowledge and Indigenous knowledge, and negative impacts on traditional diets and food security, aesthetic aspects, and marine recreational activities (medium confidence)’ (SROCC SPM B.8.4; IPCC, 2019c).‘Future shifts in fish distribution and decreases in their abundance and fisheries catch potential due to climate change are projected to affect income, livelihoods, and food security of marine resource-dependent communities medium 1 train +3898 AR6_WGII 470 16 Long-term loss and degradation of marine ecosystems compromises the ocean’s role in cultural, recreational, and intrinsic values important for human identity and well-being medium 1 train +3899 AR6_WGII 471 1 Laboratory studies show that ocean acidification decreases the fitness, growth or survival of many economically and culturally important larval or juvenile shelled mollusks high 2 train +3900 AR6_WGII 471 2 Ocean acidification alters larval settlement and metamorphosis of fish in laboratory studies high 2 train +3901 AR6_WGII 471 6 Ocean and coastal organisms will continue moving poleward under RCP8.5 (high confidence) (Section 3.4.3.1.3; Figure 3.18), and this is expected to decrease fisheries harvests in low latitudes and alter species composition and abundance in higher latitudes high 2 train +3902 AR6_WGII 471 8 Temperature will continue to be a major driver of fisheries changes globally, but other non-climate factors like organism physiology and ecosystem response (Section 3.3) and fishing pressure (Chapter 5), as well as other climate-induced drivers like acidification, deoxygenation and sea ice loss (Section 3.2), will play critical roles in future global and local fisheries changes high 2 train +3903 AR6_WGII 471 9 Warming, acidification and business-as-usual fishing policy under RCP8.5 are projected to place around 60% of global fisheries at very high risk medium 1 train +3904 AR6_WGII 471 10 Model intercomparison showed that ocean acidification and protection affect ecosystems more than fishing pressure, and ecological adaptation will significantly determine impacts on fishery biomass, catch and value until approximately 2050 medium 1 train +3905 AR6_WGII 471 14 Declining stocks of forage fish are expected to have detrimental effects on seabirds, pelagic fish and marine mammals (medium confidence) (Lindegren et al., 2018; Steiner et al., 2019), which may harm dependent human communities, including Arctic Indigenous Peoples low 0 train +3906 AR6_WGII 471 15 Modelled fishery futures and revenue depend on environmental scenario, fishing- fleet composition and management, and ocean acidification and temperature responses of harvested species high 2 train +3907 AR6_WGII 471 17 At the same time, projected hypoxic conditions of ~2 mg l–1 of oxygen will be consistently detrimental across taxonomic groups, developmental stages and climate regions high 2 train +3908 AR6_WGII 472 3 Because ooligan spawning relies on the timing of the spring freshet, and because the species has declined in the past 25 years due to fishing pressure and predation, the species may be at risk from combined climate-induced and non-climate drivers medium 1 test +3909 AR6_WGII 472 14 Decreasing the vulnerability of renewable-energy installations, particularly wind turbines, to climate risks (Table 3.26; Bindoff et al., 2019a) could include technological adaptations (Section 3.6.2.2) such as storm ‘survival mode’ settings (Penalba et al., 2018); preparation for hazards such as icing, SLR, drifting sea ice and wave activity (Neill et al., 2018; Goodale and Milman, 2019; Solaun and Cerdá, 2019); and biofouling (medium confidence) (Want and Porter, 2018; Joyce et al., 2019; Vinagre et al., 2020), which is expected to increase in response to warming and acidification medium 1 train +3910 AR6_WGII 472 23 Over the 21st century, ocean heat and CO 2 uptake will continue (WGI AR6 SPMB4.1, B5.1; IPCC, 2021b) and sea ice loss from warming will allow some additional CO 2 uptake (Armstrong et al., 2019), but the ocean will take up a smaller fraction of CO 2 emissions as atmospheric CO 2 concentrations rise high 2 train +3911 AR6_WGII 473 1 Coastal salinisation is attributed to regionally varying combinations of climate-induced drivers, like SLR and storm-related flooding by seawater, and non- climate drivers, like water withdrawal and land-use changes very high 3 train +3912 AR6_WGII 473 5 Salinisation will continue to endanger coastal water and soil quality in the future (high confidence) (Islam et al., 2019; Paldor and Michael, 2021), but the evidence assessed above shows that subsequent impacts to human health and agriculture will depend heavily on regional variations in environment and human behaviour medium 1 train +3913 AR6_WGII 473 6 Together, climate-induced and non-climate drivers can mobilise toxins and contaminants in ways that harm human and marine species health (very high confidence) (see Box 3.2), and climate change is altering these relationships high 2 train +3914 AR6_WGII 473 26 Projected increases in temperature, extreme and variable rainfall conditions, coastal flooding and SLR (Section 3.2; Cross-Chapter Box SLR in Chapter 3) strongly increase the risk of frequent and severe aquatic human pathogen outbreaks in ocean and coastal areas that will continue to harm human health and cause economic losses high 2 train +3915 AR6_WGII 473 28 Climate-driven changes in temperature, salinity (from ice melt and precipitation changes), deoxygenation and ocean acidification can alter dynamics of infectious diseases that target ocean and coastal species by increasing hosts’ susceptibility or pathogens’ abundance or virulence high 2 train +3916 AR6_WGII 473 29 Coral and urchin diseases have increased over time driven by warming-related declines in organism recovery and survival or immunity medium 1 train +3917 AR6_WGII 474 13 Climate change, especially episodic extreme rains and RSLR (Romero-Lankao et al., 2014), is challenging management and design of wastewater and stormwater systems high 2 train +3918 AR6_WGII 474 14 Without substantial adaptation that addresses projected wastewater management challenges and community needs (Section 4.2.6.1; Kirshen et al., 2018; Kirchhoff and Watson, 2019; Kool et al., 2020; Nazarnia et al., 2020; Hughes et al., 2021), coastal water quality in many areas will decrease because of more frequent or severe releases of untreated wastes (high confidence) (Flood and Cahoon, 2011; Hummel et al., 2018; Hughes et al., 2021; McKenzie et al., 2021b), and this will have harmful consequences for human and coastal ecosystem health high 2 train +3919 AR6_WGII 474 19 Non-climate drivers [e.g., invasive species (James et al., 2020), sediment-supply changes (Ganju, 2019; Ladd et al., 2019; Ilia, 2020), erosion and storm damage (Mehvar et al., 2019; Bacopoulos and Clark, 2021)], acting together with climate-induced drivers and associated impacts [e.g., SLR (Cross-Chapter Box SLR in Chapter 3), changes in plant biodiversity (Section 3.5.2; Lee Smee, 2019; Silliman et al., 2019; Schoutens et al., 2020), MHWs (Section 3.4.3.7) and acidification (Section 3.4.2.1)], compromise physical protection by coastal ecosystems very high 3 train +3920 AR6_WGII 475 17 Coastal vegetated ecosystems are vulnerable to harm from multiple climate-induced and non-climate drivers, and together these have reduced wetland area globally (high confidence) (Section 3.4.2.5) and endangered the services provided by these ecosystems high 2 train +3921 AR6_WGII 475 19 These changes also strongly determine the quantity and longevity of blue carbon storage high 2 train +3922 AR6_WGII 475 20 Specific site characteristics and ecosystem responses to climate change will determine future local blue carbon storage or loss high 2 train +3923 AR6_WGII 476 7 But recovery of coastal vegetated ecosystems is expected to bring back the full suite of ecosystem services they provide, not just carbon storage medium 1 train +3924 AR6_WGII 476 8 Successful restoration requires using appropriate plant species in suitable environmental settings (Wodehouse and Rayment, 2019; Friess et al., 2020a) with favourable geomorphology and biophysical conditions (Cameron et al., 2019; Ochoa- Gómez et al., 2019) and considering social, economic, policy and operational constraints (Section 3.6.3.2.2; Cross-Chapter Box NATURAL in Chapter 2), now and in the future high 2 train +3925 AR6_WGII 478 8 Sea level rise and storm-driven coastal erosion endanger coastal archaeological and heritage sites around the world very high 3 train +3926 AR6_WGII 478 15 Observed disruption of ocean and coastal cultural services by climate impacts, plus increasingly severe and widespread projected climate- change impacts on ocean and coastal ecosystems, imply that the risk to cultural ecosystem services will remain constant or even increase medium 1 train +3927 AR6_WGII 478 16 Recent studies assert that cultural ecosystem services are at risk from climate change high 2 train +3928 AR6_WGII 478 22 AR5 concluded that local adaptation measures would not alone be enough to offset global effects of increased climate change on marine and coastal ecosystems, and that mitigation of emissions would also be necessary high 2 train +3929 AR6_WGII 478 23 SROCC assessed that ecosystem-based adaptation, including MPAs (high confidence) (Bindoff et al., 2019a) and adaptive management, are effective to reduce climate-change impacts (IPCC, 2018; IPCC, 2019b), but that existing marine governance is insufficient to provide an effective adaptation response in the marine ecosystem high 2 train +3930 AR6_WGII 479 15 Future social responses depend on warming levels and on the institutional, socioeconomic and cultural constructs that allow or limit livelihood changes medium 1 train +3931 AR6_WGII 482 4 These types of adaptations are more effective when built on trusted relationships and effective coordination among involved parties, and are inclusive of the diversity of actors in a coastal community.Box FAQ 3.4 (continued) Table 3.27 | Conclusions from previous IPCC assessments about implemented adaptation, enablers and limits, and contribution to Sustainable Development Goals (SDGs) AR5 SR15 SROCC Degree of implementation (Section 3.6.3.1)‘The analysis and implementation of coastal adaptation towards climate-resilient and sustainable coasts has progressed more significantly in developed countries than in developing countries (high confidence)’ (Wong et al., 2014).‘Adaptation (to SLR) is already happening (high confidence) and will remain important over multi-centennial time scales’ (Hoegh-Guldberg et al., 2018a).‘A diversity of adaptation responses to coastal impacts and risks have been implemented around the world, but mostly as a reaction to current coastal risk or experienced disasters high 2 train +3932 AR6_WGII 482 5 Conservation and restoration (Section 3.6.3.2)‘With continuing climate change, local adaptation measures (such as conservation) or a reduction in human activities (such as fishing) may not sufficiently offset global-scale effects on marine ecosystems (high confidence)’ (Pörtner et al., 2014).‘Existing and restored natural coastal ecosystems may be effective in reducing the adverse impacts of rising sea levels and intensifying storms by protecting coastal and deltaic regions (medium confidence)’ (Hoegh-Guldberg et al., 2018a).‘Ecosystem restoration may be able to locally reduce climate risks (medium confidence) but at relatively high cost and effectiveness limited to low-emissions scenarios and to less-sensitive ecosystems high 2 train +3933 AR6_WGII 482 6 Enablers, barriers and limits of adaptation (Section 3.6.3.3)‘Adaptation strategies for ocean regions beyond coastal waters are generally poorly developed but will benefit from international legislation and expert networks, as well as marine spatial planning (high agreement)’ (Hoegh-Guldberg et al., 2014).‘Lower rates of change [associated with a 1.5°C temperature increase] enhance the ability of natural and human systems to adapt, with substantial benefits for a wide range of terrestrial, freshwater, wetland, coastal and ocean ecosystems (including coral reefs) (high confidence)’ (Hoegh-Guldberg et al., 2018a).‘There are a broad range of identified barriers and limits for adaptation to climate change in ecosystems and human systems high 2 train +3934 AR6_WGII 482 7 Limitations include [...] availability of technology, knowledge and financial support, and existing governance structures medium 1 train +3935 AR6_WGII 482 8 Existing ocean-governance structures are already facing multi-dimensional, scale-related challenges because of climate change [...] high 2 train +3936 AR6_WGII 485 27 NbS that contribute to climate adaptation, also known as ecosystem-based adaptations (EBA), are cross-cutting actions that harness ecosystem functions to restore, protect and sustainably manage marine ecosystems facing climate-change impacts, while also benefiting social systems and human security (Abelson et al., 2015; Barkdull and Harris, 2019) and supporting biodiversity high 2 train +3937 AR6_WGII 485 28 NbS are expected to contribute to global adaptation and mitigation goals high 2 train +3938 AR6_WGII 486 1 Marine NbS are cost-effective, can generate social, economic and cultural co-benefits, and can contribute to the conservation of biodiversity in the near- to mid- term high 2 train +3939 AR6_WGII 486 3 The feasibility and effectiveness of these adaptations are assessed in Figure 3.24.3.6.3.1 Degree of Implementation and Evidence of Effectiveness Across Sectors 3.6.3.1.1 Coastal community development and settlement Coastal adaptation often addresses the risk of flooding and erosion from SLR, changes in storm activity and degradation of coastal ecosystems and their services high 2 train +3940 AR6_WGII 487 2 Hard-engineered structures like seawalls are generally more costly than nature-based adaptations high 2 train +3941 AR6_WGII 487 6 Field and modelling studies confirm that wetland restoration and preservation are key actions to restore coastal protection and reduce community vulnerability to flooding very high 3 train +3942 AR6_WGII 487 7 Restoring coral reefs, oyster reefs and mangroves (Section 3.6.2.1) and protecting macrophyte meadows dissipates wave energy (Section 3.4.2.1; Yates et al., 2017; Beck et al., 2018; Wiberg et al., 2019; Menéndez et al., 2020), accretes sediment and elevate shorelines, which reduces exposure to waves and storm surges, and offsets erosional losses medium 1 test +3943 AR6_WGII 487 8 However, irreversible regime shifts in ocean ecosystems due to SLR and extreme events, such as MHWs, can limit or compromise restoration in the long term high 2 train +3944 AR6_WGII 487 10 Therefore, restoration and conservation are more successful when non-climate drivers are also minimised high 2 train +3945 AR6_WGII 487 16 Managed realignment is the best option to reduce risks from SLR high 2 train +3946 AR6_WGII 487 17 Implementation of protective measures varies among nations and lack of financial resources limits the options available very high 3 train +3947 AR6_WGII 487 23 Improving capacity to predict anomalous conditions in coastal and marine ecosystems (Jacox et al., 2019; Holbrook et al., 2020; Jacox et al., 2020), storm-driven flooding in reef-lined coasts (Scott et al., 2020; Winter et al., 2020) and fisheries stocks (Payne et al., 2017; Tommasi et al., 2017; Muhling et al., 2018) can improve forecasts of coastal and marine resources; these can enhance sustainability of wild-capture fisheries under climate change high 2 train +3948 AR6_WGII 488 4 Current impacts of sea level rise The rate of global mean SLR was 1.35 mm yr–1 (0.78–1.92 mm yr–1, very likely range) during 1901–1990, faster than during any century in at least 3000 years high 2 train +3949 AR6_WGII 488 5 Global mean SLR has accelerated to 3.25 mm yr–1 (2.88–3.61 mm yr–1, very likely range) during 1993–2018 high 2 train +3950 AR6_WGII 488 7 The largest observed changes in coastal ecosystems are being caused by the concurrence of human activities, waves, current-induced sediment transport and extreme storm events medium 1 test +3951 AR6_WGII 488 10 Projected risks to coastal communities, infrastructure and ecosystems Risks from SLR are very likely to increase by one order of magnitude well before 2100 without adaptation and mitigation action as agreed by parties to the Paris Agreement very high 3 train +3952 AR6_WGII 488 11 Global mean SLR is likely to continue accelerating under SSP1-2.6 and more strongly forced scenarios (Figure BoxSLR1; WGI AR6 Chapter 9; Oppenheimer et al., 2019; Fox-Kemper et al., 2021), increasing the risk of chronic coastal flooding at high tide, serious flooding during extreme events such as swells, storms and hurricanes, and erosion, and coastal ecosystem losses across many low-lying and erodible coasts very high 3 train +3953 AR6_WGII 488 18 Coastal ecosystems can migrate landward or grow vertically in response to SLR, but their resilience and capacity to keep up with SLR will be compromised by ocean warming and other drivers, depending on regions and species, for example, above 1.5°C for coral reefs high 2 train +3954 AR6_WGII 488 19 Sediments and space for landward retreat are crucial for mangroves, salt marshes and beach ecosystems high 2 train +3955 AR6_WGII 489 1 This is a reason for concern given that rapid SLR after the last glacial–interglacial transition caused a drowning of coral reefs (high confidence) (Camoin and Webster, 2015; Sanborn et al., 2017; Webster et al., 2018), extensive loss of coastal land and islands, habitats and associated biodiversity (high confidence) (AR6 WGI Chapter 9; Fruergaard et al., 2015; Fernández-Palacios et al., 2016; Hamilton et al., 2019; Helfensdorfer et al., 2019; Kane and Fletcher, 2020; Fox-Kemper et al., 2021), and triggered Neolithic migrations in Europe and Australia medium 1 train +3956 AR6_WGII 489 3 Even if climate warming is stabilised at 2°C to 2.5°C GWL, coastlines will continue to reshape over millennia, affecting at least 25 megacities and drowning low-lying areas where 0.6–1.3 billion people lived in 2010 medium 1 train +3957 AR6_WGII 489 4 Solutions, opportunities and limits to adaptation The ability to adapt to current coastal impacts, to cope with future coastal risks and to prevent further acceleration of SLR beyond 2050 depends on immediate mitigation and adaptation actions very high 3 train +3958 AR6_WGII 489 9 Risks can be anticipated, planned and decided upon, and adaptation interventions can be implemented over the coming decades considering their often long lead- and lifetimes, irrespective of the large uncertainty about SLR beyond 2050 high 2 train +3959 AR6_WGII 489 10 Adaptation capacity and governance to manage risks from projected SLR typically require decades to implement and institutionalise high 2 train +3960 AR6_WGII 489 12 Sea level rise is likely to compound social conflict in some settings high 2 train +3961 AR6_WGII 489 15 Only avoidance and relocation can remove coastal risks for the coming decades, while other measures only delay impacts for a time, have increasing residual risk or perpetuate risk and create ongoing legacy effects and virtually certain property and ecosystem losses high 2 train +3962 AR6_WGII 489 18 Effective responses to rising sea level involve locally applicable combinations of decision analysis, land-use planning, public participation and conflict resolution approaches; together these can anticipate change and help to chart adaptation pathways, over time addressing the governance challenges due to rising sea level high 2 train +3963 AR6_WGII 489 19 Ecosystem-based adaptation can reduce impacts on human settlements and bring substantial co-benefits, such as ecosystem services restoration and carbon storage, but they require space for sediment and ecosystems and have site-specific physical limits, at least above 1.5°C GWL high 2 train +3964 AR6_WGII 491 7 Transboundary agreements on shifting fisheries will reduce the risk of overharvesting medium 1 train +3965 AR6_WGII 491 10 Despite the potential for adaptive management to achieve sustainable fisheries, outcomes will very likely be inequitable (Gaines et al., 2018; Lam et al., 2020), with lower- income countries suffering the greater biomass and economic losses, increasing inequalities, especially under higher-emission scenarios high 2 train +3966 AR6_WGII 491 13 Engineered solutions, such as seawalls and revetments, have traditionally been used to address coastal erosion (Section 3.6.3.1.1), but soft infrastructure approaches, including beach nourishment, submerged breakwaters and groins, and NbS (Section 3.6.2.1), are becoming more common, partly due to demand from the tourism industry medium 1 train +3967 AR6_WGII 492 7 Increased Arctic traffic due to ice loss can benefit trade, transportation and tourism (medium confidence), but will also affect Arctic marine ecosystems and livelihoods high 2 train +3968 AR6_WGII 492 18 Forecasts facilitate preventive public health measures (World Health Organisation and United Nations Children’s Fund, 2017), or seafood harvest guidance (Maguire et al., 2016; Leadbetter et al., 2018; Anderson et al., 2019; Bolin et al., 2021), reducing risks of disease outbreaks, waste and contaminated seafood entering the market medium 1 train +3969 AR6_WGII 494 1 Current MPAs offer conservation benefits such as increases in biomass and diversity of habitats, populations and communities high 2 train +3970 AR6_WGII 494 2 But current MPAs do not provide resilience against observed warming and heatwaves in tropical-to-temperate ecosystems medium 1 train +3971 AR6_WGII 494 4 Current placement and extent of MPAs will not provide substantial protections against projected climate change past 2050 high 2 train +3972 AR6_WGII 494 18 In summary, MPAs and other marine spatial-planning tools have great potential to address climate-change mitigation and adaptation in ocean and coastal ecosystems, if they are designed and implemented in a coordinated way that takes into account ecosystem vulnerability and responses to projected climate conditions, considers existing and future ecosystem uses and non-climate drivers, and supports effective governance high 2 train +3973 AR6_WGII 495 16 Employing the natural adaptive capacity of species or individuals in active restoration for corals and kelps with current technology involves fewer risks than assisted evolution or long- distance relocation high 2 train +3974 AR6_WGII 495 19 Models show that a combination of available management approaches (restoration, reducing non-climate drivers) and speculative interventions (e.g., enhanced corals, reef shading) can contribute to sustaining some coral reefs beyond 1.5°C of global warming with declining effectiveness beyond 2°C of global warming medium 1 train +3975 AR6_WGII 498 1 The icons at the bottom show the Sustainable Development Goals to which NbS in the ocean possibly contribute.Box FAQ 3.5 (continued) Adaptations implemented at the local level that consider IKLK systems are beneficial high 2 train +3976 AR6_WGII 499 1 Recent evidence suggests that policies supporting local institutions can improve adaptation outcomes medium 1 test +3977 AR6_WGII 499 9 Economic and financing barriers to marine adaptation are often higher in low- to middle- income countries, where resources influence governance and constrain options for implementation and maintenance high 2 train +3978 AR6_WGII 499 10 Current financial flows are insufficient to meet the costs of coastal and marine impacts of climate change (very high confidence) and ocean- focused finance is unevenly distributed, with higher flows within, and to, developed countries very high 3 train +3979 AR6_WGII 499 22 Maladaptation (WGII Chapter 16; Magnan et al., 2016) is a common risk of current project-based funding due to the pressure to produce concrete results medium 1 train +3980 AR6_WGII 499 23 Maladaptation can be avoided through a focus on building adaptive capacity, community-based management, drivers of vulnerability and site-specific measures low 0 train +3981 AR6_WGII 499 24 More research is needed to identify ways that governance and financing agreements can help overcome financial barriers and sociocultural constraints to avoid maladaptation in coastal ecosystems high 2 train +3982 AR6_WGII 500 3 Communities and governments at all levels increasingly use decision- making frameworks (e.g., structured decision making) or decision- analysis tools to evaluate trade-offs between different responses, rather than applying generic best practices to different physical, technical or cultural contexts high 2 train +3983 AR6_WGII 500 7 Focusing on user engagement, relationship building and the decision-making context ensures that climate services are useful to, and used by, different stakeholders high 2 train +3984 AR6_WGII 500 21 Responding to climate-change impacts requires transformative governance high 2 train +3985 AR6_WGII 501 3 Socio-institutional marine adaptations (Section 3.6.2.2) that support current livelihoods and help develop alternatives can contribute to attainment of social SDGs by enhancing social equity and supporting societal transformation medium 1 train +3986 AR6_WGII 501 5 Marine adaptation also shows promise for helping support achievement of economic SDGs medium 1 train +3987 AR6_WGII 501 12 Developing marine adaptation pathways that offer multiple benefits requires transformational adaptation high 2 train +3988 AR6_WGII 501 14 Presently implemented adaptation activity, at the aggregate level, adversely affects multiple gender targets under SDG5 high 2 train +3989 AR6_WGII 566 3 Not surprisingly, a large share of adaptation interventions (~60%) are forged in response to water-related hazards high 2 train +3990 AR6_WGII 566 6 Approximately 163 million people live in unfamiliarly dry areas now medium 1 train +3991 AR6_WGII 566 8 Substantially more people (~709 million) live in regions where annual maximum one-day precipitation has increased than regions where it has decreased (~86 million) medium 1 train +3992 AR6_WGII 566 9 At the same time, more people (~700 million) are also experiencing longer dry spells than shorter dry spells since the 1950s medium 1 train +3993 AR6_WGII 566 12 At the same time, groundwater in aquifers across the tropics has experienced enhanced episodic recharge from intense precipitation and flooding events medium 1 train +3994 AR6_WGII 566 15 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.Extreme weather events causing highly impactful floods and droughts have become more likely and (or) more severe due to anthropogenic climate change high 2 train +3995 AR6_WGII 566 23 A significant share of those impacts are negative and felt disproportionately by already vulnerable communities high 2 train +3996 AR6_WGII 566 28 In addition, precipitation and extreme weather events are linked to increased incidence and outbreaks of water-related diseases high 2 train +3997 AR6_WGII 567 2 Between 3 and 4 billion people are projected to be exposed to physical water scarcity at 2°C and 4°C global warming levels (GWL), respectively low 0 train +3998 AR6_WGII 567 6 Modified streamflow is also projected to affect inflows to urban storage reservoirs and increase the vulnerability of urban water services to hydro-meteorological extremes, particularly in less developed countries high 2 train +3999 AR6_WGII 567 10 With RCP6.0 and SSP2, the global population exposed to extreme-to-exceptional total water storage drought is projected to increase from 3% to 8% over the 21st century medium 1 train +4000 AR6_WGII 567 12 Direct flood damages are projected to increase by four to five times at 4°C compared to 1.5°C medium 1 train +4001 AR6_WGII 567 17 Projected losses include a 1.2- to 1.8-fold increase in GDP loss due to flooding between 1.5°C and 2°C warming medium 1 train +4002 AR6_WGII 567 19 Due to the combined effects of water and temperature changes, risks to agricultural yields could be three times higher at 3°C compared to 2°C medium 1 train +4003 AR6_WGII 567 22 The number of internally displaced people in sub- Saharan Africa, South Asia and Latin America increased almost five times for RCP8.5 compared to RCP2.6 low 0 train +4004 AR6_WGII 567 26 In contrast, responses are more policy-oriented and urban-focused in developed countries high 2 train +4005 AR6_WGII 567 28 In addition, large- scale irrigation also affects local to regional climates, both in terms of temperature and precipitation change high 2 train +4006 AR6_WGII 568 1 Roughly one third and one fourth of case studies on water adaptation also documents maladaptation and co-benefits, respectively high 2 train +4007 AR6_WGII 568 2 A significant knowledge gap remains in knowing if observed adaptation benefits also translate to climate risk reduction, if so, by how much and under what conditions medium 1 train +4008 AR6_WGII 568 5 However, residual impacts remain for some options and regions at all levels of warming, and the overall effectiveness decreases at higher warming levels high 2 train +4009 AR6_WGII 568 9 SDGs cannot be met without adequate and safe water (high confidence), and water is fundamental to all systems transition high 2 train +4010 AR6_WGII 568 11 However, barriers remain for low-income countries to access funds (medium confidence), and there is insufficient evidence on benefits for marginalised groups medium 1 train +4011 AR6_WGII 568 13 The water intensity of mitigation must be managed in socially and politically acceptable ways to increase synergies with SDGs, improve water security and reduce trade-offs with adaptation medium 1 train +4012 AR6_WGII 568 16 Water adaptation policies enabled through ethical co-production between holders of Indigenous knowledge, local knowledge and technical knowledge (medium confidence), through cooperation and coordinated actions among multiple actors, including women and all marginalised groups, at various levels of governance medium 1 train +4013 AR6_WGII 569 7 First, approximately half the world’s population (~4 billion out of ~8 billion people) are assessed as being currently subject to severe water scarcity for at least some part fo the year medium 1 train +4014 AR6_WGII 569 11 Currently, many people are experiencing climate change on a day-to-day basis through water-related impacts such as the increased frequency and intensity of heavy precipitation (high confidence) (Section 4.2.1.1, Seneviratne et al., 2021); accelerated melting of glaciers (high confidence) (Section 4.2.2, Douville et al., 2021); changes in frequency, magnitude and timing of floods (high confidence) (Section 4.2.4, Seneviratne et al., 2021); more frequent and severe droughts in some places (high confidence) (Section 4.2.5, Seneviratne et al., 2021); decline in groundwater storage and reduction in recharge (medium confidence) (Section 4.2.6, Douville et al., 2021) and water quality deterioration due to extreme events medium 1 train +4015 AR6_WGII 569 16 Vulnerability to water-related impacts of climate change and extreme weather are already felt in all major sectors and are projected to intensify in the future, for example, in agriculture (high confidence) (Sections 4.3.1, 4.5.1); energy and industry (high confidence for observed drought impacts and projected impacts) (Sections 4.3.2, 4.5.2); water for health and sanitation (high confidence about links to precipitation extremes and disease outbreaks) (Sections 4.3.3, 4.5.3); water for urban, peri-urban and municipal sectors medium 1 train +4016 AR6_WGII 569 28 Third, a large majority (~60%) of all adaptation responses documented since 2014 are about adapting to water-related hazards like droughts, floods and rainfall variability (Berrang-Ford et al., 2021b) high 2 train +4017 AR6_WGII 569 32 Only ~20% of all documented case studies on observed water-related adaptations measure outcomes (positive or negative), but the link between positive outcomes and climate risk reduction is unclear and remains challenging to assess (Section 4.7.1) medium 1 train +4018 AR6_WGII 569 33 On the other hand, most of the future projected water-related adaptations are more effective at lower GWLs (1.5°C) than at higher GWLs, showing the importance of mitigation for future adaptations to remain effective high 2 train +4019 AR6_WGII 570 2 Therefore, minimising the risks to water security from climate change will require a full-systems view that considers the direct impacts of mitigation measures on water resources and their indirect effect via limiting climate change high 2 train +4020 AR6_WGII 570 11 AR5 projected an increase in meteorological, agricultural and hydrological droughts in dry regions medium 1 train +4021 AR6_WGII 570 12 The Special Report on Global Warming of 1.5°C (SR1.5) assessed that limiting global warming to 1.5°C is expected to substantially reduce the probability of extreme droughts, precipitation deficits and risks associated with water availability in some regions medium 1 train +4022 AR6_WGII 570 13 On the other hand, higher risks to natural and human systems in a 2.0°C world would mean increased vulnerability for the poor, showing that socioeconomic drivers are expected to have a more significant influence on water-related risks and vulnerabilities than changes in climate alone medium 1 train +4023 AR6_WGII 571 19 Climate change patterns of streamflow change include declines in western North America, northeast South America, the Mediterranean and South Asia medium 1 train +4024 AR6_WGII 574 3 The Special Report on Climate Change and Land (SRCCL) stated that groundwater over-extraction for irrigation is causing depletion of groundwater storage high 2 train +4025 AR6_WGII 574 4 The report also noted that precipitation changes, coupled with human drivers, will have a role in causing desertification, and water-driven soil erosion is projected to increase due to climate change medium 1 train +4026 AR6_WGII 574 6 SRCCL stated that improved irrigation techniques (e.g., drip irrigation) and moisture conservation (e.g., rainwater harvesting using indigenous and local practices) could increase farmers’ adaptive capacity high 2 train +4027 AR6_WGII 574 7 The Sixth Assessment Report (AR6) Working Group I (WGI) (Douville et al., 2021) concluded that anthropogenic climate change has increased atmospheric moisture and precipitation intensity (very likely by 2–3% per 1°C) (high confidence), increased terrestrial ET (medium confidence) and contributed to drying in dry summer climates including in the Mediterranean, southwestern Australia, southwestern South America, South Africa and western North America (medium to high confidence), and has caused earlier onset of snowmelt and increased melting of glaciers high 2 train +4028 AR6_WGII 574 21 Therefore, improving societal aspects of water management will be key in adapting to climate change-driven increases in water scarcity in the future high 2 train +4029 AR6_WGII 574 28 In summary, roughly half of the world’s population are assessed as currently subject to severe water scarcity for at least some part of the year due to climatic and non-climatic factors, and this is projected to be exacerbated at higher levels of warming medium 1 train +4030 AR6_WGII 574 29 General water insecurity issues are seen worldwide, particularly in South Asia, North China, Africa and the Middle East, due to high population densities often coupled with low water availability, accessibility, quality and governance high 2 train +4031 AR6_WGII 574 30 Areas with high water availability can also be water-insecure due to increased flood risk, deteriorated water quality, and poor governance high 2 train +4032 AR6_WGII 574 31 Future water security will depend on the evolution of all these socioeconomic and governance factors and future regional climate change high 2 train +4033 AR6_WGII 574 32 The main climate change contribution to water insecurity is the potential for reduced water availability, with a secondary contribution from increased flooding risk medium 1 train +4034 AR6_WGII 574 33 Future socioeconomic conditions are a crucial driver of water insecurity, implying the need for further adaptation to some level of future climate change medium 1 train +4035 AR6_WGII 574 34 However, policy challenges are high in many regions, with uncertainty in the regional climate outcomes being a key factor high 2 train +4036 AR6_WGII 576 12 In summary, radiative forcing by GHG and aerosols drives changes in ET and precipitation at global and regional scales, and the associated warming shifts the balance between frozen and liquid water high 2 train +4037 AR6_WGII 576 14 Land cover changes and urbanisation affect both the climate and land hydrology by altering the exchanges of energy and moisture between the atmosphere and surface high 2 train +4038 AR6_WGII 576 15 Direct human interventions in river systems and groundwater systems are non- climatic drivers with substantial impacts on the water cycle high 2 train +4039 AR6_WGII 578 9 In 2020, approximately 498 million people lived in unfamiliarly wet areas, where the long-term average precipitation is as high as previously seen in only about one in 6 years medium 1 train +4040 AR6_WGII 578 11 On the other hand, approximately 163 million people lived in unfamiliarly dry areas, mostly in low latitudes medium 1 train +4041 AR6_WGII 578 12 Due to high variability over time, the signal of long-term change in annual mean precipitation is not distinguishable from the noise of variability in many areas (Hawkins et al., 2020), implying that the local annual precipitation cannot yet be defined ‘unfamiliar’ by the above definition.Notably, many regions have seen increased precipitation for part of the year and decreased precipitation at other times high 2 train +4042 AR6_WGII 578 13 Therefore, the numbers of people seeing unfamiliar seasonal precipitation levels are expected to be higher than those quoted above for unfamiliar annual precipitation changes medium 1 train +4043 AR6_WGII 578 15 The intensity of heavy precipitation has increased in many regions high 2 train +4044 AR6_WGII 578 21 Substantially more people (~709 million) live in regions where annual maximum one-day precipitation has increased than in regions where it has decreased (~86 million) medium 1 train +4045 AR6_WGII 578 22 However, more people are experiencing longer dry spells than shorter dry spells: approximately 711 million people live in places where annual mean CDD is longer than in the 1950s, and ~404 million in places with shorter CDD medium 1 train +4046 AR6_WGII 578 24 Nearly half a billion people live in areas with historically unfamiliar wet conditions, and over 160 million in areas with historically unfamiliar dry conditions medium 1 train +4047 AR6_WGII 579 29 Changes in land cover and irrigation have also changed regional ET medium 1 train +4048 AR6_WGII 581 1 Global mean soil moisture has slightly decreased, but regional changes vary, with both increases and decreases of 20% or more in some regions medium 1 test +4049 AR6_WGII 581 2 Drying soil moisture trends are more widespread than wetting trends, not only in arid areas but also in humid and transitional areas medium 1 train +4050 AR6_WGII 581 3 Reduced dry-season water availability is driven mainly by increasing transpiration medium 1 train +4051 AR6_WGII 581 5 Tourism and recreation activities have been negatively impacted by declining snow cover, glaciers and permafrost in high mountains medium 1 train +4052 AR6_WGII 581 19 During the last two decades, the global glacier mass loss rate exceeded 0.5-meter water equivalent (m w.e.) per year compared to an average of 0.33 m w.e. yr–1 in 1950–2000. This volume of mass loss is the highest since the start of the entire observation period very high 3 test +4053 AR6_WGII 582 5 Widespread cryospheric changes are affecting humans and ecosystems in mid-to-high latitudes and the high-mountain regions high 2 train +4054 AR6_WGII 584 26 However, trends emerge on a regional level (a general increasing trend in the northern higher latitude region and mixed trend in the rest of the word) high 2 train +4055 AR6_WGII 584 27 Climatic factors contribute to these trends in most basins high 2 train +4056 AR6_WGII 584 28 They are more important than direct human influence in a larger share of major global basins (medium confidence), although direct human influence dominates in some medium 1 train +4057 AR6_WGII 584 29 Overall, anthropogenic climate change is attributed as a driver to the global pattern of change in streamflow medium 1 train +4058 AR6_WGII 584 37 Warming in the last 40–60 years has led to a 1–10-d earlier per decade spring flood occurrence depending on the location (the most frequent being 2–4 d per decade) high 2 train +4059 AR6_WGII 588 18 Nevertheless, anthropogenic climate change increased the likelihood of a number of major heavy precipitation events and floods that resulted in disastrous impacts in southern and eastern Asia, Europe, North America and South America (Table 4.3) high 2 train +4060 AR6_WGII 588 24 Although there is growing evidence on the effects of anthropogenic climate change on each event, given the relatively poor regional coverage and high model uncertainty, there is low confidence in the attribution of human- induced climate change to flood change on the global scale.In snow-dominated regions, 1~10 d earlier spring floods per decade due to warmer temperature are reported for the last decades high 2 train +4061 AR6_WGII 588 26 Despite the increase in the number of glacial lake studies (Wang and Zhou, 2017; Harrison et al., 2018; Begam and Sen, 2019; Bolch et al., 2019), changes in the frequency of occurrence of glacier-related floods associated with climate change remain unclear medium 1 train +4062 AR6_WGII 588 31 In summary, the frequency and magnitude of river floods have changed in the past several decades with high regional variations high 2 train +4063 AR6_WGII 588 32 Anthropogenic climate change has increased the likelihood of extreme precipitation events and the associated increase in the frequency and magnitude of river floods high 2 train +4064 AR6_WGII 589 4 Agricultural drought threatens food production through crop damage and yield decreases (e.g., Section 4.3.1) high 2 train +4065 AR6_WGII 589 7 Cascading effects of droughts can include health issues triggered by a lack of sanitation (Section 4.3.3); can cause human displacements and loss of social ties, sense of place and cultural identity; and migration to unsafe settlements medium 1 train +4066 AR6_WGII 590 15 In summary, droughts can have substantial societal impacts (virtually certain), and agricultural and ecological drought conditions in particular have become more frequent and severe in many parts of the world but less frequent and severe in some others high 2 train +4067 AR6_WGII 590 16 Drought- induced economic losses relative to GDP are approximately twice as high in lower-income countries compared to higher-income countries, although the gap has narrowed since the 1980s, and at the global scale there is a decreasing trend of economic vulnerability to drought medium 1 train +4068 AR6_WGII 590 17 Nevertheless, anthropogenic climate change has contributed to the increased likelihood or severity of drought events in many parts of the world, causing reduced agricultural yields, drinking water shortages for millions of people, increased wildfire risk, loss of lives of humans and other species and loss of billions of dollars of economic damages medium 1 train +4069 AR6_WGII 592 14 These estimates range from approximately 113 to 510 km3 yr−1 and variation in estimates is due to methods and spatio-temporal scales considered high 2 train +4070 AR6_WGII 592 15 Global hydrological models (Herbert and Döll, 2019) show that human-induced groundwater depletion at rates exceeding 20 mm yr–1 (2001–2010) is occurring in the major aquifers systems such as the High Plains and California Central Valley aquifers (USA), Arabian aquifer (Middle East), North-Western Sahara Aquifer System (North Africa), Indo-Gangetic Basin (India) and North China Plain (China) high 2 train +4071 AR6_WGII 592 16 Groundwater depletion at lower rates (<10 mm yr–1) is taking place in the Amazon Basin (Brazil) and Mekong River Basin (South East Asia), primarily due to climate variability and change high 2 train +4072 AR6_WGII 593 10 In Finland, a sustained reduction (almost 100 mm in 100 years) of long- term snow accumulation combined with early snowmelt has reduced spring recharge (Irannezhad et al., 2016) medium 1 train +4073 AR6_WGII 593 24 In summary, groundwater storage has declined in many parts of the world, most notably since the beginning of the 21st century, due to the intensification of groundwater-fed irrigation high 2 train +4074 AR6_WGII 593 26 In higher altitudes, warmer climates have altered groundwater regimes and may have led to reduced spring recharge due to reduced duration and snowmelt discharges medium 1 train +4075 AR6_WGII 594 13 In summary, although climate-induced water quality degradation due to increases in water and surface temperatures or melting of the cryosphere has been observed medium 1 train +4076 AR6_WGII 594 17 In addition, accelerated soil erosion and sedimentation have severe societal impacts through land degradation, reduced soil productivity and water quality (Section 4.2.7), increased eutrophication and disturbance to aquatic ecosystems (Section 4.3.5), sedimentation of waterways and damage to infrastructure (Graves et al., 2015; Issaka and Ashraf, 2017; Schellenberg et al., 2017; Hewett et al., 2018; Panagos et al., 2018; Sartori et al., 2019) medium 1 train +4077 AR6_WGII 594 19 Sedimentation increases due to soil erosion in mountainous regions burned by wildfires, as a result of warming and altered precipitation, is documented with high confidence in the USA (Gould et al., 2016; DeLong et al., 2018), Australia (Nyman et al., 2015; Langhans et al., 2016), China (Cui et al., 2014) and Greece (Karamesouti et al., 2016) and can potentially damage downstream aquatic ecosystems (Section 4.3.5) and water quality (Section 4.2.7) (Cui et al., 2014; Murphy et al., 2015; Langhans et al., 2016) medium 1 train +4078 AR6_WGII 594 26 The climate change impact on erosion and sediment load varies significantly over the world (Li et al., 2020b) high 2 train +4079 AR6_WGII 595 8 In summary, in the areas with high human activity, factors other than climate have a more significant impact on soil erosion and sediment flux high 2 train +4080 AR6_WGII 595 26 Drought has been singled out as a major driver of yield reductions globally high 2 test +4081 AR6_WGII 596 10 Livestock production has also been affected by changing seasonality, increasing frequency of drought, rising temperatures and vector-borne diseases and parasites through changes in the overall availability, as well as reduced nutritional value, of forage and feed crops (Varadan and Kumar, 2014; Naqvi et al., 2015; Zougmoré et al., 2016; Henry et al., 2018; Godde et al., 2019) medium 1 train +4082 AR6_WGII 596 14 Moreover, extreme precipitation can lead to increased surface flooding, waterlogging, soil erosion and susceptibility to salinisation high 2 train +4083 AR6_WGII 596 22 These are important because they shape farmers’ perceptions, which in turn shape the adaptation measures farmers will undertake (Caretta and Börjeson, 2015; Savo et al., 2016; Sujakhu et al., 2016; Su et al., 2017) (Section 4.8.4) high 2 train +4084 AR6_WGII 596 24 In subtropical/ tropical climates, climate-induced hazards such as floods and droughts negatively impact agricultural production high 2 train +4085 AR6_WGII 596 26 They often rely on rain-fed agriculture in marginal areas with high exposure and high vulnerability to water- related stress and low adaptive capacity high 2 train +4086 AR6_WGII 596 34 A recent global study concluded that reservoirs might emit more carbon than they bury, especially in the tropics (Keller et al., 2021) medium 1 train +4087 AR6_WGII 597 2 Thus, there is a growing body of evidence of negative impacts of extreme events on hydropower production high 2 train +4088 AR6_WGII 597 7 Globally, for the period 1981–2010, the utilisation rate of thermoelectric power was reduced by 3.8% during drought years compared to long-term average values (van Vliet et al., 2016a), and none of the studies reported increases in thermoelectric power production as a consequence of climate change high 2 train +4089 AR6_WGII 597 15 In addition, melting and thawing of snow, ice and permafrost (Section 4.2.2) have also adversely impacted water quality, security and health (high confidence) (IPCC, 2019a) (Section 4.2.7).Literature since AR5 confirms that temperature, precipitation and extreme weather events are linked to increased incidence and outbreaks of water-related and neglected tropical diseases (Colón- González et al., 2016; Levy et al., 2016; Azage et al., 2017; Harp et al., 2021) high 2 train +4090 AR6_WGII 597 23 Drinking water treatment can be compromised by degraded source water quality and extreme weather events, including droughts, storms, ice storms and wildfires that overwhelm or cause infrastructure damage (Sherpa et al., 2014; Khan et al., 2015; Howard et al., 2016; White et al., 2017) high 2 train +4091 AR6_WGII 597 30 Water insecurity and inadequate WaSH have been associated with increased disease risk high 2 train +4092 AR6_WGII 598 11 Climate extremes have profound implications for urban and peri-urban water management, particularly in an increasingly urbanised world high 2 train +4093 AR6_WGII 599 1 Introduction Some extreme weather events are increasing in frequency and (or) severity as a result of climate change (Seneviratne et al., 2021) high 2 train +4094 AR6_WGII 599 7 These disasters reflect immediate societal and political implications of rising risks (high confidence), but also provide windows of opportunity to raise awareness about climate change and to implement disaster-reduction policies and strategies high 2 train +4095 AR6_WGII 602 24 With the frequency, severity and (or) likelihood of several types of extreme weather increasing, disasters can increasingly be regarded as ‘the public face of climate change’ high 2 train +4096 AR6_WGII 602 25 Detection and attribution studies make the climate change fingerprint of several types of disasters increasingly clear high 2 train +4097 AR6_WGII 602 26 Moreover, existing vulnerabilities and exposures play an important role in turning extreme events into disasters, further exacerbating existing racial, gender and social inequalities high 2 train +4098 AR6_WGII 603 1 Further research is necessary to determine the extent and nature of water-related climate change impacts in the urban areas of developing countries high 2 train +4099 AR6_WGII 603 6 Under extreme heat, often associated with minimal rainfall or water flows, the drying of shallower areas and the migration or death of individual organisms can occur (Dell et al., 2014; Miller et al., 2014; Scheffers et al., 2016; Szekeres et al., 2016; Myers et al., 2017; FAO, 2018a) high 2 train +4100 AR6_WGII 603 14 Many wetland-dependent species have seen a long-term decline, with the Living Planet Index showing that 81% of populations of freshwater species are in decline and others being threatened by extinction (Davidson and Finlayson, 2018; Darrah et al., 2019; Diaz et al., 2019) high 2 train +4101 AR6_WGII 603 34 Freshwater ecosystems are also under extreme pressure from changes in land use and water pollution, with climate change exacerbating these, such as the further decline of snow cover (DeBeer et al., 2016) and increased consumptive use of fresh water, and leading to the decline, and possibly extinction, of many freshwater-dependent populations high 2 train +4102 AR6_WGII 604 2 The predominant key drivers are changes in land use and water pollution high 2 train +4103 AR6_WGII 604 8 However, the large majority acknowledges reduction of water availability due to climate change as having the potential to exacerbate tensions (de Stefano et al., 2017; Waha et al., 2017), especially in regions and within groups dependent on agriculture for food production (von Uexkull et al., 2016; Koubi, 2019) high 2 train +4104 AR6_WGII 604 12 However, evidence suggests that changes in rainfall patterns amplify existing tensions (Abel et al., 2019); examples include Syria, Iraq (Abbas et al., 2016; von Lossow, 2016) and Yemen (Mohamed et al., 2017) medium 1 train +4105 AR6_WGII 604 13 There is also medium evidence that in some regions of Africa (e.g., Kenya, Democratic Republic of the Congo), there are links between observed water stress and individual attitude for participating in violence, particularly for the least resilient individuals (von Uexkull et al., 2020) medium 1 train +4106 AR6_WGII 604 17 Climate change concerns also play a role in stimulating cooperative efforts, as in the case of the Ganges- Brahmaputra-Meghna River Basin (Mirumachi, 2015; Link et al., 2016) medium 1 train +4107 AR6_WGII 605 1 Furthermore, household water insecurity has also been singled out as a driver of migration, given its physical, mental health and socioeconomic effects (Stoler et al., 2021) medium 1 train +4108 AR6_WGII 605 4 The outcome is determined mainly by the socioeconomic, political and environmental context medium 1 train +4109 AR6_WGII 605 10 Climate-driven hydrological changes are affecting culturally significant terrestrial and freshwater species and ecosystems, particularly for Indigenous Peoples, local communities and traditional peoples in the Arctic, high mountain areas, and small islands high 2 train +4110 AR6_WGII 607 11 In summary, the cultural water uses of Indigenous Peoples, local communities and traditional peoples are being impacted by climate change high 2 train +4111 AR6_WGII 607 14 Continuation of projected warming and other physical mechanisms will further accelerate the melting of snow cover and glaciers and thawing of permafrost high 2 train +4112 AR6_WGII 607 21 Water cycle variability and extremes are projected to increase faster than average changes in most regions of the world and under all emission scenarios high 2 train +4113 AR6_WGII 607 25 Importantly, in most land regions, the future changes are subject to high uncertainty even in the sign of the projected change high 2 train +4114 AR6_WGII 607 27 For any given location, the range of projected changes generally increases with global warming high 2 train +4115 AR6_WGII 608 2 The geographical patterns of local agreement/disagreement in projected precipitation change remain broadly similar with increased global warming, but the range of uncertainty generally increases high 2 train +4116 AR6_WGII 609 10 Taken together, these projections of more intense precipitation and changes in the length of dry spells give a clear picture of increasingly volatile precipitation regimes, with many regions seeing both longer dry spells and heavier events when precipitation does occur high 2 train +4117 AR6_WGII 609 14 The ranges of projected precipitation changes are smaller at lower levels of global warming high 2 train +4118 AR6_WGII 609 15 Either an increase or decrease is possible in most regions, but there is an agreement among models on the increase in the far north high 2 train +4119 AR6_WGII 609 16 There is a stronger model consensus on heavy precipitation increasing with global warming over most land areas high 2 train +4120 AR6_WGII 609 17 There are widely varying projections of change in dry spell length (high confidence), but in regions with increasing projected dry spells, the potential increase is larger at higher levels of global warming high 2 train +4121 AR6_WGII 611 5 In addition, the impacts of rising CO 2 concentrations on plant stomata and leaf area play a role in model projections of ET change high 2 train +4122 AR6_WGII 612 10 In summary, projected soil moisture changes increase with levels of global warming high 2 train +4123 AR6_WGII 612 11 In the CMIP6 multi- model ensemble at 4°C global warming, decreased soil moisture of up to 40% is projected in Amazonia, southern Africa and western Europe in all models high 2 train +4124 AR6_WGII 613 3 Constraining warming to 1.5°C would prevent the thawing of a permafrost area of 1.5 to 2.5 million km2 compared to thawing under 2°C medium 1 train +4125 AR6_WGII 614 3 Projections of snow cover metrics [IPCC AR6 WGI, 2021 (Section 9.5.3.3)] suggest a further decrease in snow water equivalent (SWE) and snow cover extent (SCE), though the inter-model spread is considerable (Lute et al., 2015; Thackeray et al., 2016; Kong and Wang, 2017; Henderson et al., 2018) high 2 train +4126 AR6_WGII 614 12 In summary, in most basins fed by glaciers, runoff is projected to increase initially in the 21st century and then decline medium 1 train +4127 AR6_WGII 614 13 Projections suggest a further decrease in seasonal snow cover extent and mass in mid to high latitudes and high mountains high 2 train +4128 AR6_WGII 614 14 Permafrost will continue to thaw throughout the 21st century high 2 train +4129 AR6_WGII 614 25 Changes in streamflow could increase the number of people facing water scarcity or insecurity high 2 train +4130 AR6_WGII 616 1 Nevertheless, since projected changes typically increase with global warming, limiting warming to 1.5°C or 2°C substantially reduces the potential for either large increases or decreases in mean streamflow compared to 3°C or 4°C (Warszawski et al., 2014; Falkner, 2016; Gosling et al., 2017; Figure 4.16) high 2 train +4131 AR6_WGII 616 24 In summary, mean and extreme streamflow changes are projected over most of the ice-free land surface high 2 train +4132 AR6_WGII 616 25 The magnitude of streamflow change is projected to increase with global warming in most regions high 2 train +4133 AR6_WGII 616 27 Annual mean runoff in one third of assessed glacierised catchments is projected to decline by at least 10% by 2100 under RCP4.5, with the most significant reductions in central Asia and the Andes medium 1 train +4134 AR6_WGII 616 29 Substantial fractions of ensemble projections disagree with the multi-model mean high 2 train +4135 AR6_WGII 616 30 With 1.5 and 2°C global warming, approximately 15 and 20% of the current global population, respectively, would experience both an increase in high streamflows and a decrease in low streamflows medium 1 train +4136 AR6_WGII 616 33 Both AR5 (Jiménez Cisneros et al., 2014) and SROCC (Hock et al., 2019b) concluded that spring snowmelt floods would be earlier (high confidence), and hazards from floods involving meltwater will gradually diminish, particularly at low elevation medium 1 train +4137 AR6_WGII 618 14 In all climate scenarios projected, earlier snowmelt leads to earlier spring floods high 2 train +4138 AR6_WGII 619 12 Projected increases in flooding pose increasing risks, with a 1.2–1.8 and 4–5 times increase in global GDP loss at 2°C and 4°C compared to 1.5°C warming, respectively medium 1 train +4139 AR6_WGII 619 13 Without adaptation, projected increases in flooding are 1.4 to 2.5 and 2.5 to 3.9 times in global GDP loss at 2°C and 3°C compared to 1.5°C warming, respectively medium 1 train +4140 AR6_WGII 619 16 However, the snowmelt floods are projected to decrease (medium confidence) and occur 25–30 d earlier in the year by the end of the 21st century with RCP8.5 high 2 train +4141 AR6_WGII 619 19 Many studies focus on precipitation-based drought indices (Carrão et al., 2018), but higher evaporative demands and changes in snow cover are additional drivers of hydrological, agricultural and ecological drought medium 1 train +4142 AR6_WGII 621 10 In summary, the likelihood of drought is projected to increase in many regions over the 21st century high 2 train +4143 AR6_WGII 621 11 Different forms of drought broadly show similar patterns of projected change in many regions (high confidence), but the frequency of agricultural drought is projected to increase over wider areas than for meteorological drought medium 1 train +4144 AR6_WGII 622 5 Overall, several recent studies of climate change impacts on groundwater in different parts of the world have concluded that projected groundwater recharge could either increase or decrease, and results are often uncertain high 2 train +4145 AR6_WGII 622 10 An emerging body of studies have projected amplification of episodic recharge in the tropics and semiarid regions due to extreme precipitation under global warming medium 1 test +4146 AR6_WGII 622 11 Climate change is also projected to impact groundwater-dependent ecosystems and groundwater quality negatively medium 1 train +4147 AR6_WGII 622 27 In the tropics and semiarid regions, growing precipitation intensification under global warming may enhance the resilience of groundwater through increased episodic recharge medium 1 train +4148 AR6_WGII 622 31 In addition, SROCC reported water quality degradation due to the release of legacy contaminants in glaciers and permafrost medium 1 train +4149 AR6_WGII 622 33 Water insecurity due to water quality degradation is projected to increase under climate change due to warming, enhanced floods and sea level rise (Arnell and Lloyd-Hughes, 2014; Dyer et al., 2014; Whitehead et al., 2015) medium 1 train +4150 AR6_WGII 623 7 In summary, climate change is projected to increase water pollution incidences, salinisation and eutrophication due to increasing drought and flood events, sea level rise and water temperature rise, respectively, in some local rivers and lakes, but there is a dearth of exact quantification at a global scale medium 1 train +4151 AR6_WGII 623 18 Herewith, recent studies demonstrate increasing impact of the projected climate change (increase of precipitation, thawing permafrost) on soil erosion medium 1 train +4152 AR6_WGII 624 1 SR1.5 further reinforced AR5 conclusions in terms of projected crop yield reductions, especially for wheat and rice (high confidence), loss of livestock and increased risks for small-scale fisheries and aquaculture medium 1 train +4153 AR6_WGII 625 16 In summary, agricultural water use is projected to increase globally due to cropland expansion and intensification and climate change- induced changes in water requirements high 2 train +4154 AR6_WGII 625 17 Parts of temperate drylands may experience increases in suitability for rain- fed production based on mean climate conditions; however, risks to rain-fed agriculture increase globally because of increasing variability in precipitation regimes and changes in water availability high 2 train +4155 AR6_WGII 625 19 Regions reliant on snowmelt for irrigation purposes will be affected by substantial reductions in water availability high 2 train +4156 AR6_WGII 625 29 For example, regions like central Africa, India, central Asia and northern high-latitude areas are projected to see more than 20% increases in gross hydropower potential high 2 train +4157 AR6_WGII 625 31 The Mediterranean region is projected to see almost a 40% reduction in hydropower production high 2 train +4158 AR6_WGII 625 32 On the other hand, northern Europe and India are projected to add to their hydropower production capacity due to climate change by mid-century high 2 train +4159 AR6_WGII 626 16 Apart from climate impacts on hydropower production, climate- induced flood loads and reservoir water level change may lead to dam failure under RCP2.6 and RCP4.5 scenarios (Fluixá-Sanmartín et al., 2018; Fluixá-Sanmartín et al., 2019) medium 1 train +4160 AR6_WGII 626 30 A systematic review showed consistent decreases in mid to end of the century in thermal power production capacity due to insufficiency of cooling water in southern, western and eastern Europe (high confidence); North America and Oceania (high confidence), central, southern and western Asia (high confidence) and western and southern Africa medium 1 train +4161 AR6_WGII 626 31 Overall, apart from emissions benefits, moving away from thermal power generation to other renewable energy will also lower the chances of climate-induced curtailment of energy production high 2 train +4162 AR6_WGII 627 4 While not WaSH-specific, AR5 showed that more people would experience water scarcity and floods (high confidence) and identified WaSH failure due to climate change as an emergent risk medium 1 train +4163 AR6_WGII 627 5 In addition, both SR1.5 (IPCC, 2018a) and SRCCL (IPCC, 2019b) projected the risk from droughts, heavy precipitation, water scarcity, wildfire damage and permafrost degradation to be higher at 2°C warming than 1.5°C medium 1 train +4164 AR6_WGII 627 6 Waterborne diseases result from complex causal relationships between climatic, environmental and socioeconomic factors that are not fully understood or modelled (Boholm and Prutzer, 2017) high 2 train +4165 AR6_WGII 627 8 In addition, changes in thermotolerance and chlorine resistance of certain viruses have been observed in laboratory experiments simulating different temperatures and sunlight conditions (Carratalà et al., 2020), increasing potential health risks even where traditional water treatment exists (Jiménez Cisneros et al., 2014) low 0 train +4166 AR6_WGII 628 3 These challenges include population growth, the rapid pace of urbanisation and inadequate investment, particularly in less developed economies with limited governance capacity high 2 train +4167 AR6_WGII 628 7 Modified streamflow is projected to affect the amount and variability of inflow to urban storage reservoirs high 2 train +4168 AR6_WGII 628 13 In summary, rapid population growth, urbanisation, ageing infrastructure and changes in water use are responsible for increasing the vulnerability of urban and peri-urban areas to extreme rainfall and drought, particularly in less developed economies with limited governance capacity high 2 train +4169 AR6_WGII 628 14 In addition, modified stream flows due to climate change (Section 4.4.3) are projected to affect the amount and variability of inflows to storage reservoirs that serve urban areas and may exacerbate challenges to reservoir capacity, such as sedimentation and poor water quality high 2 train +4170 AR6_WGII 628 16 Rising water temperatures are also projected to cause shifts in freshwater species distribution and worsen water quality problems high 2 train +4171 AR6_WGII 629 14 Furthermore, according to IPCC SR1.5 (Hoegh-Guldberg et al., 2018), if the world warms by 2°C–4°C by 2050, rates of human conflict could increase, but again, the role of hydrological change in this was not explicit medium 1 train +4172 AR6_WGII 629 16 Yet, since both conflictive and cooperative events are possible under conditions of climatic variability, whether conflict arises or increases depends on several contextual socioeconomic and political factors, including the adaptive capacity of the riparian states (Koubi, 2019), the existence of power asymmetries (Dinar et al., 2019) and pre-existing social tensions medium 1 train +4173 AR6_WGII 629 18 However, to date, other factors are considered more influential drivers of conflict, including lack of natural resource use regulations (Linke et al., 2018b), societal exclusion (von Uexkull et al., 2016; van Weezel, 2019), poor infrastructures and a history of violent conflict (Detges, 2016) high 2 train +4174 AR6_WGII 629 21 On the other side, intergroup cohesion (De Juan and Hänze, 2020) and policies that improve societal development and good governance reduce the risk of conflict associated with the challenges to adaptation to climate change (Hegre et al., 2016; Witmer et al., 2017) medium 1 train +4175 AR6_WGII 629 24 In addition, recent scenario analysis in global transboundary basins supports the finding that there is more potential for conflict in areas already under water stress, such as central Asia and the northern parts of Africa (Munia et al., 2020) medium 1 train +4176 AR6_WGII 629 27 Evidence exists that climate change imposes additional pressures on regions already under water stress or fragile and conflict-prone medium 1 train +4177 AR6_WGII 630 28 In synthesis, fundamental changes in observed climate are already visible in water-related outcomes high 2 train +4178 AR6_WGII 631 16 Projected increases in hydrological extremes pose increasing risks to societal systems globally high 2 train +4179 AR6_WGII 631 18 Similarly, a near doubling of drought duration (Naumann et al., 2018) and an increasing share of the population affected by various types, durations and severity levels of drought are projected high 2 train +4180 AR6_WGII 633 18 While there are increasing potentials of ~2–6% for hydropower production by 2080 (medium confidence), risks to thermoelectric power production increase for most regions high 2 train +4181 AR6_WGII 633 20 Increasing hydrological extremes also have consequences for the maintenance and further improvement of the provision of WaSH services medium 1 train +4182 AR6_WGII 633 24 Globally, climate change will exacerbate existing challenges for urban water services, driven by further population growth, the rapid pace of urbanisation and inadequate investment, particularly in less developed economies with limited governance capacity high 2 train +4183 AR6_WGII 633 27 However, water-specific conflicts between sectors and users may be exacerbated for some regions of the world high 2 train +4184 AR6_WGII 633 34 Hydrological change, especially increasing extreme events, pose risks to the cultural uses of water of Indigenous Peoples, local communities and traditional peoples (high confidence), with implications for the physical well-being of these groups high 2 train +4185 AR6_WGII 634 1 Small islands are already regularly experiencing droughts and freshwater shortages high 2 train +4186 AR6_WGII 634 7 In addition, FWLs are threatened by climate change due to changes in rainfall patterns, extended droughts and wash-over events caused by storm surges and sea level rise high 2 train +4187 AR6_WGII 634 23 In sum, small islands are already regularly experiencing droughts and freshwater shortages high 2 train +4188 AR6_WGII 634 24 For atoll islands, freshwater availability may be severely limited as early as 2030 low 0 train +4189 AR6_WGII 634 25 The effects of temperature increase, changing rainfall patterns, sea level rise and population pressure combined with limited options available for water-related adaptation leave small islands partially water-insecure currently, with increasing risks in the near-term and at warming above 1.5°C high 2 train +4190 AR6_WGII 635 2 There is high confidence that water-related adaptation is occurring in the agricultural sector (Acevedo et al., 2020; Ricciardi et al., 2020), and water-related adaptation in the agricultural sector makes up the majority of documented local, regional and global evidence of implemented adaptation high 2 test +4191 AR6_WGII 635 3 However, while there is increasing evidence of adaptation and its benefits across multiple dimensions, the link between adaptation benefits and climate risk reduction is unclear due to methodological challenges medium 1 train +4192 AR6_WGII 635 4 On the other hand, while it is methodologically possible to measure the effectiveness of future adaptation in reducing climate risks, the main limitation here is that not all possible ranges of future adaptations can be modelled given the limitations of climate and impact models high 2 train +4193 AR6_WGII 635 10 This measure is deemed to have economic benefits and benefits for vulnerable communities who adopt this measure (high confidence) and benefits in terms of water saving and positive ecological and sociocultural benefits medium 1 train +4194 AR6_WGII 635 22 In addition, this measure is shown to have positive economic benefits (high confidence) and also benefits on other parameters medium 1 train +4195 AR6_WGII 636 14 In sum, water-related adaptation in the agricultural sector is widely documented, with irrigation, agricultural water management, crop diversification and improved agronomic practices among the most common adaptation measures adopted high 2 train +4196 AR6_WGII 636 25 Overall, freshwater withdrawals for adapted cooling systems under all scenarios are projected to decline by −3% to −63% by 2100 compared to the base year of 2000 (Fricko et al., 2016) medium 1 train +4197 AR6_WGII 636 30 Furthermore, reducing the share of thermoelectric power with solar and wind energy (Tobin et al., 2018; Arango-Aramburo et al., 2019; Emodi et al., 2019) can be synergistic from both climate and water perspectives, as solar and wind energy have lower water footprints high 2 train +4198 AR6_WGII 636 34 For instance, sharing of hydropower revenues and profits to fund local infrastructure and pay dividends to local people has been practiced in Nepal and in some countries of the Mekong basin to enhance the social acceptability of hydropower projects (Balasubramanya et al., 2014; Shrestha et al., 2016) low 0 train +4199 AR6_WGII 637 28 Further, while irrigation expansion is one of the most commonly proposed adaptation responses, there are limitations to further increases in water use, as many regions are already facing water limitations under current climatic conditions high 2 train +4200 AR6_WGII 638 6 However, residual risks will remain, especially at higher levels of warming medium 1 train +4201 AR6_WGII 638 10 WaSH interventions have been demonstrated to reduce diarrhoea risk by 25–75% depending on the specific intervention (Wolf et al., 2018) high 2 train +4202 AR6_WGII 638 11 Conversely, inadequate WaSH is associated with an estimated annual loss of 50 million daily adjusted life years (Prüss-Ustün et al., 2019), of which 89% of deaths are due to diarrhoea, and 8% of deaths from acute respiratory infections (Chapter 7 WGII 7.3.2), making universal access to WaSH (i.e., achievement of SDG 6.1, 6.2) a critical adaptation strategy high 2 train +4203 AR6_WGII 638 16 Improved integrated (urban) water resources management (Kirshen et al., 2018; Tosun and Leopold, 2019) and governance (Chu, 2017; Miller et al., 2020) and enhanced ecosystem management (Adhikari et al., 2018b) lead to policies and regulations that reduce water insecurity and, when developed appropriately, reduce inequities medium 1 train +4204 AR6_WGII 638 17 Supply (source) augmentation, including dams, storage and rainwater/fog harvesting, can increase the supply or reliability of water for drinking, sanitation and hygiene (DeNicola et al., 2015; Pearson et al., 2015; Majuru et al., 2016; Poudel and Duex, 2017; Lucier and Qadir, 2018; Goodrich et al., 2019) high 2 train +4205 AR6_WGII 638 22 Demand for water can be decreased through reductions in water loss from the system (e.g., pipe leakage) (Orlove et al., 2019) and water conservation measures (Duran-Encalada et al., 2017) medium 1 train +4206 AR6_WGII 639 22 Further analyses of co-benefits, particularly employing a gender lens, are required to improve adaptation strategies (McIver et al., 2016).In summary, ensuring access to climate-resilient WaSH infrastructure and practices represents a key adaptation strategy that can protect beneficiaries against water-related diseases induced by climate change high 2 train +4207 AR6_WGII 639 23 Better management of water resources, supply augmentation and demand management are important adaptation strategies high 2 train +4208 AR6_WGII 640 15 In summary, although water-related adaptation is underway in the urban, peri-urban and municipal sectors of some nations, governance, technical and economic barriers remain in implementing locally informed strategies, particularly in developing countries (high confidence).4.6.6 Adaptation for Communities Dependent on Freshwater Ecosystems AR5 concluded that some adaptation responses in the urban and agricultural sectors could negatively impact freshwater ecosystems medium 1 train +4209 AR6_WGII 640 17 These have been implemented in many locations around the world, yet, challenges remain, including improving the evidence base of their effectiveness, scaling up of these interventions, mainstreaming across sectors and receiving more adaptation finance medium 1 train +4210 AR6_WGII 640 25 These measures also require further financial support, mainstreaming across sectors and the scaling up of individual measures medium 1 train +4211 AR6_WGII 640 27 However, challenges remain, including improving the evidence base of their effectiveness, scaling up these interventions, mainstreaming across sectors and receiving more adaptation finance medium 1 train +4212 AR6_WGII 643 11 At the urban and peri-urban scale, the use and effectiveness of NbS is a crucial feature to build resilience in cities for urban stormwater management and heat mitigation (Depietri and McPhearson, 2017; Carter et al., 2018; Huang et al., 2020; Babí Almenar et al., 2021) high 2 train +4213 AR6_WGII 643 20 Moreover, several NbS– —for example, natural (blue and green) and grey infrastructure—can help address water-related hazards such as coastal hazards, heavy precipitation, drought, erosion and low water quality high 2 train +4214 AR6_WGII 644 4 Vulnerable populations exposed to hydrological changes may become trapped due to a lack of economic and social capital required for migration (Adams, 2016; Zickgraf, 2018) medium 1 train +4215 AR6_WGII 644 7 Migration has increased vulnerability among women and female-headed households (Patel and Giri, 2019), but has also triggered gender-positive processes such as increased female school enrolment (Gioli et al., 2014) medium 1 train +4216 AR6_WGII 644 8 Remittances, that is, transfers of money from migrants to beneficiaries in sending areas, may reduce vulnerability and increase adaptive capacity to climate-induced hydrological changes (Ng’ang’a et al., 2016; Jha et al., 2018b) medium 1 train +4217 AR6_WGII 644 11 However, they often fail to include affected populations in the process and may lead to greater impoverishment and increased vulnerability (Wilmsen and Webber, 2015) medium 1 train +4218 AR6_WGII 644 14 In summary, measures that facilitate successful migration and inclusive resettlement may facilitate adaptation to climate-induced hydrological changes medium 1 train +4219 AR6_WGII 644 28 However, relocation can be culturally, socially, financially, politically and geographically constrained due to the importance of cultural relationships with traditional, customary or ancestral lands high 2 train +4220 AR6_WGII 645 6 In sum, although some Indigenous Peoples, local communities and traditional peoples can adapt, and are adapting to climate-driven hydrological changes, and their impacts on and risks to culturally significant practices and beliefs (medium confidence), these strategies are constrained by structural barriers and adaptation limits high 2 train +4221 AR6_WGII 646 26 Most of these adaptation case studies are from Asia and Africa, and agriculture is the predominant sector where most of these adaptation responses are being implemented high 2 train +4222 AR6_WGII 646 28 Agriculture is the most important sector in all continents, except Europe and Australasia, where most adaptation occurs in the urban sector high 2 train +4223 AR6_WGII 648 6 These top four responses provide several benefits such as higher incomes and yields, better water use efficiencies and related outcomes high 2 train +4224 AR6_WGII 648 9 Responses such as migration, including spontaneous and planned relocation, are also relatively well documented (medium confidence), as are responses such as collective action, training and capacity building and economic and financial measures for increasing adaptive capacities medium 1 train +4225 AR6_WGII 654 16 In developing countries, most adaptation measures improve economic outcomes high 2 train +4226 AR6_WGII 654 17 Adaptation responses also have benefits in terms of water outcomes and environmental and ecological parameters, and these benefits are more commonly manifested in developed countries high 2 train +4227 AR6_WGII 654 18 Of the papers assessed for water-related adaptation, roughly one fourth reported adaptation co-benefits high 2 train +4228 AR6_WGII 654 19 In contrast, one third of studies reported maladaptive outcomes, now or in the future high 2 train +4229 AR6_WGII 654 20 Despite many adaptation case studies, there is a knowledge gap in understanding if the benefits of adaptation also translate into a reduction of climate impacts, and if so, to what extent, and under what conditions high 2 train +4230 AR6_WGII 656 19 Results show a range of effectiveness levels across regions and warming levels and vary depending on the tested response options (Qin et al., 2018) (Figure 4.29), with moderate to small effectiveness, large residual impacts or potential maladaptive outcomes, as well as decreasing effectiveness with increasing warming (Figure 4.28) high 2 train +4231 AR6_WGII 656 28 For all regions, a reduction in effectiveness is apparent from 1.5°C to higher levels of warming, leading to increased residual risk with increasing warming high 2 train +4232 AR6_WGII 656 29 Irrigation can increase yield relative to present day, showing co-benefits for some regions, though the share of co-benefits decreases with higher warming high 2 train +4233 AR6_WGII 658 22 Adaptation generally performs more effectively at 1.5°C, though residual damages are projected at this warming level across sectors and regions high 2 train +4234 AR6_WGII 660 8 Reduction in the effectiveness of future adaptation at higher global warming levels emphasises the need for limiting warming to 1.5°C, as space for adaptation solution starts to shrink beyond that for most options for which future projections exists high 2 train +4235 AR6_WGII 660 10 First, the nature of literature on current adaptation makes it challenging to infer their effectiveness in reducing climate risks, even though the benefits of adaptation are clear high 2 train +4236 AR6_WGII 660 15 SR1.5 also shows that water-related risks can be reduced substantially by limiting warming to 1.5°C high 2 train +4237 AR6_WGII 660 17 The SRCCL further highlighted the critical importance of water-related climate change adaptation and potential limits to adaptation in the land sector when extreme forms of desertification lead to a complete loss of land productivity high 2 train +4238 AR6_WGII 660 18 Institutional constraints, including path dependency and lengthy decision- making processes, remain major limitations to successful adaptation globally high 2 train +4239 AR6_WGII 661 12 Water-related impacts that occurred despite implemented adaptation have been documented across all world regions high 2 test +4240 AR6_WGII 663 4 In summary, institutional constraints (governance, institutions, policy), including path dependency and financial and information constraints, are the main challenge to adaptation implementation in the water sector high 2 train +4241 AR6_WGII 663 5 Water-related losses and damages that manifest despite or beyond implemented adaptation have been observed across world regions, primarily for exposed and vulnerable communities high 2 train +4242 AR6_WGII 663 20 So, while there is general agreement about negative impacts on GDP due to water-related risks in the future, the magnitude of GDP loss estimates varies substantially and depends on various model assumptions high 2 train +4243 AR6_WGII 664 1 In summary, climate change impacts on water resources are projected to lower GDP in many low-and middle-income countries without adequate adaptation measures high 2 train +4244 AR6_WGII 664 4 Still, more work needs to be done on actual benefits and costs of adaptation strategies and residual impacts and risks of delaying adaptation action medium 1 train +4245 AR6_WGII 664 5 In addition, better evidence on the costs and benefits of low-regret solutions, such as water pricing, increasing water use efficiency through technology and service improvements, and enhanced support for autonomous adaptation, is also needed for informed decision-making high 2 train +4246 AR6_WGII 664 8 Different mitigation pathways can either increase or decrease water withdrawals or water consumption (or both, or either) depending on the specific combination of mitigation technologies deployed high 2 train +4247 AR6_WGII 665 14 Some DAC technologies that include solid sorbents also produce water as a by-product, but not in quantities that can offset total water losses (Beuttler et al., 2019; Fasihi et al., 2019) medium 1 train +4248 AR6_WGII 665 16 Results from a simulation study on retrofitting coal- fired power plants built after 2000 with carbon capture and storage (CCS) technologies show an increase in global water consumption, currently at 9.66 km3 yr–1, by 31–50% (to 12.66 km3 yr–1 and 14.47 km3 yr–1, respectively) depending on the cooling and CCS technology deployed, and hence are best deployed in locations which are not water scarce (Rosa et al., 2020c) medium 1 train +4249 AR6_WGII 665 18 Carbon can be ‘scrubbed’ from thermoelectric power plant emissions and injected for storage in deep geological strata (Turner et al., 2018), but this can lead to pollution of deep aquifers (Chen et al., 2021) and have health consequences low 0 train +4250 AR6_WGII 665 32 Overall, extensive BECCS and afforestation/reforestation deployment can alter the water cycle at regional scales high 2 train +4251 AR6_WGII 666 2 Many mitigation measures have a considerable water footprint high 2 train +4252 AR6_WGII 666 34 The effectiveness of technology in reducing climate-related risks depends on its appropriateness to the local context (Biagini et al., 2014; Mfitumukiza et al., 2020) and other factors, including institutional and governance frameworks high 2 train +4253 AR6_WGII 667 2 Water-related technologies can also have adverse distributional outcomes when gains from technology adoption accrue disproportionately to a small section of the population; for example, only rich and male farmers can adopt high-cost technologies like solar irrigation pumps (Gupta, 2019) medium 1 test +4254 AR6_WGII 667 14 Such financing focuses on returns and scale (Cholibois, 2020), and as such, local needs, especially those of the poor, may not be adequately represented (Manuamorn et al., 2020; Williams, 2020) medium 1 test +4255 AR6_WGII 667 19 In summary, water garners a significant share of public and private adaptation funds high 2 train +4256 AR6_WGII 667 31 The necessity of water collection takes away time from income-generating activities and education high 2 train +4257 AR6_WGII 668 8 Hence, a lack of gender-sensitive analysis before implementing water management projects can lead to maladaptation and increase gender vulnerability (Phan et al., 2019; Eriksen et al., 2021) high 2 train +4258 AR6_WGII 668 17 SRCCL found that IKLK contribute to enhancing resilience against climate change and combating desertification medium 1 train +4259 AR6_WGII 668 25 Community-led actions and restoration measures are helping to ameliorate climate impacts and provide ‘safe havens’ to affected freshwater species high 2 train +4260 AR6_WGII 668 29 Community-led applications of IKLK in conjunction with external knowledge and funding can improve water security high 2 train +4261 AR6_WGII 669 7 In summary, IKLK are dynamic and have developed over time to adapt to climate and environmental change in culturally specific and place-based ways high 2 train +4262 AR6_WGII 669 13 Effective participation of these actors in climate change adaptation planning in the water sector can contribute to more just adaptation actions high 2 train +4263 AR6_WGII 669 28 Therefore, the legitimacy of the decisions taken by multiple decision-makers at different levels of water governance derives from the perceived fairness of the decision-making process (Baldwin et al., 2018) and the inclusion of women, Indigenous Peoples and young people (Iza, 2019) medium 1 train +4264 AR6_WGII 670 1 Polycentric governance systems require cross- scale information sharing, coordination and democratic participation to work appropriately (Pahl-Wostl and Knieper, 2014; Carlisle and Gruby, 2017; Morrison et al., 2017; Biesbroek and Lesnikowski, 2018; Frey et al., 2021) high 2 test +4265 AR6_WGII 670 4 These, in turn, can produce better environmental outcomes and improve water governance outcomes high 2 train +4266 AR6_WGII 670 8 For instance, historical inequities and injustices due to settler colonialism and top-down water policies, governance and laws (Collins et al., 2017; Arsenault et al., 2018; Johnson et al., 2018; Robison et al., 2018) have resulted in long-term water insecurity in many indigenous communities in North America (Simms et al., 2016; Medeiros et al., 2017; Conroy-Ben and Richard, 2018; Diver, 2018; Emanuel, 2018) high 2 train +4267 AR6_WGII 670 12 In summary, polycentric governance can enable improved water governance and effective climate change adaptation medium 1 train +4268 AR6_WGII 670 13 However, it can also exacerbate existing inequalities as long as less powerful actors, such as women, Indigenous Peoples and young people, are not adequately involved in the decision-making process high 2 train +4269 AR6_WGII 670 30 Evidence suggests that adaptation failure in the water sector is due to policy and regulatory failures (Keohane and Victor, 2016; Oberlack and Eisenack, 2018; Javeline et al., 2019), reflecting political myopia (Muller, 2018; Empinotti et al., 2019; Pralle, 2019) high 2 train +4270 AR6_WGII 670 31 International donors and supranational/transnational legislation (e.g., EU law) can support the capacity of national and sub-national governments to act and remove possible barriers to the effective implementation of climate change adaptation policies in the water sector, including obstacles posed due to lack of financial support for the developing countries (Massey et al., 2014; Tilleard and Ford, 2016; Biesbroek et al., 2018; Rahman and Tosun, 2018) medium 1 train +4271 AR6_WGII 728 1 Human-induced warming has slowed growth of agricultural productivity over the past 50 years in mid and low latitudes medium 1 train +4272 AR6_WGII 728 2 Crop yields are compromised by surface ozone high 2 train +4273 AR6_WGII 728 3 Methane emissions have negatively impacted crop yields by increasing temperatures and surface ozone concentrations medium 1 train +4274 AR6_WGII 728 4 Warming is negatively affecting crop and grassland quality and harvest stability high 2 train +4275 AR6_WGII 728 5 Warmer and drier conditions have increased tree mortality and forest disturbances in many temperate and boreal biomes (high confidence), negatively impacting provisioning services medium 1 train +4276 AR6_WGII 728 6 Ocean warming has decreased sustainable yields of some wild fish populations high 2 train +4277 AR6_WGII 728 7 Ocean acidification and warming have already affected farmed aquatic species high 2 train +4278 AR6_WGII 728 10 At higher latitudes, warming has expanded potential area but has also altered phenology (high confidence), potentially causing plant–pollinator and pest mismatches medium 1 train +4279 AR6_WGII 728 11 At low latitude, temperatures have crossed upper tolerance thresholds, more frequently leading to heat stress high 2 train +4280 AR6_WGII 728 14 Droughts, floods and marine heatwaves contribute to reduced food availability and increased food prices, threatening food security, nutrition and livelihoods of millions high 2 train +4281 AR6_WGII 728 15 Droughts induced by the 2015–2016 El Niño, partially attributable to human influences (medium confidence), caused acute food insecurity in various regions, including eastern and southern Africa and the dry corridor of Central America high 2 train +4282 AR6_WGII 728 16 In the northeast Pacific, a recent 5-year warm period impacted the migration, distribution and abundance of key fish resources high 2 train +4283 AR6_WGII 728 27 Increasing competition for land, energy and water exacerbates impacts of climate change on food security high 2 train +4284 AR6_WGII 728 29 Current global crop and livestock areas will increasingly become climatically unsuitable under a high- emission scenario high 2 train +4285 AR6_WGII 728 30 Increased, potentially concurrent climate extremes will periodically increase simultaneous losses in major food-producing regions medium 1 train +4286 AR6_WGII 728 32 Climate change will increase the number of people at risk of hunger in mid-century, concentrated in Sub-Saharan Africa, South Asia and Central America high 2 train +4287 AR6_WGII 728 33 Increased CO 2 concentrations will reduce nutrient density of some crops high 2 train +4288 AR6_WGII 728 36 The number of days with climatically stressful conditions for outdoor workers will increase by up to 250 workdays per year by century’s end in some parts of South Asia, tropical sub-Saharan Africa and parts of Central and South America under Shared Socioeconomic Pathway (SSP) 5-8.5, with negative consequences such as reduced food productivity, higher costs and prices medium 1 train +4289 AR6_WGII 729 1 Meat and milk productivity will be reduced medium 1 train +4290 AR6_WGII 729 3 Climate change will reduce the effectiveness of pollinator agents as species are lost from certain areas, or the coordination of pollinator activity and flower receptiveness is disrupted in some regions high 2 train +4291 AR6_WGII 729 4 Greenhouse-gas emissions will negatively impact air, soil and water quality, exacerbating direct climatic impacts on yields high 2 train +4292 AR6_WGII 729 6 Climate change will reduce marine fisheries and aquaculture productivity, altering the species that will be fished or cultured, and reducing aquaculture habitat in tropical and subtropical areas high 2 train +4293 AR6_WGII 729 7 Global ocean animal biomass will decrease by 5–17% under RCP2.6 and 8.5, respectively, from 1970 to 2100 with an average decline of 5% for every 1°C of warming, affecting food provisioning, revenue value and distribution medium 1 train +4294 AR6_WGII 729 8 Global marine aquaculture will decline under warming and acidification from 2020 to 2100, with potential short-term gains for temperate finfish and overall negative impacts on bivalve aquaculture from habitat reduction (50–100% for some countries in the Northern Hemisphere) medium 1 train +4295 AR6_WGII 729 9 Changes in precipitation, sea level, temperature and extreme climate events will affect food provisioning from inland and coastal aquatic systems high 2 train +4296 AR6_WGII 729 10 Sea level rise and altered precipitation will increase coastal inundation and water conflicts between water-dependent sectors, such as rice production, direct human use and hydropower medium 1 train +4297 AR6_WGII 729 15 In temperate and boreal regions, some productivity gains are projected, but tree mortality will increase in some areas high 2 train +4298 AR6_WGII 729 16 In tropical forests, change in species composition and forest structure will lower production medium 1 train +4299 AR6_WGII 729 17 Some models project a possible increase in global wood supply and lowering of average wood prices, but they do not account for the negative impacts of extreme events and thus possibly overestimate the wood supply medium 1 train +4300 AR6_WGII 729 19 Higher temperatures and humidity will favour toxigenic fungi, plant and animal-based pathogens, and harmful algal blooms (HABs) high 2 train +4301 AR6_WGII 729 20 More frequent and intense flood events and in- creased melting of snow and ice will increase food contamination high 2 train +4302 AR6_WGII 729 21 Incidence and severity of HABs and water-borne diseases will increase, as will indirect effects from infrastructure damage during extreme events high 2 train +4303 AR6_WGII 729 23 Autonomous responses include livestock and farm management, switching varieties/species and altered timing of key farm activities such as planting or stocking high 2 train +4304 AR6_WGII 729 24 However, because of limited adaptive capacities and non-climatic compounding drivers of food insecurity, SDG2 will not be met high 2 train +4305 AR6_WGII 729 29 Ecosystem-based approaches support long-term productivity and ecosystem services such as pest control, soil health, pollination and buffering of temperature extremes (high confidence), but potential and trade-offs vary by socioeconomic context, ecosystem zone, species combinations and institutional support medium 1 train +4306 AR6_WGII 729 31 A sustainable bioeconomy relying on bioresources will need to be supported by technology innovation and international cooperation and governance of global trade to disincentivise environmental and social externalities medium 1 train +4307 AR6_WGII 730 1 Adaptive transboundary governance and ecosystem-based management, livelihood diversification, capacity development and improved knowledge-sharing will reduce conflict and promote the fair distribution of sustainably harvested wild products and revenues medium 1 train +4308 AR6_WGII 730 2 Other options include shared quotas and access rights considering trade-offs, shifting livelihoods to follow target species, new markets for emerging species, and technology {Cross Chapter Box MOVING PLATE this chapter, 5.8.4, 5.14.3.4} Implemented adaptation in crop production will be insufficient to offset the negative effects of climate change high 2 train +4309 AR6_WGII 730 3 Currently available management options have the potential to compensate global crop production losses due to climate change up to ~2°C warming, but the negative impacts even with adaptation will grow substantially from the mid-century under high temperature change scenarios high 2 train +4310 AR6_WGII 730 4 Regionally, the negative effects will prevail sooner where current temperatures are already higher as in lower latitudes high 2 train +4311 AR6_WGII 730 6 Policies that support system transitions include shifting subsidies, removing perverse incentives, regulation and certification, green public procurement, investment in sustainable value chains, support for capacity-building, access to insurance premiums, payments for ecosystem services, and social protection, among others medium 1 train +4312 AR6_WGII 730 8 Adaptation strategies that address power inequities lead to co-benefits in equity outcomes and resilience for vulnerable groups medium 1 train +4313 AR6_WGII 730 9 Indigenous knowledge and local knowledge facilitate adaptation strategies for ecosystem provisioning, especially when combined with scientific knowledge using participatory and community-based approaches high 2 train +4314 AR6_WGII 730 11 Lacking sufficient stakeholder participation, large-scale land acquisitions have had mostly negative implications for vulnerable groups and climate change adaptation high 2 train +4315 AR6_WGII 730 12 Policy and programme appraisal of adaptation options that consider the risks of adverse effects across different groups at different scales and use inclusive rights-based approaches help avoid maladaptation medium 1 train +4316 AR6_WGII 730 13 Successful forest adaptation involves recognition of land rights and cooperation with Indigenous Peoples and other local communities who depend on forest resources high 2 train +4317 AR6_WGII 730 15 Public-sector investment in adaptation of agriculture, forestry and fisheries has grown four-fold since 2010, but adaptation costs will be much higher to meet future adaptation needs medium 1 train +4318 AR6_WGII 730 16 Expanding access to financial services and pooling climate risks will enable and incentivise climate change adaptation medium 1 train +4319 AR6_WGII 730 18 Robust analyses are needed that detail plausible path- ways to move towards more resilient, equitable and sustainable food systems in ways that are socially, economically and environmentally acceptable through time high 2 train +4320 AR6_WGII 730 19 Appropriate monitoring and rapid feedback to food system actors will be critical to the success of many current and future adaptation actions high 2 train +4321 AR6_WGII 731 21 Climate change impacts on fisheries will be particularly high in tropical regions, where reductions in catch are expected to be among the largest globally, leading to negative economic and social effects for fishing communities and with implications for the supply of fish and shellfish high 2 train +4322 AR6_WGII 735 1 Climate change has caused regionally different, but mostly negative, impacts on crop yields and quality and marketability of products high 2 train +4323 AR6_WGII 735 7 Methane emissions significantly impact crop yields by increasing temperatures as a greenhouse gas (GHG) and surface ozone concentrations as a precursor medium 1 train +4324 AR6_WGII 735 9 Although these estimates were not linked with historical yield changes, more than half of the estimated yield loss is attributable to increasing temperature and ozone concentrations from methane emissions, suggesting the importance of methane mitigation in alleviating yield losses medium 1 train +4325 AR6_WGII 735 10 Climate change is already affecting livestock production high 2 train +4326 AR6_WGII 735 13 In aquatic systems, more evidence has accumulated since AR5 on warming-induced shifts (mainly poleward) of species high 2 train +4327 AR6_WGII 735 16 The effects of climate change on aquaculture are apparent but diverse, depending on the types and species of aquaculture high 2 train +4328 AR6_WGII 736 3 Climate change will increase malnourished populations through direct impacts on food production and have cascading impacts on food prices and household incomes, all of which will reduce access to safe and nutritious food high 2 train +4329 AR6_WGII 736 4 Extreme climate events will become more frequent and force some of the current food production areas beyond the safe climatic space for production high 2 train +4330 AR6_WGII 736 18 The safety challenges arise from contamination caused by increased prevalence of pathogens, HAB and toxic inorganic bioaccumulation high 2 train +4331 AR6_WGII 739 13 Recent warming trends have generally shortened the life cycle of major crops high 2 train +4332 AR6_WGII 739 15 Conversely, in mid-to-low latitudes in Asia, a review study found that farmers favoured early maturing cultivars to reduce risks of damages due to drought, flood and/or heat (Shaffril et al., 2018), suggesting that region-specific adaptations are already occurring in different parts of the world high 2 train +4333 AR6_WGII 739 17 Plant breeding, fertilisation, irrigation and integrated pest management have been the major drivers, but many studies have found significant impacts from recent climate trends on crop yield high 2 train +4334 AR6_WGII 739 30 Nevertheless, yield variability in less productive regions has severe impacts on local food availability and livelihood high 2 train +4335 AR6_WGII 741 13 Different pest species populations respond differently to ongoing climate change, with some shifting, contracting or expanding their current distribution range and others persisting or disappearing in their current range high 2 train +4336 AR6_WGII 742 1 Large-scale sea surface temperature (SST) oscillations greatly influence global yield of major crops high 2 train +4337 AR6_WGII 742 2 Some studies showed that crop yields in different regions covaried with SST oscillations, suggesting occurrences of tele-connected yield failures (crop losses caused by related factors in distant regions; Table Box 5.1.1) medium 1 train +4338 AR6_WGII 743 6 Surface ozone concentration has increased substantially since the late 19th century (Cooper et al., 2014; Forster et al., 2021; Gulev et al., 2021; Szopa et al., 2021) and in some locations and times reaches levels that harm plants, animals and human high 2 train +4339 AR6_WGII 745 10 Social inequity How social inequity increases vulnerability to climate change in cropping systems Gender inequity can create and worsen social vulnerability to climate change impacts within cropping systems high 2 train +4340 AR6_WGII 745 14 Globally, smallholder food producers are more vulnerable than large-scale producers to climate change impacts high 2 train +4341 AR6_WGII 745 19 Farmworkers are another social group with heightened vulnerability to climate change medium 1 train +4342 AR6_WGII 746 1 Recent FACE studies found that the effects of elevated CO 2 are greater under water-limited conditions medium 1 train +4343 AR6_WGII 746 3 There are significant interactions between CO 2, temperature, cultivars, nitrogen and phosphorous nutrients (Kimball, 2016; Toreti et al., 2020): positive effects of rising CO 2 on yield are significantly reduced by higher temperatures for soybean, wheat and rice medium 1 train +4344 AR6_WGII 746 6 Elevated CO 2 reduces some important nutrients such as protein, iron, zinc and some grains, fruit or vegetables to varying degrees depending on crop species and cultivars high 2 train +4345 AR6_WGII 746 7 Elevated CO2 reduces some important nutrients such as protein, iron, zinc and some grains, fruit or vegetables to varying degrees depending on crop species and cultivars (Mattos et al., 2014; Myers et al., 2014; Dong et al., 2018; Scheelbeek et al., 2018; Zhu et al., 2018a; Jin et al., 2019; Ujiie et al., 2019). TThis is of particular relevance for fruit and vegetable crops given their importance in human nutrition high 2 test +4346 AR6_WGII 746 11 Additional research confirms that climate change will disproportionately affect crop yields among regions, with more negative than positive effects being expected in most areas, especially in currently warm regions, including Africa and Central and South America high 2 train +4347 AR6_WGII 750 3 Recent warming trends have advanced flowering, maturity and harvest high 2 train +4348 AR6_WGII 750 8 Developmental phases are projected to proceed faster in response to warming high 2 train +4349 AR6_WGII 750 18 In regions where low temperature is a limiting factor, warming will enable growers to grow a wider range of varieties and obtain better-quality wines high 2 train +4350 AR6_WGII 750 19 Subtropical and Mediterranean regions will experience major declines in fruit quality for high-quality wines high 2 train +4351 AR6_WGII 750 23 Box 5.3: Pollinators Climate change will reduce the effectiveness of pollinator agents as species are lost from certain areas, or the coordination of pollinator activity and flower receptiveness is disrupted in some regions high 2 train +4352 AR6_WGII 752 15 Increase in soil temperature will negatively impact SOC, but primarily in higher latitudes medium 1 train +4353 AR6_WGII 752 29 Soil C sequestration is an important strategy to improve crop and livestock production sustainably that could be applied at large scales and at a low cost, if there was adequate institutional support and labour, using agroforestry, conservation agriculture, mixed cropping and targeted application of fertilizer and compost high 2 train +4354 AR6_WGII 753 4 Climate change is projected to have negative impacts on CES medium 1 train +4355 AR6_WGII 753 25 There are still large uncertainties in the crop model projections (Müller et al., 2021a), but these multiple lines of evidence suggest that warming beyond +2°C (projected to be reached by mid-century under high-emission scenarios) will substantially increase the cost of adaptation and the residual damage to major crops high 2 train +4356 AR6_WGII 754 8 Plant breeding biotechnology will contribute to adaptation for large-scale producers high 2 train +4357 AR6_WGII 755 1 Genome sequencing significantly increases the rate and accuracy for identifying genes of agronomic traits that are relevant to climate change, including adaptation to stress from pests and disease, temperature and water extremes high 2 train +4358 AR6_WGII 755 9 Other breeding approaches assisted by genomics have been making steady gains in introducing traits that adapt crops to climate change high 2 train +4359 AR6_WGII 755 12 Modern biotechnology has not demonstrated the scale neutrality needed to serve smallholder-dominated agroecosystems, due to a combination of the kinds of traits and restrictions that come from the predominant intellectual property rights instruments used in their commercialisation, as well as the focus on a small number of major crop species medium 1 train +4360 AR6_WGII 757 3 Diversification improves regulating and supporting ecosystem services such as pest control, soil fertility and health, pollination, nutrient cycling, water regulation and buffering of temperature extremes (high confidence) (Barral et al., 2015; Prieto et al., 2015; Tiemann et al., 2015; Schulte et al., 2017; Beillouin et al., 2019a; Dainese et al., 2019; Kuyah et al., 2019; Tamburini et al., 2020), which can in turn mediate yield stability and reduced risk of crop loss according to socio-ecological contexts and time since adoption high 2 train +4361 AR6_WGII 757 14 Evidence is accumulating that rising temperatures are increasing heat stress in domestic species and affecting productivity high 2 train +4362 AR6_WGII 757 19 Changing seasonality, increasing frequency of drought and rising temperatures are affecting pastoral systems globally high 2 train +4363 AR6_WGII 758 15 Recent stagnation in dairy production in West Africa and China may be associated with increased periods of high daily temperatures low 0 train +4364 AR6_WGII 758 17 Escalating demand for livestock products in low-to-middle-income countries (LMICs) may necessitate considerable adaptation in the face of new thermal environments medium 1 train +4365 AR6_WGII 758 20 Under SSP5-8.5 to mid-century, land suitability for livestock production will decrease because of increased heat stress prevalence in mid and lower latitudes high 2 train +4366 AR6_WGII 758 32 Growing infectious disease burdens in domesticated animals may have wide- ranging impacts on the vulnerability of rural livestock producers in the future, particularly related to human health and projected increases in zoonoses high 2 train +4367 AR6_WGII 759 7 Gender inequities can act as a risk multiplier, with women being more vulnerable than men to climate-change- induced food insecurity and related risks high 2 train +4368 AR6_WGII 759 23 The future makeup of grasslands under climate change is uncertain, given the variation in responses of the component species, though this variation may provide a climate buffer (Jones, 2019) low 0 train +4369 AR6_WGII 761 6 The prevalence and occurrence of some livestock diseases are positively associated with extreme weather events high 2 train +4370 AR6_WGII 762 20 Climate change will negatively affect the provisioning of social benefits in many of the world’s grasslands medium 1 train +4371 AR6_WGII 762 27 Combining adaptations can result in increases in benefits in terms of production and livelihoods over and above those attainable from single adaptations high 2 train +4372 AR6_WGII 764 18 Large pulses of tree mortality were consistently linked to warmer and drier than average conditions for forests throughout the temperate and boreal biomes high 2 train +4373 AR6_WGII 764 21 A recent example of the impacts of climatic extremes is the European drought of 2018 (Buras et al., 2020), which led to a significant browning of the vegetation and resulted in widespread tree mortality high 2 train +4374 AR6_WGII 765 2 However, the observed patterns of post-disturbance recovery vary with region, with reduced tree regeneration reported for the western USA (Stevens-Rumann and Morgan, 2019; Turner et al., 2019) but robust recovery observed in Canada (White et al., 2017) and Central Europe medium 1 train +4375 AR6_WGII 765 30 Climate-induced disturbances could also reduce the temporal stability of ecosystem service supply (Albrich et al., 2018), increasing the volatility of timber markets medium 1 train +4376 AR6_WGII 766 1 On a global and regional scale, there is limited evidence and high agreement medium 1 train +4377 AR6_WGII 766 4 According to these studies, global timber supply will increase as the result of an increase in global forest growth under climate change scenarios medium 1 train +4378 AR6_WGII 768 1 Evidence emerging since the last assessment report further bolstered the notion that adapting the tree species composition to more warm-tolerant and less disturbance-prone species can significantly mitigate climate change impacts high 2 train +4379 AR6_WGII 768 3 Furthermore, increasing the diversity of tree species within stands can have positive effects on tree growth and reduce disturbance impacts high 2 train +4380 AR6_WGII 768 5 Managing for continuous forest cover can also help to maintain the forest microclimate and buffer tree regeneration and the forest floor community against climate change high 2 train +4381 AR6_WGII 768 8 However, recent evidence suggests that these measures diminish in efficiency under climate change and can have corollary effects on other important forest functions such as carbon storage and habitat quality medium 1 train +4382 AR6_WGII 768 9 Also, measures targeting landscape structure and composition have proven effective for increasing the climate resilience of forest systems medium 1 train +4383 AR6_WGII 769 33 Experimental trials have shown that drought stresses increase phytochemical content, either by decreasing biomass or increasing metabolites production high 2 train +4384 AR6_WGII 771 7 Climate-change-induced impacts of access to wild foods are also of concern in Arctic regions high 2 train +4385 AR6_WGII 771 15 Communities across other (non-Arctic) parts of North America and Europe also report declining availability of wild foods, with climate change among the perceived drivers for decline medium 1 train +4386 AR6_WGII 775 2 Marine heatwaves have increased in frequency over the 20th century, with an approximate doubling since the 1980s (high confidence), and their intensity and duration have also increased medium 1 train +4387 AR6_WGII 775 4 The surface open ocean pH has declined globally over the last 40 years by 0.003–0.026 pH per decade (virtually certain), and a decline in the ocean interior pH has been observed in all ocean basins over the past two to three decades high 2 train +4388 AR6_WGII 775 15 Observed impacts in some inland aquatic systems indicate substantial productivity reductions medium 1 train +4389 AR6_WGII 775 27 These changes are of particular concern in regions with few nutrition alternatives, such as low-income countries in Africa, Asia, Australasia, and Central and South America high 2 train +4390 AR6_WGII 775 28 Freshwater ecosystems that support most inland fisheries are under continuing threat from changes in land use, water availability and pollution and other pressures that will be exacerbated by climate change high 2 train +4391 AR6_WGII 776 6 The concern over aquatic food products’ safety due to climate change is increasing high 2 train +4392 AR6_WGII 776 9 The social vulnerability can differ largely between locations, even between relatively close coastal or inland communities (Bennett et al., 2014; Maina et al., 2016; Ndhlovu et al., 2017; Martins et al., 2019) and among inhabitants within a location, depending on factors such as access to other economic activities, education, health, adults in the household, and political connections high 2 train +4393 AR6_WGII 776 14 In general, gendered division of labour tends to cause lower salaries for women and different perception and experience of risk to climate change impacts high 2 train +4394 AR6_WGII 776 17 Coordinated fisheries management can substantially expand capacity to respond to a changing climate (Pinsky et al., 2020), but a great deal of political will, capacity building and collective action will be necessary high 2 train +4395 AR6_WGII 776 18 Today, approximately half the world’s population (~4 billion out of 7.8 billion people) are assessed as being currently subject to severe water scarcity for at least 1 month per year (medium confidence) (Box 4.1), and freshwater inland fisheries are particularly vulnerable as they are given lower priority for water resources than other sectors high 2 train +4396 AR6_WGII 777 8 Temperate tunas (albacore, Atlantic bluefin and southern bluefin) and the tropical bigeye tuna are expected to decline in the tropics and shift poleward by the end of the century under RCP8.5, while skipjack and yellowfin tunas are projected to increase abundance in tropical areas of the eastern Pacific but decrease in the equatorial western Pacific medium 1 train +4397 AR6_WGII 777 12 However, in general, where land barriers constrain the latitudinal shifts, the expected impacts of climate change are population declines and reduced productivity high 2 train +4398 AR6_WGII 777 13 Besides direct impacts on the abundance of fisheries-targeted species, climate-change-induced proliferation of invasive species could also affect fisheries’ productivity low 0 train +4399 AR6_WGII 777 14 Shifting marine fisheries will affect national economies high 2 train +4400 AR6_WGII 777 19 Ocean acidification is also expected to drive large global economic impacts medium 1 train +4401 AR6_WGII 777 23 In addition to temperature and water availability stress, climate change will bring new water quality challenges in freshwater systems, including increased dissolved organic carbon and toxic metal loads high 2 train +4402 AR6_WGII 777 26 Climate warming may enhance northward colonisation of water bodies of commercial freshwater species in the Arctic, where there are few ecological competitors medium 1 train +4403 AR6_WGII 778 2 Specifically, overfishing is the most critical non-climatic driver affecting the sustainability of fisheries, and therefore improving management could help rebuild fish stocks, reduce ecosystem impacts and increase the adaptive capacity of fishing high 2 train +4404 AR6_WGII 778 12 For Pacific Islands and Coastal Territories, fisheries adaptation will require significant investment from local governments and the private sector (Rosegrant et al., 2016), and reducing dependence on or finding alternatives to vulnerable marine resources (Johnson et al., 2020; Mabe and Asase, 2020).Adaptive capacity is strongly associated with social capital (i.e., the networks, shared norms, values and understandings that facilitate cooperation within or among groups) high 2 train +4405 AR6_WGII 778 13 Improving information flows allows for a more efficient co-management implementation medium 1 train +4406 AR6_WGII 778 16 Engaging Indigenous Peoples and local communities as partners across climate research ensures this knowledge is utilised, enhancing the usefulness of assessments (Bindoff et al., 2019) and facilitating the co-construction and implementation of sustainable solutions medium 1 train +4407 AR6_WGII 779 2 Marine, freshwater and terrestrial systems are already experiencing species shifts in response to climate change (very high confidence) (see also Sections 2.4.2.1. and 3.4.3., Figure MOVING PLATE.1 this chapter), with subsequent impacts on food provisioning services, pests and diseases high 2 train +4408 AR6_WGII 779 8 Nutritional dependency, cultural importance, livelihood, or economic reliance on shifting species will increase impacts of climate change, especially for small-scale fishers (marine and freshwater), farmers, women and communities highly dependent on local sources of food and nutrition high 2 train +4409 AR6_WGII 779 15 Shifting species have negative implications for the equitable distribution of food provisioning services, increasing the complexity of resolving sovereignty claims and climate justice high 2 train +4410 AR6_WGII 780 1 Flexible and rapid policy reform and management adaptation will help to meet sustainability targets (Nguyen et al., 2016; Pentz and Klenk, 2020), and may only be available for countries with the scientific, technical and institutional capacity to implement these high 2 train +4411 AR6_WGII 780 5 Overall, decreases in GHG emissions under future scenarios would reduce increases in global temperatures and limit species shifts, thereby lowering the likelihood of conflicts and food insecurity high 2 train +4412 AR6_WGII 780 16 Terrestrial Species Shifts There is robust evidence of shifts that terrestrial species have shifted poleward in high latitudes, with general declines of sea-ice dependent as well as some extreme-polar-adapted species high 2 train +4413 AR6_WGII 784 1 Ecosystem-based management approaches exist for terrestrial, marine and freshwater systems, but have proved successful only with early engagement of local small-scale, subsistence fishers/harvesters, utilising Indigenous knowledge and local knowledge and needs, in addition to those of larger-scale operators high 2 train +4414 AR6_WGII 784 4 Aquaculture can contribute to SDGs by reducing poverty and food insecurity, filling increasing aquatic food demand shortages from declining capture fisheries production medium 1 train +4415 AR6_WGII 784 12 Marine heatwaves have been increasing in both incidence and longevity over the past century (Frolicher and Laufkotter, 2018; Oliver et al., 2018; Bricknell et al., 2021), with productivity consequences for marine aquaculture (mariculture), carbon sequestration and local species extinctions high 2 train +4416 AR6_WGII 784 13 Temperature increases related to El Niño climatic oscillations have caused mass fish mortalities either through warming waters (e.g., Pacific threadfin in Hawaii (McCoy et al., 2017)) or associated HABs (e.g., 12% loss of Atlantic salmon as well as other fish and shellfish in Chile in 2016, with estimated USD 800 million in losses high 2 train +4417 AR6_WGII 784 14 Increases in sea lice parasite infestations on salmon are related to higher salinity and warmer waters medium 1 train +4418 AR6_WGII 784 15 Ocean acidification is having negative impacts on the sustainability of mariculture production high 2 train +4419 AR6_WGII 784 16 Ocean oxygen levels are declining due to climate change (Hoegh-Guldberg et al., 2018; IPCC, 2021), and decreased oxygen (hypoxia) has negative impacts on fish physiology (Cadiz et al., 2018; Hvas and Oppedal, 2019; Martos-Sitcha et al., 2019; Perera et al., 2021), fish growth, behaviour and sensitivity to concurrent stressors high 2 train +4420 AR6_WGII 784 17 Observed climate impacts on inland aquaculture systems have generally been site and region specific high 2 test +4421 AR6_WGII 784 18 Salinity intrusions into freshwater aquaculture systems have changed oxygen and water quality of inland ponds, resulting in mortalities in areas such as India and Bangladesh medium 1 train +4422 AR6_WGII 784 19 Rapid changes in temperature, precipitation, droughts, floods and erosion have created significant production losses for aquatic farmers in Cambodia, Laos, Myanmar, Thailand, Viet Nam and Ghana medium 1 train +4423 AR6_WGII 786 2 Common vulnerabilities to inland and marine aquaculture include increasing incidence and toxicity of HABs related to warming waters, causing fish kills and product consumption risks, negatively impacting the productivity and stability of production sectors and reliant communities high 2 train +4424 AR6_WGII 786 8 Climate uncertainty and data limitations hinder vulnerability assess- ments high 2 train +4425 AR6_WGII 786 9 Filling data gaps with mon- itoring high 2 train +4426 AR6_WGII 786 11 In the majority of lower-middle-income countries, seaweed culture is dominated by women in family-owned businesses as in Zanzibar and the Philippines (Brugere et al., 2020; Ramirez et al., 2020), where women are not always paid directly but contribute to family incomes high 2 train +4427 AR6_WGII 786 13 Women employed in aquaculture cooperatives gained adaptive capacity, which reduced gender inequities medium 1 train +4428 AR6_WGII 786 14 Women in aquaculture experience competing roles between employment, childcare and home duties (high confidence) (Morgan et al., 2015; Lauria et al., 2018; Chávez et al., 2019; see Cross- Chapter Box GENDER in Chapter 18) and differ from men in terms of perceptions of environmental risk, climate change and adaptation behaviour, with limited contributions to decision making medium 1 train +4429 AR6_WGII 787 5 Sensitivities for marine finfish may be high even under +1.5–2.0°C medium 1 train +4430 AR6_WGII 787 7 Marine heatwaves are predicted to increase in occurrence, intensity and persistence under RCP4.5 or RCP8.5 by 2100 (Oliver et al., 2019; Bricknell et al., 2021), with risk partly mitigated by husbandry medium 1 train +4431 AR6_WGII 788 1 Climate change is predicted to affect the incidence, magnitude and virulence of finfish disease such as Vibriosis (Barber et al., 2016; Mohamad et al., 2019a; Mohamad et al., 2019b), but specific host–pathogen–climate relationships are not yet established high 2 train +4432 AR6_WGII 790 2 Shellfish growth will increase with warming waters until tolerances are reached, such as through extreme El Niño events high 2 train +4433 AR6_WGII 790 3 Rising temperatures and ocean acidification will result in losses of primary productivity and farmed species from tropical and subtropical regions, and gains in higher latitudes high 2 train +4434 AR6_WGII 790 4 Shellfish Vibrio infections will increase with warming waters and extreme events, increasing shellfish mortalities medium 1 train +4435 AR6_WGII 790 5 Bivalve larvae are known to be highly vulnerable to ocean acidification (high confidence) (see Section 3.3, Bindoff et al., 2019), with projected regional and species-specific levels of impact high 2 train +4436 AR6_WGII 790 6 Ocean acidification is also projected to weaken shells, affecting productivity and processing high 2 train +4437 AR6_WGII 790 9 Projected thermal increases of 1.5°C will reduce ecosystem services, further reduced under 2°C warming, with associated increases in acidification, hypoxia, dead zones, flooding and water restrictions medium 1 train +4438 AR6_WGII 790 10 Sudden production losses from extreme climate events can exacerbate food security challenges across production sectors, including aquaculture, increasing global hunger high 2 train +4439 AR6_WGII 790 13 Mangrove reforestation efforts in Asia may have some effectiveness in re-creating important nursery grounds for aquatic species low 0 test +4440 AR6_WGII 790 14 Families are highly vulnerable to climate change where nutritional needs are being met by self-production, such as in Mozambique, Namibia (Villasante et al., 2015), Zambia (Kaminski et al., 2018) and Bangladesh high 2 train +4441 AR6_WGII 790 15 Climate change will therefore affect multiple ecosystem services where ultimately decisions on balance or trade-offs will vary with regional perceptions of service value high 2 train +4442 AR6_WGII 791 6 Early participation of stakeholders in adaptive planning has promoted action and ownership of results (high confidence), such as in India and the USA (Link et al., 2015; FAO, 2018c; Soto et al., 2018) Early outreach, education and knowledge gap assessments raise awareness, where utilisation of local knowledge and Indigenous knowledge and scientific involvement support informed adaptive planning and uptake for all stakeholders high 2 train +4443 AR6_WGII 791 7 Supporting the active involvement of women helps address gender inequity and perceived risk, particularly for smallholder farmers high 2 train +4444 AR6_WGII 791 8 However, regional and national political influences, financial and technical capacity, governance planning and policy development will ultimately support or hinder adaptation for aquaculture high 2 train +4445 AR6_WGII 791 10 Species diversification through co-culture, integrated aquaculture– agriculture (e.g., rice–fish) or integrated multi-trophic culture (e.g., shrimp–tilapia–seaweed or finfish–bivalve–seaweed) may maintain farm long-term performance and viability by: creating new aquaculture opportunities; promoting societal and environmental stability; reducing GHG emissions through reduced feed usage and waste; and carbon sequestration medium 1 train +4446 AR6_WGII 791 12 Selective breeding of species in aquaculture systems can promote climate resilience medium 1 test +4447 AR6_WGII 791 18 Projections of climate on aquaculture production traits are not well understood (Lhorente et al., 2019); therefore, genetic diversity needs to be maintained to ensure population fitness high 2 train +4448 AR6_WGII 791 20 However, land-based aquaculture requires large capital and operational costs and use of land, increasing conflicts between land and water use, have increased energy demands (increasing GHG if fossil fuels are the primary energy source), require necessary expertise and will not reduce outgrowing exposures high 2 train +4449 AR6_WGII 791 21 Geographical selection of marine farm sites may prevent climate productivity declines medium 1 train +4450 AR6_WGII 792 2 Building coastal protection, stronger cages and mooring systems, and deeper ponds and using sheltered bays can reduce escapees and mortalities related to flooding, increased storms and extreme events medium 1 train +4451 AR6_WGII 793 1 Other adaptation options for limited water supply are government equitable water allocations and water storage high 2 train +4452 AR6_WGII 793 2 Feed formulations and improved feed conversion can reduce climate- associated stress for freshwater species, significantly reducing waste and increase sustainability medium 1 train +4453 AR6_WGII 793 8 Improved farm management is a key opportunity high 2 train +4454 AR6_WGII 793 10 Specialised industry portals (Pacific shellfish) and government-established monitoring programmes (Chilean salmon) and other observational networks (e.g., Global Ocean Acidification Observing Network (GOA-ON)) can provide real-time monitoring and early-warning event alerts and facilitate aquaculture decision making medium 1 train +4455 AR6_WGII 793 11 Seasonal forecasting, downscaled models and early-warning systems provide valuable regional or farm site risk information (Hobday et al., 2018; Galappaththi et al., 2020b; Whitney et al., 2020), but monitoring will need to be useful for farmers, involve farmers, and be accurate, timely, cost-effective, reviewed and maintained in order to ensure uptake high 2 train +4456 AR6_WGII 793 12 Early-warning systems for HABs enable rapid decision making and risk mitigation medium 1 test +4457 AR6_WGII 793 13 New tools, strategies and observations are needed to predict HAB occurrences and range shifts with changing climate high 2 train +4458 AR6_WGII 793 15 Indigenous groups differ in opinions on aquaculture acceptability, implications for coastal management and territorial rights high 2 train +4459 AR6_WGII 793 18 Therefore policy, economic, knowledge and other support must ensure representation with traditional and other stakeholder ecological knowledge at national, regional and local levels to facilitate climate change adaptation and safeguard human rights for poor and vulnerable groups high 2 train +4460 AR6_WGII 794 7 Research indicates that mixed crop–livestock systems are often more resilient to climate change medium 1 train +4461 AR6_WGII 797 11 The adoption and maintenance of agroforestry practices require appropriate incentives or the removal of barriers high 2 train +4462 AR6_WGII 798 2 The relative environmental impact of hydroponic systems is lower com- pared with conventional systems owing to the significant reductions in land use and fertilizer usage high 2 train +4463 AR6_WGII 798 3 While studies indicate that aquaponics and hydroponics have higher yields and a lower environmental footprint than conventional agriculture medium 1 train +4464 AR6_WGII 798 20 The vulnerability of many crop–livestock keepers to climate change is particularly affected by property and grazing rights high 2 train +4465 AR6_WGII 799 11 Shade-grown cocoa and coffee agroforestry systems provide an array of ecosystem services, including regulating pests and diseases, maintaining soil fertility, maintaining biodiversity and carbon sequestration high 2 train +4466 AR6_WGII 799 32 Adaptation measures in such systems need to consider co-benefits and negative trade-offs, especially in vulnerable communities, to avoid widening further the inequities, rural livelihood loss, migration and marginalisation, and ensure progress towards the SDGs high 2 train +4467 AR6_WGII 800 4 Climate change can affect the growth and geographical expansion of these fungi high 2 train +4468 AR6_WGII 800 10 The occurrence of toxin-producing fungi will increase and expand from tropical and subtropical areas into new regions and where appropriate capacity for surveillance and risk management is lacking medium 1 train +4469 AR6_WGII 800 11 The increase in toxigenic fungi in crops, and consequent contamination of staple foods with mycotoxins, will increase the risks of human and animal exposure high 2 train +4470 AR6_WGII 800 12 In aquatic systems, toxins produced during HABs also cause food safety problems high 2 train +4471 AR6_WGII 800 14 Vibrio-related mortalities from finfish consumption are expected to rise with climate change (water temperature, salinity, oxygen and pH) medium 1 train +4472 AR6_WGII 800 15 For shellfish species, oxygen deficits (Mohamad et al., 2019b), sea level rise (Deeb et al., 2018) and temperature (Green et al., 2019) will be most important for food safety.Food safety is also anticipated to worsen from increased contaminant bioaccumulation under climate-induced warming high 2 train +4473 AR6_WGII 800 17 Occurrence of bacterial pathogens such as Salmonella and Campylobacter will increase with rising temperatures high 2 train +4474 AR6_WGII 800 21 Increases in rainfall intensity will have some effect on the transport of heavy metals by enhancing runoff from soil and increasing the leaching of heavy metals into water systems, with magnitudes dependent on local conditions high 2 train +4475 AR6_WGII 801 15 Climate change can alter insect damage in at least two ways: increases in reproductive rate from temperature increases and changes in pheromone effectiveness high 2 train +4476 AR6_WGII 801 40 Climate change impacts will increase most global prices relative to early 21st century levels, with varying effects on the cost of food imports high 2 train +4477 AR6_WGII 803 16 Climate change is already contributing to reduced food security and nutrition and will continue to do so high 2 train +4478 AR6_WGII 805 1 Climate change will make efforts to reduce this threat more difficult to achieve medium 1 train +4479 AR6_WGII 805 6 Several CIDs increase the number of people experiencing food insecurity high 2 train +4480 AR6_WGII 806 2 Labour capacity, supply and productivity loss in moderate outdoor work due to heat stress is estimated between 2% and 14%, depending on the location and indicator (Ioannou et al., 2017; Kjellstrom et al., 2018), with an overall estimate of 5.3% loss in productivity for outdoor work between 2000 and 2015 medium 1 train +4481 AR6_WGII 806 3 Highly vulnerable occupation groups affected by heat stress include farmers, farmworkers and livestock keepers working outdoors in low-income tropical countries high 2 train +4482 AR6_WGII 807 1 Low-income urban and rural households who are net food buyers are particularly affected by food price increases, with reduction in consumption of diverse food groups high 2 train +4483 AR6_WGII 807 3 Indigenous Peoples are often more vulnerable to climate change, due to conditions of poverty, limited resources, discrimination and marginalisation high 2 train +4484 AR6_WGII 807 7 Climate change have increased food safety risks high 2 train +4485 AR6_WGII 807 9 Weather variability and extreme events (Seneviratne et al., 2021) have reduced availability and access to diverse foods to sell and to purchase in rural markets, thereby reducing access to affordable, diverse foods for both rural small-scale producers and net consumers, particularly for landlocked and low-income countries high 2 train +4486 AR6_WGII 807 24 Without adaptive measures, heat stress impacts on agricultural labour will increase with climate change high 2 train +4487 AR6_WGII 808 19 Adaptation options needed to protect agricultural worker productivity outdoors and reduce occupational heat illnesses and deaths include cooled working environments, improved surveillance systems and education on the need to monitor high 2 train +4488 AR6_WGII 809 3 While climate change impacts, including drought impacts on food security, are important risk factors for conflict, other key drivers are often more influential, including low socioeconomic development, limited state capacity, weak governance, intergroup inequities and recent histories of conflict medium 1 train +4489 AR6_WGII 809 16 Some underline their potential in building resilience to changing climatic conditions, in the form of enhanced drought/heat tolerance, pest/disease protection and/or reduced land usage, thus serving to bolster food security and nutrition (Sainger et al., 2015; Muzhinji and Box 5.10: Food Safety Interactions with Food Security and Malnutrition Climate change significantly increases the future food safety risks high 2 train +4490 AR6_WGII 809 21 Children in low-income countries will be at greater risk of undernutrition from these multiple climate change impacts, including lower food availability, quality and safety and increased risk of diarrheal disease high 2 train +4491 AR6_WGII 810 27 A more iterative and flexible adaptation approach beyond just genomic improvement to tackle the multiplicity of factors limiting smallholder production is anticipated to increase the likelihood that these promising technologies can enhance food security and nutrition medium 1 train +4492 AR6_WGII 810 29 To make breeding technologies scale-neutral, the policy structure needs to support and protect smallholders medium 1 train +4493 AR6_WGII 811 10 UPA cannot fully feed urban dwellers within its boundaries but can make an important contribution to local food security and nutrition medium 1 train +4494 AR6_WGII 813 16 Land deals raise important social justice questions (Franco et al., 2017; Hunsberger et al., 2017; Borras and Franco, 2018b; Borras et al., 2020; Sekine, 2021) high 2 train +4495 AR6_WGII 814 6 Land use dimensions Impacts and implications References (2014 to present) ForestryDirect and indirect land use change provoked by LSLAs accelerates deforestation of tropical forests globally medium 1 train +4496 AR6_WGII 814 10 LSLAs may adversely affect local populations’ access to energy resources medium 1 train +4497 AR6_WGII 815 3 Agricultural intensification could meet short-term food security and livelihood goals, but reduces biological and landscape diversity, and ecosystem services high 2 train +4498 AR6_WGII 815 10 Land deals frequently target common land and may increase the vulnerability of customary, traditional, and Indigenous systems common property, while reducing their adaptive capacity high 2 train +4499 AR6_WGII 816 9 An increasing demand for aquaculture products intensifies competition for feed supplies medium 1 train +4500 AR6_WGII 816 10 Increases in demands for animal protein and shifts to pescatarian diets will increase the existing competition for land resources, particularly in low- and medium-income countries, with negative impacts on food security (Makkar, 2018), but may be mitigated by dietary changes, novel feeds and food waste usage for aquatic systems (Berners-Lee et al., 2018; Hua et al., 2019; Cottrell et al., 2020).Competition over use of major aquaculture feed crops (Fry et al., 2016) with terrestrial livestock (Troell et al., 2014), and fish use by terrestrial livestock, will also place pressure on fish and crop resources medium 1 train +4501 AR6_WGII 816 13 Waste fish products can supplement fish meal and oil to reduce competition for feed, as well as reducing use of fish that could go to human consumption medium 1 train +4502 AR6_WGII 817 2 Shallow and microtidal estuaries will be more vulnerable to changing river runoffs and saltwater intrusions, eutrophication and hypoxia high 2 train +4503 AR6_WGII 817 16 Adaptation options that consider adverse effects for different groups reduce the risk increasing vulnerability, negatively affecting socioeconomic factors to deal with climate impacts, or impeding efforts to implement SDGs high 2 train +4504 AR6_WGII 817 17 Adaptation methods considering historical roots of current vulnerabilities can identify viable solutions, which are difficult to undertake because of path dependencies high 2 train +4505 AR6_WGII 817 19 Inclusive planning initiatives such as community-based anticipatory adaptation combined with ‘two-way learning’ that considers future scenarios and different adaptation pathways can prevent maladaptation high 2 train +4506 AR6_WGII 820 4 A/R programmes can negatively affect a range of substantial and procedural Indigenous Peoples’ rights entrenched in international human rights law (Table 5.22) and their potential for climate change adaptation high 2 train +4507 AR6_WGII 820 14 Policies and safeguards attached to specific A/R initiatives determine their impact high 2 train +4508 AR6_WGII 821 2 These trends are connected to climate change as financial investments are influenced by the likelihood that climate change will increase commodity and farmland price variability medium 1 train +4509 AR6_WGII 822 5 Climate change will affect the food–energy–water (FEW) nexus, commonly in the form of risk multiplier high 2 train +4510 AR6_WGII 822 10 Increasing demands for food, energy and water can lead to domestic and international conflict, including political instability and migration, often in the context of drought high 2 train +4511 AR6_WGII 822 32 Several adaptation options have high to medium feasibility, with robust evidence, high agreement about the adaptive capacity resilience building potential of options in relation to climate change impact drivers high 2 train +4512 AR6_WGII 823 1 Most adaptation options have medium to high microeconomic feasibility high 2 train +4513 AR6_WGII 823 3 Higher-scored options to reduce risk included increasing biodiversity (at landscape and field level), community seed banks, conventional breeding (plant and animals), mixed systems and agroecological approaches medium 1 train +4514 AR6_WGII 823 4 Most options have high scores for enhancing social well-being and economic and environmental benefits medium 1 train +4515 AR6_WGII 823 5 There were low scores for potential maladaptation medium 1 train +4516 AR6_WGII 824 8 Improved design and delivery of climate services can enhance effectiveness medium 1 train +4517 AR6_WGII 825 2 Bundling additional services such as market information with climate information may be effective at plugging information gaps (low confidence) (Chatuphale and Armstrong, 2018; Tsan et al., 2019; Tesfaye et al., 2019) There may be inequality in access to climate services; their use may tend to benefit large-scale operations and disadvantage small- and medium- scale farmers and others who face issues of access due to social and economic inequity; also some groups such as pastoralists have not yet benefitted from climate services high 2 train +4518 AR6_WGII 826 3 To address smallholder vulnerability to climate change impacts, however, additional policy support beyond agroecology will be needed that is context specific; for example, addressing farmer capacity, limited political power to access land, water, seeds and other key natural resources, structural gender inequities, policy and market disincentives that support large-scale monocultures high 2 train +4519 AR6_WGII 827 10 Food security and nutrition: Agroecological practices can increase household food security and nutrition for producer households, with more evidence in low- and medium-income countries high 2 train +4520 AR6_WGII 828 12 Community-based participatory scenario planning can help identify multiple climate stressors and vulnerabilities to develop effective adaptation plans (Fernández- Giménez et al., 2015; Bennett et al., 2016; Cross-Chapter Box MOVING PLATE this chapter).Different dimensions of agroecological transitions as a transformative climate change adaptation strategyLinks to climate change impacts, benefits, trade-offs and constraints to imple- mentation with examples Policy tools: Investment in agroecological approaches that are designed for socio-ecological context, farmer-led schools, co-learning platforms, and networks of farmers, scientists, private sector and civil society can support agroecological transitions at a regional scale high 2 train +4521 AR6_WGII 828 15 Small to mid-sized farms can more effectively integrate agroecological methods such as increasing landscape diversity, on-farm diversity and intercrops medium 1 train +4522 AR6_WGII 828 19 Other drivers of agroecological transitions can include crises (environmental, economic or social), social movements, changing socio-cultural values, addressing social inequities, and discourse (Pérez-Marin et al., 2017; Mier y Terán Giménez Cacho et al., 2018; Anderson et al., 2019a).Further research could provide context-specific information about economic and ecological benefits of some practices and combinations, with effective policies to support their implementation high 2 train +4523 AR6_WGII 829 6 Evidence on strengthening local and regional food systems with a food sovereignty approach, in terms of access to resources (land, seeds, water), shortened food chains and CbA strategies suggest that these strategies can positively contribute to climate change adaptation in many contexts medium 1 train +4524 AR6_WGII 829 21 Adaptation strategies can have negative impacts on marginalised social groups and worsen socioeconomic inequities unless explicit efforts are made to address unequal power dynamics and differences in access to resources in agricultural, fisheries, aquaculture, livestock and forestry systems high 2 train +4525 AR6_WGII 830 1 Where Indigenous Peoples have access to and control over their lands and natural resources, food systems can potentially be more sustainably managed and more resilient high 2 train +4526 AR6_WGII 830 5 Effective adaptation requires a more holistic approach that includes the recognition of Indigenous rights, governance systems and laws high 2 train +4527 AR6_WGII 830 9 Education utilising IK and LK can help prevent maladaptation options high 2 train +4528 AR6_WGII 830 26 In general CSA programmes have tended to overlook questions of inequity medium 1 train +4529 AR6_WGII 831 14 Policies around property and grazing rights are directly linked to small-scale food producer vulnerability, and land ownership changes will pose a key challenge as climate change Box 5.13: Supporting Youth Adaptation in Food Systems Young people are key agents in agrifood systems: both a vulnerable group, and one that can foster systemic change high 2 train +4530 AR6_WGII 831 17 Rural youth in these sectors are particularly vulnerable, often with less access to land, water, capital and other resources, shaped by family and social relations, and fewer opportunities high 2 train +4531 AR6_WGII 831 18 In these vulnerable regions, climate change compounds other drivers such as poverty to increase youth out-migration to urban areas or other regions medium 1 train +4532 AR6_WGII 831 19 Young low-income rural women may be particularly marginalised and vulnerable due to systemic gender inequities in access to land, credit, employment, institutions and other resources medium 1 train +4533 AR6_WGII 831 20 Youth play a critical role in all sectors of the food system (HLPE, 2021; Figure Box 5.13.1), and some are actively pursuing work and innovation in agrifood systems medium 1 train +4534 AR6_WGII 831 22 At the same time, due to heightened awareness about climate change, youth may be more willing to apply climate adaptation strategies medium 1 train +4535 AR6_WGII 831 27 Harnessing youth innovation and vision to address climate change alongside other SDGs such as gender inequity and rural poverty will be a crucial strategy to ensure resilient economies in food systems high 2 train +4536 AR6_WGII 835 9 Financial barriers limit implementation of adaptation options in agriculture, fisheries, aquaculture and forestry high 2 train +4537 AR6_WGII 835 14 Expanding access to financial services and pooling climate risks can enable and incentivise climate change adaptation medium 1 train +4538 AR6_WGII 836 11 Government agencies and multilateral institutions –Strengthen enabling environments for sustainable production and ecosystem protection (e.g., price transparency; information exchange; international coordination) –Support demonstration projects for sustainable land and resource management (e.g., grants) –Disaster risk reduction (e.g., national disaster funds; social protection programmes; contingent credit lines; sovereign/sub-sovereign insurance (Global Commission on Adaptation, 2019) –Increase resilience through early-warning systems, infrastructure, and capacity building (e.g., climate change adaptation funds) –Increase revenues for adaptation activities (e.g., income/luxury taxes) –Reduce production risks (e.g., agricultural subsidies) –Promote advanced technology implementation (e.g., tax incentives) –Coordinate and align donor funding with national priorities (e.g., multi-donor national climate change funds) –Incentivise and de-risk commercial investments (e.g., interest rate reduction programmes, structured financing, guarantee funds) (Woodard et al., 2019) Methods to strengthen adaptation finance include updating regulations and policies to support adaptation finance instruments (e.g., climate accounting standards), requiring climate-risk disclosure, improved information-sharing among public and private sector actors and devolving funding to local actors medium 1 train +4539 AR6_WGII 837 2 Production insuranceCompensation for specified losses related to production (e.g., insurance indexed to specific weather events) or supply chains (e.g., shipping insurance) Market and price insurance Compensation for specified market-related losses (e.g., price or currency fluctuation) Grants: Concessionary funding provided by public or philanthropic entities to support climate adaptation costs or outcomes (no expectation of repayment) Direct supportFunding for provision of goods (e.g., fertilizer, seeds, nursery stock) or services (e.g., technical assistance, product storage) to producers, local companies or intermediaries (e.g., for agronomic or business management expertise); can reduce credit risk when part of blended finance arrangements Performance-based grantsGrants or other concessionary funding contingent on achievement of defined adaptation outcomes (with possible third-party verification requirement); may support development and testing of new approaches (i.e., design funding; challenges/prizes) Governmental instruments Policy incentivesPublic policies designed to stimulate adaptation action among targeted groups (e.g., producers, consumers, agri-businesses, financiers) including direct or indirect subsidies (e.g., producer payments, tax breaks, health insurance), procurement policies (e.g., low carbon and sustainability criteria; nutrition-sensitive school feeding programmes) and other fiscal measures (e.g., infrastructure development; funding R&D in climate-resilient practices or technologies) (Shukla et al., 2019) Development aidInternational or domestic programmes that directly or indirectly fund adaptation actions including financial transfers (e.g., producer support or anti-poverty programmes) and subsidised credit medium 1 train +4540 AR6_WGII 839 1 Increased technology innovation, stakeholder integration and transparent governance structures and procedures at local to global scales are key to successful bioeconomy deployment maximising benefits and managing trade-offs high 2 train +4541 AR6_WGII 839 2 Limited global land and biomass resources accompanied by growing demands for food, feed, fibre and fuels, together with prospects for a paradigm shift towards phasing out fossil fuels, set the frame for potentially fierce competition for land and biomass to meet burgeoning demands even as climate change increasingly limits natural resource potentials high 2 test +4542 AR6_WGII 839 3 Sustainable agriculture and forestry, technology innovation in bio-based production within a circular economy and international cooperation and governance of global trade in products to reflect and disincentivise their environmental and social externalities can provide mitigation and adaptation via bioeconomy development that responds to the needs and perspectives of multiple stakeholders to achieve outcomes that maximise synergies while limiting trade-offs high 2 train +4543 AR6_WGII 842 18 In summary, there is significant scope for optimising use of land resources to produce more biomass while reducing adverse effects high 2 train +4544 AR6_WGII 842 19 Context-specific prioritisation, technology innovation in bio-based production, integrative policies, coordinated institutions and improved governance mechanisms to enhance synergies and minimise trade-offs can mitigate the pressure on managed as well as natural and semi-natural ecosystems medium 1 train +4545 AR6_WGII 842 20 Yet, energy conservation and efficiency measures, and deployment of technologies and systems that do not rely on carbon-based energy and materials, are essential for mitigating biomass demand growth as countries pursue ambitious climate goals high 2 train +4546 AR6_WGII 920 2 Urbanisation processes generate vulnerability and exposure which combine with climate change hazards to drive urban risk and impacts high 2 train +4547 AR6_WGII 920 3 Globally, the most rapid growth in urban vulnerability and exposure has been in cities and settlements where adaptive capacity is limited, especially in unplanned and informal settlements in low- and middle-income nations and in smaller and medium-sized urban centres high 2 train +4548 AR6_WGII 920 10 Losses become systemic when affecting entire systems and can even jump from one system to another (e.g., drought impacting on rural food production contributing to urban food insecurity) medium 1 train +4549 AR6_WGII 920 13 Vulnerabilities are shaped by drivers of inequality, including gender, class, race, ethnic origin, age, level of ability, sexuality and non-conforming gender orientation, framed by cultural norms, diverse values and practices high 2 train +4550 AR6_WGII 920 22 The COVID-19 pandemic is estimated to have pushed an additional 119 to 124 million people into poverty in 2020, with South Asia and Sub-Saharan Africa each contributing roughly two-fifths of this total medium 1 train +4551 AR6_WGII 920 29 Sea level increase and increases in tropical cyclone storm surge and rainfall intensity will increase the probability of coastal city flooding, with more than a billion people located in low-lying cities and settlements expected to be at risk from coastal-specific climate hazards by 2050 high 2 train +4552 AR6_WGII 920 30 Sea level rise, increases in tropical cyclone storm surges and more frequent and intense extreme precipitation will increase the number of people, area of urban land, and damages from flood hazard high 2 train +4553 AR6_WGII 920 34 Many of these plans focus narrowly on climate risk reduction, missing opportunities to advance co-benefits of climate mitigation and sustainable development, compounding inequality and reducing well-being medium 1 train +4554 AR6_WGII 921 2 The greatest gaps between policy and action are in failures to manage adaptation of social infrastructure (community facilities, services and networks) and failure to address complex interconnected risks for example in the food–energy–water– health nexus or the inter-relationships of air quality and climate risk medium 1 train +4555 AR6_WGII 921 9 Governance capacity, financial support and the legacy of past urban infrastructure investment constrain how all cities and settlements are able to adapt high 2 train +4556 AR6_WGII 921 11 These include the limited ability to identify social vulnerability and community strengths; the absence of integrated planning to protect communities; and the lack of access to innovative funding arrangements and limited capability to manage finance and commercial insurance medium 1 train +4557 AR6_WGII 921 20 Private and business investment in key infrastructure, housing construction and through insurance requirements can also drive widespread adaptive action, though at times excluding the priorities of the poor medium 1 train +4558 AR6_WGII 921 23 Transnational networks of local government can also enhance city level capacity, share lessons and advocacy medium 1 train +4559 AR6_WGII 921 34 Finance de- ployed at the interface of multiple, integrated adaptation measures can support climate resilient development high 2 train +4560 AR6_WGII 922 9 Multi-level leadership and institutional capacity, together with financial resources (including climate finance) to support inclusive and sustainable adaptation in the context of multiple pressures and interconnected risks, can help to ensure that global urbanisation of an additional 2.5 billion people by 2050 reduces rather than generates climate risk medium 1 train +4561 AR6_WGII 922 11 The great- est gains in well-being in urban areas can be achieved by prioritising investment to reduce climate risk for low-income and marginalised residents and targeting informal settlements high 2 train +4562 AR6_WGII 922 14 Providing opportunities for marginalised people, including women, to take on leadership and par - ticipation in local projects can enhance climate governance and its outcomes high 2 train +4563 AR6_WGII 922 21 More comprehensive and clearly articulated global ambitions for city and community adaptation will contribute to inclusive urbanisation, by addressing the root causes of social and economic inequalities that drive social exclusion and marginalisation, so that adaptation can directly support the 2030 Sustainable Development Agenda high 2 train +4564 AR6_WGII 923 2 By 2050, over two thirds of the world’s population is expected to be urban, many living in unplanned and informal settlements and in smaller urban centres in Africa and Asia high 2 train +4565 AR6_WGII 926 12 The IPCC 1.5°C Special Report commented that ‘The extent of risk depends on human vulnerability and the effectiveness of adaptation for regions (coastal and non-coastal), informal settlements, and infrastructure sectors (energy, water, and transport) high 2 train +4566 AR6_WGII 928 15 Evidence since AR5 confirms that occupants of informal settlements are particularly exposed to climate events given low-quality housing, limited capacity to adapt, and limited or no risk-reducing infrastructure high 2 train +4567 AR6_WGII 928 19 Box 6.1 | Planetary Urbanisation and Climate Risk The scale, reach and complexity of contemporary urbanisation compounds climate risks and conditions adaptation high 2 train +4568 AR6_WGII 929 32 The fragmentation of the hinterland for extractivist purposes depletes ecosystem services and further exacerbates cascading risks high 2 train +4569 AR6_WGII 930 5 Cities can be effective change agents when supported by networked local and national institutions, including professional bodies high 2 train +4570 AR6_WGII 932 11 The conclusions of the IPCC Special Report on Global Warming of 1.5°C noted that ‘Global warming of 2°C is expected to pose greater risks to urban areas than global warming of 1.5°C medium 1 train +4571 AR6_WGII 933 4 Additionally, carbon-intensive economic growth, increasing inequalities, global pandemics, and uncontrolled or unmanaged urbanisation will exacerbate the exposure and vulnerability of urban systems modelled in existing climate scenarios and pathways high 2 train +4572 AR6_WGII 933 10 Climate change can have direct impacts on the functioning of urban systems, while the nature of those systems plays a substantial role in modifying the effects of climate change high 2 train +4573 AR6_WGII 933 28 Future urbanisation will amplify projected local air temperature increase, particularly by strong influence on minimum temperatures, which is approximately comparable in magnitude to global warming high 2 train +4574 AR6_WGII 933 31 The risks to cities, settlements and infrastructure from heatwaves will worsen high 2 train +4575 AR6_WGII 935 3 Heat risk is associated with a range of health issues for urban residents, with the consequences of higher urban temperatures being unevenly distributed across urban populations high 2 train +4576 AR6_WGII 935 11 There is an emerging risk of diminished indoor thermal comfort due to climate change, evidenced by research into negatively affected thermal comfort indices and/or increased number of overheating hours under future emissions scenarios medium 1 train +4577 AR6_WGII 935 18 Higher urban temperatures result in lower labour productivity levels and economic outputs medium 1 train +4578 AR6_WGII 935 24 Higher urban temperatures place unequal economic stresses on residents and households through higher utilities demand during warm periods, for example, electricity in regions where air conditioning is predicted to become more prevalent, and due to medical costs associated with care for heat illnesses and related health effects, missed work and other related impacts medium 1 train +4579 AR6_WGII 936 5 Social surveys from temperate and tropical cities highlight the risk of reduced quality of life during heat events, including increased incidence of personal discomfort in indoor and outdoor settings, elevated anxiety, depression and other indicators of adverse psychological health, and reductions in physical activity, social interactions, work attendance, tourism and recreation high 2 train +4580 AR6_WGII 936 8 Sea level increase and increases in tropical cyclone storm surge and rainfall intensity will increase the probability of coastal city flooding high 2 test +4581 AR6_WGII 936 9 Globally, the increase in frequencies and intensities of extreme precipitation from global warming will likely4 expand the global land area affected by flood hazards medium 1 train +4582 AR6_WGII 936 20 Urban flooding risks are also increased by urban expansion and land use and land cover change which enlarges impermeable surface areas through soil sealing, impacting drainage of floodwaters with consequent sewer overflows high 2 train +4583 AR6_WGII 936 25 Future risks of urban flooding is increasing in conjunction with continued increases in global surface temperature high 2 train +4584 AR6_WGII 937 7 Risks arising from urban water scarcity worldwide are very likely increasing due to climate drivers (e.g., warmer temperatures and droughts) and urbanisation processes (e.g., land use changes, migration to cities and changing patterns of water use including over extraction of surface and groundwater resources) affecting supply and demand high 2 train +4585 AR6_WGII 937 11 Decreased regional precipitation and associated changes in runoff and storage from droughts is exacerbating urban scarcity by impairing the quality of water available for its resource management in cities high 2 train +4586 AR6_WGII 937 15 Risks of urban water scarcity and security are compounded by vulnerabilities such as service availability and quality of infrastructure to supply water for increased urban demand from in-migration to cities medium 1 train +4587 AR6_WGII 937 16 Risks to local water security in cities are also exacerbated by drivers such as dependence on imported water resources from distant locales that may be exposed to additional drought risks high 2 train +4588 AR6_WGII 938 40 In particular, there is evidence from North American cities that tornado damage are likely fundamentally driven by growing built-environment exposure medium 1 train +4589 AR6_WGII 939 5 Future climate risk of fires at the WUI are likely medium 1 train +4590 AR6_WGII 939 9 The mortality risk for several pollutants, for example PM 2.5, is considerable high 2 train +4591 AR6_WGII 939 20 The impacts of future climate change on air quality and consequent risks on human health have been studied at urban (Knowlton et al., 2004; Physick, Cope and Lee, 2014) and national scales (Fann et al., 2015; Orru et al., 2013; Doherty, Heal and O’Connor, 2017); globally, these studies have found a likely net increased risk of climate change on air pollution-related health low 0 train +4592 AR6_WGII 940 7 While some individuals, including children, may be able to exercise agency to reduce their risk (Treichel, 2020), and some indicators are culturally specific, overall, poor, marginalised, socially isolated and informal urban households are particularly at risk high 2 train +4593 AR6_WGII 941 2 Climate-induced migration is not necessarily higher among poorer households whose mobility is more likely to be limited due to the poverty trap (i.e., lack of financial resources) high 2 train +4594 AR6_WGII 941 13 Migration can also be maladaptive for the receiving contexts, whether due to the pressure on and/or conflict over land and/or the urban resources high 2 train +4595 AR6_WGII 941 16 Regardless of the reasons and the initiators for migration, community control over resettlement both at the origin and destination leads to more positive outcomes for both the communities being resettled and the receiving communities high 2 train +4596 AR6_WGII 941 17 The protection of livelihoods contributes to ensuring the well-being (physical and mental) and the protection of the rights of communities high 2 train +4597 AR6_WGII 942 3 Current climate variability is already causing impacts on infrastructure systems around the world high 2 train +4598 AR6_WGII 947 19 Projected global compound risks will increase in the future, with significant risks across energy, food and water sectors that likely overlap spatially and temporally while affecting increasing numbers of people and regions particularly in Africa and Asia high 2 train +4599 AR6_WGII 948 23 Extending into urban areas within stable states, alienation and loss of trust between local populations and the state can be exacerbated by top-down adaptation planning and delivery; socially and spatially uneven adaptation investment; and in the economic and administrative limits of government that can lead to some places being excluded from formal planned investment high 2 train +4600 AR6_WGII 953 27 Evidence from both richer countries and the Global South reveals that conventional zoning is more effective when governance systems facilitate the implementation of land use policies for climate adaptation that preclude negative human-nature interactions and that curb spatial inequity, both of which can trigger climate gentrification and increase the vulnerability of economically disadvantaged groups to climate-related risk high 2 train +4601 AR6_WGII 953 31 Adaptation actions through zoning and land use are more effective when combined with other planning measures high 2 train +4602 AR6_WGII 975 13 Local government reform at different levels can improve local adaptation, whether this is by strengthening specific teams or building cross-departmental linkages high 2 train +4603 AR6_WGII 975 21 Multi-level governance measures that support local governments can foster robust adaptation approaches and address risks and vulnerabilities across scales high 2 train +4604 AR6_WGII 975 22 Effective action by local government requires national government’s support medium 1 train +4605 AR6_WGII 975 27 National urban adaptation directives can influence municipal governments’ action and planning, but evidence suggests that national policy alone is not sufficient to deliver action on the ground without understanding local conditions high 2 train +4606 AR6_WGII 975 31 Adaptation actions, even where financed effectively, do not always deliver positive outcomes high 2 train +4607 AR6_WGII 976 8 Individual coping strategies are generally ineffective in reducing extreme risks and they rarely address the underlying structural causes of vulnerability high 2 train +4608 AR6_WGII 976 10 However, individual coping strategies can provide foundations for the implementation of collaborative action in communities, building on people’s experiences, in ways which may have a longer-term, durable impact on developing resilience high 2 train +4609 AR6_WGII 976 25 Corporate, private sector interventions in urban risk reduction more broadly remain limited, with a mix of public and private responsibility for planning, implementing and maintaining adaptations in the built environment, and yet limited engagement of private sector actors in providing healthcare measures for heat prevention medium 1 train +4610 AR6_WGII 977 32 Local authorities are an important enabling actor that can guide the private sector and communities to take responsibility for creating policy and regulatory environments that encourage private sector participation aligned with the SDGs’ equity and ecological sustainability principles high 2 train +4611 AR6_WGII 982 17 The integration of climate adaptation in local policies in cities and settlements has often been seen as maintaining business-as-usual and not always aligned with transformative efforts to address structural drivers of vulnerability high 2 test +4612 AR6_WGII 989 17 Access to private finance can support infrastructure development through private provisioning, public–private partnerships (PPP) and public debt arrangements high 2 train +4613 AR6_WGII 999 15 Urbanisation: A Megatrend Driving Global Climate Risk and Potential for Low-Carbon and Resilient Futures Severe weather events, exacerbated by anthropogenic emissions, are already having devastating impacts on people who live in urban areas, and on the infrastructure that supports these communities and those of many other distant places high 2 train +4614 AR6_WGII 1000 1 Urban settlements are drivers of climate change, generating about 70% of global CO 2-eq emissions high 2 train +4615 AR6_WGII 1000 6 As urbanisation unfolds, its legacy continues to be the locking in of emissions and vulnerabilities high 2 train +4616 AR6_WGII 1001 9 Enabling Action Innovative governance and finance solutions are required to manage complex and interconnected risks across essential key infrastructures, networks and services and meet basic human needs in urban areas medium 1 train +4617 AR6_WGII 1001 14 Increasing investment at pace will put pressure on governance capability and transparency and accountability of decision making medium 1 train +4618 AR6_WGII 1055 1 Climate hazards are a growing driver of involuntary migration and displacement (high confidence) and are a contributing factor to violent conflict high 2 train +4619 AR6_WGII 1055 2 These impacts are often inter-connected, are unevenly distributed across and within societies, and will continue to be experienced inequitably due to differences in exposure and vulnerability very high 3 train +4620 AR6_WGII 1055 3 Cascading and compounding risks affecting health due to extreme weather events have been observed in all inhabited regions, and risks are expected to increase with further warming very high 3 train +4621 AR6_WGII 1055 4 Since AR5, new evidence and awareness of current impacts and projected risks of climate change on health, well-being, migration and conflict have emerged, including greater evidence of the detrimental impacts of climate change on mental health very high 3 train +4622 AR6_WGII 1055 5 New international agreements were reached on climate change (Paris Agreement), disaster risk reduction (DRR) (Sendai Agreement), sustainable development (the Sustainable Development Goals (SDGs)), urbanisation (The New Urban Agenda), migration (Global Compact for Safe, Orderly and Regular Migration) and refugees (Global Compact on Refugees) that, if achieved, would reduce the impacts of climate change on health, well-being, migration and conflict very high 3 train +4623 AR6_WGII 1055 6 However, the challenges with implementing these agreements are highlighted by the coronavirus disease 2019 (COVID-19) pandemic, which exposed systemic weaknesses at community, national and international levels in the ability of societies to anticipate and respond to global risks high 2 train +4624 AR6_WGII 1055 7 Incremental changes in policies and strategies have proven insufficient to reduce climate-related risks to health, well-being, migration and conflict, highlighting the value of more integrated approaches and frameworks for solutions across systems and sectors that are embodied in these new international agreements high 2 train +4625 AR6_WGII 1055 8 With proactive, timely and effective adaptation, many risks for human health and well-being could be reduced and some potentially avoided very high 3 train +4626 AR6_WGII 1055 9 A significant adaptation gap exists for human health and well-being and for responses to disaster risks very high 3 train +4627 AR6_WGII 1055 11 National planning on health and climate change is advancing, but the comprehensiveness of strategies and plans need to be strengthened, and implementing action on key health and climate change priorities remains challenging high 2 train +4628 AR6_WGII 1055 14 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.adaptation is only 0.5% of dispersed multi-lateral climate finance projects high 2 train +4629 AR6_WGII 1055 15 This level of investment is insufficient to protect population health and health systems from most climate- sensitive health risks very high 3 train +4630 AR6_WGII 1055 16 Climate resilient development has a strong potential to generate substantial co-benefits for health and well-being and to reduce risks of involuntary displacement and conflict very high 3 train +4631 AR6_WGII 1055 17 Sustainable and climate resilient development that decreases exposure, vulnerability and societal inequity and that increases timely and effective adaptation and mitigation more broadly, has the potential to reduce but not necessarily eliminate climate change impacts on health, well-being, involuntary migration and conflict high 2 train +4632 AR6_WGII 1055 18 This development includes greenhouse gas (GHG) emission reductions through clean energy and transport; climate-resilient urban planning; sustainable food systems that lead to healthier diets; universal access to healthcare and social protection systems; wide-scale, proactive adaptive capacity building for climate change; and achievement of the SDGs very high 3 train +4633 AR6_WGII 1055 20 The net global financial gains from these co-benefits to health and well- being, including avoided hospitalisations, morbidity and premature deaths, exceed the financial costs of mitigation high 2 train +4634 AR6_WGII 1055 21 As an example of co-benefits, the financial value of health benefits from improved air quality alone is projected to be greater than the costs of meeting the goals of the Paris Agreement high 2 train +4635 AR6_WGII 1055 22 All pathways to climate resilient development, including those for the health and healthcare systems, involve balancing complex synergies and trade-offs between development pathways and the options that underpin climate mitigation and adaptation pathways very high 3 train +4636 AR6_WGII 1055 23 Key transformations are needed to facilitate climate resilient development pathways (CRDPs) for health, well- being, migration and conflict avoidance high 2 test +4637 AR6_WGII 1055 24 The transformational changes will be more effective if they are responsive to regional, local and Indigenous knowledge and consider the many dimensions of vulnerability, including those that are gender- and age-specific high 2 train +4638 AR6_WGII 1055 25 A key pathway towards climate resilience in the health sector is universal access to primary healthcare, including mental healthcare high 2 train +4639 AR6_WGII 1055 26 Investments in other sectors and systems that improve upon the social determinants of health have the potential to reduce vulnerability to climate-related health risks high 2 train +4640 AR6_WGII 1055 27 Links between climate risks, adaptation, migration and labour markets highlight the value of providing better mobility options as part of transformative change medium 1 train +4641 AR6_WGII 1056 1 Observed Impacts Climate hazards are increasingly contributing to a growing number of adverse health outcomes (including communicable and non-communicable diseases (NCDs)) in multiple geographical areas very high 3 train +4642 AR6_WGII 1056 2 The net impacts are largely negative at all scales (very high confidence), and there are very few examples of beneficial outcomes from climate change at any scale high 2 train +4643 AR6_WGII 1056 3 While malaria incidence has declined globally due to non- climatic socioeconomic factors and health system responses, a shift to higher altitudes has been observed as the climate warms very high 3 train +4644 AR6_WGII 1056 4 Climate variability and change (including temperature, relative humidity and rainfall) and population mobility are significantly and positively associated with observed increases in dengue globally; chikungunya virus in Asia, Latin America, North America and Europe (high confidence); Lyme disease vector Ixodes scapularis in North America (high confidence); and Lyme disease and tick-borne encephalitis vector Ixodes ricinus in Europe medium 1 train +4645 AR6_WGII 1056 5 Higher temperatures (very high confidence), heavy rainfall events (high confidence) and flooding (medium confidence) are associated with an increase of diarrhoeal diseases in affected regions, including cholera (very high confidence), other gastrointestinal infections (high confidence) and food-borne diseases due to Salmonella and Campylobacter medium 1 train +4646 AR6_WGII 1056 6 Floods have led to increases in vector- and waterborne diseases and to disturbances of public health services high 2 train +4647 AR6_WGII 1056 7 Climate extremes increase the risks of several types of respiratory tract infections high 2 train +4648 AR6_WGII 1056 8 Climate-related extreme events such as wildfires, storms and floods are followed by increased rates of mental illness in exposed populations very high 3 train +4649 AR6_WGII 1056 9 Several chronic, non-communicable respiratory diseases are climate-sensitive based on their exposure pathways (e.g., heat, cold, dust, small particulates, ozone, fire smoke and allergens) high 2 train +4650 AR6_WGII 1056 10 Worldwide, rates of adverse health impacts associated with small particulate matter (PM) exposure have decreased steadily due to decreasing primary emissions (very high confidence), while rates of adverse health impacts from ozone air pollution exposure have increased very high 3 train +4651 AR6_WGII 1056 11 Exposure to wildland fires and associated smoke has increased in several regions very high 3 train +4652 AR6_WGII 1056 12 Spring pollen season start dates in northern mid-latitudes are occurring earlier due to climate change, increasing the risks of allergic respiratory diseases high 2 train +4653 AR6_WGII 1056 13 Heat is a growing health risk due to burgeoning urbanisation, an increase in high temperature extremes and demographic changes in countries with aging populations very high 3 train +4654 AR6_WGII 1056 14 Potential hours of work lost due to heat has increased significantly over the past two decades high 2 train +4655 AR6_WGII 1056 15 Some regions are already experiencing heat stress conditions at or approaching the upper limits of labour productivity high 2 train +4656 AR6_WGII 1056 16 A significant proportion of warm-season heat-related mortality in temperate regions is linked to observed anthropogenic climate change medium 1 train +4657 AR6_WGII 1056 17 For some heatwave events over the last two decades, associated health impacts can be at least partially attributed to observed climate change high 2 train +4658 AR6_WGII 1056 18 Extreme heat has negative impacts on mental health, well-being, life satisfaction, happiness, cognitive performance and aggression medium 1 train +4659 AR6_WGII 1056 19 Climate variability and change contribute to food insecurity, which can lead to malnutrition, including undernutrition, overweight and obesity, and to disease susceptibility in low- and middle-income countries high 2 train +4660 AR6_WGII 1056 20 Populations exposed to extreme weather and climate events may consume inadequate or insufficient food, leading to malnutrition and increasing the risk of disease high 2 train +4661 AR6_WGII 1056 21 Children and pregnant women experience disproportionately greater adverse nutrition and health impacts high 2 train +4662 AR6_WGII 1056 22 Climatic influences on nutrition are strongly mediated by socioeconomic factors very high 3 train +4663 AR6_WGII 1056 23 Extreme climate events act as both direct drivers (e.g., destruction of homes by tropical cyclones) and as indirect drivers (e.g., rural income losses during prolonged droughts) of involuntary migration and displacement very high 3 train +4664 AR6_WGII 1056 24 Most documented examples of climate-related displacement occur within national boundaries, with international movements occurring primarily within regions, particularly between countries with contiguous borders high 2 train +4665 AR6_WGII 1056 25 Global statistics collected since 2008 by the Internal Displacement Monitoring Centre (IDMC) show an annual average of over 20 million people internally displaced by weather-related extreme events, with storms and floods the most common drivers high 2 train +4666 AR6_WGII 1056 26 The largest absolute number of people displaced by extreme weather each year occurs in Asia (South, Southeast and East), followed by sub-Saharan Africa, but small island states in the Caribbean and South Pacific are disproportionately affected relative to their small population size high 2 train +4667 AR6_WGII 1056 27 Immobility in the context of climate risks can reflect vulnerability and lack of agency but can also be a deliberate choice of people to maintain livelihoods, economic considerations and social and cultural attachments to place high 2 train +4668 AR6_WGII 1056 28 Climate hazards have affected armed conflict within countries (medium confidence), but the influence of climate is small compared to socioeconomic, political and cultural factors high 2 train +4669 AR6_WGII 1056 29 Climate increases conflict risk by undermining food and water security, income and livelihoods in situations where there are large populations, weather-sensitive economic activities, weak institutions and high levels of poverty and inequality high 2 train +4670 AR6_WGII 1056 30 In urban areas, food and water insecurity and inequitable access to services has been associated with civil unrest where there are weak institutions medium 1 train +4671 AR6_WGII 1056 31 Climate hazards are associated with increased violence against women, girls and vulnerable groups, and the experience of armed conflict is gendered medium 1 train +4672 AR6_WGII 1056 32 Adaptation and mitigation projects implemented without consideration of local social dynamics have exacerbated non-violent conflict medium 1 train +4673 AR6_WGII 1057 1 An excess of 250,000 deaths yr–1 by 2050 attributable to climate change is projected due to heat, undernutrition, malaria and diarrhoeal disease, with more than half of this excess mortality projected for Africa (compared to a 1961–1991 baseline period for a mid-range emissions scenario) high 2 train +4674 AR6_WGII 1057 5 Climate change is projected to significantly increase population exposure to heatwaves (very high confidence) and heat-related morbidity and mortality high 2 train +4675 AR6_WGII 1057 7 Globally, the impact of projected climate change on temperature-related mortality is expected to be a net increase under RCP4.5 to RCP8.5, even with adaptation high 2 train +4676 AR6_WGII 1057 8 Heat related cardiovascular disease mortality is projected to increase by the end of this century high 2 train +4677 AR6_WGII 1057 9 Strong geographical differences in heat-related mortality are projected to emerge later this century, mainly driven by population growth and aging in regions with tropical and subtropical climates very high 3 train +4678 AR6_WGII 1057 10 The burdens of several climate-sensitive food-borne, waterborne, and vector-borne diseases (VBDs) are projected to increase under climate change, assuming no additional adaptation very high 3 train +4679 AR6_WGII 1057 11 The distribution and intensity of transmission of malaria is expected to decrease in some areas and increase in others, with increases projected mainly along the current edges of its geographic distribution in endemic areas of sub-Saharan Africa, Asia and South America high 2 train +4680 AR6_WGII 1057 12 Dengue risk will increase, with a larger spatio-temporal distribution in Asia, Europe and sub-Saharan Africa under RCP6.0 and RCP8.5, potentially putting another 2.25 billion people at risk high 2 train +4681 AR6_WGII 1057 13 Higher incidence rates are projected for Lyme disease in the Northern Hemisphere (high confidence) and for transmission of Schistosoma mansoni in eastern Africa high 2 train +4682 AR6_WGII 1057 14 Increasing atmospheric concentrations of carbon dioxide and climate change are projected to increase diet-related risk factors and related non-communicable diseasess globally and increase undernutrition, stunting and related childhood mortality particularly in Africa and Asia, with outcomes depending on the extent of mitigation and adaptation high 2 train +4683 AR6_WGII 1057 15 These projected changes are expected to slow progress towards eradication of child undernutrition and malnutrition high 2 train +4684 AR6_WGII 1057 16 Higher atmospheric concentrations of carbon dioxide reduce the nutritional quality of wheat, rice and other major crops, potentially affecting millions of people at a doubling of carbon dioxide (very high confidence) {7.3.1}.Climate change is expected to have adverse impacts on well-being and to further threaten mental health very high 3 train +4685 AR6_WGII 1057 18 Mental health impacts are expected to arise from exposure to high temperatures, extreme weather events, displacement, malnutrition, conflict, climate-related economic and social losses, and anxiety and distress associated with worry about climate change very high 3 train +4686 AR6_WGII 1057 19 Future climate-related migration is expected to vary by region and over time, according to future climatic drivers, patterns of population growth, adaptive capacity of exposed populations and international development and migration policies high 2 train +4687 AR6_WGII 1057 20 The wide range of potential outcomes is reflected in model projections of population displacements by 2050 in Latin America, sub- Saharan Africa and south Asia due to climate change, which vary from 31 million to 143 million people, depending on assumptions about future GHG emissions and socioeconomic development trajectories high 2 train +4688 AR6_WGII 1057 21 With every additional one degree Celsius of warming, the global risks of involuntary displacement due to flood events are projected to rise by approximately 50% high 2 train +4689 AR6_WGII 1057 22 High emissions/low development scenarios raise the potential for higher levels of migration and involuntary displacement (high confidence) and increase the need for planned relocations and support for people exposed to climate extremes but lacking the means to move high 2 train +4690 AR6_WGII 1057 23 Climate change may increase susceptibility to violent conflict, primarily intra-state conflicts, by strengthening climate-sensitive drivers of conflict medium 1 train +4691 AR6_WGII 1057 24 Future violent conflict risk is highly mediated by socioeconomic development trajectories (high confidence) and so trajectories that prioritise economic growth, political rights and sustainability are associated with lower conflict risk medium 1 train +4692 AR6_WGII 1057 25 Future climate change may exceed adaptation limits and generate new causal pathways not observed under current climate variability medium 1 train +4693 AR6_WGII 1057 26 Economic shocks are not included in many models of conflict risks currently used, and some projections do not incorporate known socioeconomic predictors of conflict medium 1 train +4694 AR6_WGII 1057 27 As such, future increases in conflict- related deaths with climate change have been estimated, but results are inconclusive medium 1 train +4695 AR6_WGII 1057 28 Solutions Since AR5, the value of cross-sectoral collaboration to advance sustainable development has been more widely recognised, but despite acknowledgement of the importance of health adaptation as a key component, action has been slow high 2 train +4696 AR6_WGII 1057 29 Building climate-resilient health systems will require multi-sectoral, multi-system and collaborative efforts at all governance scales very high 3 test +4697 AR6_WGII 1057 30 Globally, health systems are poorly resourced in general, and their capacity to respond to climate change is weak, with mental health support being particularly inadequate very high 3 train +4698 AR6_WGII 1057 31 The health sectors of some countries have focused on implementing incremental changes to policies and measures to fill the adaptation gap very high 3 train +4699 AR6_WGII 1058 1 This highlights an urgent and immediate need to address the wider interactions between environmental change, socioeconomic development and human health and well-being high 2 train +4700 AR6_WGII 1058 2 Targeted investments in health and other systems, including multi-sectoral, integrated approaches to protect against key health risks can effectively increase resilience high 2 train +4701 AR6_WGII 1058 3 Increased investment in strengthening general health systems, along with targeted investments to enhance protection against specific climate-sensitive exposures (e.g., hazard early warning and response systems, and integrated vector control programmes for VBDs) will increase resilience if implemented to at least keep pace with climate change high 2 train +4702 AR6_WGII 1058 7 These short-term responses can be complemented by longer-term urban planning and design, including nature-based solutions (NbS) that mitigate urban heat island (UHI) effects high 2 train +4703 AR6_WGII 1058 9 The COVID-19 pandemic has demonstrated the value of coordinated and multi-sectoral planning, social protection systems, safety nets and other capacities in societies to cope with a range of shocks and stresses high 2 train +4704 AR6_WGII 1058 10 The pandemic has posed a severe shock to many socioeconomic systems, resulting in substantial changes in vulnerability and exposure of people to climate risks high 2 train +4705 AR6_WGII 1058 11 The pandemic emphasises the inter-connected and compound nature of risks, vulnerabilities, and responses to emergencies that are simultaneously local and global high 2 train +4706 AR6_WGII 1058 12 Pathways to climate resilient development can be pursued simultaneously with recovering from the COVID-19 pandemic high 2 train +4707 AR6_WGII 1058 13 The COVID-19 pandemic has aggravated climate risks, demonstrated the global and local vulnerability to cascading shocks and illustrated the importance of integrated solutions that tackle ecosystem degradation and structural vulnerabilities in human societies high 2 train +4708 AR6_WGII 1058 14 Transitioning towards equitable, low-carbon societies has multiple benefits for health and well-being very high 3 train +4709 AR6_WGII 1058 15 Benefits for health and well-being can be gained from wide-spread, equitable access to affordable renewable energy (high confidence); active transport (e.g., walking and cycling) (high confidence); green buildings and nature-based solutions, such as green and blue urban infrastructure (high confidence); and by transitioning to a low-carbon, well-being-oriented and equity-oriented economy consistent with the aims of the SDGs high 2 train +4710 AR6_WGII 1058 16 Plant-rich diets consistent with international recommendations for healthy diets could contribute to lower GHG emissions while also generating health co-benefits, such as reducing ill health related to over-consumption of animal-based products high 2 train +4711 AR6_WGII 1058 17 Reducing future risks of involuntary migration and displacement due to climate change is possible through cooperative international efforts to enhance institutional adaptive capacity and sustainable development high 2 train +4712 AR6_WGII 1058 18 Institutional and cross-sectoral efforts to build adaptive capacity, coupled with policies aimed at ensuring safe and orderly movements of people within and between states, can form part of the CRDPs that reduce future risks of climate-related involuntary migration, displacement and immobility medium 1 train +4713 AR6_WGII 1058 19 In locations where permanent, government- assisted relocation becomes unavoidable, active involvement of local populations in planning and decision-making increases the likelihood of successful outcomes medium 1 train +4714 AR6_WGII 1058 20 People who live on small island states do not view relocation as an appropriate or desirable means of adapting to the impacts of climate change high 2 train +4715 AR6_WGII 1058 21 Adaptation and sustainable development build peace in conflict- prone regions by addressing the drivers of grievances that lead to conflict and vulnerability to climate change high 2 train +4716 AR6_WGII 1058 22 Environmental peacebuilding (EP) through natural resource sharing, conflict-sensitive adaptation and climate-resilient peacebuilding offer promising avenues for addressing conflict risk, but their efficacy is still to be demonstrated through effective monitoring and evaluation high 2 train +4717 AR6_WGII 1058 23 Formal institutional arrangements for natural resource management contribute to wider cooperation and peacebuilding (high confidence) and gender-based approaches provide under-utilised pathways to achieving sustainable peace medium 1 train +4718 AR6_WGII 1058 24 Inclusion, cross-issue and cross-sectoral integration in policy and programming, and approaches that incorporate different geographical scales and work across national boundaries can support climate- resilient peace high 2 train +4719 AR6_WGII 1059 5 Specific observations of current impacts included the expansion of the geographical ranges of some diseases into previously unaffected areas and changes in the distributions of some food-, water- and vector- borne diseases high 2 train +4720 AR6_WGII 1059 6 Increasing future health risks were projected from injury, disease and death due to more intense heatwaves and fires (very high confidence), undernutrition in poor regions (high confidence), food- and waterborne diseases (very high confidence) and VBDs medium 1 train +4721 AR6_WGII 1059 7 AR5 found that climate change is a multiplier of existing health vulnerabilities, including food insecurity and limited access to safe water, improved sanitation, healthcare and education, and that the most effective measures to reduce vulnerability in the near term are programmes that implement and improve basic public health very high 3 train +4722 AR6_WGII 1059 8 Opportunities for co-benefits from mitigation actions were identified through such actions as reducing local emissions of short-lived climate pollutants from energy systems (very high confidence) and expanding transport systems that promote active travel high 2 train +4723 AR6_WGII 1060 15 Access to green spaces was also closely associated with well-being high 2 train +4724 AR6_WGII 1061 9 The exposure to hazards of populations, infrastructure, ecosystem capital, socioeconomic systems and cultural assets critical to health and well-being varies considerably across and within regions high 2 train +4725 AR6_WGII 1069 5 The most vulnerable groups include smallholder farmers, pastoralists, agricultural laborers, poorer households, refugees, indigenous groups, women, children, the elderly and those who are socioeconomically marginalised (FAO et al., 2018; IPCC, 2019b) high 2 train +4726 AR6_WGII 1070 5 Observed Impacts of Climate Change on Health, Well-Being, Migration and Conflict 7.2.1 Observed Impacts on Health and Well-Being Eleven categories of diseases and health outcomes have been identified in this assessment as being climate-sensitive through direct pathways (e.g., heat and floods) and indirect pathways mediated through natural and human systems and economic and social disruptions (e.g., disease vectors, allergens, air and water pollution, and food system disruption) high 2 train +4727 AR6_WGII 1073 7 Figure 7.5 illustrates how climatic and non-climatic drivers and responses determine VBD outcomes.Evidence has increased since AR5 that the vectorial capacity has increased for dengue fever, malaria and other mosquito-borne diseases and that higher global average temperatures are making wider geographic areas more suitable for transmission very high 3 train +4728 AR6_WGII 1073 8 Transmission rates of malaria are directly influenced by climatic and weather variables such as temperature, with non-climatic socioeconomic factors and health system responses counteracting the climatic drivers very high 3 train +4729 AR6_WGII 1074 7 Changing climatic patterns are facilitating the spread of CHIKV, Zika, Japanese encephalitis and Rift Valley Fever in Asia, Latin America, North America and Europe high 2 train +4730 AR6_WGII 1074 13 Changes in temperature, precipitation, and relative humidity have been implicated as drivers of West Nile fever in southeastern Europe medium 1 train +4731 AR6_WGII 1074 16 Climate change has contributed to the spread of the Lyme disease vector Ixodes scapularis, a corresponding increase in cases of Lyme disease in North America high 2 train +4732 AR6_WGII 1075 3 Ixodes ricinus, the primary vector in Europe for both Lyme borreliosis and tick-borne encephalitis is sensitive to humidity and temperature (Daniel et al., 2018; Estrada- Peña and Fernández-Ruiz, 2020) high 2 train +4733 AR6_WGII 1075 14 Since AR5 there is a growing body of evidence that increases in temperature (very high confidence), heavy rainfall (high confidence), flooding (medium confidence) and drought low 0 train +4734 AR6_WGII 1075 21 Heavy rainfall and higher than normal temperatures are associated with increased cholera risk in affected regions very high 3 train +4735 AR6_WGII 1075 25 Heavy rainfall, warmer weather and drought are linked to increased risks for other GI infections high 2 train +4736 AR6_WGII 1076 11 A strong association exists between increases in FBDs and high air and water temperatures and longer summer seasons very high 3 train +4737 AR6_WGII 1077 2 There is a strong association observed between the increase in average ambient temperature and increases in Salmonella infections high 2 train +4738 AR6_WGII 1077 6 Significant associations exist between FBDs due to Campylobacter, precipitation and temperature medium 1 train +4739 AR6_WGII 1078 8 COVID-19 and environmental connections Infectious diseases may emerge and spread through multiple climate-related avenues, including direct effects of climatic conditions on disease reproduction and transmission and various indirect effects, often interlinked with ecosystem degradation high 2 train +4740 AR6_WGII 1078 9 Climate change is affecting the risk of emerging infectious diseases by contributing to factors that drive the movements of species, including vectors and reservoirs of diseases, into novel human populations and vice versa high 2 train +4741 AR6_WGII 1078 11 Human-to-human transmission is the prominent driver in the spread of the COVID-19 pandemic, rather than climatic drivers high 2 train +4742 AR6_WGII 1078 14 The impact of COVID-19 containment measures resulted in a temporary reduction in greenhouse gas (GHG) emissions and reduced air pollution high 2 train +4743 AR6_WGII 1079 7 During the pandemic, extreme weather and climate events such as droughts, storms, floods, wildfires and heatwaves continued, resulting in disastrous compounding impacts high 2 train +4744 AR6_WGII 1079 12 Responses and implications for adaptation and climate resilient development The pandemic emphasises the inter-connected and compound nature of risks, vulnerabilities and responses to emergencies that are simultaneously local and global high 2 train +4745 AR6_WGII 1080 1 The management of the COVID-19 pandemic has highlighted the value of scientific (including medical and epidemiological) expertise and the importance of fast, accurate and comprehensive data to inform policy decisions and to anticipate and manage risk high 2 train +4746 AR6_WGII 1080 9 There is a considerable evidence base of specific actions that have co-benefits for reducing pandemic and climate change risks while enhancing social justice and biodiversity conservation high 2 train +4747 AR6_WGII 1080 13 In particular, given their immense scale, COVID-19 recovery investments may offer an opportunity to contribute to climate resilient development pathways (CRDPs) through a green, resilient, healthy and inclusive recovery high 2 train +4748 AR6_WGII 1080 14 However, windows of opportunity to enable such transitions are only open for a limited period and need to be swiftly acted upon to effect change high 2 train +4749 AR6_WGII 1080 17 In addition, heightened societal and political attention to one crisis often comes at the cost of other policy priorities high 2 train +4750 AR6_WGII 1082 9 Climate change affects the risk of CVD through high temperatures and extreme heat (assessed in Section 7.2.4.1) and through other mechanisms medium 1 train +4751 AR6_WGII 1082 19 Several non-communicable respiratory diseases are climate sensitive based on their exposure pathways very high 3 train +4752 AR6_WGII 1082 21 Burdens of allergic disease, particularly allergic rhinitis and allergic asthma may be changing in response to climate change medium 1 train +4753 AR6_WGII 1083 15 People with chronic illnesses are at particular risk during and after extreme weather events due to treatment interruptions and lack of access to medication medium 1 train +4754 AR6_WGII 1083 23 Heat continues to pose a significant health risk due to increases in exposure, changes in the size and spatial distribution of the human population, mounting vulnerability and an increase in extreme heat events high 2 train +4755 AR6_WGII 1083 24 Some regions are already experiencing heat stress conditions approaching the upper limits of labour productivity and human survivability high 2 train +4756 AR6_WGII 1084 8 Several lines of evidence point to a possible decrease in population sensitivity to heat, albeit mainly for high-income countries high 2 train +4757 AR6_WGII 1084 15 Excess deaths during extreme heat events occur predominantly in older individuals and are overwhelmingly cardiovascular in origin very high 3 train +4758 AR6_WGII 1084 26 Heat- and cold-related health outcomes vary by location (Dialesandro et al., 2021; Hu et al., 2019; Phung et al., 2016), suggesting outcomes are highly moderated by socioeconomic, occupational and other non-climatic determinants of individual health and socioeconomic vulnerability (Åström et al., 2020; McGregor et al., 2017; McGregor et al., 2017; Schuster et al., 2017; Benmarhnia et al., 2015; Watts et al., 2019) high 2 train +4759 AR6_WGII 1086 1 It is as likely as not that climate change has increased the current burden of disease from injuries related to extreme weather, particularly in low- income settings low 0 train +4760 AR6_WGII 1087 2 Consumption of mercury-contaminated fish has been found to be linked to neurological disorders due to methyl mercury poisoning (i.e., Minamata disease) that is associated with climate change-contaminant interactions that alter the bioaccumulation and biomagnification of toxic and fat-soluble persistent organic pollutants and polychlorinated biphenyls (PCBs) (Alava et al., 2017) in seafood and marine mammals medium 1 train +4761 AR6_WGII 1087 14 There is an observable association between high temperatures and mental health decrements high 2 train +4762 AR6_WGII 1087 21 Discrete climate hazards including storms (Kessler et al., 2008; Boscarino et al., 2013; Boscarino et al., 2017; Obradovich et al., 2018), floods (Baryshnikova and Pham, 2019), heatwaves, wildfires and drought (Hanigan et al., 2012; Carleton, 2017; Zhong et al., 2018; Charlson et al., 2021) have significant negative consequences for mental health very high 3 train +4763 AR6_WGII 1089 3 Mental health impacts can emerge as result of climate impacts on economic, social and food systems high 2 train +4764 AR6_WGII 1089 9 Anxiety about the potential risks of climate change and awareness of climate change itself can affect mental health even in the absence of direct impacts low 0 test +4765 AR6_WGII 1089 21 Heat is one of the best-studied aspects of climate change observed to reduce well-being high 2 train +4766 AR6_WGII 1090 4 Both heat and air pollution also impair human capabilities through a negative effect on cognitive performance (Taylor et al., 2016b) and even impair skills acquisition, reducing the ability to learn (Park et al., 2021) and affecting marginalised groups more strongly (Park et al., 2020), although findings are inconsistent and depend in part on the nature of the task low 0 train +4767 AR6_WGII 1090 5 Systematic reviews have found an association between higher ambient levels of fine airborne particles with cognitive impairment in the elderly and with behavioural problems (related to impulsivity and attention problems) in children (Power et al., 2016; Yorifuji et al., 2017; Younan et al., 2018; Zhao et al., 2018b) medium 1 train +4768 AR6_WGII 1090 8 A general theme across studies from all regions is that climate-related migration outcomes are diverse high 2 train +4769 AR6_WGII 1090 11 The diversity of potential migration and displacement outcomes reflects (a) the variable nature of climate hazards in terms of the rate of onset, intensity, duration, spatial extent and severity of damage caused to housing, infrastructure and livelihoods and (b) the wide range of social, economic, cultural, political and other non-climatic factors that influence exposure, vulnerability, adaptation options and the contexts in which migration decisions are made (Neumann and Hermans, 2015; McLeman, 2017; Barnett and McMichael, 2018; Cattaneo et al., 2019; Hoffmann et al., 2020) high 2 train +4770 AR6_WGII 1090 12 Weather events and climate conditions can act as direct drivers of migration and displacement (e.g., destruction of homes by tropical cyclones) and as indirect drivers (e.g., rural income losses and/or food insecurity due to heat- or drought-related crop failures that in turn generate new population movements) high 2 train +4771 AR6_WGII 1090 15 The diversity of potential migration and displacement outcomes reflects the scale and physical impacts of specific climate hazard events and the wide range of social, economic, cultural, political and other non-climatic factors that influence exposure, vulnerability, adaptation options and the contexts in which migration decisions are made high 2 train +4772 AR6_WGII 1090 18 Climate-related migration and displacement outcomes display high variability in terms of migrant success, often reflecting pre-existing socioeconomic conditions and household wealth high 2 train +4773 AR6_WGII 1091 3 Climatic conditions, events and variability are important drivers of migration and displacement (high confidence) (Table MIGRATE.1 in Chapter 7), with migration responses to specific climate hazards being strongly influenced by economic, social, political and demographic processes high 2 train +4774 AR6_WGII 1091 5 Involuntary displacement occurs when adaptation alternatives are exhausted or not viable and reflects non-climatic factors that constrain adaptive capacity and create high levels of exposure and vulnerability high 2 train +4775 AR6_WGII 1091 6 There is strong evidence that climatic disruptions to agricultural and other rural livelihoods can generate migration high 2 train +4776 AR6_WGII 1091 7 Specific climate events and conditions may cause migration to increase, decrease or flow in new directions high 2 train +4777 AR6_WGII 1091 9 Most climate-related migration and displacement observed currently takes place within countries high 2 train +4778 AR6_WGII 1091 10 The climate hazards most commonly associated with displacement are tropical cyclones and flooding in most regions, with droughts being an important driver in sub-Saharan Africa, parts of south Asia and South America high 2 train +4779 AR6_WGII 1091 14 The additional impacts of climate change anticipated to generate future migration and displacement include mean sea level rise that increases flooding and saltwater contamination of soil and/or groundwater in low-lying coastal areas and small islands high 2 train +4780 AR6_WGII 1092 1 There is growing evidence about the future prospects of immobile populations: groups and individuals that are unable or unwilling to move away from areas highly exposed to climatic hazards high 2 train +4781 AR6_WGII 1092 3 Managed retreat and organised relocations of people from hazardous areas in recent years have proven to be politically and emotionally charged, socially disruptive and costly high 2 train +4782 AR6_WGII 1093 23 Policy interventions can remove barriers to and expand the alternatives for safe, orderly and regular migration that allows vulnerable people to adapt to climate change high 2 train +4783 AR6_WGII 1094 6 Reducing the future risk of large-scale population displacements, including those requiring active humanitarian interventions and organised relocations of people, requires the international community to meet the requirements of the Paris Agreement and take further action to control future warming high 2 train +4784 AR6_WGII 1094 10 Populations in low-income countries and small-island states that have historically had low greenhouse gas (GHG) emissions are at particular risk of involuntary migration and displacement due to climate change, reinforcing the urgency for industrialised countries to continue lowering GHG emissions, to support adaptive capacity-building initiatives under the UNFCCC and to meet objectives expressed in the Global Compacts regarding safe, orderly and regular migration and the support and accommodation of displaced people (Sections 4.3.7, 4.5.7, 5.12.2, 7.4.5.5, 8.4.2; Box 8.1; Cross-Chapter Box SLR in Chapter 3).Cross-Chapter Box MIGRATE (continued) Climate-related migration originates most often in rural areas in low- and middle-income countries, with migrant destinations usually being other rural areas or urban centres within their home countries (i.e., internal migration) medium 1 train +4785 AR6_WGII 1095 9 Regional distribution of displacement events has been consistent throughout the IDMC data collection period high 2 train +4786 AR6_WGII 1095 10 Relative to their absolute population size, small island states experience a disproportionate risk of climate-related population displacements (Desai et al., 2021) high 2 train +4787 AR6_WGII 1095 11 Tropical cyclones and extreme storms are a particularly significant displacement risk in East and Southeast Asia, the Caribbean region, the Bay of Bengal region and southeast Africa (IDMC 2020) high 2 train +4788 AR6_WGII 1096 1 Riverine flood displacement can lead to increases or decreases in temporary or short-distance migration flows, depending on the local context medium 1 train +4789 AR6_WGII 1096 10 The most common response to drought is an increase in short-distance, rural–urban migration medium 1 train +4790 AR6_WGII 1097 10 These latter may be attributable to a sense of loss or fear (Schwerdtle et al., 2018; Torres and Casey, 2017) as well as due to the interruption of healthcare; occupational injuries; sleep deprivation; non-hygienic lodgings and insufficient sanitary facilities; heightened exposure to vector- and WBDs; vulnerability to psychosocial, sexual and reproductive issues; behavioural disorders; substance abuse; and violence (Farhat et al., 2018; Wickramage et al., 2018) high 2 train +4791 AR6_WGII 1097 15 The chapter concluded that some of the factors that increase the risk of violent conflict within states are sensitive to climate change (medium evidence, medium agreement), that people living in places affected by violent conflict are particularly vulnerable to climate change medium 1 train +4792 AR6_WGII 1101 10 Under the high emissions scenario, climate change was projected to result in approximately 85 deaths equivalents per 100,000 population.Temperature increases are projected to exceed critical risk thresholds for six key climate-sensitive health outcomes, highlighting the criticality of building adaptive capacity in health systems and in other sectors that influence health and well-being high 2 train +4793 AR6_WGII 1102 4 Population heat exposure will increase under climate change very high 3 train +4794 AR6_WGII 1102 10 Further, for North America and Europe, where rural depopulation is projected, the predominant driver of increases in exposure is urban growth (Jones et al., 2018).Comparisons of heatwave exposure for 1.5°C and 2.0°C warming for different SSPs indicate strong geographical contrasts in potential heatwave risk high 2 train +4795 AR6_WGII 1102 16 Regional level assessments of changes in population heat exposure for Africa, Europe, the USA, China and India corroborate general findings at the global level, that the impact of warming is amplified under divergent regional development pathways (e.g., SSP4 – inequality) compared to those fostering sustainable development (e.g., SSP1 – sustainability) high 2 train +4796 AR6_WGII 1102 17 For some regions, such as Europe, changes in exposure are projected to be largely a consequence of climate change, while for others, such as Africa and to a lesser extent Asia, Oceania, North America and South America, the interactive effects of demographic and climate change are projected to be important (Jones et al., 2018; Liu et al., 2017; Russo et al., 2016; Ma and Yuan, 2021) medium 1 train +4797 AR6_WGII 1103 1 Estimates of heat-related mortality based solely on changes in temperature point to elevated levels of global and regional level mortality compared to the present, with the magnitude of this increasing from RCP4.5 through to RCP8.5 high 2 train +4798 AR6_WGII 1103 2 Further support comes from the projection that heat-related health impacts for a 2°C increase in global temperatures will be greater than those for 1.5°C warming very high 3 train +4799 AR6_WGII 1103 3 Estimates of future mortality that incorporate adaptation in addition to temperature change point to increases in heat-related mortality under global warming, albeit at lower levels than the case of no adaptation high 2 train +4800 AR6_WGII 1103 4 Whether adaptation is considered or not, the consensus is Central and South America, southern Europe, southern and Southeast Asia and Africa will be the most affected by climate change in terms of heat-related mortality high 2 train +4801 AR6_WGII 1103 5 Similarly, projections of the impacts of future heat on occupational health, worker productivity and workability point to these regions as problematic under climate change high 2 train +4802 AR6_WGII 1103 6 This accords with the findings from independent projections of population heat exposure as outlined above high 2 train +4803 AR6_WGII 1103 7 The effect of climate change on productivity is projected to reduce GDP at a range of geographical scales high 2 train +4804 AR6_WGII 1103 12 Future increases in heat-related deaths are expected to outweigh those related to cold high 2 train +4805 AR6_WGII 1103 16 Heat risks are expected to be greater in urban areas due to changes in regional heat exacerbated by ‘heat island’ effects high 2 train +4806 AR6_WGII 1103 25 There is a high likelihood that climate change will contribute to increased distributional range and vectorial capacity of malaria vectors in parts of sub-Saharan Africa, Asia and South America high 2 train +4807 AR6_WGII 1104 3 Malaria infections have significant potential to increase in parts of sub-Saharan Africa and Asia, with risk varying according to the warming scenario medium 1 train +4808 AR6_WGII 1105 4 Rising temperatures are likely to cause poleward shifts and overall expansion in the distribution of mosquitoes Aedes aegypti and Aedes albopictus, the principal vectors of dengue, yellow fever, chikungunya and Zika high 2 train +4809 AR6_WGII 1105 10 Climate change is expected to increase dengue risk and facilitate its global spread, with the risk being greatest under high emissions scenarios high 2 train +4810 AR6_WGII 1105 19 Zika virus transmits under different temperature optimums than does dengue, suggesting environmental suitability for Zika transmission could expand with future warming low 0 train +4811 AR6_WGII 1105 20 Climate change can be expected to continue to contribute to the geographical spread of the Lyme disease vector Ixodes scapularis (high confidence) and the spread of tick-borne encephalitis and Lyme disease vector Ixodes ricinus in Europe medium 1 train +4812 AR6_WGII 1105 24 Climate change is projected to increase the incidence of Lyme disease and tick-borne encephalitis in the Northern Hemisphere high 2 train +4813 AR6_WGII 1105 27 Climate change is projected to change the distribution of schistosomiasis in Africa and Asia (high confidence), with a possible increase in global land area suitable for transmission medium 1 train +4814 AR6_WGII 1106 7 Climate change is expected to increase future health risks associated with a range of other WBDs and parasites, with effects varying by region medium 1 train +4815 AR6_WGII 1106 12 The risk of Campylobacteriosis and other enteric pathogens could rise in regions where heavy precipitation events or flooding are projected to increase medium 1 train +4816 AR6_WGII 1106 21 Demographic trends in aging and more vulnerable population are likely to be important determinants of future air quality—a human health climate penalty high 2 train +4817 AR6_WGII 1107 1 Projections indicate that emission reduction scenarios consistent with stabilisation of global temperature change at 2°C or below would yield substantial co-benefits for air quality-related health outcomes (Chowdhury et al., 2018b; von Schneidemesser et al., 2020; Silva et al., 2016c; Markandya et al., 2018; Orru et al., 2019; Shindell et al., 2018) high 2 train +4818 AR6_WGII 1107 21 The burden of disease associated with aeroallergens is anticipated to grow due to climate change high 2 train +4819 AR6_WGII 1109 4 The nutritional density, including protein content, micronutrients and B-vitamins, of wheat, rice, barley and other important food crops is negatively affected by higher CO 2 concentrations very high 3 train +4820 AR6_WGII 1109 15 Climate change impacts on oceans could generate increased risks of ciguatera poisoning in some regions medium 1 train +4821 AR6_WGII 1109 26 However, changes (Hayes and Poland, 2018) in extreme events due to climate change, including floods (Baryshnikova, 2019), droughts (Carleton, 2017) and hurricanes (Kessler et al., 2008; Boscarino et al., 2013, Boscarino et al., 2017; Obradovich et al., 2018), which are projected to increase due to climate change, directly worsen mental health and well-being and increase anxiety high 2 train +4822 AR6_WGII 1110 2 Human behaviours and systems will be disrupted by climate change in a myriad of ways, and the potential consequences for mental health and well-being are correspondingly large in number and complex in mechanism high 2 train +4823 AR6_WGII 1110 7 Broad societal outcomes such as economic unrest, political conflict or governmental dysfunction assessed in Section 7.3.5 may undermine the mental health of populations in the future medium 1 train +4824 AR6_WGII 1110 13 The most common drivers of observed climate- related migration and displacement are extreme storms (particularly tropical cyclones), floods and droughts high 2 test +4825 AR6_WGII 1110 16 In low-lying coastal areas of most regions, future increases in mean sea levels will amplify the impacts of coastal hazards on settlements, including erosion, inland penetration of storm surges and groundwater contamination by salt water, and eventually lead to inundation of very low-lying coastal settlements high 2 train +4826 AR6_WGII 1112 10 Recent evidence adds further support for such conclusions high 2 train +4827 AR6_WGII 1113 5 Future violent conflict risk is highly mediated by socioeconomic development trajectories high 2 train +4828 AR6_WGII 1114 6 As documented across this chapter, there is a large adaptation deficit for health and well-being, with climate change causing avoidable injuries, illnesses, disabilities, diseases and deaths high 2 train +4829 AR6_WGII 1114 8 Current global investments in health adaptation are insufficient to protect the health of populations and communities high 2 train +4830 AR6_WGII 1114 15 There is increased understanding of exposure and vulnerabilities to climate variability and change, the capacities to manage the health risks, the effectiveness of adaptation (including a growing number of lessons learned and best practices), and the co- benefits of mitigation policies and technologies high 2 train +4831 AR6_WGII 1114 16 Effectively preparing for and managing the health risks of climate change requires considering the multiple interacting sectors that affect population health and the effective functioning of health systems high 2 train +4832 AR6_WGII 1114 25 Effective health risk management incorporates the magnitude and pattern of future climate risks as well as potential changes in factors that determine vulnerability and exposure to climate hazards, such as determinants of healthcare access, demographic shifts, urbanisation patterns and changes in ecosystems very high 3 train +4833 AR6_WGII 1116 13 Increased investment in strengthening general health systems, along with targeted investments to enhance protection against specific climate-sensitive exposures (e.g., hazard early warning and response systems and integrated vector control programmes for VBDs) will increase resilience if implemented to at least keep pace with climate change high 2 train +4834 AR6_WGII 1116 14 Investments to address the social determinants of health can reduce inequities and increase resilience high 2 train +4835 AR6_WGII 1118 2 Vulnerability, adaptation and capacity assessments include consideration of the feasibility and effectiveness of priority health adaptation options and can help decision makers identify strategies for enhancing adaptation feasibility in specific contexts.7.4.2.3 Adaptation Options for Vector-borne, Water-borne and Food-Borne Diseases Integrated vector control approaches are crucial to effectively manage the geographic spread, distribution and transmission of VBDs associated with climate change high 2 train +4836 AR6_WGII 1119 6 Adaptation options for climate-related risks for WBDs and FBDs are strongly associated with wider, multi-sectoral initiatives to improve sustainable development in low-income communities high 2 train +4837 AR6_WGII 1119 17 While well-designed and operationalisable HAPs possess the potential to reduce the likelihood of mortality from extreme heat events medium 1 train +4838 AR6_WGII 1119 18 Evaluations of heatwave early warning systems as a component within HAPs show inconsistent results in terms of their impact on predicting mortality rates (Nitschke et al., 2016; Benmarhnia et al., 2016; Heo et al., 2019a; Heo et al., 2019b; Ragettli and Roosli, 2019; Martinez et al., 2019; De’Donato et al., 2018; Weinberger et al., 2018b), indicating climate-based heat warning systems, which use a range of heat stress metrics (Schwingshackl et al., 2021), are not sufficient as a stand-alone approach to heat risk management high 2 train +4839 AR6_WGII 1119 19 To support HAP and heat risk-related policy development, identification and mapping of heat vulnerability ‘hot spots’ within urban areas have been proposed (Chen et al., 2019; Hatvani-Kovacs et al., 2018) A multi-sectoral approach, including the engagement of a range of stakeholders will likely benefit the response to longer-term heat risks through the implementation of measures such as climate- sensitive urban design and planning that mitigates UHI effects high 2 train +4840 AR6_WGII 1121 13 Adaptation actions include access to healthy, affordable diverse diets from sustainable food systems (high confidence); a combination of access to health— including maternal, child and reproductive health— and nutrition services, water and sanitation (high confidence); access to nutrition- sensitive and shock-responsive social protection (high confidence); and early warning systems (high agreement), risk sharing, transfer, and risk reduction schemes such as index-based weather insurance medium 1 train +4841 AR6_WGII 1121 14 Common enablers across adaptation actions that enhance the effectiveness and feasibility of the adaptation include: education, women’s and girls’ empowerment (high confidence), rights-based governance and peacebuilding social cohesion initiatives such as the framework of the Humanitarian Development and Peace Nexus medium 1 train +4842 AR6_WGII 1121 15 Nutrition-sensitive and integrated agroecological farming systems offer opportunities to increase dietary diversity at household levels while building local resilience to climate-related food insecurity high 2 train +4843 AR6_WGII 1124 3 Because mental health is fundamentally inter-twined with social and economic well-being, adaptation for climate-related mental health risks benefits from wider multi-sectoral initiatives to enhance well- being, with the potential for co-benefits to emerge high 2 train +4844 AR6_WGII 1124 7 Adaptive urban design that provides access to healthy natural spaces—an option for reducing risks associated with heat stress—also promotes social cohesion and mitigates mental health challenges high 2 train +4845 AR6_WGII 1124 13 An effective early warning system for malaria was implemented in the Amhara region of Ethiopia (Merkord et al., 2017).Early warning systems are effective at detecting and potentially reducing food security and nutrition risks high 2 train +4846 AR6_WGII 1124 16 Financial investments to develop early warning systems are cost-effective and reduce human suffering (Choularton and Krishnamurthy, 2019) high 2 train +4847 AR6_WGII 1125 16 No universal standardised approach exists for monitoring or evaluating adaptation activities in the health sector high 2 train +4848 AR6_WGII 1125 23 Successful adaptation to the health impacts of climate change in Indigenous Peoples requires recognition of their rights to self-determination, focusing on indigenous conceptualisations of well-being, prioritising Indigenous knowledge and understanding the broader agenda of decolonisation, health and human rights high 2 train +4849 AR6_WGII 1126 8 Climate resilient development pursued in these other sectors, and in cooperation with the health sector, simultaneously increases the potential for adaptation and climate resilience in terms of health and well-being high 2 train +4850 AR6_WGII 1127 6 Efficacy beliefs, social norms and subjective resilience also affect adaptation behaviour medium 1 train +4851 AR6_WGII 1127 19 Subsequent research indicates that the circumstances under which migration occurs and the degree of agency under which household migration decisions are made are important determinants of whether migration outcomes are successful in terms of advancing the well- being of the household and providing benefits to sending and receiving communities high 2 train +4852 AR6_WGII 1128 12 Attitudes of residents in migrant-receiving areas with respect to climate-related migration warrant consideration when formulating adaptation policy medium 1 train +4853 AR6_WGII 1128 23 Disruptive and expensive relocations of low-lying coastal settlements in many regions would become increasingly necessary in coming decades under high levels of warming high 2 train +4854 AR6_WGII 1129 16 Formal institutional arrangements for natural resource management can contribute to transnational cooperation high 2 train +4855 AR6_WGII 1134 10 Recognising that conflict results from underlying vulnerabilities, development that reduces vulnerability offers the best win-win option for building sustainable, climate- resilient peace rather than specific security-focused interventions high 2 train +4856 AR6_WGII 1135 1 Given the overlap in sources of greenhouse gases (GHGs) and co-pollutants in energy systems, strategies that pursue GHG emission reductions and improvements in energy efficiency hold significant potential health co-benefits through air pollution emission reductions high 2 train +4857 AR6_WGII 1135 10 Transitioning to affordable clean energy sources for all presents opportunities for substantial well-being, health, and equity co-benefits high 2 train +4858 AR6_WGII 1135 17 Transformative approaches that reduce climate-related risks and deliver enhanced social inclusion and development opportunities for the urban poor are most likely where local governments act in partnership with local communities and other civil society actors high 2 train +4859 AR6_WGII 1135 19 Multi-level leadership, institutional capacity and financial resources to support inclusive adaptation in the context of multiple pressures and inter-connected risks can help ensure that the additional 2.5 billion people projected to live in urban areas by 2050 are less exposed to climate-related hazards and contribute less to global warming high 2 train +4860 AR6_WGII 1135 22 Stimulating active mobility (walking and bicycling) can bring physical and mental health benefits high 2 train +4861 AR6_WGII 1136 2 Urban green and blue spaces contribute to climate change adaptation and mitigation and improve physical and mental health and well-being high 2 train +4862 AR6_WGII 1136 7 Climate adaptation and mitigation policies in the building sector offer multiple well-being and health co-benefits high 2 train +4863 AR6_WGII 1136 11 Shifting to sustainable food systems that provide affordable, diverse and plant-rich diets with moderate quantities of GHG-intensive animal protein can bring health co-benefits and substantially reduce GHG emissions, especially in high income countries and where ill health related to overconsumption of animal-based products is prevalent very high 3 train +4864 AR6_WGII 1136 12 Transforming the food system by limiting the demand for GHG-intensive animal foods, reducing food over-consumption and transitioning to nutritious, plant-rich diets can have significant co-benefits to health high 2 train +4865 AR6_WGII 1185 2 These impacts disproportionately affect margin- alised groups, amplifying inequalities and undermining sustainable development across all regions high 2 train +4866 AR6_WGII 1185 6 Climate change increases the threat of chronic and sudden onset development challenges, such as poverty traps and food insecurity high 2 train +4867 AR6_WGII 1185 7 Adaptation interventions and transformative solutions that prioritise inclusive and wide-ranging climate resilient development and the reduction of poverty and inequality are increasingly seen as necessary to minimise loss and damage from climate change high 2 train +4868 AR6_WGII 1185 9 Observed average mortality from floods, drought and storms is 15 times higher for regions and countries ranked as very high vulnerable, such as Mozambique, Somalia, Nigeria, Afghanistan and Haiti compared to very low vulnerable regions and countries, such as UK, Australia, Canada and Sweden in the last decade high 2 train +4869 AR6_WGII 1185 10 Over 3.3 billion people are living in countries classified as very highly or highly vulnerable, while around 1.8 billion people live in countries with low or very low vulnerability high 2 train +4870 AR6_WGII 1185 11 Approximately 3.6 billion people live in low and lower middle-income countries, which are most vulnerable and disproportionally bear the human costs of dis- asters due to extreme weather events and hazards high 2 train +4871 AR6_WGII 1185 13 Vulnerability is a result of many interlinked issues concerning poverty, migration, in- equality, access to basic services, education, institutions and govern- ance capacities, often made more complex by past developments, such as histories of colonialism high 2 train +4872 AR6_WGII 1185 22 If future climate change under high emissions scenarios continues and increases risks, without strong adaptation measures, losses and damages will likely2 be concentrated among the poorest vulnerable populations high 2 train +4873 AR6_WGII 1185 25 Under an inequality scenario (Shared Socioeconomic Pathway (SSP) 4) the projected number of people living in extreme poverty may increase by 122 million by 2030 medium 1 train +4874 AR6_WGII 1185 28 Gender inequality and discrimination are among the barriers to adaptation high 2 train +4875 AR6_WGII 1185 31 Even with moderate climate change3 people in vulner - able regions will experience a further erosion of livelihood security that can interact with humanitarian crises, such as displacement and forced migration (high confidence) and violent conflict, and lead to social tipping points medium 1 train +4876 AR6_WGII 1185 34 The most vulnerable regions are particularly located in East, Central and West Africa, South Asia, Micronesia and Melanesia and in Central America high 2 train +4877 AR6_WGII 1185 36 Areas of high human vulnerability are characterised by larger transboundary regional clusters high 2 train +4878 AR6_WGII 1186 1 Greater investments are required under higher levels of global warming and of inequality (Relative Concentration Pathway (RCP) 4.5; RCP8.5 and SSP4) high 2 train +4879 AR6_WGII 1186 3 The COVID-19 pandemic is expected to increase the adverse consequences of climate change since the financial consequences have led to a shift in priorities and constrain vulnerability reduction medium 1 train +4880 AR6_WGII 1186 4 Moreover, the COVID-19 pandemic is also influencing the capacities of governmental institutions in developing nations to support planned adaptation and poverty reduction of most vulnerable people/groups, since the crisis also means significant reductions in tax revenues high 2 train +4881 AR6_WGII 1186 6 Enabling environments that sup- port sustainable development are essential for adaptation and climate resilient development high 2 train +4882 AR6_WGII 1186 7 Enabling and supportive environments for adaptation share common governance characteristics, including multiple actors and assets, multiple centres of power at different levels and an effective vertical and horizontal integration between levels high 2 test +4883 AR6_WGII 1186 8 Enabling conditions can support livelihood strategies that do not undermine human well-being medium 1 train +4884 AR6_WGII 1186 10 Improving coherence between adaptations of different social groups and sectors at different scales can reduce maladaptation, enable mitigation and advance progress towards climate resilience medium 1 train +4885 AR6_WGII 1186 13 Climate justice and rights-based approaches are increasingly recognised as key principles within mitigation and adaptation strategies and projects medium 1 train +4886 AR6_WGII 1186 14 Narrowing gender gaps can play a transformative role in pursuing climate justice medium 1 train +4887 AR6_WGII 1186 16 Synergies between adaptation and mitigation exist, and these can have benefits for the poor medium 1 train +4888 AR6_WGII 1186 19 Appropriate governance, in- cluding mainstreaming and policy coherence, supported by adaptation finance that targets the poor and marginalised, is essential for adaptation and climate compatible development medium 1 train +4889 AR6_WGII 1189 11 Recent research shows that climate change impacts may exacerbate poverty indirectly through increasing cost of food, housing and healthcare, among other rising costs borne by the poor (Islam et al., 2014; Ebi et al., 2017; Hallegatte et al., 2018) high 2 train +4890 AR6_WGII 1189 16 One of the key factors that drives disproportionate impacts among poor households globally is lost agricultural income high 2 train +4891 AR6_WGII 1189 22 Figure 8.2 reflects the fundamental threat that climate hazards pose to the survival of plants, livestock and fish, as well as the people on which livelihoods depend high 2 train +4892 AR6_WGII 1189 25 It is revealed that warming trends and droughts pose greatest risks to the widest array of livelihood resources, and are particularly detrimental to crops and human health, a long-term requirement for livelihoods and well-being high 2 train +4893 AR6_WGII 1190 1 Salinity is a secondary hazard related to droughts, coastal flooding and sea level rise, and poses a fundamental risk to agriculture high 2 train +4894 AR6_WGII 1190 2 There is also robust evidence for rainfall variability driving short-term impacts to agricultural productivity as well as permanent loss of agriculture high 2 train +4895 AR6_WGII 1191 3 The hazards most prevalent in all regions include warming trends, droughts and sea level rise (Figure 8.2c), and undermine crop productivity, crop varieties, and cropland in most regions high 2 train +4896 AR6_WGII 1191 4 Along coastlines, climate hazards threaten livelihoods particularly exposed to extreme weather, flooding and sea level rise, and where poor populations are heavily dependent on agriculture and fisheries high 2 train +4897 AR6_WGII 1191 28 Increasingly, intersections of age, gender, socioeconomic class, ethnicity and race are recognised as important to the climate risks and differential impacts and losses experienced by vulnerable, marginal and poor in societies high 2 train +4898 AR6_WGII 1193 4 Migration and displacement are directly induced by the impacts of climate change high 2 train +4899 AR6_WGII 1196 15 First, climate change impacts may undermine progress toward various SDGs medium 1 train +4900 AR6_WGII 1196 20 At the same time, there is increasing evidence that successful adapta- tion depends on equitable development and climate justice; for exam- ple, gender inequality (SDG 5) and discrimination (SDG 16) are among the barriers to effective adaptation high 2 train +4901 AR6_WGII 1196 24 They do not have uniform climate risk profiles, and not all adaptations are equally appropriate in all contexts (Nurse et al., 2014) high 2 train +4902 AR6_WGII 1198 7 Recent literature provides evidence that impacts of climate change together with non-climatic drivers can create poverty– environment traps that may increase the probability of long-term and chronic poverty (Figure 8.4; Hallegatte et al., 2015; Djalante et al., 2020; Malhi et al., 2020; McCloskey et al., 2020) high 2 train +4903 AR6_WGII 1205 5 This literature underscores that approaches to assess resilience, vulnerability and human well-being include global assessments that can inform strategies and priority settings for adaptation and risk reduction in the context of climate change high 2 train +4904 AR6_WGII 1212 12 There is high agreement that, with global warming of about 3°C, such undermining of poverty reduction efforts will intensify and more regions will face development setbacks due to the spatial and temporal expansion of climate hazards, including the further erosion of capital that enables people to develop adaptive capacities high 2 train +4905 AR6_WGII 1212 19 Reducing vulnerability to climate change is therefore indispensable for climate justice and just transitions high 2 train +4906 AR6_WGII 1217 3 The sale of property particularly reduces the asset base, creates long-term vulnerabilities to future events and can trigger chronic poverty high 2 train +4907 AR6_WGII 1217 5 In South Asia, there is robust evidence of economic impacts of climate change (Cao et al., 2021), for example in the Sundarbans (a transboundary ecosystem with components in both India and Bangladesh, with the problem of unproductive livelihoods being common across residents of both countries) observations show local livelihoods are rapidly becoming unproductive (loss of fish, and increasing salination making agriculture increasingly difficult) (Ghosh, 2018); conditions that are exacerbated by climate change impacts high 2 train +4908 AR6_WGII 1218 27 Non-economic losses are particularly relevant for understanding adverse consequences of climate change on the poor and most vulnerable population groups high 2 train +4909 AR6_WGII 1220 4 Examples of livelihood shifts across Asia and Southeast Asia (e.g., Bangladesh, India, Philippines, Vietnam) include rural communities in coastal areas, urban settlements that are experiencing economic losses high 2 train +4910 AR6_WGII 1220 25 Climate change-related extreme weather events, such as typhoons, floods, and droughts, can have detrimental impacts on crop production high 2 train +4911 AR6_WGII 1220 28 In Africa, many communities already experience drought- and flood- related disasters high 2 train +4912 AR6_WGII 1220 30 Food security and agriculture productivity are examples of livelihood resources most at risk to climate hazards (see Figure 8.2) high 2 train +4913 AR6_WGII 1222 10 The multiple shocks of extreme events reduce crop yields, destroy homes, and lead to loss of infrastructure and displacement high 2 train +4914 AR6_WGII 1230 8 In the Himalayas (part of the Hindu Kush Himalaya, HKH) temperature warming is expected to increase up to 2°C by 2050 high 2 train +4915 AR6_WGII 1230 22 Likewise, in the Amazon basin, climate change hazards of severe droughts and floods high 2 train +4916 AR6_WGII 1232 25 The increase in global temperature level up to 2°C will exacerbate food and water insecurity in the Amazon (Betts et al., 2018; Hoegh- Guldberg et al., 2018) medium 1 train +4917 AR6_WGII 1233 15 The categories in Table 8.5 also represent important future compounding and complex risks that can emerge due to maladaptation high 2 train +4918 AR6_WGII 1236 12 In SRCCL (IPCC, 2019a), land degradation and climate change compounded to highly expose the livelihoods of the poor to climate hazards and caused food insecurity (high confidence), migration, conflict and loss of cultural heritage low 0 train +4919 AR6_WGII 1246 2 Some authors suggest including Indigenous knowledge in the IPCC assessment process should be of high priority, as it is becoming increasingly relevant for climate services high 2 train +4920 AR6_WGII 1252 7 CSA, carbon-forestry programmes and the water–energy–climate nexus show trade-offs across levels and sectors with identified winners and losers high 2 train +4921 AR6_WGII 1254 20 Depending on the farming and pastoral systems and level of development, reductions in the emissions intensity of livestock products may lead to absolute reductions in GHG emissions (IPCC, 2019a) medium 1 train +4922 AR6_WGII 1254 21 Significant synergies exist between adaptation and mitigation, for example, through SLM approaches high 2 train +4923 AR6_WGII 1260 10 Meanwhile, the livelihood resources that are globally at greatest risk include people’s bodily health, food security and agricultural productivity high 2 train +4924 AR6_WGII 1260 11 Evidence suggests that the fundamental challenge of climate change to livelihoods is that rising temperatures, drought and other hazards endanger human life, and the lives of plants and animals that humans rely on to survive high 2 train +4925 AR6_WGII 1261 6 This chapter builds on AR5 and the IPCC SR 1.5°C on key limits to the adaptation of natural and social systems that are compounded by the effects of poverty and inequality, such as on water scarcity, ecosystems alteration and degradation, coastal cities in relation to sea level rise, cyclones and coastal erosion, food systems and human health high 2 train +4926 AR6_WGII 1261 7 Climate change risks could have substantial negative impacts on climate-sensitive livelihoods of smallholder farmers, fisheries communities, Indigenous People, urban poor and informal settlements, with limits to adaptation evidenced in the loss income, ecosystems, health and increasing migration high 2 train +4927 AR6_WGII 1261 9 For instance, a hard limit associated with losses of coral reefs in a 1.5°C warmer world will lead to substantial loss of income and livelihoods for coastal communities (high confidence), including loss of culture- and place-based attachment medium 1 train +4928 AR6_WGII 1261 11 Their threshold will affect residents of Arctic regions dependent on hunting and fishing livelihoods high 2 train +4929 AR6_WGII 1261 14 Thus, without sustainable, equitable and urgent adaptation measures, maladaptation risks are likely to further increase vulnerability, marginalisation and ecological tipping points among the poor within countries medium 1 train +4930 AR6_WGII 1300 2 Limiting global warming to 1.5°C is expected to substantially reduce damages to African economies, agriculture, human health, and ecosystems compared to higher levels of global warming high 2 train +4931 AR6_WGII 1300 16 Finance has not targeted more vulnerable countries high 2 train +4932 AR6_WGII 1300 19 Increasing public and private finance flows by billions of dollars per year, increasing direct access to multilateral funds, strengthening project pipeline development and shifting more finance to project implementation would help realise transformative adaptation in Africa high 2 train +4933 AR6_WGII 1300 20 Concessional finance will be required for adaptation in low-income settings high 2 train +4934 AR6_WGII 1300 24 For example, climate change adaptation benefits pandemic preparedness, ‘One Health’ approaches benefit human and ecosystem health, and ecosystem-based adaptation can deliver adap- tation and emissions mitigation high 2 train +4935 AR6_WGII 1301 18 This Report also uses the term ‘likely range’ to indicate that the assessed likelihood of an outcome lies within the 17–83% probability range.Africa), increasing exposure to pluvial and riverine flooding high 2 train +4936 AR6_WGII 1302 20 Although yield declines for some crops may be partially compensated by increasing atmospheric CO 2 concentrations, global warming above 2°C will result in yield reductions for staple crops across most of Africa compared to 2005 yields (e.g., 20–40% decline in west African maize yields), even when considering adaptation options and increasing CO 2 medium 1 train +4937 AR6_WGII 1302 32 Above 2°C of global warming, distribution and seasonal transmission of vector-borne diseases is expected to increase, exposing tens of millions more people, mostly in west, east and southern Africa high 2 train +4938 AR6_WGII 1302 33 Above 1.5°C risk of heat-related deaths rises sharply medium 1 train +4939 AR6_WGII 1302 34 Above 2°C global warming, thousands to tens of thousands of additional cases of diarrhoeal disease are projected, mainly in west, central and east Africa medium 1 train +4940 AR6_WGII 1302 38 High population growth and urbanisation in low-elevation coastal zones will be a major driver of exposure to sea level rise in the next 50 years high 2 train +4941 AR6_WGII 1303 19 Further increasing CO 2 concentrations could increase woody plant cover, but increasing aridity could counteract this, destabilising forest and peatland carbon stores in central Africa low 0 train +4942 AR6_WGII 1303 20 Changes in vegetation cover could occur rapidly if tipping points are crossed {9.6.1, 9.6.2, 9.8.2} African biodiversity loss is projected to be widespread and escalating with every 0.5°C increase above present-day global warming high 2 train +4943 AR6_WGII 1303 30 Most African heritage sites are neither prepared for, nor adapted to, future climate change high 2 train +4944 AR6_WGII 1304 19 Integrated water management measures including sub-national financing, demand management through subsidies, rates and taxes, and sustainable water technologies can reduce water insecurity caused by either drought or floods medium 1 train +4945 AR6_WGII 1304 22 Climate information services, institutional capacity building, secure land tenure, and strategic financial investment can help overcome these barriers to adaptation medium 1 train +4946 AR6_WGII 1305 3 The contribution of Africa is among the lowest of historical greenhouse gas (GHG) emissions responsible for human-induced climate change and it has the lowest per capita GHG emissions of all regions currently high 2 train +4947 AR6_WGII 1305 4 Yet Africa has already experienced widespread impacts from human-induced climate change high 2 train +4948 AR6_WGII 1305 10 Hot days, hot nights and heatwaves have become more frequent; heatwaves have also become longer. Drying is projected particularly for west and southwestern Africa high 2 test +4949 AR6_WGII 1309 15 Climate-related research in Africa faces severe funding constraints with unequal funding relationships between countries and with research partners in Europe and North America high 2 train +4950 AR6_WGII 1310 2 These unequal funding relations influence inequalities in climate-related research design, participation and dissemination between African researchers and researchers from high-income countries outside Africa, in ways that can reduce adaptive capacity in Africa very high 3 train +4951 AR6_WGII 1311 6 In scenarios with low adaptation (that is largely localised and incremental), the transition to high risk—widespread and severe impacts—has already begun at the current level of global warming for biodiversity loss (high confidence), and begins below 1.5°C global warming for both food production (medium confidence) and mortality and morbidity from heat and infectious disease high 2 train +4952 AR6_WGII 1316 7 Adaptation costs in Africa are expected to rise rapidly as global warming increases high 2 train +4953 AR6_WGII 1316 23 The amounts of finance being mobilised internationally to support adaptation in African countries are billions of US dollars less than adaptation cost estimates, and finance has targeted mitigation more than adaptation high 2 train +4954 AR6_WGII 1322 3 In Africa, placing cross-sectoral approaches at the core of CRD provides significant opportunities to deliver large benefits and/or avoided damages across multiple sectors including water, health, ecosystems and economies very high 3 train +4955 AR6_WGII 1327 1 Despite these challenges, the inclusive nature of co-production has had a positive influence on the uptake of CS into decision making where it has been applied (Table 9.4; Figure 9.12; Vincent et al., 2018; Vogel et al., 2019; Carter et al., 2020; Chiputwa et al., 2020) medium 1 train +4956 AR6_WGII 1331 20 Most African countries are expected to experience high temperatures unprecedented in their recent history earlier in this century than generally wealthier, higher latitude countries high 2 test +4957 AR6_WGII 1336 19 A shift to a later onset and end of the west African monsoon is also reported in west Africa and Sahel low 0 train +4958 AR6_WGII 1337 2 Meteorological, agricultural and hydrological drought in the region has increased in frequency since the 1950s medium 1 train +4959 AR6_WGII 1337 6 Although there are uncertainties in rainfall projections over the Sahel (Klutse et al., 2018; Gutiérrez et al., 2021), CMIP6 models project monsoon rainfall amounts to increase by approximately 2.9% per degree of warming (Jin et al., 2020; Wang et al., 2020a), therefore, at higher levels of warming and towards the end of the century, a wetter monsoon is projected in the eastern Sahel medium 1 train +4960 AR6_WGII 1337 16 According to CMIP6 and CORDEX (Coordinated Regional Climate Downscaling Experiment) models, the annual average number of days with maximum temperature exceeding 35°C will increase between 14–27 days at GWL 2°C and 33–59 days at GWL 3°C above the 61–63 days for 1995 –2014 (Gutiérrez et al., 2021; Ranasinghe et al., 2021) high 2 train +4961 AR6_WGII 1337 21 There is some evidence of drying since the mid-20th century through decreased mean rainfall and increased precipitation deficits (Gutiérrez et al., 2021), as well as increases in meteorological, agricultural and ecological drought medium 1 train +4962 AR6_WGII 1338 3 At GWL 3°C and GWL 4.4°C, an increased mean annual rainfall of 10–25% is projected by regional climate models (Coppola et al., 2014; Pinto et al., 2015) and the intensity of extreme precipitation will increase high 2 train +4963 AR6_WGII 1339 5 Heavy rainfall events are projected to increase over the region at global warming of 2°C and higher high 2 train +4964 AR6_WGII 1339 15 Since the 1960s, decreasing precipitation trends have been detected over the South African winter rainfall region (high confidence) and the far eastern parts of South Africa low 0 train +4965 AR6_WGII 1339 19 Mean annual rainfall in the summer rainfall region of South Afria is projected to decrease by 10–20%, accompanied by an increase in the number of consecutive dry days during the rainy season under RCP8.5. The western parts of the region are projected to become drier, with increasing drought frequency, intensity and duration likely under RCP8.5 high 2 test +4966 AR6_WGII 1340 2 There is a projected decrease in the number of tropical cyclones making landfall in the region at 1°C, 2°C and 3°C of global warming, however, they are projected to become more intense with higher wind speeds so when they do make landfall the impacts are expected to be high medium 1 train +4967 AR6_WGII 1340 25 Reducing emissions and limiting warming to lower levels reduces risk to these systems high 2 train +4968 AR6_WGII 1341 4 In Africa, IKLK are exceptionally rich in ecosystem-specific knowledge, with the potential to enhance the management of natural hazards and climate variability high 2 train +4969 AR6_WGII 1344 18 Increased temperature, changes in rainfall and reduced wind speed altered the physical and chemical properties of inland water bodies, affecting water quality and productivity of algae, invertebrates and fish high 2 train +4970 AR6_WGII 1344 27 Mangroves, seagrasses and coral reefs support nursery habitats for fish, sequester carbon, trap sediment and provide shoreline protection (Ghermandi et al., 2019). Climate change is compromising these ecosystem services medium 1 test +4971 AR6_WGII 1345 11 The magnitude and extent of local extinctions predicted across Africa increase substantially under all future GWLs high 2 train +4972 AR6_WGII 1347 2 With increasing warming, there is a lower likelihood species can migrate rapidly enough to track shifting climates, increasing global extinction risk and biodiversity loss across more of Africa high 2 train +4973 AR6_WGII 1347 12 At 1.5°C of global warming, mangroves will be exposed to sedimentation and sea level rise, while seagrass ecosystems will be most affected by heat extremes high 2 train +4974 AR6_WGII 1347 14 Over 90% of east African coral reefs are projected to be destroyed by bleaching at 2°C of global warming very high 3 train +4975 AR6_WGII 1349 8 This declining carbon storage may be offset by CO 2 fertilization low 0 train +4976 AR6_WGII 1349 13 Climate change is projected to change patterns of invasive species spread high 2 train +4977 AR6_WGII 1349 21 Climate change is already negatively affecting tourism in Africa high 2 train +4978 AR6_WGII 1350 8 Climate change and land use change will interact to influence the effectiveness of African protected areas high 2 train +4979 AR6_WGII 1352 9 Maintaining existing indigenous forest and indigenous forest restoration is a win–win, maximising benefits to biodiversity, adaptation and mitigation (Griscom et al., 2017; Watson et al., 2018; Lewis et al., 2019) high 2 train +4980 AR6_WGII 1352 10 Yet many areas targeted by AFR100 erroneously mark Africa’s open ecosystems (grasslands, savannas, shrublands) as degraded and suitable for afforestation (Figure Box 9.3.1; (Veldman et al., 2015; Bond et al., 2019) high 2 train +4981 AR6_WGII 1353 1 General principles for this type of monitoring were developed for Lake Tanganyika (Plisnier et al., 2018) and could be applied to develop harmonised, regional monitoring of African lakes, rivers and wetlands (Tamatamah and Mwedzi, 2020) 9.6.4.3 Marine and Coastal Ecosystems Marine and coastal ecosystems such as mangroves, seagrass and coral reefs provide storm protection and food security for coastal communities high 2 train +4982 AR6_WGII 1353 18 However, in southern Africa as a whole, river flows have mostly decreased high 2 train +4983 AR6_WGII 1353 22 Observed climate changes in Africa (see Section 9.5) have led to changes in river flow and runoff (Dallas and Rivers-Moore, 2014; Wolski et al., 2014) and high fluctuations in lake levels high 2 train +4984 AR6_WGII 1355 15 In some higher altitude regions, like the Niger Inland Delta in west Africa, river flows and water levels are expected to increase medium 1 train +4985 AR6_WGII 1355 17 In the Volta River system, increasing wet season river flows (+36% by 2090s) and Volta lake outflow (+5% by 2090s) are anticipated under RCP8.5 medium 1 train +4986 AR6_WGII 1356 9 Changes in the amplitude, timing and frequency of extreme events such as droughts and floods will continue to affect lake levels, rates of river discharge and runoff and groundwater recharge high 2 train +4987 AR6_WGII 1361 10 Pressure on water demand due to climate change and variability is threatening income, development processes and food security in the region high 2 train +4988 AR6_WGII 1361 16 Extreme climate events have been key drivers in rising acute food insecurity and malnutrition of millions of people requiring humanitarian assistance in Africa high 2 train +4989 AR6_WGII 1361 18 Children and pregnant women experience disproportionately greater adverse health and nutrition impacts very high 3 train +4990 AR6_WGII 1362 27 Climate change is projected to have overall positive impacts on sugarcane and Bambara nuts in southern Africa, oil palm in Nigeria and chickpea in Ethiopia low 0 train +4991 AR6_WGII 1367 29 However, adaptation limits for crops in Africa are increasingly reached for global warming above 2°C (high confidence), and in tropical Africa may already be reached at current levels of global warming low 0 train +4992 AR6_WGII 1371 15 Approximately 59% of sub-Saharan Africa’s urban population resides in informal settlements (in some cities up to 80%), and the population in informal settlements is expected to increase very high 3 train +4993 AR6_WGII 1371 16 These urbanisation trends are compounding increasing exposure to climate hazards, particularly floods and heatwaves high 2 train +4994 AR6_WGII 1371 17 Globally, the highest rates of population growth and urbanisation are taking place in Africa’s coastal zones high 2 train +4995 AR6_WGII 1374 18 Africa’s large population of urban poor and marginalised groups and informal sector workers, further contribute to high vulnerability to extreme weather and climate change in many settlements high 2 train +4996 AR6_WGII 1374 19 Other non-climatic stressors which exacerbate vulnerabilities, especially in urban areas, include poor socioeconomic development, weak municipal governance, poor resource and institutional capacities, together with multi-dimensional, location-specific inequalities high 2 train +4997 AR6_WGII 1375 9 Future rapid coastal development is expected to increase existing high vulnerabilities to sea level rise (SLR) and coastal hazards, particularly in east Africa high 2 train +4998 AR6_WGII 1378 15 This portends increased vulnerability to risk of heat stress in big cities of central, east and west Africa very high 3 train +4999 AR6_WGII 1378 27 With increasing societal demands on limited water resources and future climate change, it is expected that there will be an intensification of WEF competition and trade-offs high 2 train +5000 AR6_WGII 1380 16 Modelling suggests that proactive adaptation of road designs to account for temperature increases is a ‘no regret’ option in all cases, but accounting for precipitation increases should be assessed on a case-by-case basis medium 1 train +5001 AR6_WGII 1380 19 These interventions simultaneously reduce the vulnerability of low- income residents to climate shocks, prevent lock-ins into carbon- intensive development pathways and reduce poverty high 2 train +5002 AR6_WGII 1380 20 The combined mitigation–adaptation interventions in the land use transport systems of African cities are also expected to have sufficient short-term co- benefits (reducing air pollution, congestion and traffic fatalities) to be ‘no regret’ investments very high 3 train +5003 AR6_WGII 1383 1 Climate change is already challenging the health and well-being of African communities, compounding the effects of underlying inequalities high 2 train +5004 AR6_WGII 1383 13 In east Africa, there has been an expansion of the Anopheles vector into higher altitudes (Gone et al., 2014; Carlson et al., 2019) and increasing incidence of infection with Plasmodium falciparum with higher temperatures high 2 train +5005 AR6_WGII 1383 22 In east and southern Africa and the Sahel, malaria vector hotspots and prevalence are projected to increase under RCP4.5 and RCP8.5 by 2030 (1.5°C–1.7°C global warming) high 2 train +5006 AR6_WGII 1383 27 With continued GHG emissions, these include: west Africa by 2030 (1.7°C global warming) (high confidence) (Yamana et al., 2016; Semakula et al., 2017b; Ryan et al., 2020), parts of southern central Africa and dryland regions in east Africa by 2050 (2.5°C global warming) high 2 train +5007 AR6_WGII 1384 3 The ENSO cycle currently contributes to seasonal epidemic malaria in epidemic-prone areas (high confidence), and is projected to shift the malaria epidemic fringe southward and into higher altitudes by mid- to end-century high 2 train +5008 AR6_WGII 1386 17 Emerging and future pandemic threats Future influenza pandemics are highly likely, as are regional epidemics and pandemics of novel zoonotic viruses (including coronaviruses and flaviviruses) high 2 train +5009 AR6_WGII 1394 12 Contextualised risk studies on local drivers of transmission are still lacking and present a major gap in developing appropriate adaptation strategies high 2 train +5010 AR6_WGII 1398 19 Substantial avoided economic damages to African countries are projected from ambitious, near-term global mitigation limiting global warming well below 2°C above pre-industrial levels high 2 train +5011 AR6_WGII 1399 13 Nevertheless, climate change impacts on poverty in Africa will depend on how socioeconomic development unfolds over the coming decades medium 1 train +5012 AR6_WGII 1400 2 In rural Africa, poor and female-headed households face greater livelihood risks from climate hazards high 2 train +5013 AR6_WGII 1402 8 Urbanisation in Africa is affected by climate conditions in rural agricultural areas high 2 train +5014 AR6_WGII 1402 14 Migration is an important and potentially effective climate change adaptation strategy in Africa and must be considered in adaptation planning high 2 train +5015 AR6_WGII 1470 1 Significant warming has intensified the threat to social and economic sustainability medium 1 train +5016 AR6_WGII 1470 2 Rising temperatures increase the likelihood of the threat of heatwaves across Asia, droughts in arid and semiarid areas of West, Central and South Asia, delays and weakening of the monsoon circulation in South Asia, floods in monsoon regions in South, Southeast and East Asia, and glacier melting in the Hindu Kush Himalaya region medium 1 train +5017 AR6_WGII 1470 4 Decrease in precipitation influences energy demand as well as desalination, underground water pumping and other energy- intensive methods are increasingly used for water supply high 2 train +5018 AR6_WGII 1470 6 Among 13 developing countries with large energy consumption in Asia, 11 are exposed to high-energy insecurity and industrial-systems risk high 2 train +5019 AR6_WGII 1470 9 This can be explained by site-specific complex interaction of positive effect of warming on tree growth, drought stress, change in snow precipitation, land-use change (especially grazing) and other factors high 2 train +5020 AR6_WGII 1470 10 The increased considerable changes in biomes in Asia are a response to warming medium 1 train +5021 AR6_WGII 1470 12 Climate change, human activity and lightning have caused the increase in wildfire severity and area burned in North Asia after the 1990s medium 1 train +5022 AR6_WGII 1470 13 Length of plant growth season has increased in some parts of East and North Asia, while the opposite trend, or no change at all, has been observed in other parts high 2 train +5023 AR6_WGII 1470 14 Observed biodiversity or habitat losses of animals plants have been linked to climate change in some parts of Asia high 2 train +5024 AR6_WGII 1470 15 There is evidence that climate change can alter species interaction or spatial distribution of invasive species in Asia high 2 train +5025 AR6_WGII 1470 17 Across Asia, under a range of representative concentration pathways and other scenarios, rising temperatures are expected to contribute to a northward shift of biome boundaries and an upwards shift of mountain treeline medium 1 train +5026 AR6_WGII 1470 20 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.Coastal habitats of Asia are diverse, and the impacts of climate change including rising temperatures, ocean acidification and sea level rise (SLR) has brought negative effects to the services and the livelihoods of people depending on it high 2 train +5027 AR6_WGII 1470 22 The risk of irreversible loss of coral reefs, tidal marshes, seagrass meadows, plankton community and other marine and coastal ecosystems increases with global warming, especially at 2°C temperature rise or more high 2 train +5028 AR6_WGII 1470 26 By mid- 21st Century, the international transboundary river basins of Amu Darya, Indus, Ganges could face severe water scarcity challenges due to climatic variability and changes acting as stress multipliers high 2 train +5029 AR6_WGII 1470 27 Due to global warming, Asian countries could experience an increase in drought conditions (5–20%) by the end of this century high 2 train +5030 AR6_WGII 1470 31 The total amount and area of glacier lakes have increased during the past decade high 2 train +5031 AR6_WGII 1470 33 Glacier lake outburst flood (GLOF) will threaten the securities of the local and downstream communities high 2 train +5032 AR6_WGII 1470 34 Snowmelt water contributed 19% of the increase change in runoff of arid regions’ rivers in Xinjiang, China, and 10.6% of the upper Brahmaputra River during 2003–2014 medium 1 train +5033 AR6_WGII 1470 37 Climate-related risks to agriculture and food systems in Asia will progressively escalate with the changing climate, with differentiated impacts across the region medium 1 train +5034 AR6_WGII 1471 8 Coastal cities, especially in South and Southeast Asia, are expected to see significant increases in average annual economic losses between 2005 and 2050 due to flooding, with very high losses in East Asian cities under the high- emissions scenario high 2 train +5035 AR6_WGII 1471 10 Under the high-emissions scenario, higher risks from extreme temperature and precipitation are projected for almost all cities medium 1 train +5036 AR6_WGII 1471 11 By 2080, 940 million to 1.1 billion urban dwellers in South and Southeast Asia could be affected by extreme heat lasting more than 30 d yr–1 high 2 train +5037 AR6_WGII 1471 13 At higher warming, key infrastructures, such as power lines, transport by roads and railways, and built infrastructures, such as airports and harbours, are more exposed to climate-induced extreme events, especially in coastal cities medium 1 train +5038 AR6_WGII 1471 15 Adaptation actions tend to be in the initial stages and more reactive (57% of urban adaptations focus on preparatory interventions, such as capacity building, and 43% of cities report implemented adaptation interventions) medium 1 test +5039 AR6_WGII 1471 16 The degree of implementation of urban adaptation is uneven with large cities receiving more funding and priority, and smaller cities and towns, and peri-urban spaces, seeing relatively lower adaptation action medium 1 train +5040 AR6_WGII 1471 19 Under the medium-to-high emissions scenario, rising temperatures and extreme climate events will have an increasing impact on human health and well-being with varying types and magnitudes of impact across Asia high 2 train +5041 AR6_WGII 1471 28 Factors motivating adaptation actions include risk perception, perceived self-efficacy, sociocultural norms and beliefs, previous experiences of impacts, levels of education and awareness high 2 train +5042 AR6_WGII 1471 31 Non-material losses and damages are reported to a lesser degree, but this is due to under-reporting and methodological issues with detection and attribution to climate change high 2 train +5043 AR6_WGII 1472 2 Climate risks, vulnerability and adaptation measures need to be factored into decision making across all levels of governance high 2 train +5044 AR6_WGII 1472 4 More accurate forecasting of extreme events, risk awareness and prioritising individual and collective decision making also need to be addressed high 2 train +5045 AR6_WGII 1472 7 Some Asian countries and regions offer solutions to overcome these barriers: through use of advanced technologies (in situ observation and remote sensing, a variety of new sensor technologies, citizen science, artificial intelligence and machine learning tools); regional partnerships and learning; improved forecasting capabilities; and better risk awareness high 2 train +5046 AR6_WGII 1474 16 Surface temperature has increased in the past century all over Asia very high 3 train +5047 AR6_WGII 1474 17 Elevation-dependent warming (i.e., the warming rate is different across elevation bands is observed in HMA) medium 1 train +5048 AR6_WGII 1475 1 Large increases in temperature extremes are observed in West and Central Asia high 2 test +5049 AR6_WGII 1475 4 In 2016 and 2018, extreme warmth was observed in Asia for which an event-attribution study revealed that this would not have been possible without anthropogenic global warming medium 1 train +5050 AR6_WGII 1475 5 There are considerable regional differences in observed annual pre- cipitation trend medium 1 train +5051 AR6_WGII 1475 6 Observations show a decreas- ing trend of the South Asian summer monsoon precipitation during the second half of the 20th century high 2 train +5052 AR6_WGII 1475 8 Increase in heavy precipita- tion occurred recently in South Asia (high confidence), and in South- east and East Asia medium 1 train +5053 AR6_WGII 1476 1 Annual surface wind speeds have been decreasing in Asia since the 1950s high 2 train +5054 AR6_WGII 1476 2 The observed changes in the frequency of sand and dust storms vary from region to region in Asia medium 1 train +5055 AR6_WGII 1476 5 In contrast, West Asia has witnessed more frequent and intensified dust storms affecting Iran and Persian Gulf countries in recent decades medium 1 train +5056 AR6_WGII 1476 6 There is no significant long-term trend during 1951–2017 in the numbers of tropical cyclones (TCs) with maximum winds of 66.37 km h–1 or higher forming in the western North Pacific and the South China Sea medium 1 train +5057 AR6_WGII 1476 12 There has been a significant northwestward shift in TC tracks since the 1980s, and a detectable poleward shift since the 1940s in the average latitude where TCs reach their peak intensity in the western North Pacific medium 1 train +5058 AR6_WGII 1476 15 Their report also summarised that there is increased agreement between coupled model simulations of anthropogenic climate change and observations of changes in ocean heat content high 2 train +5059 AR6_WGII 1476 18 Ocean acidification continues with surface seawater pH values having shown a clear decrease by 0.01–0.09 from 1981–2011 along the Pacific coasts of Asia high 2 train +5060 AR6_WGII 1476 24 This Report also uses the term ‘likely range’ to indicate that the assessed likelihood of an outcome lies within the 17–83% probability range.10.3.1.2 Projected Climate Change Rising temperatures increase the likelihood of the threat of heatwaves across Asia, droughts in arid and semiarid areas of West, Central and South Asia, floods in monsoon regions in South, Southeast and East Asia, and glacier melting in the HKH region high 2 train +5061 AR6_WGII 1476 26 Projections of future changes in annual mean surface air temperature in Asia are qualitatively similar to those in the previous assessments with greater warming at higher latitudes (i.e., North Asia) high 2 train +5062 AR6_WGII 1476 27 Projected surface air temperature changes in the Tibetan Plateau, Central Asia and West Asia are also significant high 2 train +5063 AR6_WGII 1476 28 The highest levels of warming for extremely hot days are expected to occur in West and Central Asia with increased dryness of land high 2 train +5064 AR6_WGII 1476 29 Over mountainous regions, elevation-dependent warming will continue medium 1 train +5065 AR6_WGII 1476 30 Glaciers will generally shrink, but rates will vary among regions high 2 train +5066 AR6_WGII 1476 32 Temperature rise will be strongest in winter in most regions, while it will be the strongest on summer in the northern part of West Asia and some parts of South Asia where a desert climate prevails high 2 train +5067 AR6_WGII 1476 33 The wet-bulb globe temperature, which is a measure of heat stress, is likely2 to approach critical health thresholds in West and South Asia under the RCP4.5 scenario, and in some other regions, such as East Asia, under the RCP8.5 scenario high 2 train +5068 AR6_WGII 1476 35 Projections show that a sizeable part of South Asia will experience heat stress conditions in the future high 2 train +5069 AR6_WGII 1476 38 A very likely large percentage increase in annual precipitation is projected in South and North Asia high 2 train +5070 AR6_WGII 1476 39 Precipitation is projected to decrease over the northwest part of the Arabian Peninsula and increase over its southern part medium 1 train +5071 AR6_WGII 1476 40 Both heavy and intense precipitation are projected to intensify and become more frequent in South, Southeast and East Asia high 2 train +5072 AR6_WGII 1477 2 Monsoon land precipitation likely will increase in East, Southeast and South Asia mainly due to increasing moisture convergence by elevated temperature high 2 train +5073 AR6_WGII 1477 3 Increasing land–sea thermal contrast and resultant lower tropospheric circulation changes, together with increasing moisture, are projected to intensify the South Asian summer monsoon precipitation medium 1 train +5074 AR6_WGII 1477 5 Monsoonal winds will generally become weaker in a future warming world with different magnitudes across regions medium 1 train +5075 AR6_WGII 1477 9 Models suggest a reduction in TC frequency but an increase in the proportion of very intense TCs over the western North Pacific in the future; however, some individual studies project an increase in western North Pacific TC frequency medium 1 train +5076 AR6_WGII 1477 10 In the western North Pacific, some models project a poleward expansion of the latitude of maximum TC intensity, leading to a future increase in intense TC frequency south of Japan medium 1 train +5077 AR6_WGII 1481 27 With global warming, the energy consumption for heating in winter decreases, while the energy consumption for cooling in summer significantly increases, but the overall energy demand shows an upwards trend high 2 train +5078 AR6_WGII 1482 16 Since the 1960s, the total solar radiation on the ground in Asia has shown a downwards trend as a whole, which is consistent with the change in global total solar radiation on the ground, and has experienced a phased change process of ‘first darkening and then brightening’ high 2 train +5079 AR6_WGII 1482 18 However, wind speed over most Asian regions is obviously decreasing high 2 train +5080 AR6_WGII 1482 25 At the same time, with the increase in the proportion of renewable energy in the power system, the power system will be more vulnerable to climate change and extreme weather and climate events, and the vulnerability and risk of the power system will greatly increase medium 1 train +5081 AR6_WGII 1483 32 Alpine treeline position in Asian mountains in recent decades either moves upwards in North Asia or demonstrates multi-directional shifts in Himalaya high 2 train +5082 AR6_WGII 1484 24 The observed loss of biodiversity and habitat of animals and plants has been linked to climate change in some parts of Asia high 2 train +5083 AR6_WGII 1484 28 There is evidence that climate change can alter species interaction or spatial distribution of invasive species in Asia high 2 train +5084 AR6_WGII 1485 32 In Asia, the date of arrival of migrant birds to nesting areas and the date of departure from winter areas are changing consistently with climate change medium 1 train +5085 AR6_WGII 1487 3 Future climate change would cause biodiversity and habitat loss in many parts of Asia using modelling approaches high 2 train +5086 AR6_WGII 1487 10 The impact of future climate change on invasive species may be species- or region specific medium 1 train +5087 AR6_WGII 1489 8 The risk of irreversible loss of many marine and coastal ecosystems increases with global warming, especially at 2°C or more high 2 train +5088 AR6_WGII 1495 21 By mid-21st Century, the international transboundary river basins of Amu Darya, Indus, Ganges could face severe water scarcity challenges due to climatic variability and changes acting as stress multipliers high 2 train +5089 AR6_WGII 1496 16 The changes in snowmelt water can explain 19% of the variations in rivers of arid regions like Xinjiang, China (Bai et al., 2018) (medium confidence), and the 10.6% of the runoff of the upper Brahmaputra River was contributed by snow during 2003–2014 (Chen et al., 2017c) medium 1 train +5090 AR6_WGII 1497 17 Bhambri et al., 2017; Mukherjee et al., 2017; Ding et al., 2018), threatening the security of the local and down streaming societies (high confidence).The total amount and area of glacier lakes increased during last decade (Zhang et al., 2015; Chen et al., 2017c) high 2 train +5091 AR6_WGII 1498 14 Researchers have found that the southern Tibetan Plateau has been consistently melting from 1998–2007 and is projected to continue melting until 2050 (Lutz et al., 2014b) high 2 train +5092 AR6_WGII 1499 6 Increase in extreme precipitation events is likely to cause more flash-flood events in the future medium 1 test +5093 AR6_WGII 1511 18 By 2050, it is likely that 69% of fundamental human infrastructure in the Pan Arctic will be at risk (RCP 4.5 scenario) medium 1 test +5094 AR6_WGII 1515 17 In many Asian cities, land subsidence control can serve as an adaptation strategy since it is estimated to significantly reduce relative SLR high 2 train +5095 AR6_WGII 1518 2 In addition to all-cause mortality (Dang et al., 2016; Chen et al., 2018e), deaths related to circulatory, respiratory, diabetic (Li et al., 2014b) and infectious diseases (Ingole et al., 2015), as well as infant mortality (Son et al., 2017), are increased with high temperature high 2 train +5096 AR6_WGII 1518 3 Increased hospital admissions (Giang et al., 2014; Lin et al., 2019) and ambulance transport (Onozuka and Hagihara, 2015) coincide with increased ambient temperature high 2 train +5097 AR6_WGII 1518 6 Individuals with lower degrees of education and socioeconomic status, older individuals and individuals living in communities with less green space are more susceptible to heat-related mortality high 2 train +5098 AR6_WGII 1518 7 These heat effects have been attenuating over recent decades in East Asian countries, although the driving force behind this remains unknown high 2 train +5099 AR6_WGII 1518 8 Rising ambient temperature accelerates pollutant formation reactions and may modify air-pollution-related health effects medium 1 test +5100 AR6_WGII 1518 11 Mortality and hospital admissions for circulatory and respiratory diseases are increased after exposures to Asian dust events high 2 train +5101 AR6_WGII 1518 14 Ambient temperature is associated with the risk of an outbreak of mosquito-borne disease in South and Southeast Asia high 2 train +5102 AR6_WGII 1519 1 The duration and survival rate of dengue mosquito development, mosquito density, mosquito biting activity, mosquito spatio-temporal range and distribution, and mosquito flying distance are all affected by temperature high 2 train +5103 AR6_WGII 1519 2 Temperature, precipitation, humidity and air pressure are major weather factors associated with dengue fever transmission high 2 train +5104 AR6_WGII 1519 3 Climate change alters the hydrological cycle by increasing the frequency of extreme weather events such as excess precipitation, storm surges, floods and droughts high 2 train +5105 AR6_WGII 1519 4 Water-borne diseases, such as diarrhoea, leptospirosis and typhoid fever, can increase in incidence following heavy rainfall, tropical cyclones and flooding events high 2 train +5106 AR6_WGII 1519 9 Crop destruction due to tropical cyclones can include salt damage from tides blowing inland medium 1 train +5107 AR6_WGII 1519 14 Extreme weather events, such as storms, floods, hurricanes and cyclones, increase injuries and mental disorders (post- traumatic stress disorder and depressive disorders) (Rataj et al., 2016), thereby negatively affecting well-being high 2 train +5108 AR6_WGII 1519 16 The association between variations in ambient temperature and the occurrence of asthma has been reported in several Asian countries/regions such as Japan (Yamazaki et al., 2015), the Republic of Korea (Kwon et al., 2016), China (Li et al., 2016a) and Hong Kong SAR of China medium 1 train +5109 AR6_WGII 1520 3 South and Southeast Asia are projected to be among the highest-risk regions for reduced dietary iron intake among women of childbearing age and children under five years due to elevated CO 2 concentrations medium 1 train +5110 AR6_WGII 1523 19 For nearly five decades, integrated coastal management (ICM), advocated by several international organisations (e.g., IMO, UNEP , WHO, FAO) and adopted by over 100 countries, has been acknowledged as a holistic coastal governance approach aimed at achieving coastal sustainability and reducing the vulnerability of coastal communities in the face of multiple environmental impacts high 2 train +5111 AR6_WGII 1537 24 Some Asian countries and regions offer solutions to overcome these barriers: through use of advanced technologies (in situ observation and remote sensing, a variety of new sensor technologies, citizen science, AI and machine learning tools); regional partnerships and learning; improved forecasting capabilities; and better risk awareness high 2 train +5112 AR6_WGII 1594 28 Ongoing warming is projected, with more hot days and fewer cold days very high 3 train +5113 AR6_WGII 1594 29 Further sea level rise (SLR), ocean warming and ocean acidification are projected very high 3 train +5114 AR6_WGII 1594 30 Less winter and spring rainfall is projected in southern Australia, with more winter rainfall in Tasmania, less autumn rainfall in southwestern Victoria and less summer rainfall in western Tasmania medium 1 train +5115 AR6_WGII 1594 31 In New Zealand, more winter and spring rainfall is projected in the west and less in the east and north, with more summer rainfall in the east and less in the west and central North Island medium 1 train +5116 AR6_WGII 1594 32 In New Zealand, ongoing glacier retreat is projected very high 3 train +5117 AR6_WGII 1594 33 More extreme fire weather is projected in southern and eastern Australia (high confidence) and over northern and eastern New Zealand medium 1 train +5118 AR6_WGII 1594 34 Increased drought frequency is projected for southern and eastern Australia and northern New Zealand medium 1 train +5119 AR6_WGII 1594 35 Increased heavy rainfall intensity is projected, with fewer tropical cyclones and a greater proportion of severe cyclones medium 1 train +5120 AR6_WGII 1595 2 Three marine heatwaves on the Great Barrier Reef (GBR) during 2016–2020 caused significant bleaching and loss very high 3 test +5121 AR6_WGII 1595 4 For example loss of alpine vegetation communities (snow patch Feldmark and short alpine herb-fields) and increased stress on snow-dependent plant and animal species high 2 train +5122 AR6_WGII 1595 6 For example declining rainfall in southern Australia over the past 30 years, has led to drought- induced canopy dieback across a range of forest and woodland types and death of fire-sensitive tree species due to unprecedented wildfires high 2 train +5123 AR6_WGII 1595 8 For example less than 10% of giant kelp in Tasmania was remaining by 2011 due to ocean warming high 2 train +5124 AR6_WGII 1595 10 For example for 0.5 m sea level rise (SLR), the value of buildings in New Zealand exposed to 1-in-100-year coastal inundation could increase by NZ$12.75 billion and the current 1-in-100-year flood in Australia could occur several times a year high 2 train +5125 AR6_WGII 1595 12 For example by 2050, a decline in median wheat yields of up to 30% in southwestern Australia and up to 15% in South Australia and increased heat stress in livestock by 31–42 days per year high 2 train +5126 AR6_WGII 1595 14 For example heat-related excess deaths in Melbourne, Sydney and Brisbane are projected to increase by about 300/year (low emission pathway) to 600/year (high emission pathway) during the 2031–2080 period relative to 142/year in the period 1971–2020 high 2 train +5127 AR6_WGII 1595 16 For example in New Zealand, extreme snow, heavy rainfall and wind events have combined to impact road networks, power and water supply, interdependent wastewater and stormwater services and business activities high 2 train +5128 AR6_WGII 1595 18 For example the scale and scope of projected climate impacts overwhelm the capacity of institutions, organisations and systems to provide necessary policies, services, resources and coordination to address socioeconomic impacts high 2 train +5129 AR6_WGII 1595 19 There are important interactions between mitigation and adaptation policies and their implementation high 2 train +5130 AR6_WGII 1597 7 Existing vulnerabilities expose and exacerbate inequalities between rural, regional and urban areas, Indigenous and non-Indigenous Peoples, those with health and disability needs, and between generations, incomes and health status, increasing the relative climate change risk faced by some groups and places high 2 train +5131 AR6_WGII 1597 11 In the absence of proactive adaptation, climate change impacts are projected to worsen inequalities between Indigenous and non-Indigenous peoples and other vulnerable groups (Green et al., 2009; Manning et al., 2014; Ambrey et al., 2017) high 2 train +5132 AR6_WGII 1597 14 Changes to the composition and location of different demographic groups in the region contribute to increased exposure or vulnerability to climate change medium 1 train +5133 AR6_WGII 1598 1 Conclusions Report Our regional climate is changing very high 3 train +5134 AR6_WGII 1598 3 Benefits are projected for some sectors and locations high 2 train +5135 AR6_WGII 1598 4 Adaptation is occurring and becoming mainstreamed in some planning processes high 2 train +5136 AR6_WGII 1598 5 Adaptive capacity is considered generally high in many human systems, but adaptation implementation faces major barriers, especially for transformational responses high 2 train +5137 AR6_WGII 1598 6 Some synergies and trade-offs exist between different adaptation responses and between mitigation and adaptation, with interactions occurring both within and outside the region very high 3 test +5138 AR6_WGII 1598 7 Vulnerability remains uncertain due to incomplete consideration of socioeconomic dimensions very high 3 train +5139 AR6_WGII 1601 5 Further climate change is inevitable, with the rate and magnitude largely dependent on the emission pathway very high 3 train +5140 AR6_WGII 1602 4 Further warming is projected, with more hot days, fewer cold days, reduced snow cover, ongoing sea level rise (SLR) and ocean acidification very high 3 train +5141 AR6_WGII 1602 5 Winter and spring rainfall and soil moisture are projected to decrease, with higher evaporation rates, decreased wind over southern mainland Australia, increased wind over Tasmania, and more extreme fire weather in southern and eastern Australia high 2 train +5142 AR6_WGII 1602 6 Heavy rainfall intensity is projected to increase, with more droughts over southern and eastern Australia medium 1 train +5143 AR6_WGII 1602 7 Increased winter rainfall is projected over Tasmania, with decreased rainfall in southwestern Victoria in autumn and in western Tasmania in summer, fewer tropical cyclones with a greater proportion of severe cyclones and decreased soil moisture in the north medium 1 train +5144 AR6_WGII 1602 8 Hailstorm frequency may increase low 0 train +5145 AR6_WGII 1602 10 Further warming is projected, with more hot days, fewer cold days, less snow and glacial ice, ongoing sea level rise (SLR) and ocean acidification very high 3 train +5146 AR6_WGII 1602 11 Increases in winter and spring rainfall are projected in the west of the North and South Islands, with drier conditions in the east and north, caused by stronger westerly winds medium 1 train +5147 AR6_WGII 1602 12 In summer, wetter conditions are projected in the east of both islands, with drier conditions in the west and central North Island medium 1 train +5148 AR6_WGII 1602 13 Fire weather indices are projected to increase over northern and eastern New Zealand medium 1 train +5149 AR6_WGII 1602 14 Heavy rainfall intensity is projected to increase over most regions, with increased extreme wind speeds in eastern regions, especially in Marlborough and Canterbury, and reduced relative humidity almost everywhere, except for the west coast in winter medium 1 train +5150 AR6_WGII 1602 15 Drought frequency may increase in the north medium 1 train +5151 AR6_WGII 1602 18 Risk is considered in terms of vulnerability, hazards (impact driver), exposure, reasons for concern and complex and cascading risks (Chapter 1; Figure 1.2).11.3.1 Terrestrial and Freshwater Ecosystems 11.3.1.1 Observed Impacts Widespread and severe impacts on ecosystems and species are now evident across the region very high 3 train +5152 AR6_WGII 1602 25 Natural forest and woodland ecosystem processes are experiencing differing impacts and responses depending on the climate zone high 2 train +5153 AR6_WGII 1605 5 In southern Australia, some forest ecosystems (alpine ash, snow gum woodland, pencil pine, northern jarrah) are projected to transition to a new state or collapse due to hotter and drier conditions with more fires high 2 train +5154 AR6_WGII 1605 8 Most New Zealand native plants are not fire resistant and are projected to be replaced by fire-resistant introduced species following climate-change- related fires (Perry et al., 2014).A loss of alpine biodiversity in the southeast Australian Alps bioregion is projected in the near-term as a result of less snow on snow patch feldmark and short alpine herb fields as well as increased stress on snow- dependent plant and animal species high 2 train +5155 AR6_WGII 1605 9 In Australia, invasive plants’ and weeds’ response rates are expected to be faster than for native species, and climate change could foster the appearance of a new set of weed species, with many bioregions facing increased impacts from non-native plants medium 1 train +5156 AR6_WGII 1605 10 In New Zealand, climate change is projected to enable invasive species to expand to higher elevations and southwards medium 1 train +5157 AR6_WGII 1605 15 Improved coastal modelling, experiments and in situ studies are reducing uncertainties at a local scale about the impact of future sea level rise (SLR) on coastal freshwater terrestrial wetlands medium 1 train +5158 AR6_WGII 1607 10 Best practice conservation adaptation planning is informed by data on key habitats, including refugia, and restoration that facilitates species movements and employs adaptive pathways very high 3 train +5159 AR6_WGII 1607 14 Adaptation planning for ecosystems and species requires monitoring and evaluation to identify trigger points and thresholds for new actions to be implemented high 2 train +5160 AR6_WGII 1609 2 There remains a gap between the knowledge generated, potential adaptation strategies and their incorporation into conservation instruments medium 1 train +5161 AR6_WGII 1610 2 Increased fire activity in southeast Australia associated with climate change has been observed since 1950 (Abram et al., 2021), though trends vary regionally medium 1 train +5162 AR6_WGII 1610 3 In New Zealand, there has been an increased frequency of major wildfires in plantations (FENZ, 2018) and at the rural–urban interface medium 1 train +5163 AR6_WGII 1610 5 In Australia, the frequency and severity of dangerous fire weather conditions is increasing, with partial attribution to climate change very high 3 train +5164 AR6_WGII 1610 9 Fire weather is projected to increase in frequency, severity and duration for southern and eastern Australia (high confidence) and most of New Zealand medium 1 train +5165 AR6_WGII 1613 8 Ocean carbon storage and acidification has led to decreased surface pH in the region (Table 11.2), including the sub-Antarctic waters off the East Coast of New Zealand’s South Island very high 3 train +5166 AR6_WGII 1613 9 The depth of the Aragonite Saturation Horizon has shallowed by 50–100 m over much of New Zealand, which may limit and/or increase the energetic costs of growth of calcifying species low 0 test +5167 AR6_WGII 1613 11 Extensive changes in the life history and distribution of species have been observed in Australia’s (very high confidence) (Gervais et al., 2021) and New Zealand’s marine systems medium 1 train +5168 AR6_WGII 1613 15 Extreme climatic events in Australia from 2011 to 2017 led to abrupt and extensive mortality of key habitat-forming organisms — corals, kelps, seagrasses and mangroves — along over 45% of the continental coastline of Australia high 2 train +5169 AR6_WGII 1613 16 In 2016 and 2017, the GBR experienced consecutive occurrences of the most severe coral bleaching in recorded history very high 3 train +5170 AR6_WGII 1613 19 Coral reefs in Australia are at very high risk of continued negative effects on ecosystem structure and function (very high confidence) (Hughes et al., 2019b), cultural well-being (very high confidence) (Goldberg et al., 2016; Lyons et al., 2019), food provision (medium confidence) (Hoegh-Guldberg et al., 2017), coastal protection (high confidence) (Ferrario et al., 2014) and tourism high 2 train +5171 AR6_WGII 1614 1 Acidification may alter sex determination (e.g., in the oyster Saccostrea glomerate), resulting in changes in sex ratios (Parker et al., 2018), and may thus affect reproductive success low 0 test +5172 AR6_WGII 1616 3 The GBR is already severely impacted by climate change, particularly ocean warming, through more frequent and severe coral bleaching very high 3 train +5173 AR6_WGII 1616 7 In 2017, the central third of the reef was the most severely affected and the back-to-back regional-scale bleaching events has led to an unprecedented shift in the composition of GBR coral assemblages, transforming the northern and middle sections of the reef system (Hughes et al., 2018c) to a highly degraded state very high 3 train +5174 AR6_WGII 1616 10 Increased heat exposure also affects the abundance and distribution of associated fish, invertebrates and algae high 2 train +5175 AR6_WGII 1616 17 Recovery of coral reefs following repeated disturbance events is slow (Hughes et al., 2019b; IPCC, 2019b), and it takes at least a decade after each bleaching event for the very fastest growing corals to recover high 2 train +5176 AR6_WGII 1616 18 Estimates of future levels of thermal stress, measured as degree heating months, which incorporates both the magnitude and duration of warm season SST anomalies, suggest that achieving the 1.5°C Paris Agreement target would be insufficient to prevent more frequent mass bleaching events very high 3 train +5177 AR6_WGII 1618 2 The pH of surface waters around New Zealand is projected to decline by 0.33 under RCP 8.5 by 2090 (Tait et al., 2016), and the depth at which carbonate dissolves is projected to be significantly shallower (Mikaloff-Fletcher et al., 2017), affecting the distribution of some species of calcifying cold water corals medium 1 train +5178 AR6_WGII 1619 12 The runoff decline in southern Australia is projected to be further accentuated by higher temperature and potential evapotranspiration (Potter and Chiew, 2011; Chiew et al., 2014), transpiration from tree regrowth following more frequent and severe wildfires (Brookhouse et al., 2013) (Box 11.1), interceptions from farm dams (Fowler et al., 2015) and reduced surface–groundwater connectivity (limiting groundwater discharge to rivers) in long dry spells high 2 train +5179 AR6_WGII 1619 15 Groundwater recharge across southern Australia has decreased in recent decades (Fu et al., 2019), and this trend is expected to continue high 2 train +5180 AR6_WGII 1619 16 Climate change is also projected to impact water quality in rivers and water bodies, particularly through higher temperature and low flows (Jöhnk et al., 2008) (Box 11.5) and increased sediment and nutrient load following wildfires high 2 train +5181 AR6_WGII 1619 21 Adaptation initiatives and mechanisms, like significant government investment to enhance the Bureau of Meteorology online water information (Vertessy, 2013; BoM, 2016), funding to improve agricultural water use and irrigation efficiency (Koech and Langat, 2018), enhanced supply through inter-basin transfers and upgrading water infrastructure and an active water trading market (Wheeler et al., 2013; Kirby et al., 2014; Grafton et al., 2016) are helping to buffer regional systems against droughts and facilitating some adaptation to climate change medium 1 train +5182 AR6_WGII 1619 25 The impact of declining water resources on agricultural, ecosystems and communities in southeastern Australia would escalate with ongoing climate change medium 1 train +5183 AR6_WGII 1621 2 In New Zealand, many water supplies are at risk from drought, extreme rainfall events and sea level rise (SLR), exacerbated by underinvestment in existing water infrastructure (in part due to funding constraints) and urban densification high 2 train +5184 AR6_WGII 1621 16 Climate change is projected to substantially reduce water resources in the MDB high 2 train +5185 AR6_WGII 1621 17 This reduction, plus increased demand for water in hot and dry conditions, would increase the already intense competition for water high 2 train +5186 AR6_WGII 1622 2 Despite contestation, the reforms have resulted in some substantive achievements, including returning an equivalent of about one-fifth of consumptive water to the environment through the purchase of irrigation water entitlements and infrastructure projects medium 1 train +5187 AR6_WGII 1622 4 Reform initiatives such as water markets, improving agriculture water use efficiency (Koech and Langat, 2018), and increasing environmental water are helping buffer the system against droughts medium 1 train +5188 AR6_WGII 1622 8 The intense drought conditions in 2017–2019 (BoM, 2021b), the South Australian Royal Commission investigation into the MDB reforms (SA Government, 2019b) and major fish kills in the lower Darling River in the summer of 2018/2019 (AAS, 2019; Vertessy et al., 2019) have increased concerns about the Basin Plan’s climate adaptation deficit medium 1 train +5189 AR6_WGII 1623 7 Extreme rainfall intensity in northern Australia and New Zealand has been increasing, particularly for shorter (sub-daily) duration and more extreme high rainfall high 2 train +5190 AR6_WGII 1623 9 Extreme rainfall is projected to become more intense high 2 train +5191 AR6_WGII 1623 11 In urban areas, extreme rainfall intensity is projected to increase pluvial flood risk high 2 train +5192 AR6_WGII 1623 14 Modelling studies project increases in flood magnitudes in northern and eastern Australia and in western and northern New Zealand high 2 train +5193 AR6_WGII 1623 17 There is some recognition of the need for flood management and planning to adapt to climate change medium 1 train +5194 AR6_WGII 1623 23 Adaptation to changing flood risks is currently mostly reactive and incremental in response to flood and heavy rainfall events high 2 train +5195 AR6_WGII 1623 25 Nevertheless, adaptation planning that is pre-emptive and incorporates uncertainties into flood projections is emerging medium 1 train +5196 AR6_WGII 1624 6 Impacts resulting from climate change are observed across sectors and the region high 2 train +5197 AR6_WGII 1624 12 In the longer term, transformative adaptation, including land use change, will be required (Cradock-Henry et al., 2020a), both as a result of sectoral adaptations and mitigation medium 1 train +5198 AR6_WGII 1624 14 Future adaptive capacity may be limited by declining institutional and community capacity resulting from high debt, unavailability of insurance, increasing regulatory requirements and funding mechanisms that lock in ongoing exposure to climate risk, creating mental health impacts (Rickards et al., 2014; Wiseman and Bardsley, 2016; McNamara and Buggy, 2017; McNamara et al., 2017; Moyle et al., 2017; Robinson et al., 2018; Ma et al., 2020; Yazd et al., 2020).11.3.4.1 Field Crops and Horticulture 11.3.4.1.1 Observed impacts Drought, heat and frost in recent decades have shown the vulnerability of Australian field crops and horticulture to climate change (Cai et al., 2014; Howden et al., 2014; CSIRO and BOM, 2015; Lobell et al., 2015; Hughes and Lawson, 2017; King et al., 2017; Webb et al., 2017; Harris et al., 2020) as recognised by policymakers (CoA, 2019a) high 2 train +5199 AR6_WGII 1624 26 Winegrape maturity is projected to occur earlier due to warmer temperatures high 2 train +5200 AR6_WGII 1625 17 Climate change is projected to shift agro-ecological zones high 2 train +5201 AR6_WGII 1625 21 Shifts in agro-ecological zones present some opportunities, for example warming is projected to be beneficial for wine production in Tasmania (Harris et al., 2020).11.3.4.1.3 Adaptation Some farmers are adapting to drier and warmer conditions through more effective capture of non-growing-season rainfall (e.g., stubble retention to store soil water), improved water use efficiency and matching sowing times and cultivars to the environment high 2 train +5202 AR6_WGII 1625 36 While there is potential for a greater proportion of agriculture to be located to northern Australia, there are significant and complex agronomic, environmental, institutional, financial and social challenges for successful transformation, including the risk of disruption medium 1 test +5203 AR6_WGII 1626 9 Extreme climatic events (droughts, floods and heatwaves) are projected to adversely impact productivity for livestock systems medium 1 train +5204 AR6_WGII 1626 12 Increased heat stress in livestock is projected to decrease milk production and livestock reproduction rates high 2 train +5205 AR6_WGII 1626 14 In New Zealand, an extra 5 (RCP2.6) to 7 (RCP8.5) moderate heat stress days per year are projected for 2046– 2060 high 2 train +5206 AR6_WGII 1627 1 Elevated CO 2 is projected to increase forest growth if other biophysical factors are not limiting medium 1 train +5207 AR6_WGII 1627 2 Forestry plantations are projected to be negatively impacted from increases in fire weather (Box 11.1), particularly in southern Australia high 2 train +5208 AR6_WGII 1627 6 Effective management of the interactions between mitigation and adaptation policies can be achieved through governance and institutions, including Māori tribal organisations and sectoral adaptation, to ensure effective and continued carbon sequestration and storage as the climate changes medium 1 train +5209 AR6_WGII 1627 20 Changes in ocean temperature and acidification and the downstream impacts on species distribution, productivity and catch are projected concerns medium 1 train +5210 AR6_WGII 1627 22 For wild fisheries, multi- model projections suggest temperate and demersal systems, especially invertebrate shallow-water species, would be more strongly affected by climate change than tropical and pelagic systems medium 1 train +5211 AR6_WGII 1628 4 Given the value of that infrastructure and the rising damage costs, this represents a large knowledge gap that has led to an adaptation investment deficit.11.3.5.1 Observed Impacts Critical infrastructure, cities and settlements are being increasingly affected by chronic and acute climate hazards, including heat, drought, fire, pluvial and fluvial flooding and sea level rise (SLR), with consequent effects on many sectors high 2 train +5212 AR6_WGII 1628 5 Risks and impacts vary with physical characteristics, location, connectivity and socioeconomic status of settlements because of the ways these influence exposure and vulnerability high 2 train +5213 AR6_WGII 1628 10 Such tensions will be further challenged as temperatures rise and extreme events intensify beyond what has been experienced, thus stressing current adaptive capacities high 2 train +5214 AR6_WGII 1628 13 While investment in irrigation infrastructure may reduce climate change impacts in the short term, maladaptive outcomes cannot be ruled out longer term, which means that focusing attention now on adaptive and transformational measures can help increase climate resilience in areas exposed to increasing drought and climate extremes that disrupt production medium 1 train +5215 AR6_WGII 1629 4 Seaports, airports, water treatment plants, desalination plants, roads and railways are increasingly exposed to sea level rise (SLR) (very high confidence), impacting their longevity and levels of service and maintenance high 2 train +5216 AR6_WGII 1629 6 Extreme heat events exacerbate problems for vulnerable people and infrastructure in urban Australia, where urban heat is superimposed upon regional warming, and there are adverse impacts for population and vegetation health, particularly for socioeconomically disadvantaged groups (Tapper et al., 2014; Heaviside et al., 2017; Filho et al., 2018; Gebert et al., 2018; Rogers et al., 2018; Longden, 2019; Marchionni et al., 2019; Tapper, 2021) (11.3.6), energy demand, energy supply and infrastructure very high 3 train +5217 AR6_WGII 1629 13 Climatic extremes are exacerbating existing vulnerabilities high 2 train +5218 AR6_WGII 1630 4 Effective adaptations to urban heat include spatial planning, expanding tree canopy and greenery, shading, sprays and heat-resistant and energy-efficient building design, including cool materials and reflective or green roofs very high 3 train +5219 AR6_WGII 1630 14 Climate risk management is evolving, but adaptive capacity, implemen- tation, monitoring and evaluation are uneven across all scales of cities, settlements and infrastructure very high 3 train +5220 AR6_WGII 1630 25 Infrastructure planning is lagging behind international standards for climate resilience evaluation and guidance for adaptation to climate risk high 2 train +5221 AR6_WGII 1632 3 Projected SLR will cause more frequent flooding in Australia and New Zealand before mid-century (very high confidence)(Hunter, 2012; McInnes et al., 2016; Stephens et al., 2017; Stephens et al., 2020); (Steffen et al., 2014; PCE, 2015; MfE, 2017a; Hague et al., 2019; Paulik et al., 2020) Squeeze in intertidal habitats (high confidence)(Steffen et al., 2014; Peirson et al., 2015; Mills et al., 2016a; Mills et al., 2016b; Pettit et al., 2016; Rouse et al., 2017; Rayner et al., 2021) Significant property and infrastructure exposure (high confidence)(Steffen et al., 2014; PCE, 2015; Harvey, 2019; LGNZ, 2019; Paulik et al., 2020) (Table Box 11.5.2 and Table Box 11.6.2) Loss of significant cultural and archaeological sites and projected to compound with several hazards over this century (medium confidence)(Bickler et al., 2013; Birkett-Rees et al., 2020; NZ Archaeological Association, 2020) Increasing flood risk and water insecurity with health and well-being impacts on Torres Strait Islanders (high confidence)(Steffen et al., 2014; McInnes et al., 2016; McNamara et al., 2017) Degradation and loss of freshwater wetlands high 2 train +5222 AR6_WGII 1632 4 RSLR, to date, is a secondary factor influencing shoreline stability medium 1 train +5223 AR6_WGII 1632 5 The primary impacts of rising mean sea level (Table Box 11.6.1) are being compounded by climate-related changes in waves, storm surge, rising water tables, river flows and alterations in sediment delivery to the coast medium 1 train +5224 AR6_WGII 1632 8 The cumulative direct and residual risk from RSLR and associated impacts are projected to continue for centuries, necessitating ongoing adaptive decisions for exposed coastal communities and assets high 2 train +5225 AR6_WGII 1632 10 Rapid coastal development has increased exposure of coastal communities and infrastructure (high confidence) (Helman and Tomlinson, 2018; Paulik et al., 2020), reinforcing perceptions of safety (Gibbs, 2015; Lawrence et al., 2015) and creating barriers to retreat and nature-based adaptations very high 3 train +5226 AR6_WGII 1633 9 Remaining adaptation barriers are social or cultural (the absence of licence and legitimacy) and institutional (the absence of regulations, policies and processes that support changes to existing property rights and the funding of retreat) high 2 train +5227 AR6_WGII 1633 10 Legacy development, competing public and private interests, trade-offs among development and conservation objectives, policy inconsistencies, short- and long-term objectives and the timing and scale of impacts compound to create contestation over implementation of coastal adaptation high 2 train +5228 AR6_WGII 1633 11 Legal barriers to coastal adaptation remain (Schumacher, 2020) with a risk that the courts will become decision makers (Iorns Magallanes et al., 2018) due to legislative fragmentation, status quo leadership, lack of coordination between governance levels and agreement about who pays for what adaptation very high 3 train +5229 AR6_WGII 1633 13 Risk signalling through land use planning, flooding events and changes in insurance availability and costs is projected to increase recognition of coastal risks medium 1 train +5230 AR6_WGII 1633 15 Adopting ‘fit for purpose’ decision tools that are flexible as sea levels rise (11.7.3) can build adaptive capacity in communities and institutions high 2 train +5231 AR6_WGII 1634 20 Vulnerability to detrimental effects of climate change will vary with socioeconomic conditions high 2 train +5232 AR6_WGII 1635 1 Underlying health and economic trends affect the vulnerability of the population to extreme weather high 2 train +5233 AR6_WGII 1635 7 Heatwave responses, from public education to formal heat-warning systems, are the best-developed element of adaptation planning for health in Australia, but many metropolitan centres are still not covered high 2 train +5234 AR6_WGII 1636 11 Impacts of climate change are being observed across the tourism system high 2 train +5235 AR6_WGII 1636 24 Glacier tourism, a multi-million-dollar industry in New Zealand, is potentially under threat because glacier volumes are projected to decrease very high 3 train +5236 AR6_WGII 1636 29 Snow skiing faces significant challenges from climate change high 2 train +5237 AR6_WGII 1638 2 Climate adaptation finance is not evident medium 1 train +5238 AR6_WGII 1639 4 For a 4°C global warming, the changes are −48%, −14%, +135%, +213% and +350% respectively.11.3.10.3 Adaptation Options to manage risks include adaptation of energy markets, integrated planning, improved asset design standards, smart- grid technologies, energy generation diversification, distributed generation (e.g., roof-top solar, microgrids), energy efficiency, demand management, pumped hydro storage, battery storage and improved capacity to respond to supply deficits and balance variable energy resources across the network (Table 11.8) high 2 train +5239 AR6_WGII 1640 11 Fundamental shifts in the structure and composition of some ecosystems are partly due to anthropogenic climate change high 2 train +5240 AR6_WGII 1640 12 In human systems, the costs of droughts and floods in New Zealand, and heat-related mortality and fire damage in Australia, are partly attributed to anthropogenic climate change medium 1 train +5241 AR6_WGII 1640 15 Changing climate conditions are expected to exacerbate many of the social, economic and health inequalities faced by Aboriginal and Torres Strait Islander Peoples in Australia and Māori in New Zealand high 2 train +5242 AR6_WGII 1640 16 As a consequence, effective policy responses are those that take advantage of the interlinkages and dependencies between mitigation, adaptation and Indigenous Peoples’ well-being (Jones, 2019) and those that address the transformative change needed from colonial legacies high 2 train +5243 AR6_WGII 1640 17 There is a central role for Indigenous Peoples in climate change decision-making that helps address the enduring legacy of colonisation through building opportunities based on Indigenous governance regimes, cultural practices to care for land and water and intergenerational perspectives very high 3 test +5244 AR6_WGII 1640 21 Climate-related impacts on Aboriginal and Torres Strait Islander Peoples, countries (traditional estates) and cultures have been observed across Australia and are pervasive, complex and compounding high 2 train +5245 AR6_WGII 1641 8 Large proportions of collectively owned land already suffer from high rates of erosion (Warmenhoven et al., 2014; Awatere et al., 2018), which are projected to be exacerbated by climate-change-induced extreme rainfalls high 2 train +5246 AR6_WGII 1641 10 Climate-related impacts on Aboriginal and Torres Strait Islander Peoples, country and cultures: loss of biocultural diversity (land, water and sky) medium 1 test +5247 AR6_WGII 1641 12 Climate change impacts can exacerbate and/or accelerate existing threats of habitat degradation and biodiversity loss and create challenges for traditional stewardship of landscapes (Mackey and Claudie, 2015) Climate-driven loss of native title and other customary lands medium 1 train +5248 AR6_WGII 1641 14 Changing availability of traditional foods and forced diet change (medium confidence)Human health impacts can be exacerbated by climate change through the changing availability of traditional foods and medicines, while outages and the high costs of electricity can limit the storage of fresh food and medication (Kingsley et al., 2013; Spurway and Soldatic, 2016; Hall and Crosby, 2020) Changing climatic conditions for subsistence food harvesting (medium confidence)Climate-change-induced SLR and saltwater intrusion can limit the capacity for traditional Indigenous floodplain pastoralism and affect food security, access to and affordability of healthy, nutritional food (Ligtermoet, 2016; Spurway and Soldatic, 2016) Extreme weather events triggering disasters (high confidence)Increasing frequency or intensity of extreme weather events (floods, droughts, cyclones, heatwaves) can cause disaster responses in remote communities, including infrastructure damage of essential water and energy systems and health facilities (TSRA, 2018; Hall and Crosby, 2020) Heatwave impacts on human health high 2 train +5249 AR6_WGII 1641 16 For example, the Torres Strait Islands are already categorised under the U.S. National Oceanic and Atmospheric Administration (NOAA) Heat Index as a danger zone for extreme human health risk during summer (TSRA, 2018) Health impacts from changing conditions for vector-borne diseases high 2 train +5250 AR6_WGII 1641 17 For example, in the Torres Strait Islands the changing climate is affecting the range and extension of the Aedes albopictus and Aedes aegypti mosquitoes that can carry and transmit dengue and other viruses (Horwood et al., 2018; TSRA, 2018) Unadaptable infrastructure for changing environmental conditions high 2 train +5251 AR6_WGII 1641 18 Essential community-scale water and energy service infrastructure, unpaved roads, sea walls and stormwater drains can fail in extreme weather events (McNamara et al., 2017) Drinking water security medium 1 train +5252 AR6_WGII 1642 1 Further, many Māori-owned lands and cultural assets, such as marae and urupa, are located on coastal lowlands vulnerable to sea level rise (SLR) impacts high 2 train +5253 AR6_WGII 1642 2 Māori tribal investment in fisheries and aquaculture faces substantial risks from changes in ocean temperature and acidification and the downstream impacts on species distribution, productivity and yields medium 1 train +5254 AR6_WGII 1642 4 Changing climate conditions are projected to exacerbate health inequities faced by Māori medium 1 train +5255 AR6_WGII 1642 5 The production and ecology of some keystone cultural flora and fauna may be impacted by projected warming temperatures and reductions in rainfall medium 1 train +5256 AR6_WGII 1642 6 Obstruction of access to keystone species is expected to adversely impact customary practice, cultural identity and well-being medium 1 train +5257 AR6_WGII 1642 8 Māori tribal organisations have a critical role in defining climate risks and policy responses (Bargh et al., 2014; Parsons et al., 2019), as well as entering into strategic partnerships with business, science, research and government to address these risks high 2 train +5258 AR6_WGII 1642 9 More integrated assessments of climate change impacts, adaptation and socioeconomic risk for different Māori groups and communities, in the context of multiple stresses, inequities and different ways of knowing and being (King et al., 2013; Schneider et al., 2017; Henwood et al., 2019), would assist those striving to evaluate impacts and risks and how to integrate these assessments into adaptation plans high 2 train +5259 AR6_WGII 1643 1 Intergenerational approaches to climate change planning will become increasingly important, elevating political discussions about conceptions of rationality, diversity and the rights of non-human entities high 2 train +5260 AR6_WGII 1644 3 Anticipatory governance and agile decision-making can build resilience to cascading, compounding and aggregate impacts high 2 train +5261 AR6_WGII 1645 9 Delaying adaptation to climate risks may result in higher overall costs in future when adaptation is more urgent and impacts more extreme medium 1 train +5262 AR6_WGII 1647 6 Loss and degradation of tropical shallow coral reefs and associated biodiversity and ecosystem service values in Australia due to ocean warming and marine heatwaves very high 3 train +5263 AR6_WGII 1647 16 Loss of alpine biodiversity in Australia due to less snow high 2 train +5264 AR6_WGII 1647 21 Transition or collapse of alpine ash, snow gum woodland, pencil pine and northern jarrah forests in southern Australia due to hotter and drier conditions with more fires high 2 train +5265 AR6_WGII 1647 25 Loss of kelp forests in southern Australia and southeast New Zealand due to ocean warming, marine heatwaves and overgrazing by climate-driven range extensions of herbivore fish and urchins high 2 train +5266 AR6_WGII 1647 32 Loss of human and natural systems in low-lying coastal areas from ongoing SLR high 2 train +5267 AR6_WGII 1648 19 Increase in heat-related mortality and morbidity for people and wildlife in Australia high 2 train +5268 AR6_WGII 1648 31 Cascading, compounding and aggregate impacts on cities, settlements, infrastructure, supply chains and services due to extreme events high 2 train +5269 AR6_WGII 1649 5 Inability of institutions and governance systems to manage climate risks high 2 train +5270 AR6_WGII 1651 1 Short-term benefits from climate change may include reduced winter mortality, reduced energy demand for winter heating, increased agriculture productivity and forest growth in south and west New Zealand and increased forest and pasture growth in southern Australia, except where rainfall and soil nutrients are limiting (11.3.4, 11.3.6, 11.3.10) medium 1 train +5271 AR6_WGII 1651 4 Large gaps remain, especially in effective implementation, monitoring and evaluation (Supplementary Material SM 11.1) (CCATWG, 2017; Warnken and Mosadeghi, 2018), and current adaptation is largely incremental and reactive very high 3 train +5272 AR6_WGII 1651 15 Opportunities for integrated adaptation and mitigation planning in regional policies and plans have arisen through the Resource Management Amendment Act 2020 (Dickie, 2020), the National Policy Statement on Freshwater Management (MfE, 2020b) and the revised national coastal guidance (MfE, 2017a), but rely on funding instruments to be in place and statutes are aligned for their effectiveness very high 3 train +5273 AR6_WGII 1651 16 There is growing awareness of the need for more proactive adaptation planning at multiple scales and across sectors, and a better understanding of future risks and limits to adaptation is emerging medium 1 train +5274 AR6_WGII 1651 22 New tools are available in the region (Table 11.17), but uptake cannot be assumed high 2 train +5275 AR6_WGII 1651 23 Resilience and adaptation approaches are beginning to converge (White and O’Hare, 2014; Aldunce et al., 2015) (Supplementary Material SM 11.1) but widespread ‘bounce-back’ resilience-driven responses that lock in risk by discounting ongoing and changing climate risk (Leitch and Bohensky, 2014; O’Hare et al., 2016; Wenger, 2017; Torabi et al., 2018) can create maladaptation and impede long- term adaptation goals high 2 train +5276 AR6_WGII 1651 25 Nature-based adaptations (Colloff et al., 2016; Lavorel et al., 2019; Della Bosca and Gillespie, 2020) and ‘green infrastructure’ medium 1 train +5277 AR6_WGII 1653 2 Effective cooperation and a positive innovation culture can contribute to the collaborative development of climate change adaptation pathways medium 1 train +5278 AR6_WGII 1653 7 Learning and experimentation across governance boundaries and between agencies and local communities enable adaptation to be better aligned with changing climate risks and community high 2 train +5279 AR6_WGII 1653 8 There is increasing focus on improving adaptive capacity for transitional and transformational responses, but reactive responses dominate very high 3 train +5280 AR6_WGII 1655 2 There are many barriers to starting adaptation pre-emptively very high 3 train +5281 AR6_WGII 1655 11 Communities’ vulnerabilities are dynamic and uneven high 2 train +5282 AR6_WGII 1655 17 In some human systems, fundamental limits to adaptation include thermal thresholds and safe freshwater (Alston et al., 2018) (Table 11.14) and the inability of some low-lying coastal communities to adapt in place (Box 11.6) very high 3 train +5283 AR6_WGII 1655 19 A lack of robust and timely adaptation means key risks will increasingly manifest as impacts, and numerous systems, communities and institutions are projected to reach limits (Table 11.14, Figure 11.6), compounding current adaptation deficits and undermining society’s capacity to adapt to future impacts very high 3 train +5284 AR6_WGII 1655 25 Decision makers face the challenge of how to adapt when there are ongoing knowledge gaps and uncertainties about when some climate change impacts will occur and their scale, for example coastal flooding (Box 11.6) or extreme rainfall events and their cascading effects (Box 11.4) very high 3 train +5285 AR6_WGII 1656 2 More inclusive, collaborative and learning-oriented community engagement processes are fundamental to effective adaptation outcomes (11.7.3.2) very high 3 test +5286 AR6_WGII 1656 5 Regular monitoring, evaluation, communication and coordination of adaptation are essential for accelerating learning and adjusting to dynamic climate impacts and changes in socioeconomic and cultural conditions high 2 train +5287 AR6_WGII 1657 1 Nevertheless, New Zealanders have a tendency to overestimate the amount of sea level rise (SLR), especially among those most concerned about climate change and incorrectly associate it with melting sea ice, which has implications for engagement and communication strategies (Priestley et al., 2021).The use of more systemic, collaborative and future-oriented engagement approaches is facilitating adaptation in local contexts high 2 train +5288 AR6_WGII 1658 5 All of these approaches depend on adequate resourcing very high 3 train +5289 AR6_WGII 1658 20 Current global emissions reduction policies are projected to lead to a global warming of 2.1°C–3.9°C by 2100 (Liu and Raftery, 2021), leaving many of the region’s human and natural systems at very high risk and beyond adaptation limits high 2 train +5290 AR6_WGII 1659 2 Building cities and settlements that are resilient to the impacts of climate change requires the simultaneous consideration of infrastructural, ecological, social, economic, institutional and political dimensions of resilience, including political will, leadership, commitment, community support, multi-level governance and policy continuity (Torabi et al., 2021).11.8.2 Challenges for Climate Resilient Development Pathways Implementing enablers can help drive adaptation ambition and action consistent with climate resilient development very high 3 train +5291 AR6_WGII 1663 4 Whatever the outcome, adaptation and mitigation are essential and urgent very high 3 train +5292 AR6_WGII 1702 2 High levels of widespread poverty, weak water governance, unequal access to safe water and sanitation services and lack of infrastructure and financing reduce adaptation capacity, increasing and creating new population vulnerabilities high 2 train +5293 AR6_WGII 1702 5 This resulted in high tree mortality rates and basin-wide reductions in forest productivity, momentarily turning pristine forest areas from a carbon sink into a net source of carbon to the atmosphere high 2 train +5294 AR6_WGII 1702 7 The combined effect of anthropogenic land use change and climate change increases the vulnerabilities of terrestrial ecosystems to extreme climate events and fires medium 1 train +5295 AR6_WGII 1702 9 Species have shifted upslope, leading to range contractions for highland species and range contractions and expansions for lowland species, including crops and vectors of diseases very high 3 train +5296 AR6_WGII 1702 15 Glacier retreat, temperature increase and precipitation variability, together with land use changes, have affected ecosystems, water resources and livelihoods through landslides and flood disasters very high 3 train +5297 AR6_WGII 1702 16 In several areas of the Andes, flood and landslide disasters have increased, and water availability and quality and soil erosion have been affected by both climatic and non-climatic factors high 2 train +5298 AR6_WGII 1702 19 On average, people in the region were more exposed to high fire danger between 1 and 26 additional days depending on the sub-region for the years 2017–2020 compared to 2001–2004 high 2 train +5299 AR6_WGII 1702 22 Conversely, reduced precipitation and altered rainfall at the start and end of the rainy season and during the mid-summer drought (MSD) is impacting rainfed subsistence farming, particularly in the Dry Corridor in CA and in the tropical Andes, compromising food security high 2 train +5300 AR6_WGII 1702 25 Examples are the effects of warming temperatures on increasing the suitability of transmission of vector-borne diseases, including endemic and emerging arboviral diseases such as dengue fever, chikungunya and Zika medium 1 train +5301 AR6_WGII 1702 26 The reproduction potential for the transmission of dengue increased between 17% and 80% for the period 1950–1954 to 2016–2021, depending on the sub-region, as a result of changes in temperature and precipitation high 2 train +5302 AR6_WGII 1703 3 Often they have less capacity to adapt, further widening structural gender gaps high 2 train +5303 AR6_WGII 1703 6 The adaptation measures in place, however, are insufficient to safeguard terrestrial and freshwater ecosystems in the CSA from the negative impacts of climate change high 2 train +5304 AR6_WGII 1703 8 The main adaptation measures are ocean zoning, the prohibition of productive activities (e.g., fisheries, aquaculture, mining and tourism) on marine ecosystems, the improvement of research and education programmes and the creation of specific national policies high 2 train +5305 AR6_WGII 1703 10 Inclusive water regimes that overcome social inequalities and approaches including nature-based solutions, such as wetland restoration and water storage and infiltration infrastructure, with synergies for ecosystem conservation and disaster risk reduction, have been found to be more successful for adaptation and sustainable development high 2 train +5306 AR6_WGII 1703 13 Adaptation requires governance improvements and new strategies to address the changing climate; nevertheless, barriers limiting adaptive capacity persist such as lack of educational programmes for farmers, adequate knowledge of site-specific adaptation and institutional and financial constraints high 2 train +5307 AR6_WGII 1703 15 Regulation, planning and control systems are central tools for reducing risk associated with the security of buildings and their location and the proper supply of basic urban services and transport high 2 train +5308 AR6_WGII 1703 16 The adoption of nature-based solutions (e.g., urban agriculture and river restoration) and hybrid (grey-green) infrastructure is still in the early stages, with weak connections to poverty and inequality reduction strategies medium 1 train +5309 AR6_WGII 1703 17 Focusing on risk reduction encompasses upgrading informal and precarious settlements, built environments and housing conditions, which offer an important but still limited contribution to urban adaptation high 2 train +5310 AR6_WGII 1703 20 Political, institutional and financial barriers reduce the feasibility of implementing these tools high 2 train +5311 AR6_WGII 1703 22 IKLK can contribute to reducing the vulnerability of local communities to climate change medium 1 train +5312 AR6_WGII 1703 24 Climate change is projected to convert existing risks in the region into severe key risks medium 1 train +5313 AR6_WGII 1703 36 Small fisheries and farming of seafood will be negatively affected as El Niño Southern Oscillation (ENSO) events become more frequent and intense and ocean warming and acidification continues medium 1 train +5314 AR6_WGII 1704 2 Disruption in water flows will significantly degrade ecosystems such as high-elevation wetlands and affect farming communities, public health and energy production high 2 train +5315 AR6_WGII 1704 4 This can happen through expanded distribution of vectors, especially viral infectious diseases of zoonotic origin in transition areas between urban and suburban, or rural settings, and upslope in the mountains medium 1 train +5316 AR6_WGII 1704 6 The combined effect of both impacts will lead to a long-term decrease in carbon stocks in forest biomass, compromising Amazonia’s role as a carbon sink, largely conditioned on the forest’s responses to elevated atmospheric CO 2 medium 1 train +5317 AR6_WGII 1704 7 The southern portion of the Amazon has become a net carbon source to the atmosphere in the past decade high 2 train +5318 AR6_WGII 1704 9 Available studies focus mainly on vertebrates and plants of the Atlantic Forest and Cerrado in Brazil and in CA, with a large knowledge gap on freshwater ecosystems {12.3, 12.5.1, CCP1} Ocean and coastal ecosystems in the region will continue to be highly impacted by climate change high 2 train +5319 AR6_WGII 1704 11 In the Representative Concentration Pathway (RCP) 4.5 and RCP8.5 scenarios, by 2050, virtually every coral reef will experience at least one severe bleaching event per year high 2 train +5320 AR6_WGII 1704 12 Under all RCP scenarios of climate change, there will be changes in the geographical distribution of marine species and ocean and coastal ecosystems such as mangroves, estuaries and rocky shores, as well as those species held in fisheries medium 1 train +5321 AR6_WGII 1704 16 Socioeconomic and political factors that provide some level of safety and continuity of policies and actions are critical enablers of adaptation high 2 train +5322 AR6_WGII 1704 18 Conflicts in which the direct biophysical impacts of climate change play a major role can unleash protests and strengthen social movements medium 1 train +5323 AR6_WGII 1704 28 Initiatives to improve informal and precarious settlement, guaranteeing access to land and decent housing, are aligned with comprehensive adaptation policies that include the development and reduction of poverty, inequality and disaster risk medium 1 train +5324 AR6_WGII 1705 1 Adaptation policies often address climate impact drivers, but seldom include the social and economic underpinnings of vulnerability. This narrow scope limits adaptation results and compromises their continuity in the region high 2 test +5325 AR6_WGII 1705 3 Dialogue and agreement that include multiple actors are mechanisms to acknowledge trade-offs and promote dynamic, site-specific adaptation options medium 1 train +5326 AR6_WGII 1706 4 Yet it is a region of substantial social inequality including the highest inequality in land tenure, where a large percentage of the population remains below the poverty line, unequally distributed between rural and urban areas and along aspects like gender and race; these groups are highly vulnerable to climate change and natural extreme events that frequently affect the region high 2 train +5327 AR6_WGII 1706 10 Indigenous Peoples and smallholder families are lacking adequate climate policies combined with institutions to protect their property rights; this could result in a more sustainable process of agricultural expansion, without substantially increasing greenhouse gas (GHG) emissions and the vulnerability of those populations high 2 train +5328 AR6_WGII 1708 3 Inadequate governance and lack of participation escalates the vulnerability and risk to climate variability and change in the region high 2 train +5329 AR6_WGII 1708 4 Increasing trends in precipitation had been observed in SES (Figure 12.1), in contrast to decreasing trends in CA and central- southern Chile high 2 train +5330 AR6_WGII 1708 8 Significant dependency on rainfed agriculture (>30% in Guatemala, Honduras and Nicaragua) indicates high sensitivity to climatic variability and change and represents a challenge for food security high 2 train +5331 AR6_WGII 1708 13 Though reduced suitability and yield for beans, coffee, maize, plantain and rice are expected in CA (SRCCL Chapter 5, Mbow et al., 2019), limiting the warming to 1.5°C, compared with 2°C, are projected to result in smaller net reductions in yields of maize, rice, wheat and other cereal crops for CSA high 2 train +5332 AR6_WGII 1708 20 Projections of potential impacts of climate change on malaria confirm that weather and climate are among the drivers of geographic range, intensity of transmission, and seasonality; the changes of risk become more complex with additional warming very high 3 train +5333 AR6_WGII 1708 24 This reduction has modified the frequency, magnitude and location of related natural hazards, while the exposure of people and infrastructure has increased because of growing population, tourism and economic development high 2 train +5334 AR6_WGII 1708 26 This conversion of natural ecosystems is the main cause of biodiversity and ecosystem loss and is an important source of GHG emissions high 2 train +5335 AR6_WGII 1708 27 The combination of continued anthropogenic disturbance, particularly deforestation, with global warming may result in dieback of forest in Central and South America medium 1 test +5336 AR6_WGII 1708 30 However, agricultural expansion results in large conversions in tropical dry woodlands and savannahs in SA (Brazilian Cerrado, Caatinga and Chaco) high 2 train +5337 AR6_WGII 1709 3 The magnitude and frequency of extreme precipitation events have increased, but droughts have mixed signals low 0 train +5338 AR6_WGII 1709 9 Massive heatwave events and increase in the frequency of warm extremes are projected at the end of the 21st century high 2 train +5339 AR6_WGII 1709 13 The main climate impact drivers like extreme heat, drought, relative SLR, coastal flooding, erosion, marine heatwaves, ocean aridity (high confidence) and aridity, drought and wildfires will increase by mid- century medium 1 train +5340 AR6_WGII 1709 14 The rainy season in CA will likely experience more pronounced MSD by the end of this century, with a signal for reduced minimum precipitation by mid-century for the June July August (JJA) and September October November (SON) quarters, and a broader second peak is projected, consistent with the future south displacement of the Intertropical Convergence Zone (ITCZ) high 2 train +5341 AR6_WGII 1709 18 CA is one of the regions most exposed to climatic phenomena; with long coastlines and lowland areas, the region is repeatedly affected by drought, intense rains, cyclones and ENSO events high 2 train +5342 AR6_WGII 1709 23 Vulnerability in most sectors is considered high or very high high 2 train +5343 AR6_WGII 1710 6 Hydro-meteorological events, such as storm surges and TCs, are the most frequent extreme events and have the highest impact high 2 train +5344 AR6_WGII 1710 9 Food insecurity is a serious impact of climate change in a region where 10% of the GDP depends on agriculture, livestock and fisheries very high 3 train +5345 AR6_WGII 1710 15 Studies have shown that the incidence of some vector-borne and zoonotic diseases in CA is correlated to climatic variables, particularly temperature and rainfall high 2 train +5346 AR6_WGII 1710 20 Heat stress is another health concern in this already warm and humid part of the world high 2 train +5347 AR6_WGII 1710 23 Decreasing water availability is another impact of climate change high 2 train +5348 AR6_WGII 1710 26 Impacts on rural livelihoods, particularly for small and medium-sized farmers and Indigenous Peoples in mountains, include an overall reduction in production, yield (Table 12.4), suitable farming area and water availability high 2 train +5349 AR6_WGII 1710 30 The extent and quality of suitable areas for basic grains are expected to contract high 2 test +5350 AR6_WGII 1711 6 Mesoamerica, a biodiversity hotspot spanning across CA and southern Mexico, is a global priority for terrestrial biodiversity conservation, and it is projected to be negatively impacted by climate change, especially through the contraction of distribution of native species as the area becomes increasingly dryer high 2 train +5351 AR6_WGII 1711 15 Significant increases in the intensity and frequency of hot extremes and significant decreases in the intensity and frequency of cold extremes (Dereczynski et al., 2020; Dunn et al., 2020) were likely observed. Insufficient data coverage and trends in available data are generally not significant for heavy precipitation low 0 test +5352 AR6_WGII 1712 6 Nevertheless, models project inconsistent changes in the region for extreme precipitation low 0 train +5353 AR6_WGII 1712 11 The Eastern Tropical Pacific, particularly Sector Niño 3.4, will see the worst increase in SST, affecting industrial and small-scale fisheries very high 3 train +5354 AR6_WGII 1712 14 The Andean piedmont (500–1200 metres above sea level [MASL]) ecosystems and crops and elevation ranges above the treeline are more exposed to thermal anomalies very high 3 test +5355 AR6_WGII 1712 15 Temperature rise, combined with precipitation and floods, leaves people more exposed to epidemics very high 3 train +5356 AR6_WGII 1712 16 A more significant exposure is related to lower socioeconomic conditions, poor health and marginalisation (Oliver- Smith, 2014).12.3.2.3 Vulnerability Local economies reliant on limited and specialised resources, highly dependent on ecosystem services such as water and soil fertility, such as alpaca and llama herders or small-scale fishers, are among the more vulnerable very high 3 train +5357 AR6_WGII 1712 17 Their vulnerabilities increase as a result of unequal chains of value, incomplete transfers of technology and other socioeconomic and environmental drivers high 2 train +5358 AR6_WGII 1712 18 Informal housing and settlements, usually located in areas exposed to the highest level of risk, exacerbates vulnerability very high 3 train +5359 AR6_WGII 1712 20 Most cities and infrastructure are considered highly vulnerable to climate change high 2 train +5360 AR6_WGII 1712 22 Also biodiversity and water-dependent activities where seasonality and rainfall patterns are changing and where other non-climatic sources of change, such as land use, affect the capacity of ecosystems to provide hydrological services very high 3 train +5361 AR6_WGII 1712 25 Scale studies indicate an increase of flood risk during the 21st century, consistent with more frequent floods, with the risk being worse in higher emission scenarios high 2 train +5362 AR6_WGII 1712 26 Those living on riverbanks and in slums built on steep slopes are among the most affected by floods of all kinds high 2 train +5363 AR6_WGII 1712 28 Increased SST, coupled with stronger ENSO events, will affect marine life and fisheries by loss of productive habitat, disruption of nutrient structure, productivity and alteration of species migration patterns, leading to changes in fishing rates, which will impact coastal livelihoods high 2 train +5364 AR6_WGII 1713 4 Irrigation, potable water, health and education infrastructures, as well as roads, bridges, cities and residential constructions, are frequently damaged or destroyed by extreme precipitation events, which also impact sediment transport, river erosion and annual discharge very high 3 train +5365 AR6_WGII 1713 5 The increasing variability of precipitation has compromised rainfed agriculture and power generation, particularly in the dry season high 2 train +5366 AR6_WGII 1713 6 For the Amazon–Andes transition zone, the impacts of hydrological variability and transport of sediments have been noticed in riparian agriculture and biodiversity high 2 train +5367 AR6_WGII 1713 8 Increases in vector-borne diseases can be related to increases in rainfall and minimum temperatures during ENSO events (Stewart- Ibarra and Lowe, 2013) and the expansion of the diseases’ altitudinal distribution high 2 train +5368 AR6_WGII 1713 13 Climate change will contribute to increased malaria vectorial capacity high 2 train +5369 AR6_WGII 1713 18 Glacier volume loss and permafrost thawing will continue in all scenarios high 2 train +5370 AR6_WGII 1713 21 Changing glaciers, snow and permafrost (Figure 12.13), in synergy with land use change, have implications for the occurrence, frequency and magnitude of derived floods and landslides (high confidence) (Huggel et al., 2007; Iribarren Anacona et al., 2015; Emmer, 2017; Mark et al., 2017), as well as for landscape transformation through lake formation or drying and for alterations in hydrological dynamics, with impacts on water for human consumption, agriculture, industry, hydroelectric generation, carbon sequestration and biodiversity high 2 train +5371 AR6_WGII 1714 1 Water flow has decreased in several basins, such as the Shullcas River in the Cordillera Huaytapallana in Peru, and is expected to decrease in the near future in places such as the Cordillera Blanca in Peru very high 3 train +5372 AR6_WGII 1714 2 Disruptions in water flows will significantly degrade or eliminate high-elevation wetlands high 2 train +5373 AR6_WGII 1714 8 Glacier retreat impacts Andean pastoralists high 2 train +5374 AR6_WGII 1714 12 Species distribution is changing upslope due to increasing air temperature, leading to range contraction and local extinctions of highland species, whereas lowland species are experiencing range contractions at the rear end and expansions in the front end, including vectors of diseases high 2 train +5375 AR6_WGII 1714 15 Shifts in tree species distribution may result in decreased above-ground carbon stocks and productivity in tropical mountain forests high 2 train +5376 AR6_WGII 1714 17 The agricultural frontier of crops, such as potatoes or maize, is moving upwards high 2 train +5377 AR6_WGII 1714 18 Modelling exercises agree with the observed impacts in species, ecosystem processes, crop impacts and related pests and diseases high 2 train +5378 AR6_WGII 1715 1 Droughts presented mixed trends between sub-regions, but evidence indicates an increasing length of dry periods low 0 train +5379 AR6_WGII 1715 22 The LECZ of Venezuela, Guyana and Suriname are highly vulnerable to climate change due to SLR high 2 train +5380 AR6_WGII 1716 2 Deforestation, temperature increase and any factor affecting forest ecosystem dynamics will have an impact on atmospheric CO 2 concentrations and, hence, on the global climate (Ruiz-Vásquez et al., 2020; Sullivan et al., 2020).There is robust scientific evidence of the high vulnerability of Amazon rainforests to increasing temperature and repeated extreme drought events high 2 train +5381 AR6_WGII 1717 4 Nevertheless, the combined effect of increasing tree mortality with variations in growth results in a long- term decrease in C stocks in forest biomass, compromising the role of these forests as a C sink high 2 train +5382 AR6_WGII 1717 23 The occurrence of extreme droughts has affected the carbon and water cycles in large areas of the Amazon rainforest high 2 train +5383 AR6_WGII 1718 5 Protected areas—Indigenous lands included—have significantly reduced forest clear-cutting in the Amazon deforestation arc (most of which is inside SAM) high 2 train +5384 AR6_WGII 1718 8 The agriculture in the region is highly dependent on the climate high 2 train +5385 AR6_WGII 1718 14 The rise of a large-scale soybean agroindustry in the early 2000s led to a faster increase in human development indicators in some regions, tightly linked to the agricultural production chain high 2 train +5386 AR6_WGII 1718 19 The populations of some of these cities are reported as being highly vulnerable considering the enormous social inequalities embedded in these cities high 2 train +5387 AR6_WGII 1718 20 Inequalities and uneven access to infrastructure, housing and healthcare increase populations’ vulnerability to atmospheric pollution and drier conditions high 2 train +5388 AR6_WGII 1718 22 The Brazilian Cerrado is one of the most diverse savannah in the world, with more than 12,600 plant species, with 35% being endemic high 2 train +5389 AR6_WGII 1718 23 Historic land cover change and concurrent climate change in the region strongly impacted the biodiversity and led to the extinction of 657 plant species for the Cerrado, which is more than four times the global recorded plant extinctions high 2 train +5390 AR6_WGII 1718 29 Due to the tight relationship between drought and fire occurrence, an increase of 39% to 95% of burned area is modelled to impact the Cerrado region under RCP4.5 and RCP8.5, while under RCP2.6, a 22% overshoot in temperature is estimated to impact the area in 2050 decreasing to 11% overshoot by 2100 (Silva et al., 2019d), leading to a high impact on agricultural production high 2 test +5391 AR6_WGII 1719 1 Extreme floods in the southern Amazon and Bolivian Amazon floodplains were described and related to the exceptionally warm sub-tropical South Atlantic ocean (high confidence) (Espinoza et al., 2014), causing high economic impacts (losses in crop and livestock production and infrastructure) and number of fatalities very high 3 train +5392 AR6_WGII 1719 6 The climatic connection of Pantanal regions to the Amazon, and the influence of deforestation in local precipitation (Marengo et al., 2018) has implications for conservation of ecosystem services and water security in Pantanal high 2 train +5393 AR6_WGII 1719 8 The projected impacts of climate change will lead to profound changes in the annual flood dynamics for Pantanal wetlands, altering ecosystem functioning and severely affecting biodiversity high 2 train +5394 AR6_WGII 1719 9 Soybean and corn yields in the Cerrado region will suffer one of the strongest negative impacts under the estimates of the RCP4.5 and RCP8.5 scenarios and will require high levels of investment in adaptation should they continue to be cultivated in the same areas as currently high 2 train +5395 AR6_WGII 1719 11 Thus, the future socioeconomic vigour of the region will be, to a large extent, connected to an unlikely stability of the regional climate and eventual fluctuations of global markets potentially affecting the agricultural supply chain high 2 train +5396 AR6_WGII 1719 13 Northwest SAM had an approximately 54% increase in the incidence of respiratory diseases associated with forest fires during the 2005 drought compared to a no-drought 10-year mean high 2 train +5397 AR6_WGII 1719 15 Increases in hospital admissions, asthma, DNA damage and lung cell death due to the inhalation of fine particulate matter represent an increase in public health system costs high 2 train +5398 AR6_WGII 1720 1 Global-scale studies indicate an increase of flood risk for the SAM region during the 21st century (consistent with floods that are more frequent) high 2 train +5399 AR6_WGII 1720 10 A decrease in total precipitation is projected with high confidence, with an increase in heavy precipitation events and an increase in dryness medium 1 train +5400 AR6_WGII 1720 12 Fire will significantly increase high 2 train +5401 AR6_WGII 1720 14 People are exposed to intense drought and famine high 2 train +5402 AR6_WGII 1720 16 People, infrastructure and economic activities are exposed to SLR in the 3800 km of coastline medium 1 train +5403 AR6_WGII 1721 10 Caatinga’s high vulnerability to climate change is further increased by the extensive conversion of native vegetation high 2 train +5404 AR6_WGII 1721 11 Studies with terrestrial animals show that habitat loss increases the vulnerability of species to climate change high 2 train +5405 AR6_WGII 1721 12 NES’s coral reefs have shown some resilience to bleaching, but vulnerability is intensified by the synergy between chronic heat stress caused by increased SST (Teixeira et al., 2019) and other well-documented stressors, such as coastal runoff, urban development, marine tourism, overexploitation of reef organisms and oil extraction high 2 train +5406 AR6_WGII 1721 19 More than 10 million people have been impacted by the drought of 2012/2014 in the region, which was responsible for water shortage and contamination, increasing death by diarrhoea (Marengo and Bernasconi, 2015; Government of Brazil, 2020).There is growing evidence of the impacts of climate change on human health in NES, mostly linked to food and water insecurity caused by recurrent long droughts (e.g., gastroenteritis and hepatitis) high 2 train +5407 AR6_WGII 1721 22 Recent studies predict strong negative impacts of climate change on NES’s agriculture high 2 train +5408 AR6_WGII 1722 2 Over SES most stations have registered an increase in annual rainfall, largely attributable to changes in the warm season; this is one of few sub-regions where a robust positive trend in precipitation and significant intensification of heavy precipitation have been detected since the early 20th century high 2 train +5409 AR6_WGII 1722 7 Increasing trends in mean air temperature and extreme heat and decreasing cold spells are projected high 2 train +5410 AR6_WGII 1722 9 Increases in mean precipitation (high confidence), pluvial floods and river floods are projected medium 1 train +5411 AR6_WGII 1722 10 Droughts in the River Plate basin will be more frequent in the medium term (2011–2040) and the distant future (2071–2100) (with respect to the 1979–2008 period), but also shorter and more severe, for the more extreme emission scenario (RCP8.5) low 0 train +5412 AR6_WGII 1722 12 All coastal and oceanic climate impact drivers (relative sea level, coastal flood and erosion, marine heatwaves and ocean aridity) are expected to increase by mid-century in the RCP8.5 scenario high 2 train +5413 AR6_WGII 1722 13 In the River Plate basin, urban floods have become more frequent, causing infrastructure damage and sometimes substantial mortality high 2 train +5414 AR6_WGII 1722 14 A large increase in landslides and flash floods is also predicted for the Brazilian portion of SES, where they are responsible for the majority of deaths related to disasters in the country high 2 train +5415 AR6_WGII 1722 17 Urban heat islands are already a reality in large cities in the region, such as Buenos Aires (high confidence) (Wong et al., 2013; Sarricolea and Meseguer-Ruiz, 2019; Wu et al., 2019; Mettler-Grove, 2020), Rio de Janeiro (high confidence) (Ceccherini et al., 2016; Neiva et al., 2017; Geirinhas et al., 2018; Peres et al., 2018; Sarricolea and Meseguer-Ruiz, 2019; Wu et al., 2019; de Farias et al., 2021) and São Paulo high 2 train +5416 AR6_WGII 1722 25 As sea level rises as a result of global climate change, storm surge floods will become more frequent in this densely populated area, particularly in low-lying areas high 2 train +5417 AR6_WGII 1723 1 Extreme events, including storm surges and coastal inundation/ flooding, cause injuries and economic/environmental losses on the urbanised coastline of Southern Brazil (States of São Paulo and Santa Catarina) high 2 train +5418 AR6_WGII 1723 13 Heat islands affect ecosystems by increasing the energy consumption for cooling, the concentration of pollutants and the incidence of fires high 2 train +5419 AR6_WGII 1723 15 Warming temperatures have been implicated in the emergence of dengue in temperate latitudes, increasing populations of Aedes aegypti high 2 train +5420 AR6_WGII 1723 36 Cerrado savannahs are projected to be the hotspot most negatively impacted by climate change within SA, mostly though range contraction of plant species very high 3 train +5421 AR6_WGII 1724 3 In particular, a significant increment in the duration and frequency of heatwaves mainly in central Chile from 1961 to 2016 has been observed (Piticar, 2018).A robust drying trend for Chile (30°S–48°S) has been recorded medium 1 train +5422 AR6_WGII 1724 4 However, inconsistent trends over the region in the magnitude of precipitation extremes with both decreases and increases (Chou et al., 2014; Giorgi et al., 2014; Heidinger et al., 2018; Meseguer-Ruiz et al., 2018) (WGI AR6 Table 11.14) (Seneviratne et al., 2021) have been observed low 0 train +5423 AR6_WGII 1724 8 From 1946 to 2017, the number of fires and areas burned have increased significantly in Chile high 2 train +5424 AR6_WGII 1725 1 The glaciers of the southern Andes (including the SWS and SSA regions) show the highest glacier mass loss rates worldwide high 2 train +5425 AR6_WGII 1725 3 Four sets of downscaling simulations based on the Eta Regional Climate Model forced by two global climate models (Chou et al., 2014) projected warmer conditions (more than 1°C) for the entire sub-region by 2050 under the RCP4.5 scenario medium 1 train +5426 AR6_WGII 1725 5 Drier conditions (medium confidence), by means of a decrease in total annual and extreme precipitation, are expected to increase for southern Chile, but inconsistent changes are expected in the sub-region low 0 train +5427 AR6_WGII 1725 8 Profound social inequalities, urban expansion and inadequate city planning (e.g., drainage network) increase exposure to flooding events and landslides (high confidence) (Müller and Höfer, 2014; Rojas et al., 2017; Lara et al., 2018), heat hazards such as heatwaves high 2 train +5428 AR6_WGII 1725 10 In particular, human communities living in coastal cities show a negative safety perception about the performance of the infrastructure and coastal defences to flood events low 0 train +5429 AR6_WGII 1725 13 Terrestrial ecosystems dominated by exotic species (e.g., pine) with lower landscape heterogeneity and degraded soils and that are close to settlements and roads are highly vulnerable to wildfires in comparison to forests dominated by native trees high 2 train +5430 AR6_WGII 1725 16 The water sector shows a very high vulnerability (high confidence) (Figure 12.7) mainly due to weak water governance focused on market aspects (e.g., inter-sectoral water transactions, setting rates, granting concessions, waiving the water right) high 2 train +5431 AR6_WGII 1725 17 Potable water and adequate sanitation are available in SWS; however, water availability in Chile is unevenly distributed in rural communities high 2 train +5432 AR6_WGII 1725 18 Spatial differences in water availability are enhanced by strong population growth, economic development, mining activities and the high dependence of agriculture on irrigation high 2 train +5433 AR6_WGII 1725 19 Droughts in SWS are a major threat to water security high 2 train +5434 AR6_WGII 1726 2 Urban and agriculture sectors are vulnerable to climate change (medium confidence: medium evidence, high agreement) (Figure 12.7), increasing problems and demand for water high 2 train +5435 AR6_WGII 1726 3 Important health problems (e.g., pathogenic infections, changes in vector-borne diseases, heat-related mortality, lower neurobehavioural performance) have been associated with agriculture, mining and thermal power production activities in SWS high 2 train +5436 AR6_WGII 1726 4 Large-scale agricultural growth has increased vulnerability to climate change by disfavouring traditional agriculture, the homogenisation of the biophysical landscape and the replacement of traditional crops and native forests with exotic species like pines and eucalyptus (high confidence) (Torres et al., 2015), where farmers’ perceptions of climate change are highly dependent on educational level and access to meteorological information low 0 test +5437 AR6_WGII 1726 5 Agricultural systems owned by Indigenous Peoples (i.e., Mapuche, Quechua and Aymara farmers) seem to pose a lower level of vulnerability to drought and higher response capacity than non-Indigenous farmers thanks to the use of the traditional knowledge of specific management techniques and the tendency to conserve species or varieties of crops tolerant to water scarcity low 0 train +5438 AR6_WGII 1726 6 Fishery- and aquaculture- related livelihoods are vulnerable to climate and non-climate drivers (medium confidence: medium evidence, high agreement), such as sea surface warming and precipitation reduction (Handisyde et al., 2017; Soto et al., 2019; González et al., 2021), changes in upwelling intensity low 0 train +5439 AR6_WGII 1726 8 The Chilean healthcare system has become more equitable and responsive to the population’s needs (e.g., the Bono AUGE healthcare reform programme); however, the high relative inequalities in terms of income (OECD, 2018), education level and rural–urban factors are determinants of quality of care, health system barriers and differential access to healthcare high 2 train +5440 AR6_WGII 1726 9 Exposure and vulnerability to psychosocial risks in SWS show significant inequalities in times of disasters such as earthquakes according to socioeconomic, geographic and gender factors high 2 train +5441 AR6_WGII 1726 10 Indigenous Peoples have the highest levels of vulnerability in Chile in terms of income, basic needs and access to services to climate change low 0 train +5442 AR6_WGII 1726 12 Increasing temperatures and decreasing precipitation have increased the impacts of wildfires on terrestrial ecosystems high 2 train +5443 AR6_WGII 1726 13 Future projections show important changes in the productivity, structure and biogeochemical cycles of SWS temperate and rainforests (medium confidence: medium evidence, high agreement) (Gutiérrez et al., 2014; Correa-Araneda et al., 2020) and their fauna low 0 train +5444 AR6_WGII 1726 15 Species are projected to suffer changes in their distribution, including a decrease in climatic refugia for vertebrates low 0 test +5445 AR6_WGII 1726 16 Increasing temperatures have enlarged the number and areal extent of glacier lakes in the central Andes, northern Patagonia and southern Patagonia (high confidence) (Wilson et al., 2018), while decreased rainfall and rapid glacier melting have provoked changes in the environmental, biogeochemical and biological properties of central- southern and Andes Chilean lakes low 0 train +5446 AR6_WGII 1726 19 Increasing temperatures, decreasing precipitation regimes and an unprecedented long-term drought have decreased the annual average river streamflows that supply SWS megacities such as Santiago high 2 train +5447 AR6_WGII 1727 1 Drought has affected wetlands low 0 train +5448 AR6_WGII 1727 2 There is low evidence on shoreline retreat attributed to climate change (Martínez et al., 2018; Ministerio de Medio Ambiente de Chile, 2019), although increasing wind intensity along the central Chilean coast has caused serious damage in coastal infrastructure and buildings (Winckler et al., 2017) and changes in seawater properties and processes low 0 train +5449 AR6_WGII 1727 3 Ocean and coastal ecosystems in SWS are sensitive to upwelling intensity, which affects the abundance, diversity, physiology and survivorship of coastal species high 2 train +5450 AR6_WGII 1727 4 Increasing radiation and temperatures and reduced precipitation, in conjunction with increased nutrient load, have increased HAB events, producing massive fauna mortalities high 2 train +5451 AR6_WGII 1727 5 Multiple resources subjected to fisheries and aquaculture are highly vulnerable to storms, alluvial disasters, ocean warming, ocean acidification, increasing ENSO extreme events and lower oxygen availability high 2 train +5452 AR6_WGII 1727 6 Ocean and coastal ecosystems, especially EEZs, will be highly impacted by climate change in the near and long term high 2 train +5453 AR6_WGII 1727 9 Increasing temperatures and decreasing precipitation are expected to impact the agriculture sector (i.e., fruits crops and forests) across the entire sub-region, with the largest impacts in the northern and central zone high 2 train +5454 AR6_WGII 1727 10 Observed impacts and future projections warn that increasing temperatures and decreasing precipitation will largely impact water demand by agricultural sectors high 2 train +5455 AR6_WGII 1727 13 Increasing temperature and heat extreme events in cities have increased the demand for water, damage to urban infrastructure (Monsalves- Gavilán et al., 2013) and accelerated ageing and death of trees high 2 train +5456 AR6_WGII 1728 11 An increase in the intensity and frequency of hot extremes and a decrease in the intensity and frequency of cold extremes are projected to be likely (WGI AR6 Table 11.13, Seneviratne et al., 2021); CMIP6 models project an increase in the intensity and frequency of heavy precipitation medium 1 train +5457 AR6_WGII 1728 12 It is expected that an increase in the intensity of heavy precipitation, droughts and fire weather will intensify through the 21st century in SSA, but mean wind will decrease medium 1 train +5458 AR6_WGII 1728 15 The observed area and the elevation changes indicate that the Echaurren Norte glacier may disappear in the coming years if negative mass balance rates prevail medium 1 train +5459 AR6_WGII 1728 20 Thus, if evapotranspiration and drought stress increase as temperature increases and rainfall decreases in water-limited ecosystems, a greater exposure of ranchers to a reduction in stocking rate and, therefore, family income would be expected medium 1 train +5460 AR6_WGII 1728 23 Vaca Muerta is the major region in SA, where those techniques are used to extract oil and gas, and this will lead to an exacerbation of current water scarcity issues and to competition with irrigated agriculture (Rosa and D’Odorico, 2019), which in the context of drought may exacerbate socioenvironmental conflicts medium 1 train +5461 AR6_WGII 1729 13 Most calving glaciers in the southern Patagonia ice field retreated during the last century high 2 train +5462 AR6_WGII 1729 21 Suitable areas for meadows (very productive areas for livestock production) will decrease by 7.85% by 2050 given predicted changes in climate low 0 train +5463 AR6_WGII 1729 26 The Patagonian Steppe may suffer pronounced expansion in invasive species’ ranges under climate change low 0 train +5464 AR6_WGII 1734 2 A synthesis of these impacts (Figure 12.10) indicates the following: Climate change has or will have a major impact on the observed and future decline of Andean glaciers and snow high 2 train +5465 AR6_WGII 1734 6 Many aspects and assets of ocean and coastal ecosystems (e.g., mangroves, coral reefs, saltmarshes) were identified as being strongly impacted by climate change, both for observed and future periods high 2 train +5466 AR6_WGII 1734 11 Assessment of key observed and projected impacts and risks shows that in the CSA region several systems are already approaching critical thresholds under current warming levels, in particular glaciers in the Andes, coral reefs in CA high 2 train +5467 AR6_WGII 1734 12 Some systems could cross these thresholds with different levels of reversibility depending on the degree of future warming, that is, glaciers in the Andes and coral reefs in CA, which will show partial but irreversible loss already under low levels of warming (RCP2.6) high 2 train +5468 AR6_WGII 1736 23 In particular, the high vulnerability of large populations, infrastructure and service systems, such as health, food and energy production and supply, are important factors, along with high levels of inequality and poor governance, for creating and increasing key risks high 2 train +5469 AR6_WGII 1737 1 IKLK play an important role in adaptation and are vital components of many socioecological systems, while also being threatened by climate change high 2 train +5470 AR6_WGII 1737 8 Investing in actions aimed at protection, restoration and the sustainable use of biodiversity and ecosystems represents a good approach to maintaining critical ecosystem services and constitutes part of a common strategy for adaptation, mitigation and disaster risk reduction in the region high 2 train +5471 AR6_WGII 1738 1 Protected area in CSA are underfunded very high 3 train +5472 AR6_WGII 1738 12 In their NCs, NDCs and/or NAPs, most countries identified inadequate financing and access to technology as barriers to adaptation relevant to terrestrial and freshwater socioecosystems high 2 train +5473 AR6_WGII 1738 28 Poverty is a driver of climate-change risk, while the sustainable use of ecosystems fosters adaptation (Kasecker et al., 2018) high 2 train +5474 AR6_WGII 1739 5 There is, therefore, great potential for terrestrial and freshwater ecosystem adaptation to climate change in CSA, provided the right incentives and sociocultural protective measures are in place high 2 train +5475 AR6_WGII 1739 14 Nevertheless, research on Latin American forests tend to focus on mitigation, while studies on agriculture are usually oriented towards adaptation high 2 train +5476 AR6_WGII 1739 17 They increase ecosystems’ adaptive potential, reducing climate risk and delivering numerous ecosystem services and sustainable development benefits while playing an important role in climate-change mitigation high 2 train +5477 AR6_WGII 1739 21 In addition to better managing and expanding protected area networks, other effective area-based conservation measures (OECMs), recently defined by the Parties to the Convention on Biological Diversity (Dudley et al., 2018), could also enhance ecosystem resilience low 0 train +5478 AR6_WGII 1739 24 Large-scale restoration is also seen as pivotal to limiting both climate change (IPCC, 2019a) and species extinction (IPBES, 2018a) very high 3 train +5479 AR6_WGII 1739 28 Projections for CSA ocean and coastal ecosystems warn about significant and negative impacts high 2 train +5480 AR6_WGII 1740 11 Adaptation measures in mangrove ecosystems are mainly focused on the application of EbA strategies high 2 train +5481 AR6_WGII 1740 13 The use of territorial planning tools, the promotion of sustainable resource exploitation, the adherence to certification schemes and the implementation of management instruments, such as ecosystem-based management (EbM), followed by the use of an integrated coastal zone management, coastal marine spatial planning and capacity building, ecological risk assessments, have been the main strategies used to ensure the sustainability of marine resources in fisheries across EEZs of CSA high 2 train +5482 AR6_WGII 1740 16 However, some important advances are visible in Colombian Pacific areas with coral reefs (new conservation plans, research monitoring and conservation practices) low 0 train +5483 AR6_WGII 1740 17 In Panama, actions taken have allowed the protection of a high number of marine areas with coral reefs, as well as the incorporation of management approaches that include several sectors such as fisheries, tourism, coral protection and coral conservation low 0 train +5484 AR6_WGII 1740 18 In the case of Costa Rica, 80% of coral habitats are located inside of MPAs, multiple research coral-related activities have been performed, and several training activities have favoured the engagement of the local community in their protection against climate and non-climate hazards low 0 train +5485 AR6_WGII 1740 20 In Brazil, between 75–84% of mangroves are under some level of protection which has improved the forest structures, and multiple research programmes (e.g., Mangrove Dynamics and Management, MADAM, and ‘GEF-Mangle’) have been developed medium 1 train +5486 AR6_WGII 1741 1 In Colombia, research projects (e.g., Mangroves of Colombia Projects, MCP), the installation of a geographic information system for mangroves (e.g., SIGMA Sistema de Información para la Gestión de los Manglares en Colombia), surveillance monitoring plans (e.g., EGRETTA Herramientas para el Control y Vigilancia de los Manglares), and the establishment of protected areas have contributed to decrease loss of the mangrove forest high 2 train +5487 AR6_WGII 1741 6 In addition, MPAs in Chile play a key role in climate-change adaptation for fisheries medium 1 train +5488 AR6_WGII 1741 9 Uruguay is incorporating stakeholders in its climate- change adaptation strategies low 0 train +5489 AR6_WGII 1742 3 However, the amount and type of adaptation goals differ enormously from country to country (Figure 12.12).12.5.2.4 Limits and Barriers to Adaptation in Ocean and Coastal Ecosystems Although current NAPs and many other actions and strategies focus on improving the conservation and restoration of ocean and coastal ecosystems, as well as the suitability of marine resources throughout CSA, these measures are still not able to reduce the vulnerability and sensitivity of these ecosystems to climate-change hazards high 2 train +5490 AR6_WGII 1742 4 There is high confidence that sandy beach ecosystems of CSA countries have suffered significant losses of dunes as a consequence of the construction of infrastructures that have caused interruptions in the natural dynamic of beaches, reducing protection against tides, waves, extreme events or tsunamis high 2 train +5491 AR6_WGII 1743 19 These include the Dry Corridor in CA, coastal areas of Peru (SWS) and northern Chile (SWS), the Bolivian- Peruvian Altiplano (NWS, SAM), the Dry Andes of Central Chile (SWS), Western Argentina and Chaco in northwestern Paraguay (SES) and Sertão in northeastern Brazil (NES) high 2 train +5492 AR6_WGII 1743 20 In NWS and SWS, downstream areas are increasingly affected by decreasing and unreliable river runoff due to rapid glacier shrinkage high 2 train +5493 AR6_WGII 1743 24 River fragmentation and corresponding loss of habitat connectivity due to dam constructions have been described for, for example, the NSA, SAM, NES and SES (high confidence) (Grill et al., 2015; Anderson et al., 2018a), with important implications for freshwater biota, such as fish migration medium 1 train +5494 AR6_WGII 1743 25 Furthermore, examples in, for instance, NWS (Carey et al., 2012; Duarte-Abadía et al., 2015; Hommes and Boelens, 2018) and SWS (Muñoz et al., 2019b) showcase unresolved water-related conflicts between local villagers, peasant communities, hydropower operators and governmental institutions in a context of distrust and lack of water governance high 2 train +5495 AR6_WGII 1743 27 Declining water quality can be observed, for example, due to intense agricultural and industrial activities in SWS, SES and SSA (medium confidence) (Mekonnen et al., 2015; Gomez et al., 2021), mining in Andean headwaters (NWS, SWS and Western SAM) and tropical lowlands (eastern SAM and NSA) medium 1 train +5496 AR6_WGII 1744 1 Water scarcity risks are projected to affect a growing number of people in the near and mid-term future in view of growing water demand in most regions (medium confidence: medium evidence, high agreement) (Veldkamp et al., 2017; Schoolmeester et al., 2018; Viviroli et al., 2020), expected precipitation reductions in western and northern SAM and SWS (medium confidence: medium evidence, medium agreement) (Neukom et al., 2015; Schoolmeester et al., 2018), substantial vanishing of glacier extent in NWS, SAM and SWS (Table SM12.6; Rabatel et al., 2018; Vuille et al., 2018; Cuesta et al., 2019; Drenkhan et al., 2019) and increasing evaporation rates in CA medium 1 train +5497 AR6_WGII 1744 2 Furthermore, flood risk is a serious concern (Arnell et al., 2016) and expected to increase, especially in NWS, SAM, SES and SWS in the mid- and long-term future high 2 train +5498 AR6_WGII 1744 4 In Central and South America, about 26% (130 million people) of the population have no access to safe drinking water, and strong disparities prevail regarding its spatial distribution; for example, in Chile, 99% of the population have access, compared to 50% in Peru, 73% in Colombia, 52% in Nicaragua or 56% in Guatemala high 2 test +5499 AR6_WGII 1744 6 The most vulnerable people belong to low- income groups in rural areas and informal settlements of large urban areas high 2 train +5500 AR6_WGII 1744 8 The combination of (seasonally) reduced water supply, growing water demand, declining water quality, ecosystem deterioration and habitat loss and low water governance could lead to increasing competition and conflict associated with high economic losses high 2 train +5501 AR6_WGII 1744 9 This situation threatens human water security in the long term and poses an increasing risk to adaptation success in CSA high 2 train +5502 AR6_WGII 1744 13 However, ‘hard path’ interventions are now strongly contested because negative effects exacerbate local water conflicts (Carey et al., 2012; Boelens et al., 2019; Drenkhan et al., 2019), potentially leading to increasing water demand, vulnerabilities and water shortage risks (Di Baldassarre et al., 2018), thereby limiting adaptive capacity high 2 train +5503 AR6_WGII 1744 24 In recent years, the inclusion of IKLK in current adaptation baselines has attracted increasing attention, particularly in regions with a high share of Indigenous Peoples (NWS, SAN, SWS, NSA) high 2 train +5504 AR6_WGII 1746 6 However, current realities in all regions show major challenges in implementing IWRM mechanisms and policies, related but not limited to political and institutional instabilities, governance structures, fragmented service provision, lack of economies of scale and scope, corruption and social conflicts high 2 train +5505 AR6_WGII 1746 7 Many water-related conflicts in CSA are rooted in inequitable water governance that excludes water users from decisions on water allocation high 2 train +5506 AR6_WGII 1746 12 Nevertheless, only a few experiences have been evaluated as successful due to insufficient implementation, low decision-making ability of some stakeholder groups and poor evidence-based approaches medium 1 train +5507 AR6_WGII 1746 18 These funding deficits set important limitations on future water provision, adaptation to changing water resources and the achievement of the SDGs by 2030 high 2 train +5508 AR6_WGII 1746 20 In many parts of CSA the level of success of adaptation measures depends largely on the governance of projects and stakeholder-based processes and is closely related to their effectiveness, efficiency, social equity and sociopolitical legitimacy high 2 train +5509 AR6_WGII 1746 21 Several PES experiences across CSA have been described as successful measures for watershed conservation and adaptation high 2 train +5510 AR6_WGII 1746 29 Most barriers to advance adaptation in CSA correspond to soft limits associated with missing links of science–society–policy processes, institutional fragilities, pronounced hierarchies, unequal power relations and top-down water governance regimes high 2 train +5511 AR6_WGII 1747 10 The direct social and economic consequences for the sector are evident in CA’s so-called Dry Corridor, with a growing dependence on food imports (Porkka et al., 2013), and these degrees of dependency make the region more vulnerable to price variability, climatic conditions (Bren d’Amour et al., 2016; ECLAC, 2018) and, therefore, to food insecurity in the absence of adaptation actions high 2 train +5512 AR6_WGII 1747 12 Meanwhile, the more temperate part of SA in the south is projected to have agricultural production surpluses low 0 train +5513 AR6_WGII 1748 5 Limited information regarding cost-benefit analyses of adaptation is available in the region and regarding avoiding maladaptation effects and promoting site-specific and dynamic adaptation options considering available technologies medium 1 train +5514 AR6_WGII 1748 6 Climate information services has an important role in climate-change adaptation and there is a recognised gap between climate science and farmers high 2 train +5515 AR6_WGII 1748 10 Barriers to financial access are present in the region, restricting effective adaptation to extreme weather events high 2 train +5516 AR6_WGII 1748 15 Lack of finance and proper infrastructure is compounded by limited knowledge of sustainable farming practices and high rates of financial illiteracy high 2 train +5517 AR6_WGII 1748 21 Agricultural technologies are not necessarily changing, but the economic activity is shifting to accommodate increasing climate variation and adapt to changes in water availability and ideal growing conditions high 2 train +5518 AR6_WGII 1750 3 Crop–livestock–forestry systems are also important for climate-change adaptation as they provide multiple benefits, including the coproduction of food, animal feed, organic fertilizers and soil organic carbon sequestration (Sharma et al., 2016; Rodríguez et al., 2021), achieving mitigation and adaptation goals high 2 train +5519 AR6_WGII 1750 5 Agroforestry has also proven to have protective benefits to obtain more stable, less fluctuating yields due to climate-related damage in coffee production high 2 train +5520 AR6_WGII 1750 10 On one hand, they ensure the conservation of a very rich agrobiodiversity that is likely to meet the challenges of climate change high 2 train +5521 AR6_WGII 1750 11 In the Andes, ancient technologies increased the quantity of crops produced and made it possible to cope with climatic changes and water scarcity, while nutrition conditions were improved high 2 train +5522 AR6_WGII 1751 6 The ability to enrol stakeholders and include community-based initiatives can determine adaptation success, particularly considering their impact in the decision-making arena high 2 train +5523 AR6_WGII 1751 26 Slum upgrading and built-environment interventions (housing and infrastructure improvement and provision) in informal settlements can enhance adaptation high 2 train +5524 AR6_WGII 1752 1 The climate adaptation plans of several large CSA cities include efficient land use and occupation planning and urban control systems (comprising regulation, monitoring), fostering the articulation with housing and environmental policy (by means of intersectoral and multi- level governance), inhibiting and reducing the occupation of risk areas (mainly flooding and landslides risks); increasing population density in areas already served by infrastructure; expanding slum urbanisation and technical assistance programmes to improve and expand social housing high 2 train +5525 AR6_WGII 1752 4 Those initiatives, associated with a housing policy that guarantees access to land and decent housing, represent a comprehensive intervention in vulnerable neighbourhoods for their adaptation to climate change, and CbA (community-based adaptation) strategies, including housing self-management and the participation of cooperatives, demonstrate the need and opportunity to transition to a transformative urban agenda that encompasses sustainable development, poverty reduction, disaster-risk reduction, climate- change adaptation and climate-change mitigation high 2 train +5526 AR6_WGII 1752 5 Several large cities are implementing municipal risk management plans and management and restoration plans for hydrologically relevant areas, considering threats of drought and heat waves, integrated watershed management and flood control programmes high 2 train +5527 AR6_WGII 1752 7 EWS and the use of mapping tools as undertaken in La Paz proved to be an effective adaptation measure in the face of increasing hydro-climatic extreme events (Aparicio-Effen et al., 2018).12.5.5.3.2 Green and Grey Infrastructure Hybrid solutions, combining green and grey infrastructure (GGI), have been adopted for better efficiency in flood control (Ahmed et al., 2019; Drosou et al., 2019; Romero-Duque et al., 2020), sanitation, water scarcity, landslide prevention and coastal protection high 2 train +5528 AR6_WGII 1752 12 Managing water in cities in an adaptive way has been central to reducing impacts such as floods and contributes to water security high 2 train +5529 AR6_WGII 1752 13 Many cities facing frequent heavy storms that impact mostly underprivileged communities, slums and vulnerable areas could benefit from integrated NbS for disaster risk reduction and adaptation high 2 train +5530 AR6_WGII 1752 15 In several municipal climate plans, NbSs were introduced mainly to enhance rainwater management, reduce energy consumption and urban heat areas, improve water quality, prevent landslides and set aside green areas high 2 train +5531 AR6_WGII 1753 7 The main proposed actions still consider the traditional approach in improving the hydraulic capacity of urban drainage systems as an adaptive measure high 2 train +5532 AR6_WGII 1754 1 More recently, climate services, such as EWSs and forecast models, have been promoted for the health sector (WHO and WMO, 2012, 2016; WMO, 2014; Thomson and Mason, 2018) and are an important adaptation measure to reduce the impacts of climate on health high 2 train +5533 AR6_WGII 1754 6 Epidemic forecast tools are an example of an adaptation measure being developed and/or implemented in this region high 2 train +5534 AR6_WGII 1754 11 One of the most promising climate services for the health sector are heat and cold early-warning and alert systems medium 1 train +5535 AR6_WGII 1754 15 The public dissemination of climate–health warnings via bulletins, websites and other outlets can be an adaptation measure to address climate change and weather variability to reduce health risks high 2 train +5536 AR6_WGII 1754 29 Climate and health observatories represent a promising strategy that is being developed at sub-national, national (e.g., Brazil, Argentina) and regional levels high 2 train +5537 AR6_WGII 1755 6 A Climate Change Vulnerability Index was used to generate vulnerability maps for countries of the Latin American and Caribbean region (Vörösmarty et al., 2013; CAF , 2014).12.5.6.4 Other Adaptation Actions Diverse adaptation measures are being implemented through public policies, private household responses and communal management that directly or indirectly reduce the impacts of climate change on human health high 2 train +5538 AR6_WGII 1755 10 Adaptation strategies implemented by the public sector include a diverse suite of strategies ranging from the creation of green spaces in urban areas, relocation of families located in disaster-prone areas, ecosystem restoration and improved access to clean water, among many others high 2 train +5539 AR6_WGII 1756 9 Sustainable financing and political support are needed to ensure the functioning of disaster EWSs high 2 train +5540 AR6_WGII 1757 3 Those already suffering are losing their livelihoods and reducing their development options; poor populations and countries are more vulnerable and have lower adaptive capacity to climate change compared to rich ones very high 3 train +5541 AR6_WGII 1757 6 Climate change has differential impacts, and even within a household there may be important differences in relation to age, gender, health and disability; these factors may intersect with one another high 2 train +5542 AR6_WGII 1757 9 Poor and vulnerable groups are disproportionately affected in negative ways by climate change (Section 8.2.1.4; Section 8.2.2.3; SR15 Section 5.2 and Section 5.2.1, Roy et al., 2018) due to physical exposure derived from their place of residence or work, illiteracy, low income and skills, political and institutional marginalisation tied to a lack of recognition of informal settlements and employment, poor access to good-quality services and infrastructure, resources and information and other factors very high 3 train +5543 AR6_WGII 1757 11 Researchers argue that poverty is mischaracterised and has multiple dimensions (Castán Broto and Bulkeley, 2013) (Section 8.1.1), that biodiversity loss, climate change and pollution will undermine efforts on 80% of assessed SDG targets, that biodiversity and climate change must be tackled together (Pörtner et al., 2021; United Nations Environment Programme, 2021) and due to the COVID crisis LAC countries have made uneven progress in terms of meeting SDGs high 2 train +5544 AR6_WGII 1757 14 Climate change will be a major obstacle in reducing poverty high 2 train +5545 AR6_WGII 1758 1 This suggests that caution in framing climate adaptation and resilience related interventions equally benefits everyone (high confidence) (Brown, 2014; Chu et al., 2016; Connolly, 2019; Romero-Lankao and Gnatz, 2019; Johnson et al., 2021) and that equality and justice dimensions should be incorporated into decision- making very high 3 train +5546 AR6_WGII 1758 2 Poor rural households in marginal territories that have a low productive potential and/or that are far from markets and infrastructure are highly vulnerable to climate-change impacts and could easily fall into poverty-environment traps high 2 train +5547 AR6_WGII 1758 12 Poverty and disaster risk reduction interlinked with climate-change adaptation share a focus on identifying and acting on local risks and their root causes, even though they view risk through different lenses very high 3 train +5548 AR6_WGII 1758 13 Construction of climate knowledge and risk perceptions affect decision-making to define implementation priorities, but the poor are less able to cope with and adapt so as to avoid so-called adaptation injustices high 2 train +5549 AR6_WGII 1758 18 Risks are seldom distributed equally, highlighting socioeconomic inequalities and governance failures high 2 train +5550 AR6_WGII 1760 7 In Brazil, for example, it would be difficult to clearly separate climate-change adaptation and urban policies high 2 train +5551 AR6_WGII 1760 11 Because there are asymmetries among populations regarding the vulnerability and benefits of adaptation, along the lines of gender, age, socioeconomic conditions and ethnicity, it has been noticed that adaptation policies and programmes must be adequate to diverse conditions and actors very high 3 train +5552 AR6_WGII 1760 17 IKLK participation is thought to be more considered in adaptation policies because it yields good results high 2 train +5553 AR6_WGII 1760 19 Approaches integrating IK with more formal sciences, to address research and policies, have improved adaptation processes, but they carry their own complications high 2 train +5554 AR6_WGII 1761 3 In contrast, top-down approaches without citizen or community participation have shown to be less effective high 2 train +5555 AR6_WGII 1761 9 In CSA, the awareness of climate change as a threat is increasing, a situation related to growth in climate justice activism and to the occurrence of extreme weather events of all kinds high 2 train +5556 AR6_WGII 1761 16 In rural areas, it largely relates to temperature rise and changes in rainfall patterns, changes in agriculture (pests, calendars), biodiversity loss, solar radiation or changes in the oceans, and their impacts are sometimes related or even more attributed to socioeconomic and environmental drivers, as well as to negative financial outcomes high 2 train +5557 AR6_WGII 1761 26 Often, for several economic and social reasons, women have less capacity to adapt, further widening structural gender gaps high 2 train +5558 AR6_WGII 1762 2 Recent studies emphasise that a gender approach to social inequalities ought to move beyond just looking at men and women as experiencing impacts in a differentiated manner; rather, an intersectional analysis illuminates how different individuals and groups relate differently to climate change due to their situatedness in power structures based on context-specific and dynamic social categorisations high 2 train +5559 AR6_WGII 1762 5 Particular constellations of race, gender, class, age or nationality reveal more complex realities high 2 train +5560 AR6_WGII 1762 21 Migration to cities can mean opportunities for migrants and for urban areas, but it can also worsen existing problems, as urban poor people can become even more exposed and vulnerable, and the pressure on urban capacities may not be well absorbed high 2 train +5561 AR6_WGII 1762 29 Some areas are more likely to generate climatic migration: the Andes, the dry areas of Amazonia, northern Brazil and northern countries in CA high 2 train +5562 AR6_WGII 1763 32 Climate financing in CSA is mainly focused on mitigation actions high 2 train +5563 AR6_WGII 1766 6 Therefore, there is a need to design adequate programmes and allocate resources for agroforestry system implementation and technical assistance and training medium 1 train +5564 AR6_WGII 1766 10 These included the legal and regulatory feasibility, institutional capacity and administrative feasibility, transparency and political acceptability high 2 train +5565 AR6_WGII 1766 21 This may be a potential barrier to the adaptive approach of multi-use water storage medium 1 train +5566 AR6_WGII 1767 2 However, limitations in technical assistance and funding for specific key technologies and training may act as a barrier for EbA adoption medium 1 train +5567 AR6_WGII 1767 19 Many of the new neighbourhoods were established through land invasions or informal markets, in many cases over steep slopes, in water sources and agricultural or conservation areas high 2 train +5568 AR6_WGII 1767 22 Two impacts related to warmer air conditions are the displacement of the freezing line currently placed at 5100 MASL (Basantes-Serrano et al., 2016), followed by glacier retreat and the upward displacement of mountainous ecosystems very high 3 train +5569 AR6_WGII 1767 24 Other important climatic hazards and impacts are the increase of solar radiation, the heat island effect and fires high 2 train +5570 AR6_WGII 1768 2 The programme has direct impacts on nutrition, generation of work for women, production of healthy food, reduction of runoff, recycling of organic waste and social cohesion, among others very high 3 train +5571 AR6_WGII 1768 9 Along the same lines, these areas constitute a key element to increase the adaptive capacity of rural livelihoods and contribute to mitigating climate change through landscape restoration, sustainable production and forest conservation high 2 train +5572 AR6_WGII 1768 12 A major pending environmental issue is air pollution; a high level of pollutants affects the city in general and especially the most vulnerable groups high 2 train +5573 AR6_WGII 1768 21 ADEs are characterised by their increased fertility in relation to adjacent soils, with high contents of organic carbon (C) (mainly as charcoal) as well as inorganic nutrients, especially phosphorus (P) and calcium (Ca) and high carbon/nitrogen ratios high 2 train +5574 AR6_WGII 1768 23 Charcoal content is a key indicator of pre-Hispanic fire activity and sedentary occupation, which is evidence of the anthropic origin of these soils high 2 train +5575 AR6_WGII 1769 18 Located on the Pacific coast with more than 10 million inhabitants, it suffers from flooding, mudslide disasters and water stress, and is more frequently affected by heavy rain peak events (1970, 1987, 1998, 2012, 2014, 2015 and 2017) very high 3 train +5576 AR6_WGII 1769 20 Climate-change scenarios were ignored or denied, particularly when budget allocations for preventive actions were necessary high 2 train +5577 AR6_WGII 1770 9 Impacts on human settlements, agriculture and ecosystems result from both excess and reduced precipitation high 2 train +5578 AR6_WGII 1770 14 Its effects were felt in different parts of the world, and Guatemala and the rest of CA experienced intense water scarcity due to a significant reduction in rainfall high 2 train +5579 AR6_WGII 1770 19 A lack of regulation, coordination mechanisms, information and other elements of water governance was the root cause of the problem, exacerbated by the drier conditions during the intense El Niño event, resulting in the intensification of an existing conflict high 2 train +5580 AR6_WGII 1770 30 Local governance was also essential for the implementation of flood risk management actions high 2 train +5581 AR6_WGII 1771 1 Although some may argue that water governance is in the realm of development, it has made contributions in reducing direct and indirect impacts of climate events and, therefore, can be seen as a key element for climate adaptation high 2 train +5582 AR6_WGII 1774 7 Poverty and extreme poverty rates are higher among children, young people, women, Indigenous Peoples and migrant and rural populations, but urban extreme poverty is also growing very high 3 train +5583 AR6_WGII 1774 9 Most countries in CA are already ranked as the highest risk level worldwide due to the region’s high vulnerability to climate change and low adaptive capacity; the lack of climate data and proper downscaling are challenging the adaptation process high 2 train +5584 AR6_WGII 1774 10 Many extreme events are already impacting the region and are projected to intensify; such events include warming temperatures and dryness, SLR, coastal erosion and ocean and lake acidification, resulting in coral bleaching and an increasing frequency and severity of droughts in some regions, with a concomitant decrease in water supply, which impact agricultural production, traditional fishing, food security and human health high 2 train +5585 AR6_WGII 1774 15 Urban areas in the region are vulnerable for many reasons, notably high rates of poverty and informality, poor and unevenly distributed infrastructure, housing deficits and the recurrent occupation of risk areas high 2 train +5586 AR6_WGII 1774 22 The high poverty level increases the region’s vulnerability to droughts, both in cities and rural areas, where people already suffer from natural water scarcity high 2 train +5587 AR6_WGII 1774 24 IKLK play an important role in adaptation but are also threatened by climate change high 2 train +5588 AR6_WGII 1774 26 Inadequate access to finance and technology is widely identified as an adaptation barrier high 2 train +5589 AR6_WGII 1774 28 Subsistence farmers and the urban poor are expected to be the most impacted by droughts and variable rainfall in the region high 2 train +5590 AR6_WGII 1774 32 Internal and international migrations and displacements are expected to increase high 2 train +5591 AR6_WGII 1774 33 Climatic drivers, such as droughts, tropical storms and hurricanes, heavy rains and floods, interact with social, political, geopolitical and economic drivers high 2 train +5592 AR6_WGII 1774 41 The lack of climate data and proper downscaling, weak governance, obstacles to financing, and inequality constrain the adaptation process high 2 train +5593 AR6_WGII 1775 3 The number and type of adaptation measures for ocean and coastal ecosystems and their contributions to humans are very different among CSA countries, which highlights the number of measures related to increasing scientific research and monitoring followed by the conservation of biodiversity and changes in legislation high 2 train +5594 AR6_WGII 1775 5 In the water sector, a lack of systematic analysis and evaluation of adaptation measures predominates, although important progress has been made since AR5 in terms of understanding the interlinkages among climate change, human vulnerabilities, governance, policies and adaptation success high 2 train +5595 AR6_WGII 1775 7 Climate change poses several challenges to the agri-food sector, impacting agricultural production and productivity and posing a risk to food security and the economy high 2 train +5596 AR6_WGII 1775 8 Adapting agriculture while conserving the environment represents a challenge for sustainable and resilient food production high 2 train +5597 AR6_WGII 1775 9 Adaptation in the region presents persistent barriers and limitations (Table 12.8) associated with investments and knowledge gaps medium 1 train +5598 AR6_WGII 1775 11 Urban adaptation is limited by financing constraints, weak intersectoral and multi-level governance and deficits in the housing and infrastructure sectors, the overcoming of which represents an opportunity for transformative adaptation high 2 train +5599 AR6_WGII 1775 12 Short- term interventions are more common than long-term planning high 2 train +5600 AR6_WGII 1775 14 Initiatives in social housing focus on reducing risk and overcoming urgent deficits but also adding to a transformative adaptation pathway high 2 train +5601 AR6_WGII 1775 15 Hybrid (green- grey) infrastructure has been adopted for better efficiency in flood control, sanitation, water scarcity and landslide prevention and coastal protection high 2 train +5602 AR6_WGII 1775 16 NbS, including GI and EbA, are increasing in urban areas high 2 train +5603 AR6_WGII 1775 17 The integration of transport and land use plans and the improvement of public transport are key to urban adaptation; mitigation prevails over adaptation in the sector (high confidence).There is a growing body of evidence that climate variability and climate change are causing harm to human health in CSA—including the increasing transmission of vector-borne and zoonotic diseases, heat stress, respiratory illness associated with fires, food and water insecurity associated with drought, among others medium 1 train +5604 AR6_WGII 1775 19 However, institutional barriers (limited resources, administrative feasibility and political mandates) need to be addressed to ensure the sustained implementation of adaptation strategies high 2 train +5605 AR6_WGII 1775 20 Poor and vulnerable groups exert limited political influence; the fewer channels and opportunities that exist to participate in decision-making and policymaking make these groups less able to leverage government support to invest in adaptation measures very high 3 train +5606 AR6_WGII 1775 23 Social organisation, participation and governance reconfiguration are essential for building climate resilience very high 3 train +5607 AR6_WGII 1775 24 Social organisation, participation, governance, education and communications to increase perception and knowledge are essential for building the resilience to adapt and overcome expected and unexpected climate impacts very high 3 train +5608 AR6_WGII 1775 25 The focus on inclusion and enrolling of the full range of actors in adaptation processes, including vulnerable populations, have yielded good results in the region high 2 train +5609 AR6_WGII 1775 26 However, existing poverty and inequality, imbalances in power relations, corruption, weak governance and institutions, structural problems and high levels of risk tolerance may reinforce poverty and inequality cycles high 2 train +5610 AR6_WGII 1775 29 It is widely recognised that climate adaptation measures, if carefully selected considering coupled human-environment systems, will provide significant contributions to the sustainable development pathways of the region and to achieve the SDGs if implemented together with comprehensive strategies to reduce poverty, inequality and risks high 2 train +5611 AR6_WGII 1830 2 Since AR5, there has been a substantial increase in detected or attributed impacts of climate change in Europe, including extreme events high 2 train +5612 AR6_WGII 1830 3 Impacts of compound hazards of warming and precipitation have become more frequent medium 1 train +5613 AR6_WGII 1830 4 Climate change has resulted in losses of, and damages to, people, ecosystems, food systems, infrastructure, energy and water availability, public health and the economy very high 3 train +5614 AR6_WGII 1830 5 As impacts vary both across and within European regions, sectors, and societal groups (high confidence), inequalities have deepened medium 1 train +5615 AR6_WGII 1830 7 Traditional lifestyles, for example in the European Arctic, are threatened already high 2 train +5616 AR6_WGII 1830 8 Poor households have lower capacity to adapt to, and recover from, impacts medium 1 train +5617 AR6_WGII 1830 9 The range of options available to deal with climate-change impacts has increased in most of Europe since AR5 high 2 train +5618 AR6_WGII 1830 10 Growing public perception and adaptation knowledge in public and private sectors, the increasing number of policy and legal frameworks, and dedicated spending on adaptation are all clear indications that the availability of options has expanded high 2 train +5619 AR6_WGII 1830 13 Many cities are taking adaptation action, but with large differences in level of ambition and implementation high 2 train +5620 AR6_WGII 1830 14 Observed adaptation actions are largely incremental with only a few examples of local transformative action; adaptation actions have demonstrated different degrees of effectiveness in reducing impacts and feasibility of implementation high 2 test +5621 AR6_WGII 1830 18 A gap remains between planning and implementation of adaptation action high 2 train +5622 AR6_WGII 1830 22 Warming in Europe will continue to rise faster than the global mean, widening risk disparities across Europe in the 21st century high 2 train +5623 AR6_WGII 1830 24 Four key risks (KR) have been identified for Europe, with most becoming more severe at 2°C global warming levels (GWL) compared with 1.5°C GWL in scenarios with low to medium adaptation high 2 train +5624 AR6_WGII 1830 25 From 3°C GWL and even with high adaptation, severe risks remain for many sectors in Europe high 2 train +5625 AR6_WGII 1830 27 KR1: The number of deaths and people at risk of heat stress will increase two- to threefold at 3°C compared with 1.5°C GWL high 2 train +5626 AR6_WGII 1830 28 Risk consequences will become severe more rapidly in Southern and Western Central Europe and urban areas high 2 train +5627 AR6_WGII 1830 29 Thermal comfort hours during summer will decrease significantly high 2 train +5628 AR6_WGII 1830 30 Above 3°C GWL, there are limits to the adaptation potential of people and existing health systems, particularly in Southern Europe, Eastern Europe and areas where health systems are under pressure high 2 train +5629 AR6_WGII 1830 31 KR1: Warming will decrease suitable habitat space for current terrestrial and marine ecosystems and irreversibly change their composition, increasing in severity above 2°C GWL very high 3 train +5630 AR6_WGII 1830 32 Fire-prone areas are projected to expand across Europe, threatening biodiversity and carbon sinks medium 1 train +5631 AR6_WGII 1830 34 Trade-offs between adaptation and mitigation options (e.g., coastal infrastructure and NbS) will result in risks for the integrity and function of ecosystems medium 1 train +5632 AR6_WGII 1830 35 KR2: Due to a combination of heat and drought, substantive agricultural production losses are projected for most European areas over the 21st century, which will not be offset by gains in Northern Europe high 2 train +5633 AR6_WGII 1830 37 Yields of some crops (e.g., wheat) may increase in Northern Europe if warming does not exceed 2°C medium 1 train +5634 AR6_WGII 1831 1 KR3: Risk of water scarcity will become high at 1.5°C and very high at 3°C GWL in Southern Europe (high confidence), and increase from moderate to high in Western Central Europe medium 1 train +5635 AR6_WGII 1831 2 In Southern Europe, more than a third of the population will be exposed to water scarcity at 2°C GWL; under 3°C GWL, this risk will double, and significant economic losses in water- and energy-dependent sectors may arise medium 1 train +5636 AR6_WGII 1831 5 KR4: Due to warming, changes in precipitation and sea level rise (SLR), risks to people and infrastructures from coastal, riverine and pluvial flooding will increase in Europe high 2 train +5637 AR6_WGII 1831 6 Risks of inundation and extreme flooding will increase with the accelerating pace of SLR along Europe’s coasts high 2 train +5638 AR6_WGII 1831 8 Coastal flood damage is projected to increase at least tenfold by the end of the 21st century, and even more or earlier with current adaptation and mitigation high 2 train +5639 AR6_WGII 1831 10 European cities are hotspots for multiple risks of increasing temperatures and extreme heat, floods and droughts high 2 train +5640 AR6_WGII 1831 11 Warming beyond 2°C GWL is projected to result in widespread impacts on infrastructure and businesses high 2 train +5641 AR6_WGII 1831 12 These impacts include increased risks for energy supply (high confidence) and transport infrastructure (medium confidence), increases in air conditioning needs (very high confidence) and high water demand high 2 train +5642 AR6_WGII 1831 13 European regions are affected by multiple key risks, with more severe consequences in the south than in the north high 2 train +5643 AR6_WGII 1831 15 There is high confidence that consequences for socioeconomic and natural systems will be substantial: the number of people exposed to KRs and economic losses are projected to at least double at 3°C GWL compared with 1.5°C GWL medium 1 train +5644 AR6_WGII 1831 16 The risks resulting from changes in climatic and non-climatic drivers in many sectors is a key gap in knowledge high 2 train +5645 AR6_WGII 1831 20 Assessed likelihood is typeset in italics (e.g., very likely).Climate risks from outside Europe are emerging due to a combination of the position of European countries in the global supply chain and shared resources high 2 train +5646 AR6_WGII 1831 22 Exposure of European countries to inter-regional risks can be reduced by international governance and collaboration on adaptation in other regions medium 1 train +5647 AR6_WGII 1831 24 There are a growing range of adaptation options available today to deal with future climate risks high 2 train +5648 AR6_WGII 1831 26 Nature-based Solutions for flood protection and heat alleviation are themselves under threat from warming, extreme heat, drought and SLR high 2 train +5649 AR6_WGII 1831 27 In many parts of Europe, existing and planned adaptation measures are not sufficient to avoid the residual risk, especially beyond 1.5° C GWL high 2 train +5650 AR6_WGII 1831 28 Residual risk can result in losses of habitat and ecosystem services, heat related deaths (KR1), crop failures (KR2), water rationing during droughts in Southern Europe (KR3) and loss of land (KR4) medium 1 train +5651 AR6_WGII 1831 29 At 3°C GWL and beyond, a combination of many, maybe even all, adaptation options are needed, including transformational changes, to reduce residual risk medium 1 train +5652 AR6_WGII 1831 31 Although adaptation is happening across Europe, it is not implemented at the scale, depth and speed needed to avoid the risks high 2 train +5653 AR6_WGII 1831 33 Forward-looking and adaptive planning can prevent path dependencies and maladaptation, and ensure timely action high 2 train +5654 AR6_WGII 1832 2 Most of the adaptation options to the key risks depend on limited water and land resources, creating competition and trade-offs, also with mitigation options and socioeconomic developments high 2 train +5655 AR6_WGII 1832 6 Several windows of opportunity emerge to accelerate climate resilient development (CRD) medium 1 train +5656 AR6_WGII 1832 8 These windows can be used to accelerate action through mainstreaming and transformational actions medium 1 train +5657 AR6_WGII 1832 10 Private-sector adaptation takes place mostly in response to extreme events or regulatory, shareholder or consumer pressures and incentives medium 1 train +5658 AR6_WGII 1832 11 Closing the adaptation gap requires moving beyond short-term planning and ensuring timely and adequate implementation high 2 train +5659 AR6_WGII 1832 14 The success of adaptation will depend on our understanding of which adaptation options are feasible and effective in their local context high 2 train +5660 AR6_WGII 1832 16 To close the adaptation gap, political commitment, persistence and consistent action across scales of government, and upfront mobilisation of human and financial capital, is key high 2 train +5661 AR6_WGII 1835 8 Mean and maximum temperatures, frequencies of warm days and nights, and heatwaves have increased since 1950, while the corresponding cold indices have decreased high 2 train +5662 AR6_WGII 1835 9 Average warming will be larger than the global mean in all of Europe, with largest winter warming in NEU and EEU and largest summer warming in MED high 2 train +5663 AR6_WGII 1835 11 Projections suggest a substantial reduction in European ice glacier volumes and in snow cover below elevations of 1500–2000 m, as well as further permafrost thawing and degradation, during the 21st century, even at a low GWL high 2 train +5664 AR6_WGII 1835 12 The assessment of climate change in WGI AR6 concludes that during recent decades mean precipitation has increased over NEU, WCE and EEU, while magnitude and sign of observed trends depend substantially on time period and study region in MED medium 1 train +5665 AR6_WGII 1835 13 Precipitation extremes have increased in NEU and EEU high 2 train +5666 AR6_WGII 1837 3 For >2°C GWL, of mean precipitation in NEU in winter is increasing and decreasing in MED in summer high 2 train +5667 AR6_WGII 1837 4 A widespread increase of precipitation extremes is projected for >2°C GWL for all sub-regions high 2 train +5668 AR6_WGII 1837 6 MED is projected to be most affected within Europe with all types of droughts increasing for 1.5°C (medium confidence) and 4°C GWL high 2 train +5669 AR6_WGII 1837 7 At 4°C GWL, hydrological droughts in NEU, WCE and EEU will increase medium 1 train +5670 AR6_WGII 1837 8 Projections for the 21st century show increases in storms across all of Europe medium 1 test +5671 AR6_WGII 1838 1 Salinity has increased in the SEUS and decreased in NEUS and is projected to continue medium 1 train +5672 AR6_WGII 1838 4 This SLR will very likely continue to increase during the 21st century (Figure 13.4k,l) (high confidence), with regional deviations from global mean SLR low 0 train +5673 AR6_WGII 1838 5 Extreme water levels, coastal floods and sandy coastline recession are projected to increase along many European coastlines high 2 train +5674 AR6_WGII 1838 7 Without further adaptation (Section 13.2.2), flood risks along Europe’s low-lying coasts and estuaries will increase due to SLR compounded by storm surges, rainfall and river runoff high 2 train +5675 AR6_WGII 1838 14 SLR will increase coastal erosion of sandy shorelines high 2 train +5676 AR6_WGII 1838 19 Projections indicate a continuation of the observed trends of river flood hazards in WCE (high confidence) of 10% at 2°C GWL and 18% at 4.4°C GWL, and a decrease in NEU and SEU medium 1 train +5677 AR6_WGII 1838 27 The intensity and frequency of heavy rainfall events is projected to increase high 2 train +5678 AR6_WGII 1841 8 This path dependency limits the solution space and may hamper implementation of transformative measures, such as land-use change, to accommodate the water system medium 1 train +5679 AR6_WGII 1841 21 A combination with structural defences could reduce risk in urbanised coastal regions high 2 train +5680 AR6_WGII 1841 27 Retreat is rarely applied in Europe medium 1 train +5681 AR6_WGII 1843 6 While measures taken at household level can reduce the risk of flooding, there is often insufficient investment medium 1 train +5682 AR6_WGII 1844 4 Early warning systems, insurance and behaviour change can complement protect and accommodate measures to limit residual risk high 2 train +5683 AR6_WGII 1844 9 Increasing future flood risks due to both climatic and socioeconomic change could overburden government budgets medium 1 train +5684 AR6_WGII 1844 16 Wastewater reuse is considered a low-cost and effective measure where wastewater is available (Lavrnic et al., 2017; De Roo et al., 2020), but public acceptance for domestic reuse is presently limited high 2 train +5685 AR6_WGII 1844 22 Increased irrigation efficiency has reduced water scarcity, particularly in SEU (Section 13.5; De Roo et al., 2020), and occur at farm level in WCE and NEU (Papadaskalopoulou et al., 2015b; van Duinen et al., 2015; Rey et al., 2017) but come with increasing path dependency on supply and trade-offs which may not be sustainable in the long term high 2 train +5686 AR6_WGII 1845 2 Interacting with climate change are non-climatic hazards, such as habitat loss and fragmentation, overexploitation, water abstraction, nutrient enrichment and pollution, all of which reduce resilience of biotas and ecosystems very high 3 train +5687 AR6_WGII 1845 6 Extirpation (e.g., local losses of species) have been observed in response to climate change in Europe medium 1 train +5688 AR6_WGII 1846 1 Range shifts are leading to northward and upwards expansions of warm-adapted taxa very high 3 train +5689 AR6_WGII 1846 4 Microclimatic variability in some locations can buffer warming impacts medium 1 train +5690 AR6_WGII 1846 6 The timing of many processes, including spring leaf unfolding, autumn senescence and flight rhythms, have changed in response to changes in seasonal temperatures, water and light availability very high 3 train +5691 AR6_WGII 1846 9 Land- use changes will increase extirpation and extinction risk very high 3 train +5692 AR6_WGII 1846 11 Warming <1.5°C GWL would limit risks to biodiversity, while 4°C GWL and intensive land use could lead to a loss of suitable climate and habitat space for most species low 0 train +5693 AR6_WGII 1846 13 Average wetland area is not projected to change at 1.7°C GWL across Europe, while for >4°C GWL expanding sites in NEU are not sufficient to balance losses in SEU and WCE high 2 train +5694 AR6_WGII 1846 16 The largest losses of suitable climatic conditions are projected for plants and insects, with different taxon-specific regions of highest risk, while proportions of species projected to lose suitable climates are lower for other groups medium 1 train +5695 AR6_WGII 1846 17 Temperatures >1.5°C GWL will lead to a progressive subtropicalisation in SEU, expanding into WCE at >3°C GWL, a northward shift in the temperate domain into NEU medium 1 train +5696 AR6_WGII 1846 21 Risks emerging from climate change for phenology are uncertain, given asynchrony between species, taxa and trophic responses (Thackeray et al., 2016; Posledovich et al., 2018; Keogan et al., 2021) and the complexity of phenological events and their cues medium 1 train +5697 AR6_WGII 1846 27 Fire hazard conditions, including heatwaves (Boer et al., 2017), increased throughout Europe from 1980 to 2019 (Figure 13.10), with substantive increases in SEU and WCE high 2 train +5698 AR6_WGII 1848 1 Forest expansion in boreal regions results in net warming (Bright et al., 2017), possibly influencing cloud formation and rainfall patterns medium 1 train +5699 AR6_WGII 1848 3 If not managed through increased reforestation and/or revegetation or peatland restoration, future climate-change impacts will progressively limit the climate regulation capacity of European terrestrial ecosystems medium 1 train +5700 AR6_WGII 1848 4 Predominantly positive CO 2 fertilisation effects at current warming will change into increasingly negative effects of warming and drought on forests at higher temperatures medium 1 train +5701 AR6_WGII 1848 6 Declines in pollinator ranges in response to climate change are occurring for many groups in Europe high 2 train +5702 AR6_WGII 1848 9 Projected climate impacts on pollinators show mixed responses across Europe but are greater under 3°C GWL medium 1 train +5703 AR6_WGII 1848 14 Soil erosion varies across Europe, with higher rates in parts of SEU and WCE, but lower rates in NEU high 2 train +5704 AR6_WGII 1849 4 Projected increase in rainfall could increase soil erosion, while warming enhances vegetation cover, leading to overall mixed responses medium 1 test +5705 AR6_WGII 1849 10 Lowering vulnerability by reducing other anthropogenic impacts (Gillingham et al., 2015), such as land-use change, habitat fragmentation (Eigenbrod et al., 2015; Oliver et al., 2017; Wessely et al., 2017), pollution and deforestation (Chapter 2), enhances adaptation capacity and biodiversity conservation high 2 train +5706 AR6_WGII 1849 11 Protected areas, such as the EU Natura 2000 network, have contributed to biodiversity protection medium 1 train +5707 AR6_WGII 1849 12 Most protected areas are static and thus do not take species migration into consideration high 2 train +5708 AR6_WGII 1849 15 Their success will depend on consideration of the future climate niche when restoring peatlands (Bellis et al., 2021) or long-lived species with limited mobility high 2 train +5709 AR6_WGII 1849 19 The capacity to implement and maintain these options remains limited, however medium 1 train +5710 AR6_WGII 1849 22 Ecosystem-based adaptations (EbA) and NbS that restore or recreate ecosystems, build resilience and produce synergies with adaptation and mitigation in other sectors are increasingly used in Europe high 2 train +5711 AR6_WGII 1850 1 Appropriately implemented ecosystem-based mitigation, such as reforestation with climate-resilient native species (Section 13.3.1.4), peatland and wetland restoration, and agroecology (Section 13.5.2), can enhance carbon sequestration or storage medium 1 train +5712 AR6_WGII 1850 3 Trade-offs between ecosystem protection, their services and human adaptation and mitigation needs can generate challenges, such as loss of habitats, increased emissions from restored wetlands (Günther et al., 2020) and conflicts between carbon capture services, and provisioning of bioenergy, food, timber and water medium 1 train +5713 AR6_WGII 1850 4 The solution space for responding to climate-change risks for terrestrial ecosystems has increased in parts of Europe medium 1 train +5714 AR6_WGII 1850 9 Despite an expanding solution space, widespread implementation and monitoring of natural and planned adaptation across Europe is currently limited, due to high management costs, undervaluation of nature, and conservation laws and regulations that do not consider species shifts under future socioeconomic and climatic changes high 2 train +5715 AR6_WGII 1850 11 Limited financial resources prevent widespread implementation of large-scale and connected conservation areas high 2 train +5716 AR6_WGII 1850 13 Risks to terrestrial and freshwater ecosystems are rarely integrated into regional and local land-use planning, land development plans, and agro-system management (medium confidence) (Nila et al., 2019; Heikkinen et al., 2020a).13.3.3 Knowledge Gaps Despite growing evidence of climate-change impacts and risks, including attributed changes to terrestrial ecosystems (Section 13.10.1), this information is geographically not equally distributed, leaving clear gaps for some processes or regions high 2 train +5717 AR6_WGII 1850 14 For processes such as wildfire, the Fire Weather index (Section 13.3.1.3) suggests increasing risk of fires in Europe, but robust projections on incidents and magnitudes of wildfire and their impacts on ecosystems and other sectors is currently limited, particularly for NEU, EEU and WCE high 2 train +5718 AR6_WGII 1850 16 This creates uncertainty about the emergence of extinctions and the magnitudes of impacts for European ecosystems and the services they provide high 2 train +5719 AR6_WGII 1850 20 Furthermore, adaptation actions will depend on local implementation and benefit from being assessed using cultural and Indigenous knowledge where applicable, but this is hardly studied medium 1 train +5720 AR6_WGII 1850 23 Particularly habitat loss in shallow coastal waters and at the coasts themselves, and northward distribution shifts of populations and communities, are evident across all European marine sub-regions high 2 train +5721 AR6_WGII 1850 24 Marine heatwaves have had severe ecological impacts in SEUS high 2 train +5722 AR6_WGII 1850 25 Range contractions, extirpations (medium confidence) (Smale, 2020) and species redistributions have been observed high 2 train +5723 AR6_WGII 1851 1 Reductions in growth and reproductive success of calcifying species are not yet unambiguously detected and attributed in European seas medium 1 train +5724 AR6_WGII 1851 3 Biodiversity changes depend on region, habitat and taxon (medium confidence) (Figure 13.11) overall resulting in the redistribution of biodiversity in Europe (García Molinos et al., 2016), and biodiversity declines in some sub-regions high 2 train +5725 AR6_WGII 1851 4 In TEUS, increased water-column stratification (Section 13.1) and decreasing eutrophication, result in reduced primary production (high confidence) (Figure 13.11; Capuzzo et al., 2018) and productivity at higher trophic levels (high confidence) (Free et al., 2019), while in NEUS sea ice decline has resulted in primary production increase by 40–60% high 2 train +5726 AR6_WGII 1851 5 Climate-related deoxygenation impacts are small in most European waters medium 1 train +5727 AR6_WGII 1851 6 Here warming and eutrophication have altered ecosystem functioning high 2 train +5728 AR6_WGII 1852 3 Since the capacity of natural systems for autonomous adaptation is limited medium 1 train +5729 AR6_WGII 1852 4 At 1.5°C GWL, particularly in winter, Mediterranean coastal fish communities are projected to lose ~10% of species, increasing to ~60% at 4°C GWL (Dahlke et al., 2020), exacerbating regime shifts linked to overexploitation medium 1 train +5730 AR6_WGII 1852 7 Marine primary production is projected to further decrease by 2100 in most European seas between 0.3% at 1.5°C GWL to 2.7% at 4°C GWL high 2 train +5731 AR6_WGII 1852 12 Ocean acidification and its biological and ecological risks are projected to rise in European waters by impeding growth and reproductive success of vulnerable calcifying organisms medium 1 train +5732 AR6_WGII 1852 13 Coralline algae are projected to reduce skeletal performance at 3°C GWL, with negative consequences for habitat formation medium 1 train +5733 AR6_WGII 1852 14 Regionally (Brodie et al., 2014), differences in species-specific vulnerability will result in community shifts from calcifying macroalgae (medium confidence) (Ragazzola et al., 2013) to non-calcifying macroalgae high 2 train +5734 AR6_WGII 1852 16 However, if not supported by sufficient food availability (Thomsen et al., 2013; Clements and Darrow, 2018), such energy reallocation will negatively impact growth or reproduction medium 1 train +5735 AR6_WGII 1852 17 This suggests that acidification risks will be amplified by increased stratification and reduced primary production medium 1 train +5736 AR6_WGII 1852 18 The emergence of harmful algal blooms and pathogens at higher GWLs is unclear across all European seas low 0 train +5737 AR6_WGII 1852 20 Elevated CO2 levels predicted at 4°C GWL will affect the C/N ratio of organic-matter export and, hence, the efficiency of the biological pump low 0 test +5738 AR6_WGII 1852 21 Atlantic herring (Clupea harengus) will benefit with enhanced larval growth and survival from indirect food- web effects (Sswat et al., 2018a), whereas Atlantic cod (Gadus morhua) will face overall negative impacts medium 1 train +5739 AR6_WGII 1852 24 Losses are projected for Posidonia oceanica seagrass habitats in the Mediterranean by up to 75% at 2.5°C GWL low 0 train +5740 AR6_WGII 1852 26 For the Dutch Wadden Sea, the critical rate of 6–10 mm yr–1, at which intertidal flats will start to ‘drown’, will be reached by 2030 at 1.5°C GWL medium 1 train +5741 AR6_WGII 1852 27 European coastal zones provided a total of 494 billion EUR of ecosystem services in 2018, and 4.2–5.1% of this value will be lost due to coastal erosion by 2100 at 2.5°C and 4.6°C GWL, respectively medium 1 train +5742 AR6_WGII 1853 6 These MPAs provide protection from local stressors, such as commercial exploitation, and enhance the resilience of marine and coastal ecosystems, thus lessening the impacts of climate change medium 1 train +5743 AR6_WGII 1853 11 In some partially protected MPAs, local stressors, such as fishing, are higher than adjacent unprotected areas medium 1 train +5744 AR6_WGII 1854 4 Conservation approaches (e.g., MPAs, climate refugia), habitat restoration efforts (Bekkby et al., 2020) and further ecosystem-based management policies do support alleviation of, or adaptation to, climate-change impacts medium 1 train +5745 AR6_WGII 1854 8 While rising sea levels will also directly threaten intertidal and beach ecosystems, coastal wetlands will benefit medium 1 train +5746 AR6_WGII 1854 10 The ‘Blue Growth’ strategy of the European Commission with the aim to increase offshore activities (European Comission, 2012) will increase the pressures on the marine environments medium 1 train +5747 AR6_WGII 1854 12 The introduction of novel hard-substrate intertidal habitats has, and will continue to have, profound ecological ramifications for marine systems, including hydrodynamic changes, stepping stones for non-native species, noise and vibration, and changes in the food web high 2 train +5748 AR6_WGII 1854 22 Observed climate change has led to a northward movement of agro- climatic zones in Europe and earlier onset of the growing season high 2 train +5749 AR6_WGII 1854 23 Warming and precipitation changes since 1990 explain continent-wide reductions in yield of wheat and barley, as well as increases in maize and sugar beet high 2 train +5750 AR6_WGII 1854 25 Drought, excessive rain and the compound hazards of drought and heat (Sections 13.2.1, 13.3.1, 13.10.2) have increased costs and cause economic losses in forest productivity (Schuldt et al., 2020), annual and permanent crops, and livestock farming (Stahl et al., 2016), including losses in wheat production in the EU (van der Velde et al., 2018) and EEU high 2 train +5751 AR6_WGII 1855 1 Regionally, warming caused increases in yields of field-grown fruiting vegetables, decreases in root vegetables, tomatoes and cucumbers (Potopová et al., 2017) and earlier flowering of olive trees high 2 train +5752 AR6_WGII 1855 3 Evidence for growing regional differences of projected climate risks is increasing since AR5 high 2 train +5753 AR6_WGII 1855 4 While there is high agreement of the direction of change, the absolute yield losses are uncertain due to differences in model parameterisation and whether adaptation options are represented high 2 train +5754 AR6_WGII 1855 6 Growing regions will shift northward or expand for melons (Bisbis et al., 2019), tomatoes and grapevines reaching NEU and EEU in 2050 under 1.5°C GWL (high confidence) (Hannah et al., 2013; Litskas et al., 2019), while warming would increase yields of onions, Chinese cabbage and French beans (Bisbis et al., 2019) medium 1 train +5755 AR6_WGII 1855 9 Reductions in agricultural yields will be higher in the south at 4°C GWL, with lower losses or gains in the north high 2 train +5756 AR6_WGII 1855 10 The largest impacts of warming are projected for maize in SEU high 2 train +5757 AR6_WGII 1855 11 Use of longer-season varieties can compensate for heat stress on maize in WCE and lead to yield increases for NEU, but not SEU for 4°C GWL medium 1 train +5758 AR6_WGII 1855 14 Warming causes range expansion and alters host pathogen association of pests, diseases and weeds affecting the health of European crops high 2 train +5759 AR6_WGII 1855 21 Climate change also impacts grassland production, fodder composition and quality, particularly in SEU (Dumont et al., 2015) and EEU (Bezuglova et al., 2020), as well as alters the prevalence, distribution and load of pathogens and their vectors high 2 train +5760 AR6_WGII 1855 23 Warming increases the pasture growing season and farming period in NEU and at higher altitudes (Fuhrer et al., 2014), while longer drought periods and thunderstorms can influence abandonment of remote Alpine pastures, reducing cultural and landscape ecosystem services and losing traditional farming practices high 2 train +5761 AR6_WGII 1855 24 At 2–4°C GWL grassland biomass production for forage-fed animals will increase in NEU and the northern Alps, while forage production will decrease in SEU and the southern Alps due to heat and water scarcity (Gauly et al., 2013; Jäger et al., 2020), causing regional reductions of cow milk production in WCE and SEU high 2 train +5762 AR6_WGII 1855 28 Climate change has impacted European marine food production high 2 train +5763 AR6_WGII 1855 31 In the North Sea, cuttlefish (van der Kooij et al., 2016; Oesterwind et al., 2020) and tuna (Bennema, 2018; Faillettaz et al., 2019) have become new target species medium 1 train +5764 AR6_WGII 1856 1 European countries are assessed to be globally among the least vulnerable to the impacts of climate change on fisheries-related food security risks high 2 train +5765 AR6_WGII 1857 2 Assuming MSY management, projections suggest reduced abundance of most commercial fish stocks in European waters of 35% (up to 90% for individual stocks) between 1.5°C and 4.0°C GWL medium 1 train +5766 AR6_WGII 1857 4 Ocean acidification (Section 13.4; Chapter 4) will develop into a major risk for marine food production in Europe under 4°C GWL high 2 train +5767 AR6_WGII 1857 5 Acidification is also projected to negatively affect marine shellfish production and aquaculture in Europe with 4°C GWL (medium confidence) (Fernandes et al., 2017; Narita and Rehdanz, 2017; Mangi et al., 2018).13.5.1.4 Forestry and Forest Products Climate change is altering the structure and function of European forests via changes in temperature, precipitation and atmospheric CO 2, as well as through interaction with pests and fire high 2 train +5768 AR6_WGII 1857 7 While warming and extended growing seasons have positive impacts on forest growth in cold areas in WCE and NEU (Pretzsch et al., 2014; Matskovsky et al., 2020), EEU (Tei et al., 2017) and higher altitude (Sedmáková et al., 2019), drought stress across Europe has been increasing high 2 train +5769 AR6_WGII 1858 2 Water stress exacerbates the incidence from and effects of fire and other natural disturbances (Section 13.3.1), resulting in forest productivity declines or cancelling out productivity gains from CO 2 high 2 train +5770 AR6_WGII 1858 10 Extensive droughts during the past two decades have caused many irrigated systems in SEU to cease production (Stahl et al., 2016) indicating limited adaptive capacity to heat and drought medium 1 train +5771 AR6_WGII 1858 14 Changes to cultivars and sowing dates can reduce yield losses (Figure 13.15) but are insufficient to fully ameliorate losses projected >3°C GWL, with an increase of risk from north to south and for crops growing later in the season such as maize and wheat high 2 train +5772 AR6_WGII 1858 15 Adaptations for early maturing reduce yield loss by moving the cycle towards a cooler part of year, and also constrains the increases in irrigation water demands, but reduce the period for photosynthesis and grain filling high 2 train +5773 AR6_WGII 1858 19 These options are used in indoors- reared species (Gauly et al., 2013) but are limited in mountain pastures high 2 train +5774 AR6_WGII 1858 21 Dairy systems that maximise the use of grazed pasture are considered more environmentally sustainable but are not fully supported by policy and markets medium 1 train +5775 AR6_WGII 1858 26 Agroforestry, integrating trees with crops (silvoarable), livestock (silvopasture), or both (agrosilvopasture), can enhance resilience to climate change (Chapter 5), but implementation in Europe needs improved training programmes and policy support high 2 train +5776 AR6_WGII 1858 28 Agricultural policy, market prices, new technology and socioeconomic factors play a more impor - tant role in short-term farm-level investment decisions than climate- change impacts high 2 train +5777 AR6_WGII 1860 7 Inflexible and non-adaptive allocation schemes can result in conflicts among European countries medium 1 train +5778 AR6_WGII 1860 8 The development of adaptation strategies for seafood production since the Paris Agreement is insufficient in Europe high 2 train +5779 AR6_WGII 1860 13 Successful adaptation strategies include altering the tree species composition to enhance the resilience of European forests high 2 train +5780 AR6_WGII 1860 14 Greater diversity of tree species reduces vulnerability to pests and pathogens (Felton et al., 2016), and increases resistance to natural disturbances high 2 train +5781 AR6_WGII 1860 15 Depending on forest successional history (Sheil and Bongers, 2020), tree composition change can increase carbon sequestration high 2 train +5782 AR6_WGII 1860 16 Conservation areas can also help climate-change adaptation by keeping the forest cover intact, creating favourable microclimates and protecting biodiversity low 0 train +5783 AR6_WGII 1860 17 Reforestation reduces warming rates (Zellweger et al., 2020) and extremely warm days (Sonntag et al., 2016) inside forests, reducing natural disturbances and fires high 2 train +5784 AR6_WGII 1860 20 Consumer demand for food and timber products can adapt to productivity changes and be mediated by price (e.g., in response to production changes or policies on food-related taxation), reflect changes in preferences (e.g., towards plant-based foods motivated by environmental, ethical or health concerns) or reductions in food waste high 2 train +5785 AR6_WGII 1860 21 Although mitigation potentials of dietary changes have received increasing attention, evidence is lacking on potential for adaptation through changes in European food consumption and trade, despite these socioeconomic factors being a strong driver for change medium 1 train +5786 AR6_WGII 1860 25 Effectiveness of adaptation options is predominantly qualitatively mentioned but not assessed, and the effectiveness of combinations of measures is rarely assessed high 2 train +5787 AR6_WGII 1861 2 The assessment of irrigation needs and the impact of CO 2 and O 3 tend to focus on individual species and processes hindering upscaling to multiple stressors and mixed production high 2 train +5788 AR6_WGII 1862 1 Risks of rutting and blow-ups of roads (particularly in low altitudes) due to high summer temperatures are expected to increase in WCE and EEU at 3°C GWL medium 1 train +5789 AR6_WGII 1863 3 Current damages are mainly related to river floods and storms, but heat and drought will become major drivers in the future medium 1 train +5790 AR6_WGII 1863 6 Indirect effects via supply chains, transport and electricity networks can be as high as, or substantially higher than, direct effects medium 1 train +5791 AR6_WGII 1863 9 Due to reduced snow availability and hotter summers, damages are projected for the European tourism industry, with larger losses in SEU (high confidence) and some smaller gains in the rest of Europe medium 1 train +5792 AR6_WGII 1871 3 A GWL of 1.5°C could result in 30,000 annual deaths due to extreme heat, with up to threefold the number under 3°C GWL high 2 train +5793 AR6_WGII 1871 5 Heat stress risks will be lower under SSP1 than the SSP3 or SSP4 scenarios high 2 train +5794 AR6_WGII 1871 11 In large European cities, stabilising climate warming at 1.5°C GWL would decrease premature deaths by 15–22% in summer compared with stabilisation at 2°C GWL high 2 train +5795 AR6_WGII 1871 20 Climate change could increase air pollution health effects, with the size of the effect differing across European regions and pollutants medium 1 train +5796 AR6_WGII 1871 23 At 2.5°C GWL, mortalities due to exposure to PM2.5 are projected to increase by up to 73% in Europe medium 1 train +5797 AR6_WGII 1871 24 At 2°C GWL, annual premature mortalities due to exposure to near-surface ozone are projected to increase up to 11% in WCE and SEU and to decrease up to 9% in NEU (under RCP4.5) medium 1 train +5798 AR6_WGII 1873 15 There has been a temperature-dependent range expansion of ticks that is projected to expand further north in Sweden, Norway and the Russian Arctic (Jaenson et al., 2012; Jore et al., 2014; Tokarevich et al., 2017; Waits et al., 2018), and to higher elevations in Austria and the Czech Republic medium 1 test +5799 AR6_WGII 1873 20 Projections for Europe show the West Nile virus risk to expand: by 2025, the risk is projected to increase in SEU and southern and eastern parts of WCE medium 1 train +5800 AR6_WGII 1873 21 Although climatic suitability for malaria transmission in Europe is increasing and will lead to a northward spread of the occurrences of Anopheles vectors, the risk from malaria to human health in Europe remains low due to economic and social development as well as access to health care medium 1 train +5801 AR6_WGII 1874 3 Under further warming, the number of months with risk of Vibrio transmission increases and the seasonal transmission window expands, thereby increasing the risk to human health in the future high 2 train +5802 AR6_WGII 1874 5 Allergies and Pollen The main drivers of allergies are predominantly non-climatic (e.g., increased urbanisation, adoption of westernised lifestyles, social and genetic factors), but climate change strongly contributes to the spread of some allergenic plants, thus exacerbating existing allergies and causing new ones in people across Europe high 2 train +5803 AR6_WGII 1874 17 Each of these major European heatwaves led to considerable economic losses in agriculture and construction high 2 train +5804 AR6_WGII 1874 27 Mental Health and Well-Being Extreme weather events can trigger post-traumatic stress disorder (PTSD), anxiety and depression; this is well-documented for flooding in Europe high 2 train +5805 AR6_WGII 1875 3 It shows that substantial social–cultural and institutional barriers complicate widespread implementation of measures; studies on the implementation of new blue–green spaces in existing urban structures in, for example, Sweden (Wihlborg et al., 2019), the UK (Carter et al., 2018) and the Netherlands (Aalbers et al., 2019), point to important feasibility challenges (e.g., access to financial resources, societal opposition, competition for space) high 2 train +5806 AR6_WGII 1875 6 Existing health measures, such as monitoring and early warning systems, play an important role in detecting and communicating emerging climate risks and weather extremes high 2 train +5807 AR6_WGII 1875 13 Combining multiple types of adaptation options into a consistent policy portfolio may have an amplifying effect in reducing risks, particularly at higher GWL medium 1 train +5808 AR6_WGII 1876 1 Coordination of health adaptation actions across scales and between public sectors is needed to ensure timely and effective responses for a diversity of health impacts high 2 train +5809 AR6_WGII 1876 16 Urban poor and ethnic minorities often settle in more vulnerable settlement zones, and are therefore impacted more by flooding medium 1 train +5810 AR6_WGII 1876 23 While it strengthens resilience for disadvantaged households (Church et al., 2015; Boost and Meier, 2017; Promberger, 2017; Vávra et al., 2018; Ančić et al., 2019; Pungas, 2019) and renews their local knowledge, it can become a risk in regions with projected crop yield reductions high 2 train +5811 AR6_WGII 1876 25 While energy poverty is much more prevalent in SEU and EEU (Bouzarovski and Petrova, 2015; Pye et al., 2015; Atsalis et al., 2016; Monge-Barrio and Sánchez-Ostiz Gutiérrez, 2018), climate change will also exacerbate energy poverty in European regions where heating thus far has been the major share of energy costs medium 1 train +5812 AR6_WGII 1879 4 Temperatures in Arctic and sub-Arctic regions have increased on average by 2°C over the past 30 years very high 3 train +5813 AR6_WGII 1879 5 Future warming is expected to further increase winter precipitation high 2 train +5814 AR6_WGII 1879 7 Warming and CO 2 increase result in higher plant productivity (Section 13.3), changes in plant community composition and higher parasite harassment; unstable ice conditions affect migration; extreme weather conditions during critical winter months, more frequent forest fires and changes in plant community composition reduce pasture quality medium 1 train +5815 AR6_WGII 1879 8 High snow depth and rain-on-snow events impede reindeer access to ground lichen in winter and delay spring green-up during the critical calving period; both cause malnutrition and negative impacts on reindeer health, mortality and reproductive success medium 1 train +5816 AR6_WGII 1879 9 Lower slaughter weights and increased mortality reduce the income of herders high 2 train +5817 AR6_WGII 1879 10 Reindeer herders already autonomously adapt to changing conditions through flexible use of pastures and supplementary feeding high 2 train +5818 AR6_WGII 1879 11 However, adaptive herding practices have themselves added significant burden through increased workload, costs and stress high 2 train +5819 AR6_WGII 1879 12 Supplementary feeding increases the risk of infectious diseases and implies culturally undesirable herding practices low 0 train +5820 AR6_WGII 1879 13 Rapid land-use change reduces the ability to adapt high 2 train +5821 AR6_WGII 1879 14 National and EU policies expand land uses for mining, wind energy and bioeconomy in the area, causing loss, fragmentation and degradation of pastures, and increasing human disturbance to animals medium 1 train +5822 AR6_WGII 1879 16 Herding communities face strong barriers to protecting their rights and halting further degradation of pastures medium 1 train +5823 AR6_WGII 1879 18 Combined with land-use conflicts, climate impacts cause reduced psycho-social health and increase suicidal thoughts among herders low 0 train +5824 AR6_WGII 1879 20 The cumulative effects of land-use and climate change have already increased vulnerability and reduced the adaptive capacity of reindeer herding to the extent that its long-term sustainability is threatened medium 1 train +5825 AR6_WGII 1879 22 Lack of control over land use is the biggest and most urgent threat to the adaptive capacity of reindeer herding and the right of Sámi to their culture high 2 test +5826 AR6_WGII 1882 1 The exposure of European countries to trans-European climate impact and risk pathways varies depending on their territorial settings, national policies and position in the global supply chain high 2 train +5827 AR6_WGII 1882 6 Simultaneous breadbasket failures, and trade restrictions, increase risks to food supply medium 1 train +5828 AR6_WGII 1884 2 There is emerging evidence that supporting adaptation outside Europe may generate economic co-benefits for Europe (Román et al., 2018).13.10 Detection and Attribution, Key Risks and Adaptation Pathways 13.10.1 Detection and Attribution of Impacts Since AR5, scientific documentation of observed changes attributed to global warming have proliferated high 2 train +5829 AR6_WGII 1884 3 These include ecosystem changes detected in previous assessments, such as earlier annual greening and onset of faunal reproduction processes, relocation of species towards higher latitudes and altitudes (high confidence), and impacts of heat on human health and productivity high 2 train +5830 AR6_WGII 1886 1 In recent decades (2000–2015), economic losses intensified in SEU (high confidence) and were detected for parts of WCE and NEU medium 1 train +5831 AR6_WGII 1886 13 The risk of human heat stress and mortality is largely influenced by underlying socioeconomic pathways, with consequences being more severe under SSP3, SSP4 and SSP5 scenarios than SSP1 very high 3 train +5832 AR6_WGII 1886 15 The impact of warming in marine systems are often synergistic with SLR in coastal systems and ocean acidification driven by the rise in CO 2, while habitat fragmentation and land use have important synergies in terrestrial systems high 2 train +5833 AR6_WGII 1886 22 The transition to moderate risk is currently happening as warming already results in changes in timing of development, species migration northward and upwards, and desynchronisation of species interactions, especially at the range limits, with cascading and cumulative impacts through ecosystems and food webs high 2 train +5834 AR6_WGII 1886 25 The transition is happening at slightly higher warming in terrestrial systems due to a higher number of thermal refugia in terrestrial systems causing relocation but not already severe impacts medium 1 train +5835 AR6_WGII 1886 28 Whether or not adaptation measures are effective to reduce risk severity for people’s health depends on local context high 2 train +5836 AR6_WGII 1886 29 Some adaptation options are found to be highly effective across Europe irrespective of warming levels, including air conditioning and urban planning high 2 train +5837 AR6_WGII 1887 3 To reach high adaptation, a combination of low, medium and high effectiveness measures in different sectors and sub-regions is needed, many of which entail systems’ transformations (e.g., heat-proof land management) (Chapter 16) and remain effective at higher warming levels medium 1 train +5838 AR6_WGII 1887 6 A reduction of non-climatic stressors, such as nutrient loads, resource extraction, habitat fragmentation or pesticides on land, are considered important adaptation options to increase the resilience to climate-change impacts high 2 train +5839 AR6_WGII 1887 7 A major governance tool to reduce climatic and non-climatic impacts is the establishment of networks of protected areas (Sections 13.3.2, 13.4.2) especially when aggregated, zoned or linked with corridors for migration high 2 train +5840 AR6_WGII 1887 8 Reforestation, rewilding and habitat restoration are long-term strategies for reducing risk for biodiversity loss supported by assisted migration and evolution (Section 13.3.2, 13.4), though current laws and regulations do not include species migration high 2 train +5841 AR6_WGII 1887 11 Projected SLR will strongly impact coastal ecosystems high 2 train +5842 AR6_WGII 1887 20 There will be also broader adverse impacts such as reduction of grassland biomass production for fodder, increases in weeds and reduction in pollination medium 1 train +5843 AR6_WGII 1887 26 Under high adaptation, the use of irrigation can substantially reduce risk by both reducing canopy temperature and drought impacts high 2 train +5844 AR6_WGII 1887 29 Crop production is a major consumer of water in agriculture (Gerveni et al., 2020), yet a potentially scarcer supply of water in some regions must be distributed across many needs (KR3, Section 13.10.2.3), limiting availability to agriculture which is currently the main user of water in many regions of Europe high 2 train +5845 AR6_WGII 1887 30 Where the ability to irrigate is limited by water availability, other adaptation options are insufficient to mitigate crop losses in some sub-regions, particularly at 3°C GWL and above, with an increase in risk from north to south and higher risk for late-season crops such as maize high 2 train +5846 AR6_WGII 1887 31 Under these conditions, land abandonment is projected low 0 train +5847 AR6_WGII 1888 3 Evidence from the detected changes and attribution assessment suggests that the risk is already moderate in SEU (e.g., 48 million people exposed to moderate water scarcity between 1981 and 2010) high 2 train +5848 AR6_WGII 1888 7 Socioeconomic conditions contributing to severe consequences are when more residents settle in drought-prone regions, or when the share of agriculture in GDP declines high 2 train +5849 AR6_WGII 1888 8 For Europe, risks of water scarcity will be higher under SSP5 and SSP3 than under SSP1 medium 1 train +5850 AR6_WGII 1888 10 This transition will happen at higher warming in WCE since risks are projected to increase less rapidly (transition between 2°C and 3°C GWL) medium 1 train +5851 AR6_WGII 1888 11 At 3°C GWL and beyond, water scarcity will become much more widespread and severe in already water-scarce areas in SEU (high confidence) and will expand to currently non-water-scarce regions in WCE medium 1 train +5852 AR6_WGII 1889 13 Coastal areas have already started to be affected by SLR (see Box 13.1; Section 13.10.1) and human exposure to coastal hazards is projected to increase in the next decades (high confidence), but less under SSP1 (20%) than SSP5 (50%) by the end of the century medium 1 train +5853 AR6_WGII 1889 14 Under low adaptation (i.e., coastal defences are maintained but not further strengthened), severe consequences include an increase in expected annual damage by a factor of at least 20 for 1.5°C–2.1°C GWL (i.e., high risks) and by two to three orders of magnitude between 2°C and 3°C GWL in EU-28 (i.e., very high risk) medium 1 train +5854 AR6_WGII 1891 1 Pluvial and riverine flood events in Europe have been attributed to climate change, but the associated damages and losses also depend on land-use planning and flood risk management practices medium 1 train +5855 AR6_WGII 1891 8 Soft limits to protection have been identified under high GWL, in particular due to the rate of change and delayed impacts of long-term SLR medium 1 train +5856 AR6_WGII 1891 9 Ecosystem-based solutions, such as wetlands, can reduce waves’ propagation, provide co-benefits for the environment and climate mitigation, and reduce costs for flood defences medium 1 train +5857 AR6_WGII 1891 15 Both protection and flood retention are effective in reducing inland flooding risk across Europe, but with regional variation in the benefit-to-cost ratio medium 1 train +5858 AR6_WGII 1891 20 While there is a wide range in quantifications, there is high agreement that the consequences for socioeconomic and natural systems can be substantial, with more severe consequences in the south than in the north (very high confidence); and there is some indication also for a west-to-east gradient, with higher uncertainty in eastern WCE and EEU, which makes adaptation more challenging medium 1 train +5859 AR6_WGII 1891 21 Furthermore, the food–water–energy–land nexus plays an important role in amplifying overall risk levels in Europe medium 1 train +5860 AR6_WGII 1891 22 Southern Europe, European cities and coastal areas are projected to become hotspots of multiple risks high 2 train +5861 AR6_WGII 1891 23 The number of people exposed to multiple KRs in Europe are projected to at least double at 3°C GWL compared with 1.5°C GWL (Forzieri et al., 2017; Byers et al., 2018; Arnell et al., 2019), but risk levels are already higher at 1.5°C GWL than today for a number of KRs medium 1 train +5862 AR6_WGII 1891 24 Economic losses and damages for European economies from multiple KRs are projected to increase high 2 train +5863 AR6_WGII 1891 26 The main driver for this increase in economic losses and damages is mortality due to heat stress (me- dium confidence), followed by reduced labour productivity, coastal and inland flooding, water scarcity and drought medium 1 train +5864 AR6_WGII 1891 28 Adaptation is projected to reduce macroeconomic costs, but residual costs will remain particularly for warming above 3°C GWL medium 1 train +5865 AR6_WGII 1894 29 Progress is also observed at the level of the EU with the adoption of the new EU strategy on adaptation to climate change in 2021 (European Comission, 2021a), and regionally, particularly in federalist and decentralised states (Steurer and Clar, 2018; EEA, 2020b; Pietrapertosa et al., 2021), and locally, with an increasing number of European cities planning for climate risks high 2 train +5866 AR6_WGII 1894 30 There is evidence of action across sectors and scales, even in European countries where national adaptation frameworks are absent medium 1 train +5867 AR6_WGII 1895 4 The availability of knowledge, human and financial resources appears important for proactive adaptation (Termeer et al., 2012; Sanderson et al., 2018), while adaptation is also strongly dependent on economic and social development high 2 train +5868 AR6_WGII 1895 13 Yet, high GWL scenarios beyond 2100 are often not considered in climate-change adaptation planning due to a lack of perceived usability, missing socioeconomic information, constraining institutional settings and conflicting decision-making timeframes medium 1 train +5869 AR6_WGII 1895 17 Adaptive planning and decision making are still limited across Europe high 2 train +5870 AR6_WGII 1895 24 Since AR5, progress has been made to increase coordinated adaptation actions, but so far this is limited to a few sectors (mostly water management and agriculture) and European countries and regions (mostly SEU, and WCE depending on impact) high 2 train +5871 AR6_WGII 1895 25 Despite evidence of emerging bottom-up (e.g., citizens and business) and top-down initiatives (e.g., governmental plans and instruments to ensure action), there are considerable barriers to mainstreaming adaptation high 2 train +5872 AR6_WGII 1896 12 With higher warming levels, financing needs are likely to increase high 2 train +5873 AR6_WGII 1896 21 Whereas some sectors, such as flood management, banking and insurance, and energy (Bank of England, 2015; Gasbarro and Pinkse, 2016; Bank of England, 2019; Botzen et al., 2019), have generally made moderate progress on adaptation planning across Europe, there are key vulnerable economic sectors that are in earlier stages, including aviation (Burbidge, 2018), ports and shipping (Becker et al., 2018; Ng et al., 2018), and ICT high 2 train +5874 AR6_WGII 1897 10 Perceived personal responsibility for tackling climate change remains low across the EU (Figure 13.35) and partly explains why household adaptation remains limited high 2 train +5875 AR6_WGII 1897 11 Householders’ risk perception and concern about climate change fluctuates in response to media coverage and significant weather or sociopolitical events high 2 train +5876 AR6_WGII 1897 12 On average across Europe, and particularly in relation to gradual change, compared with experts, non-experts continue to underestimate climate-change risks medium 1 train +5877 AR6_WGII 1898 4 As well as temporal trends in climate-change risk perception, the lit- erature since AR5 continues to show much heterogeneity (both within and between nations) among householders in respect of risk percep- tion high 2 train +5878 AR6_WGII 1898 6 Stronger evidence exists since AR5 that experience of extreme weather events can shape climate-change risk perceptions, if these events are attributed to climate change or evoke negative emotions high 2 train +5879 AR6_WGII 1898 10 Adaptation responses across European regions and sectors are more often incremental than transformative medium 1 train +5880 AR6_WGII 1898 13 This includes extreme weather events, financial crises, for example in Malmö (Anderson, 2014; Isaksson and Heikkinen, 2018), and the COVID-19 pandemic (e.g., Milan), all of which have disrupted the status quo and accelerated innovation and implementation (e.g., Milan; see Box 13.3; Cross-Chapter Box COVID in Chapter 7).Considerable barriers exist that prevent system transitions from taking place in Europe, including institutional and behavioural lock-ins such as administrative routines, certain types of legislation and dominant paradigms of problem solving high 2 train +5881 AR6_WGII 1942 2 Addressing these risks has been made more urgent by delays due to misinformation about climate science that has sowed uncertainty and impeded recognition of risk high 2 train +5882 AR6_WGII 1942 5 Immediate, widespread and coordinated implementation of adaptation measures aimed at reducing risks and focused on equity have the greatest potential to maintain and improve the quality of life for North Americans, ensure sustainable livelihoods and protect the long-term biodiversity, and ecological and economic productivity, in North America high 2 train +5883 AR6_WGII 1942 8 Despite scientific certainty of the anthropogenic influence on climate change, misinformation and politicisation of climate-change science has created polarisation in public and policy domains in North America, particularly in the USA, limiting climate action high 2 train +5884 AR6_WGII 1942 9 Vested interests have generated rhetoric and misinformation that undermines climate science and disregards risk and urgency medium 1 train +5885 AR6_WGII 1942 10 Resultant public misperception of climate risks and polarised public support for climate actions is delaying urgent adaptation planning and implementation high 2 train +5886 AR6_WGII 1942 13 High temperatures have increased mortality and morbidity (very high confidence), with impacts that vary by age, gender, location and socioeconomic conditions very high 3 train +5887 AR6_WGII 1942 16 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.increased risk of vector-borne (very high confidence), water-borne (high confidence) and food-borne diseases very high 3 train +5888 AR6_WGII 1942 17 Changes in climate and extreme events have been linked to wide- ranging negative mental health outcomes high 2 train +5889 AR6_WGII 1942 18 The loss of access to marine and terrestrial sources of protein has impacted the nutrition of subsistence-dependent communities across North America high 2 train +5890 AR6_WGII 1942 19 Climate change has increased the extent of warmer and drier conditions favourable for wildfires (medium confidence) that increase respiratory distress from smoke very high 3 train +5891 AR6_WGII 1942 21 Climate change and extreme weather events have impacted North American agroecosystems high 2 train +5892 AR6_WGII 1942 22 Climate change has generally reduced agricultural productivity by 12.5% since 1961, with progressively greater losses moving south from Canada to Mexico and in drought-prone rain-fed systems high 2 train +5893 AR6_WGII 1942 23 Loss of availability and access to marine and terrestrial sources of protein has impaired food security and nutrition of subsistence-dependent Indigenous Peoples across North America high 2 train +5894 AR6_WGII 1942 24 Climate change has impacted aquaculture (high confidence) and induced rapid redistribution of species (very high confidence), and population declines of multiple key fisheries high 2 train +5895 AR6_WGII 1942 26 Reduced snowpack and earlier runoff high 2 train +5896 AR6_WGII 1942 27 Recent severe droughts, floods and harmful algal and pathogen events have caused harm to large populations and key eco- nomic sectors high 2 train +5897 AR6_WGII 1942 28 Heavy exploitation of limited water supplies, especially in the western USA and northern Mexico, and deteriorating freshwater management infrastructure, have heightened the risks high 2 train +5898 AR6_WGII 1942 29 Effective examples of freshwater resource adaptation planning are already underway, but coordinated adapta- tion implementation across multiple conflicting interests and users is complicated and time-consuming high 2 train +5899 AR6_WGII 1942 31 Larger losses and adaptation costs are observed for sectors with high climate exposures, including tourism, fisheries, and agriculture (high confidence) and outdoor labour medium 1 train +5900 AR6_WGII 1942 32 Disaster planning and spending, insurance, markets, and individual and household-level adaptation have acted to moderate effects to date medium 1 train +5901 AR6_WGII 1943 2 Impacts are particularly apparent for Indigenous Peoples for whom culture, identity, commerce, health and well-being are closely connected to a resilient environment very high 3 train +5902 AR6_WGII 1943 3 Higher temperatures have been associated with violent and property crime in the USA medium 1 train +5903 AR6_WGII 1943 5 Rising air, water, ocean and ground temperatures have restructured ecosystems and contributed to the redistribution (very high confidence) and mortality high 2 train +5904 AR6_WGII 1943 6 Extreme heat and precipitation trends on land have increased vegetation stress and mortality, reduced soil quality and altered ecosystem processes including carbon and freshwater cycling very high 3 train +5905 AR6_WGII 1943 7 Warm and dry conditions associated with climate change have led to tree die-offs (high confidence) and increased prevalence of catastrophic wildfire (medium confidence) with an increase in the size of severely burned areas in western North America medium 1 train +5906 AR6_WGII 1943 8 Nature-based Solutions (NbS) and ecosystem-based management have been effective adaptation approaches in the past but are increasingly exceeded by climate extremes medium 1 train +5907 AR6_WGII 1943 10 Climate change has contributed to cascading environmental and sociocultural impacts in the Arctic (high to very high confidence) that have adversely, and often irreversibly, altered Northern livelihoods, cultural activities, essential services, health, food and nutritional security, community connectivity and well-being high 2 train +5908 AR6_WGII 1943 13 Humidity-enhanced heat stress, aridification and extreme precipitation events that lead to severe flooding, erosion, debris flows and ultimately loss of ecosystem function, life and property are projected to intensify high 2 train +5909 AR6_WGII 1943 15 Warming is projected to increase heat-related mortality (very high confidence) and morbidity medium 1 train +5910 AR6_WGII 1943 16 Vector-borne disease transmission, water-borne disease risks, food safety risks and mental health outcomes are projected to increase this century high 2 train +5911 AR6_WGII 1943 17 Available adaptation options will be less effective or unable to protect human health under high-emission scenarios high 2 train +5912 AR6_WGII 1943 19 Climate change will continue to shift North American agricultural and fishery suitability ranges (high confidence) and intensify production losses of key crops (high confidence), livestock (medium confidence), fisheries (high confidence) and aquaculture products medium 1 train +5913 AR6_WGII 1943 20 In the absence of mitigation, incremental adaptation measures may not be sufficient to address rapidly changing conditions and extreme events, increasing the need for cross-sectoral coordination in implementation of mitigation and adaptation measures high 2 train +5914 AR6_WGII 1943 21 Combining sustainable intensification, approaches based on Indigenous knowledge and local knowledge, and ecosystem-based methods with inclusive and self- determined decision making, will result in more equitable food and nutritional security high 2 train +5915 AR6_WGII 1943 23 Hotter droughts and progressive loss of seasonal water storage in snow and ice will tend to reduce summer season stream flows in much of western North America, while population growth, extensive irrigated agriculture and the needs of threatened and endangered aquatic species will continue to place high demands on those flows high 2 train +5916 AR6_WGII 1943 26 Hard limits to adaptation may be reached for outdoor labour (medium confidence) and nature-based winter tourism activities very high 3 train +5917 AR6_WGII 1943 27 At higher levels of warming, climate impacts may pose systemic risks to financial markets through impacts on transportation systems, supply chains and major infrastructure, as well as global-scale challenges to trade medium 1 train +5918 AR6_WGII 1944 3 Supporting Indigenous self-determination, recognising Indigenous Peoples’ Rights, and supporting adaptation underpinned by Indigenous knowledge are critical to reducing climate-change risks to achieve adaptation success very high 3 train +5919 AR6_WGII 1944 6 Widespread adoption of these practices and tools for infrastructure planning, disaster risk reduction, ecosystem management, budgeting practices, insurance, and climate risk reporting supports planning for a future with more climate risks high 2 train +5920 AR6_WGII 1944 7 Increased capacity to support the equitable resolution of existing and emerging resource disputes (local to international) will reduce climate impacts on livelihoods and improve the effectiveness of resource management high 2 train +5921 AR6_WGII 1944 9 Recognition of the need for adaptation across North America is increasing, but action has been mostly gradual, incremental and reactive high 2 train +5922 AR6_WGII 1944 10 Current practices will be increasingly insufficient without coordination and integration of efforts through equitable policy focused on modifying land-use impacts, consumption patterns, economic activities and emphasising NbS high 2 test +5923 AR6_WGII 1944 11 Transformational, long-term adaptation action that reduces risk and increases resilience can address rapidly escalating impacts in the long-term, especially if coupled with moderate to high mitigation measures high 2 train +5924 AR6_WGII 1947 10 Annual precipitation has increased in recent decades in northern and eastern areas (CA-PR, CA-QU, US-NP , US-SP , US-MW, US-NE, US-AK) (high confidence), and has decreased across the western part of the continent (CA-BC, US-SW, US-NW, MX-NW) medium 1 train +5925 AR6_WGII 1947 17 This Report also uses the term ‘likely range’ to indicate that the assessed likelihood of an outcome lies within the 17–83% probability range.Snowpack and snow extent across much of Canada and the western USA have declined as temperatures have increased very high 3 train +5926 AR6_WGII 1947 30 High Arctic sea ice retreat since 1971 and increases in open- water duration in the most recent decade are unprecedented (Box et al., 2019) and most pronounced in the Chukchi, Bering and Beaufort seas (US-AK, CA-NW) high 2 train +5927 AR6_WGII 1947 31 Warming of North American offshore waters is significant and attributable to human activities, particularly along the Atlantic coast, contributing to sea level rise (SLR) through thermal expansion very high 3 train +5928 AR6_WGII 1948 12 Acidification of North American coastal waters has occurred in conjunction with increased atmospheric CO 2 concentration (Mathis et al., 2015; Jewett and Romanou, 2017; Claret et al., 2018) combined with other local acidifying inputs such as nitrogen and sulphur deposition (Doney et al., 2007) and freshwater nutrient input very high 3 train +5929 AR6_WGII 1948 19 Total precipitation is projected to increase across the northern half of North America (very high confidence) and decrease in southwest North America (MX-SW, MX-NW, US-SW) medium 1 train +5930 AR6_WGII 1949 2 As temperatures rise, snow extent, duration of snow cover and accumulated snowpack are virtually certain to decline in subarctic regions of North America (Gutierrez et al., 2021a; McCrary and Mearns, 2019; Mudryk et al., 2021), with corresponding effects on snow- related hydrological changes high 2 train +5931 AR6_WGII 1949 4 Climate change is projected to magnify the impact of tropical cyclones in US-NE, MX-NE, US-SP, and US-SE by increasing rainfall (Patricola and Wehner, 2018) and extreme wind speed high 2 train +5932 AR6_WGII 1949 5 The coastal region at severe risk from tropical storms is projected to expand northward within US- NE (medium confidence) (Kossin et al., 2017).Additional reduction in polar sea ice is virtually certain (Ranasinghe et al., 2021; Mudryk et al., 2021), with the North American Arctic projected to be seasonally ice free at least once per decade under 2°C of global warming high 2 train +5933 AR6_WGII 1949 6 Duration of freshwater lake ice across the northern USA and southern Canada is projected to diminish high 2 train +5934 AR6_WGII 1949 11 Ocean acidification (OA) along North American coastlines is projected to increase very high 3 train +5935 AR6_WGII 1950 22 Despite expert scientific consensus on anthropogenic climate change, there is polarisation and an ongoing debate over the reality of anthropogenic climate change in the public and policy domains, with attendant risks to society high 2 train +5936 AR6_WGII 1950 25 Rhetoric and misinformation on climate change and the deliberate undermining of science have contributed to misperceptions of the scientific consensus, uncertainty, disregarded risk and urgency, and dissent high 2 train +5937 AR6_WGII 1951 1 Vocal groups can affect public discourse and weaken public support for climate mitigation and adaptation policies medium 1 train +5938 AR6_WGII 1951 17 Political affiliation and partisan group identity contribute to polarisation on the causes and state of climate change, most notably in the USA medium 1 train +5939 AR6_WGII 1951 22 Communicating to educate or enhance knowledge on climate-change science or consensus can, but does not necessarily lead individuals to revise their beliefs medium 1 train +5940 AR6_WGII 1952 12 Psychological distancing–the perception that the greatest impacts occur sometime in the distant future and to people and places far away–can lead to discounting of risk and the need for adaptation medium 1 train +5941 AR6_WGII 1953 4 Defining coherent groups based on variations in beliefs, risk perceptions and policy preferences offers opportunities for effectively engaging with segments of the population instead of using the same approach for everyone low 0 train +5942 AR6_WGII 1953 20 Current and projected climate-change impacts disproportionately harm Indigenous Peoples’ livelihoods and economies very high 3 train +5943 AR6_WGII 1953 27 Climate-change impacts have harmful effects on Indigenous Peoples’ public health, physical health and mental health, including harmful effects connected to the cultural and community foundations of health very high 3 train +5944 AR6_WGII 1955 5 Such IK underpins successful climate-change adaptation and mitigation very high 3 train +5945 AR6_WGII 1956 8 Indigenous self-determination and self-governance are the foundations of adaptive strategies that improve understanding and research on climate change, develop actionable community plans and policies on climate change, and have demonstrable influence in improving the design and allocation of national, regional and international programmes relating to climate change very high 3 train +5946 AR6_WGII 1956 17 Ranges and abundances of species continue to shift in response to warming throughout North America very high 3 train +5947 AR6_WGII 1956 18 Future climate change will continue to affect species and ecosystems high 2 train +5948 AR6_WGII 1956 20 Climate-induced shifts in the timing of biological events (phenology) continue to be a well-documented ecological response very high 3 train +5949 AR6_WGII 1956 23 Severe ecosystem consequences of warming and drying are well documented very high 3 train +5950 AR6_WGII 1956 24 Significant ecosystem changes are expected from projected climate change high 2 train +5951 AR6_WGII 1957 2 Climate-change impacts on natural disturbances have affected ecosystems (very high confidence) (Table 14.2; see Box 14.2), and these impacts will increase with future climate change medium 1 train +5952 AR6_WGII 1957 7 Effects include widespread tree mortality (Allen et al., 2015; Kane et al., 2017; van Mantgem et al., 2018) and accelerated ecosystem transformation medium 1 train +5953 AR6_WGII 1957 10 Projected climate change will cause habitat loss, alter physical and bio- logical processes, and decrease water quality in freshwater ecosystems high 2 train +5954 AR6_WGII 1957 23 Effective climate-informed ecosystem management requires a well- coordinated suite of adaptation efforts (e.g., assessment, planning, funding, implementation and evaluation) that is co-produced among stakeholders, Indigenous Peoples and across sectors high 2 train +5955 AR6_WGII 1958 37 Higher ocean temperatures have directly affected food-web structure (Gibert, 2019) and altered physiological rates, distribution, phenology and behaviour of marine species with cascading effects on food-web dynamics very high 3 train +5956 AR6_WGII 1958 38 Pacific coastal waters from Mexico to Canada and US mid-Atlantic coastal waters have a high proportion of species (>5% of all marine species) near their upper thermal limit, representing hotspots of risk from MHWs medium 1 train +5957 AR6_WGII 1958 40 Climate change has induced phenological and spatial shifts in primary productivity with cascading impacts on food webs high 2 test +5958 AR6_WGII 1959 1 In response, increased burned area in recent decades in western North America has been facilitated by anthropogenic climate change medium 1 train +5959 AR6_WGII 1959 14 Climate change is projected to increase fire activity in many places in North America during the coming decades (see also AR6, WGI, Chapter 12, Ranasinghe et al., 2021) (Boulanger et al., 2014; Williams et al., 2016; Halofsky et al., 2020), via longer fire seasons (Wotton and Flannigan, 1993; USGCRP , 2017), long-term warming (Villarreal et al., 2019; Wahl et al., 2019) and increased lightning frequency in some areas of the USA and Canada medium 1 train +5960 AR6_WGII 1959 17 Impacts on Natural Systems Although fire is a natural process in many North American ecosystems, increases in burned area and severity of wildland fires have had significant impacts on natural ecosystems medium 1 train +5961 AR6_WGII 1959 24 Projected future fire activity will continue to affect ecosystems and alter their structure and function medium 1 train +5962 AR6_WGII 1959 27 Impacts on Human Systems Increased fire activity, partly attributable to anthropogenic climate change, has had direct and indirect effects on mortality and morbidity, economic losses and costs, key infrastructure, cultural resources and water resources medium 1 train +5963 AR6_WGII 1960 1 Poor air quality from fires caused increased respiratory distress very high 3 train +5964 AR6_WGII 1960 18 Adaptation Wildland fire risks are not equitably distributed as they intersect with exposure and socioeconomic attributes (e.g., age, income, ethnicity) to influence vulnerability and adaptive capacity medium 1 train +5965 AR6_WGII 1961 4 Although innovative, holistic approaches to wildland fire management are becoming more common across North America, broader application is necessary to address the growing risks medium 1 train +5966 AR6_WGII 1961 10 In North American Arctic marine systems, rapid warming is significant, with cascading impacts beyond polar regions (CCP6), and presents limited opportunities (tourism, shipping, extractive) but high risks (shipping, fishing industries, Indigenous subsistence and cultural activities) high 2 train +5967 AR6_WGII 1961 14 Coral reefs in the Gulf of Mexico and along the coasts of Florida and the Yucatan Peninsula are facing increasing risk of bleaching and mortality from warming ocean waters interacting with non-climate stressors very high 3 train +5968 AR6_WGII 1961 17 Without mitigation to keep surface temperatures below a 2°C increase by the end of the century, up to 99% of coral reefs will be lost; however, 95% of reefs will still be lost even if warming is kept below 1.5°C high 2 train +5969 AR6_WGII 1961 19 Sea level rise has led to flooding, erosion and damage to infrastructure along the western Gulf of Mexico, the southeast US coasts and the southern coast of the Gulf of St Lawrence very high 3 train +5970 AR6_WGII 1961 23 Future seawater CO 2 levels have been shown in laboratory studies to negatively impact Pacific and Atlantic squid, bivalve, crab and fish species (Pacific cod), and indirectly alter food-web dynamics high 2 train +5971 AR6_WGII 1962 2 Open ocean oxygen minimum zones (OMZ) are expanding in the North Atlantic, the North Pacific California Current and tropical oceans due to warming waters, stratification and changes in precipitation medium 1 train +5972 AR6_WGII 1962 5 The OMZs and hypoxic events are projected to increase over the next century and may limit where fish can move medium 1 train +5973 AR6_WGII 1962 6 Favourable conditions for harmful algal blooms (HABs) have expanded due to warming, more frequent extreme weather events (Gobler et al., 2017; Pershing et al., 2018; Trainer et al., 2019) and increased stratification, CO 2 concentration and nutrient inputs high 2 train +5974 AR6_WGII 1962 12 Attribution of Sargassum blooms to climate change is still tenuous and complicated by multiple drivers and few observational data sources low 0 train +5975 AR6_WGII 1963 2 Ocean management that utilises a portfolio of nested, multi-scale, climate-informed and ecosystem-based management approaches in North American waters can increase the resilience of marine ecosystems by addressing multiple stressors simultaneously high 2 train +5976 AR6_WGII 1963 4 Dynamic ocean management policies may improve resilience of marine species and ecosystems to climate medium 1 train +5977 AR6_WGII 1963 15 Adaptation will be impeded in cases where there are conflicts over competing interests or unintended consequences of uncoordinated efforts, heightening the importance of cooperative, scenario-based water resource planning and governance high 2 train +5978 AR6_WGII 1963 21 Damages stem from extremity of the event and prior land-use and infrastructure decisions high 2 train +5979 AR6_WGII 1964 10 Droughts have intensified tensions among competing water-use interests and accelerated depletion of groundwater resources high 2 train +5980 AR6_WGII 1964 11 Climate trends are affecting riverine, lake and reservoir water quality medium 1 train +5981 AR6_WGII 1964 16 Households and communities dependent on substandard wells, unimproved water sources or deficient water provision systems are more exposed than others to experience climate- related impairment of drinking water quality (Section 14.5.6.5; Allaire et al., 2018; Baeza et al., 2018; California State Water Resources Control Board, 2021; Navarro-Espinoza et al., 2021; Water and Tribes Initiative, 2021).14.5.3.2 Projected Impacts and Risks Climate change is projected to amplify current trends in water resource impacts, potentially reducing water supply security, impairing water quality and increasing flood hazards to varying degrees across North America high 2 train +5982 AR6_WGII 1964 18 Projected long-term reduction in water availability in the southwest US and northern Mexico (e.g., from the Colorado and Rio Grande rivers) will have substantial ecological and economic impacts given the region’s heavy water demands high 2 test +5983 AR6_WGII 1964 24 Other aquifers, especially those farther north, face uncertain or possibly increasing recharge medium 1 train +5984 AR6_WGII 1964 25 Projected changes in temperature and precipitation present direct risks to North American water quality, varying with regional and watershed contexts (Chapra et al., 2017; Coffey et al., 2019; Paul et al., 2019a), and related to streamflow, population growth (Duran-Encalada et al., 2017) and land-use practices medium 1 test +5985 AR6_WGII 1967 11 These changes directly influence crop productivity, quality and market price high 2 train +5986 AR6_WGII 1967 14 Some crop loss events are partially attributed to climate change high 2 train +5987 AR6_WGII 1967 17 Without adaptation, climate change is projected to reduce overall yields of important North American crops (e.g., wheat, maize, soybeans) high 2 train +5988 AR6_WGII 1967 19 Warming and heat extremes will delay or prevent chill accumulation, affecting perennial crop development (e.g., fruit set failure), yield (e.g., walnuts, pistachios, stone fruit) and quality (e.g., grapes) medium 1 train +5989 AR6_WGII 1967 20 Warming will alter the length of growing seasons of cold- season crops (e.g., broccoli, lettuce) and will shift suitability ranges of warm-season California crops (e.g., tomatoes) medium 1 test +5990 AR6_WGII 1967 21 Increasing atmospheric CO 2 will enhance yields yet reduce nutrient content of many crops high 2 train +5991 AR6_WGII 1967 22 Crop pest and pathogen outbreaks are expected to worsen under climate change high 2 train +5992 AR6_WGII 1967 23 Climate change is anticipated to cause declines in livestock production across North America high 2 train +5993 AR6_WGII 1967 25 Projected aridification reduces forage production in the southwest USA and northern Mexico (high confidence) (Polley et al., 2013; Reeves et al., 2014; Cooley, 2016; Bradford et al., 2020) and transforms grasslands into woody shrublands (Briske et al., 2015; Murray-Tortarolo et al., 2018), while warmer and wetter conditions in the northern regions (CA-PR, US-NW, US-NP) may enhance rangeland production by extending growing seasons high 2 train +5994 AR6_WGII 1967 26 Increased CO 2 will enhance production (medium confidence) but reduce forage quality high 2 train +5995 AR6_WGII 1967 27 Climate-change impacts on forests (Section 14.5.1; see Box 14.2) may affect timber production by altering tree species distributions, productivity, and wildfire and insect disturbances medium 1 train +5996 AR6_WGII 1968 7 Climate-driven declines in productivity are widespread (high confidence) (Figure 14.6), although a few increases are observed in northern regions medium 1 train +5997 AR6_WGII 1968 8 Redistribution of species has increased travel distance to fishing grounds, shifted stocks across regulatory and international boundaries, and increased interactions with protected species very high 3 train +5998 AR6_WGII 1968 9 Climate shocks have reduced yield and increased instability in fishery revenue high 2 train +5999 AR6_WGII 1968 10 Declines in yield and poleward stock redistributions (an average of ~20.6 km per decade) are expected to continue under climate change and increase in magnitude with atmospheric carbon high 2 train +6000 AR6_WGII 1968 12 Expanding HABs, pathogens and altered ocean chemistry (OA and dissolved oxygen) will reduce yields and increase closures of fisheries along all North American coasts medium 1 train +6001 AR6_WGII 1968 14 Warming waters and OA have impacted aquaculture production in North America high 2 train +6002 AR6_WGII 1969 4 Ecosystem-based practices and sustainable intensification (increasing yields while minimising resource demand and ecosystem impacts) (Cassman and Grassini, 2020; Rockström et al., 2021) will help the sector meet food production demands under climate change (medium confidence), but effectiveness generally declines and is less certain after 2050 in scenarios without carbon mitigation high 2 train +6003 AR6_WGII 1969 5 Across the sector, successful adaptation is underpinned by approaches that meaningfully consider the coupled social–ecological networks around food and fibre production and value IK very high 3 train +6004 AR6_WGII 1969 6 Integrated modelling, participatory planning and inclusive decision making promote effective and equitable adaptation responses very high 3 train +6005 AR6_WGII 1970 5 Irrigation is an effective adaptation strategy in key agricultural areas (Miller, 2017; Lund et al., 2018) and could stabilise food security in rain-fed regions (e.g., southeast Mexico) (Spring, 2014); water allocation must balance multiple needs and rights medium 1 train +6006 AR6_WGII 1973 2 Climate-informed and standardised aquaculture governance, and increased coordination with fishery and coastal management, is needed for climate resilience high 2 train +6007 AR6_WGII 1973 13 These environmental conditions also stress natural assets (e.g., urban forests, wetlands, household gardens, green walls) and performance of green infrastructure leading to higher operation and maintenance costs high 2 train +6008 AR6_WGII 1974 6 Climate change (RCP8.5) interacting with urban form, development and systemic racism (Schell et al., 2020; Hsu et al., 2021) could worsen risks from extreme heat in North American cities, especially where there is limited adaptation high 2 train +6009 AR6_WGII 1974 23 In Canada, SLR is expected to increase the frequency and magnitude of extreme high-water-level events (Greenan et al., 2018) and to create widespread impacts on natural and human systems high 2 train +6010 AR6_WGII 1975 3 Barriers to adaptation include challenges related to the local physical and environmental setting, effects of colonialism and racism, socioeconomic attributes of the population, institutional frameworks and competing interests of city stakeholders medium 1 train +6011 AR6_WGII 1975 4 The current scale of adaptation is generally not commensurate with reducing risks from projected climatic hazards, although resources exist that provide guidance and examples of effective adaptation medium 1 train +6012 AR6_WGII 1975 10 Public health measures to address extreme heat events are more common across North America, with a focus on vulnerable populations (e.g., City of Toronto, 2019) and innovative approaches for reaching at-risk populations with an overarching intent of prevention medium 1 train +6013 AR6_WGII 1975 12 Other adaptation responses to reduce temperature effects include modifying structures (roofs, engineered materials) and the urban landscape through green infrastructure (e.g., urban trees, wetlands, green roofs), which increases climate resilience and quality of life by reducing urban heat island effects, while additionally improving air quality, capturing stormwater and delivering other co-benefits to the community (e.g., access to food, connection to nature, social connectivity) high 2 train +6014 AR6_WGII 1975 23 Adaptation planning and implementation to address SLR and coastal flooding have been initiated across many cities and settlements in North America, but preparedness varies high 2 train +6015 AR6_WGII 1978 11 Adaptation to the risks of wildland–urban interface fire is underway (see Box 14.2; Kovacs et al., 2020), but the scope of adaptation required to sufficiently minimise wildfire risks for cities and settlements across North America has not been assessed medium 1 train +6016 AR6_WGII 1978 13 Cities and settlements in North America can be susceptible to multiple flooding hazards (i.e., coastal SLR, pluvial or fluvial flooding); each presents unique adaptation challenges that can be addressed through structural (e.g., armouring coastlines, reservoirs, levees, floodgates; New York City commuter tunnels) and non-structural approaches (e.g., land- use planning and zoning, expanding green infrastructure; Chetumal, Mexico) high 2 train +6017 AR6_WGII 1978 20 The civil engineering profession is playing an active role in facilitating an understanding of risks and prioritisation of adaptation investments in communities (Tye and Giovannettone, 2021).The high concentration of valuable assets in cities requires mechanisms to facilitate replacement of assets including use of existing and proposed insurance mechanisms medium 1 train +6018 AR6_WGII 1979 11 Rising temperatures are projected to increase heat-related mortality across emission scenarios this century in North America very high 3 train +6019 AR6_WGII 1979 13 Warming temperatures are also projected to increase heat-related morbidity medium 1 train +6020 AR6_WGII 1979 15 While heat-related mortality is projected to increase across emissions scenarios and shared socioeconomic pathways, fewer deaths are projected under both lower-emissions scenarios and higher-adaptation scenarios in North America very high 3 train +6021 AR6_WGII 1979 18 It is important to differentiate between Table 14.5 | A summary of adaptation options for different health outcomes in North America Health outcome Adaptation options Heat-related mortality and morbidityFuture temperature-related health impacts can be reduced by adaptation measures (Petkova et al., 2014; Wu et al., 2014; Mills et al., 2015b; Kingsley et al., 2016; Anderson et al., 2018b; Marsha et al., 2018; Morefield et al., 2018), including more effective warning and response systems and building designs, enhanced pollution controls, urban planning strategies and resilient health infrastructure very high 3 train +6022 AR6_WGII 1979 23 Water-borne diseaseClimate change is projected to increase water-borne disease risks (medium confidence), particularly in areas with ageing water and wastewater infrastructure in North America high 2 train +6023 AR6_WGII 1979 26 Food-borne diseaseFood safety programmes play important roles in reducing the risk of climate-related food-borne disease high 2 train +6024 AR6_WGII 1980 1 Warmer temperatures do not always equate to lower winter mortality: many cold-related deaths do not occur during the coldest times of year or in the coldest places high 2 train +6025 AR6_WGII 1980 20 Climate change is projected to increase disease spread into new geographic regions, lengthen the season of disease transmission and increase tick-borne disease risk in North America across emissions scenarios throughout this century very high 3 train +6026 AR6_WGII 1980 23 Climate change is projected to impact the distribution, abundance and infection rates of mosquitoes in North America (high confidence), which will increase risk of mosquito-borne diseases including West Nile virus, chikungunya and dengue medium 1 train +6027 AR6_WGII 1981 9 In North America, stormwater and water treatment infrastructure play important roles in reducing water-borne disease risk during precipitation events high 2 train +6028 AR6_WGII 1981 20 Climate change is projected to increase food safety risks (medium confidence); however, the actual burden of food-borne disease will depend on the efficacy of public health interventions high 2 train +6029 AR6_WGII 1981 25 Climate change is projected to increase human food-borne exposure to chemical contaminants medium 1 train +6030 AR6_WGII 1981 28 Climate-related food-borne disease risks vary temporally, and are influenced, in part, by food availability, accessibility, preparation and preferences medium 1 test +6031 AR6_WGII 1982 7 Substitution of seafood with non-traditional foods (e.g., chicken, canned tuna) would not replace the projected nutrients lost (Marushka et al., 2019), challenging assumptions that market food substitutions could be effective adaptation strategies for Indigenous Peoples 14.5.6.8 Mental Health and Wellness Climate change has had, and will continue to have, negative impacts on mental health in North America high 2 train +6032 AR6_WGII 1982 19 Season length for snowmobiling and cross-country skiing is projected to decrease more dramatically high 2 train +6033 AR6_WGII 1984 9 The risk of a major accident or incident among Arctic-going yachts and some expedition passenger vessels is very high relative to other ships high 2 train +6034 AR6_WGII 1984 24 However, the cascading nature of climate impacts related to trade (see Box 14.5), labour productivity (Section 14.5.8.1.5) and infrastructure (Section 14.5.8.1.2) means that there is no economic sector in North America that will be unaffected by climate change very high 3 train +6035 AR6_WGII 1986 3 Operational efficiency and human safety at mining and energy production sites is expected to be adversely affected by increases in extreme events (Section 14.2), including storms, heavy rains, riverine flooding and wildfires high 2 train +6036 AR6_WGII 1986 4 General remoteness of many mining sites (especially in the North American Arctic) exacerbates risks related to emergency responses to extreme events such as wildfire medium 1 train +6037 AR6_WGII 1986 18 Permafrost thaw in northern North America will result in increased construction and reconstruction needs medium 1 train +6038 AR6_WGII 1988 5 Climate risks may create shocks to the trade system by damaging infrastructure and disrupting supply chains in North America medium 1 train +6039 AR6_WGII 1988 8 Due to the transnational nature of trade, extreme weather disruptions in one region are likely to lead to cascading effects in other regions high 2 train +6040 AR6_WGII 1988 10 Climate-change impacts may alter current trade practices and patterns with implications for regional economic development in North America, especially in the Arctic medium 1 train +6041 AR6_WGII 1988 15 Effective and equitable trade policies can act as important adaptation strategies medium 1 train +6042 AR6_WGII 1988 22 These differences have been often underpinned by social and economic inequalities and have been observed between households, social groups, rural and urban communities, and Indigenous Peoples high 2 train +6043 AR6_WGII 1988 23 These vulnerabilities have also been observed to contribute to maladaptation medium 1 train +6044 AR6_WGII 1988 24 Social and economic trends and development will determine near-term impacts on livelihoods from projected climate hazards; livelihoods will also adapt to the risks and opportunities high 2 train +6045 AR6_WGII 1988 25 Actions to enhance the livelihoods of the most vulnerable social groups in North America will lessen the impacts of climate hazards on them high 2 train +6046 AR6_WGII 1988 28 Past and current patterns of development in North America have propagated and perpetuated vulnerabilities that have created differential impacts on livelihoods from climate hazards high 2 train +6047 AR6_WGII 1989 1 Direct, indirect and non-market economic damages from extreme events have increased in some parts of North America high 2 train +6048 AR6_WGII 1989 11 The effect of climate change has been identified in aggregate measures of economic performance, such as GDP, in North America and globally (medium confidence), although the magnitude of these changes is difficult to constrain medium 1 train +6049 AR6_WGII 1989 12 Climate change has been observed to affect national GDP level and economic growth low 0 test +6050 AR6_WGII 1989 15 Projected Risks Projections of market and non-market economic damages demonstrate the substantial economic risks of climate impacts associated with high-temperature pathways (RCP8.5) high 2 train +6051 AR6_WGII 1989 21 Market and non-market risks and costs will not be experienced equally across countries, sectors and regions in North America high 2 train +6052 AR6_WGII 1989 26 Economics of Adaptation Opportunities Economic analysis can help reveal where the avoided economic damages are greater than the costs of adaptation, improving decision making for adaptation planning and efforts in North America high 2 train +6053 AR6_WGII 1990 3 Climate impacts have damaged livelihoods across North America, especially those of marginalised people (high confidence) and deepened inequalities for these groups medium 1 train +6054 AR6_WGII 1990 10 Migration and mobility have been an important part of livelihoods in North America high 2 train +6055 AR6_WGII 1990 13 Temporary or semi- permanent labour migration, generally followed by remittances, has been an important part of livelihoods for rural areas in Mexico high 2 train +6056 AR6_WGII 1990 17 Pre-existing social vulnerabilities have also led to forced displacement from extreme weather events low 0 test +6057 AR6_WGII 1990 25 Improving projections of future economic risk and damages facilitates the development of tools that can be used for economic analysis of climate policies high 2 train +6058 AR6_WGII 1991 8 Livelihoods, however, can be undermined by many of the projected climate risks with the impacts depending on adaptive capacity and adaptation limits high 2 train +6059 AR6_WGII 1991 12 Future climate hazards will deepen patterns of social inequality as vulnerable groups may also experience intersecting impacts that adversely affect their livelihoods medium 1 train +6060 AR6_WGII 1991 15 Displacement, migration and resettlement will increase along higher- emission pathways medium 1 train +6061 AR6_WGII 1991 22 Many actions that enhance and promote resilient livelihoods can have substantial benefit for adaptation to climate hazards medium 1 train +6062 AR6_WGII 1991 29 Migration is a common adaptation strategy to maintain and diversify people’s livelihoods and will continue to play an important role when households manage climate and social risks high 2 train +6063 AR6_WGII 1991 34 If social relationships prevailing now and in the recent past continue, projections show future crime rates in the USA and Mexico increasing with increasing temperatures low 0 train +6064 AR6_WGII 1993 21 While there is expert scientific consensus on anthropogenic climate change, rhetoric, misinformation and politicisation of science have contributed to misperceptions, polarisation on the severity of impacts and risks to society, indecision and delayed action high 2 test +6065 AR6_WGII 1993 23 KR2: Risk to life, safety and property from intensifying extreme events Human life and safety across North America, and especially along the coasts of Mexico, the Hawaiian Islands, Gulf of Mexico, Atlantic Canada and southeast USA, will be placed at risk from SLR and severe storms and hurricanes, even at 1.5°C GWL very high 3 train +6066 AR6_WGII 1995 1 KR3: Cumulative damages from climate hazards which pose a substantial risk to economic well-being and shared prosperity Climate-change impacts are projected to cause large market and non- market damages high 2 train +6067 AR6_WGII 1995 6 KR4: Risk of degradation of marine and coastal ecosystems, including loss of biodiversity, function and related services with cascading effects for communities and livelihoods Ocean warming will increase the frequency and intensity of MHWs (see Box 14.3), accelerate unprecedented rates of sea ice loss, and alter ocean circulation, chemistry and nutrient cycling in ways that profoundly impact marine productivity, biodiversity and food webs very high 3 train +6068 AR6_WGII 1995 7 Collectively these impacts pose a risk to nearshore ecological and human systems (high confidence), increasing the probability of phenological mismatches, large-scale redistribution of species, and species population declines (Section 14.5.4) with cascading impacts that strain cultural and economic systems reliant on marine productivity across North America high 2 train +6069 AR6_WGII 1995 9 KR5: Risk to major terrestrial ecosystems leading to disruptions of species, ecosystems and their services Major risks to terrestrial ecosystems across North America, such as semiarid landscapes, rangelands, boreal and temperate forests, and Arctic tundra, include significant ecosystem transformations and shifts in species abundances and ranges, and major vegetation types (e.g., transitions from forests to grasslands), with cascading implications for regional biodiversity very high 3 train +6070 AR6_WGII 1995 10 Warming increases the risk of permafrost thaw with propagating impacts on species and communities in the Canadian and US Arctic high 2 train +6071 AR6_WGII 1995 12 These changes will reduce services provided by terrestrial ecosystems, including timber yields and carbon sequestration medium 1 train +6072 AR6_WGII 1995 14 Streams in North America are expected to continue to warm, with important ramifications for aquatic ecosystems high 2 train +6073 AR6_WGII 1995 15 Warming and drying coupled with other stressors (e.g., pollutants, nutrients and invasive species) pose a risk to ecosystem structure and function in lakes, streams and reservoirs across many parts of North America high 2 train +6074 AR6_WGII 1995 17 KR7: Risk to human health and well-being, including mental health Heat-related human mortality is projected to increase in North America as a result of climate change and ageing populations, poverty, chronic diseases and inadequate public health systems very high 3 train +6075 AR6_WGII 1995 18 Gradual changes to temperature and precipitation are impacting urban ecosystems and creating ecosystem regime changes resulting in the poleward expansion among insects that bring risks related to vector-borne diseases such as West Nile virus and Lyme disease high 2 train +6076 AR6_WGII 1995 19 Climate change is expected to lead to wide-ranging mental health challenges related to an increase in the psychological burdens of climate change high 2 train +6077 AR6_WGII 1995 20 KR8: Risk to food and nutritional security through changes in agriculture, livestock, hunting, fisheries and aquaculture productivity and access Cascading and interacting impacts of climate change threatens food systems as well as food and nutritional security for many North Americans, especially those already experiencing food and nutritional scarcity, women and children with high nutritional needs and Indigenous Peoples reliant on subsistence resources high 2 train +6078 AR6_WGII 1995 21 In agricultural regions experiencing aridification and where water scarcity precludes substantial expansion of irrigation, warming and extreme heat pose a risk to food and forage crop and livestock production high 2 train +6079 AR6_WGII 1995 22 Ocean warming and MHWs will continue to disrupt commercial capture fisheries through species redistribution and changes to yield (high confidence), and warming waters and OA will increasingly impact aquaculture production high 2 train +6080 AR6_WGII 1996 1 KR9: Risks to major infrastructure supporting commerce and trade with implications for sustainable economic development, regional connections and livelihoods Climate change and extreme events are expected to increase risks to the North American economy via infrastructure damage and deterioration (high confidence), disruption to operations, unsafe conditions for workers (medium confidence) and interruptions to international and inter-regional supply chains medium 1 train +6081 AR6_WGII 1996 2 These climatic impacts will have cascading implications for local livelihoods, sustainable economic development pathways and regional connectivity, and will reinforce pre-existing social inequities medium 1 test +6082 AR6_WGII 1996 3 Infrastructure damage will also disrupt economic activities, including manufacturing, tourism, fisheries, natural resource extraction and energy production high 2 train +6083 AR6_WGII 1996 4 KR10: Risk to the quality of life in North American communities, cities and towns In major North American cities and settlements, vulnerability to climate change has increased and is projected to continue to rise medium 1 train +6084 AR6_WGII 1996 6 Coastal, riverine and urban flooding displacing communities and coastal ecosystems (Section 14.5.5.2) will become a dominant risk to urban centres high 2 train +6085 AR6_WGII 1996 13 Under low- mitigation scenarios, compounding risks and higher-carbon-emission scenarios increase the potential that amplifying feedback loops and fatal synergies across sectors could lead to existential threats to the social–ecological systems of North America medium 1 train +6086 AR6_WGII 1996 16 Multiple lines of evidence across sectors assessed in this chapter suggest that after mid-century and without carbon mitigation, climate-driven changes to ecological and social boundary conditions may rapidly push many systems into disequilibrium medium 1 train +6087 AR6_WGII 2000 1 Support for, and implementation of, adaptation policies, plans and measures have not been equal across the public and private sectors, regions or varying levels of governance high 2 train +6088 AR6_WGII 2000 2 To date, reactive (coping-based) and incremental adaptations have helped North Americans avoid greater damages from observed climate impacts medium 1 train +6089 AR6_WGII 2000 3 There is increasing agreement that worsening impacts and expanding risk conditions may exceed current adaptation capacities by mid-century under high-emissions scenarios (RCP8.5) medium 1 train +6090 AR6_WGII 2000 10 These efforts largely have focused on planning and less on implementation high 2 train +6091 AR6_WGII 2000 11 Some subnational governments, namely states and provinces, have engaged in advanced adaptation planning efforts high 2 train +6092 AR6_WGII 2002 19 The absence of evidence about the current effectiveness of proposed adaptation actions to guide future actions and investments presents a serious risk to North America, especially at higher GWLs medium 1 train +6093 AR6_WGII 2002 22 At current warming levels, social–ecological systems have been reaching limits to adaptation in regions with high exposure and high sensitivity medium 1 train +6094 AR6_WGII 2004 2 Adaptation actions in one place or sector can have adverse side effects elsewhere medium 1 train +6095 AR6_WGII 2004 15 Incorporating different values and knowledge systems, consideration of equity and justice as core objectives and addressing underlying vulnerabilities are principles that can guide transformational adaptation and resilience medium 1 train +6096 AR6_WGII 2007 2 Successful nature-based adaptation draws from existing adaptation approaches (Borsje et al., 2011; Temmerman et al., 2013; Law et al., 2018; Reguero et al., 2018; Buotte et al., 2019) and is applied across ecological and human systems high 2 train +6097 AR6_WGII 2007 4 Nature-based adaptation is generally less expensive and strengthens over time, as compared with built infrastructure which erodes with time medium 1 train +6098 AR6_WGII 2008 2 Effective nature-based adaptation requires a well-coordinated suite of adaptation efforts (e.g., assessment, planning, funding, implementation and evaluation) that is co- produced among stakeholders and across sectors high 2 train +6099 AR6_WGII 2056 2 The observed impacts of climate change differ between urban and rural contexts, island types and tropical and non-tropical islands high 2 train +6100 AR6_WGII 2056 4 Climate change is also affecting settlements and infrastructure, health and well-being, water and food security, and economies and culture, especially through compound events high 2 train +6101 AR6_WGII 2056 6 TCs are severely impacting small islands high 2 train +6102 AR6_WGII 2056 15 Scientific evidence has confirmed that globally and in small islands tropical corals are presently at high risk high 2 train +6103 AR6_WGII 2056 16 Severe coral bleaching, together with declines in coral abundance, has been observed in many small islands, especially those in the Pacific and Indian oceans high 2 train +6104 AR6_WGII 2056 29 Due to the large range of insular- related vulnerabilities, almost 50% of terrestrial species presently considered at risk of global extinction also occur on islands high 2 test +6105 AR6_WGII 2056 31 Projected Impacts Projected climate and ocean-related changes will significantly affect marine and terrestrial ecosystems and ecosystem services, which will in turn have cascading impacts across both natural and human systems high 2 train +6106 AR6_WGII 2056 32 Changes in wave climate superimposed on SLR will significantly increase coastal flooding high 2 train +6107 AR6_WGII 2056 33 The frequency, extent, duration and consequences of coastal flooding will significantly increase from 2050 (high confidence), unless coastal and marine ecosystems are able to naturally adapt to SLR through vertical growth low 0 train +6108 AR6_WGII 2056 35 Projected changes in the wave climate superimposed on SLR will rapidly increase flooding in small islands, despite highly contrasting exposure profiles between ocean sub-regions high 2 train +6109 AR6_WGII 2057 2 Modelling of both temperature and ocean acidification effects under future climate scenarios (RCP4.5 and RCP8.5) suggest that some small islands will experience severe coral bleaching on an annual basis before 2040 medium 1 train +6110 AR6_WGII 2057 4 Intact coral reefs, seagrass meadows and mangroves provide a variety of ecosystem services that are important to island communities high 2 train +6111 AR6_WGII 2057 5 These include provisioning services, regulating services, cultural services and those that support community resilience high 2 train +6112 AR6_WGII 2057 6 If coastal ecosystems are degraded and lost, then the benefits they provide cannot be easily replaced medium 1 train +6113 AR6_WGII 2057 7 Projected changes in aridity are expected to impose freshwater stress on many small islands, especially SIDS high 2 train +6114 AR6_WGII 2057 13 SIDS with high projected population growth rates are expected to experience the most severe freshwater stress by 2030 under a 2°C warming threshold scenario {15.3.3.2} The continued degradation and transformation of terrestrial and marine ecosystems of small islands due to human- dominated will amplify the vulnerability of island peoples to the impacts of climate change high 2 train +6115 AR6_WGII 2057 19 Reef island and coastal area habitability in small islands is expected to decrease because of increased temperature, extreme sea levels and degradation of buffering ecosystems, which will increase human exposure to sea-related hazards high 2 train +6116 AR6_WGII 2057 21 On small islands, coastal land loss attributable to higher sea level, increased extreme precipitation and wave impacts and increased aridity have contributed to food and water insecurities that are likely to become more acute in many places high 2 train +6117 AR6_WGII 2057 25 Future Risks The reduced habitability of small islands is an overarching significant risk caused by a combination of several key risks facing most small islands even under a global temperature scenario of 1.5°C high 2 train +6118 AR6_WGII 2057 29 Moreover, it can distil the benefits and/or disadvantages and long-term implications of choosing such options high 2 train +6119 AR6_WGII 2057 30 The vulnerability of communities in small islands, especially those relying on coral reef systems for livelihoods, may exceed adaptation limits well before 2100 even for a low greenhouse gas emission pathway high 2 train +6120 AR6_WGII 2057 31 The impacts of climate change on vulnerable low-lying and coastal areas present serious threats to the ability of land to support human life and livelihood high 2 train +6121 AR6_WGII 2058 6 Options, Limits and Opportunities of Adaptation Some island communities are resilient with strong social safety nets and social capital that support responses and actions already occurring, but there is limited information on the effectiveness of the adaptation practices and the scale of action needed high 2 train +6122 AR6_WGII 2058 9 In small islands, despite the existence of adaptation barriers several enablers can be used to improve adaptation outcomes and to build resilience high 2 train +6123 AR6_WGII 2058 10 These enablers include better governance and legal reforms; improving justice, equity and gender considerations; building human resource capacity; increased finance and risk transfer mechanisms; education and awareness programmes; increased access to climate information; adequately downscaled climate data and embedding Indigenous knowledge and local knowledge (IKLK) as well as integrating cultural resources into decision-making high 2 train +6124 AR6_WGII 2058 11 Small islands present the most urgent need for investment in capacity building and adaptation strategies high 2 train +6125 AR6_WGII 2058 13 Additionally, institutional and legal systems are often inadequately prepared for managing adaptation strategies such as large-scale settlement relocation and other planned and/or autonomous responses to climate risks high 2 train +6126 AR6_WGII 2058 14 Adaptation strategies are already being implemented on some small islands although barriers are encountered including inadequate up-to-date and locally relevant information, limited availability of finance and technology, lack of integration of IKLK in adaptation strategies, and institutional constraints (high confidence) {15.5.3, 15.5.4, 15.6.3, 15.6.4, 15.6.5}.For many small islands, adaptation actions are often incremental and do not match the scale of extreme or compounding events high 2 train +6127 AR6_WGII 2058 16 To address these shortcomings, enablers are being integrated into National Adaptation Plans and Disaster Risk Reduction Plans high 2 train +6128 AR6_WGII 2058 17 Although international climate finance has increased in magnitude, small islands face challenges in accessing adaptation finance to cope with slow- and rapid-onset events high 2 train +6129 AR6_WGII 2059 19 Also, scientific evidence since AR5 has confirmed that tropical corals are presently at high risk very high 3 train +6130 AR6_WGII 2059 20 Even achieving emission reduction targets consistent with the ambitious goal of 1.5°C of global warming under the Paris Agreement will result in the further loss of 70–90% of reef-building corals compared to today, with 99% of corals being lost under warming of 2°C or more above the pre-industrial period high 2 train +6131 AR6_WGII 2061 5 Stronger evidence confirms that education and awareness-raising enhance household and community adaptation high 2 train +6132 AR6_WGII 2061 6 Knowledge has improved on limits to adaptation, including projected timeframes of limits for hard protection (high confidence) and EbA medium 1 train +6133 AR6_WGII 2061 7 There is also a better understanding that barriers and governance challenges vary by island and island groups high 2 train +6134 AR6_WGII 2061 12 In small islands the methods and mechanisms to assess climate-induced losses and damages remain undeveloped medium 1 train +6135 AR6_WGII 2061 18 This is true on all types of islands (Figure 15.2).15.3.1 Synthesis of Observed and Projected Changes in the Physical Basis There is increased evidence of warming in the small islands, particularly in the latter half of the 20th century high 2 train +6136 AR6_WGII 2064 9 Reconstructions of past storm surges and modelling studies assessing storm surge risk similarly highlighted high variations of risk along island coasts, due to variations in exposure, topography and bathymetry high 2 train +6137 AR6_WGII 2066 33 Larger-scale studies confirmed that projected changes in the wave climate superimposed on SLR will rapidly increase flooding in small islands, despite highly contrasting exposure profiles between ocean sub-regions high 2 train +6138 AR6_WGII 2066 38 Since the 1950s–1970s, and even in regions exhibiting higher than global-averaged SLR rates, atoll islands maintained their land area high 2 train +6139 AR6_WGII 2067 1 The rates of change did not correlate with SLR rates, suggesting that the impact of SLR on island land area was obscured by other climate drivers and human disturbances on some islands high 2 train +6140 AR6_WGII 2067 3 Despite important knowledge gaps on coastal erosion in high tropical islands, recent studies confirmed increasing shoreline retreat and beach loss over the past decades, mainly due to TC and ETC waves and human disturbances high 2 train +6141 AR6_WGII 2067 4 Despite storm-induced erosion prevailing along some shoreline sections, recent studies reaffirmed the contribution of TC and ETC waves to coastal and reef island vertical building through massive reef- to-island sediment transfer high 2 train +6142 AR6_WGII 2067 8 Similarly, El Niño and La Niña were involved in rapid and highly contrasting shoreline changes high 2 train +6143 AR6_WGII 2067 13 Small reef islands and narrow coastal systems affected by human disturbances will increasingly be at risk of disappearance due to SLR (KR2 in Figure 15.5), enhanced sediment loss caused by extreme events (Duvat et al., 2019a) and/or human activities high 2 train +6144 AR6_WGII 2067 17 Severe coral bleaching, together with declines in coral abundance have been documented in many small islands, especially those in the Pacific Ocean and Indian Ocean (e.g., Guam, Fiji, Palau, Vanuatu, Chagos, Comoros, Mauritius, Seychelles, and the Maldives high 2 train +6145 AR6_WGII 2067 19 Median return time between two severe bleaching events has diminished steadily since 1980 and is now only 6 years (e.g., Hughes et al., 2017b; Hughes et al., 2018) and is often associated with warm phase of ENSO events high 2 train +6146 AR6_WGII 2067 20 Modelling of both bleaching and ocean acidification effects under future climate scenarios suggested that some Pacific small islands (e.g., Nauru, Guam, Northern Marianas Islands) will experience conditions that cause severe bleaching on an annual basis before 2040 and that 90% of the world reefs are projected to experience conditions that result in severe bleaching annually by 2055 medium 1 train +6147 AR6_WGII 2067 22 Even achieving emission reduction targets consistent with the ambitious goal of 1.5°C of global warming under the Paris Agreement will result in the further loss of 70–90% of reef-building corals compared to today, with 99% of corals being lost under warming of 2°C or more above the pre-industrial period high 2 train +6148 AR6_WGII 2068 8 Despite their vital social and ecological value, substantial declines in seagrass communities have been documented in many small islands (Section 3.4.2.5; Arias-Ortiz et al., 2018; Kendrick et al., 2019; Brodie et al., 2020), including Fiji (Joseph et al., 2019), Reunion Island (Cuvillier et al., 2017), Bermuda, Cayman Islands, US Virgin Islands (Waycott et al., 2009), Kiribati (Brodie et al., 2020), Federated States of Micronesia, and Palau (Short et al., 2016), but attribution of such declines to climatic influences remains weak low 0 train +6149 AR6_WGII 2068 13 In the Mediterranean, seagrass meadows are already showing signs of regression, which may have been aggravated by climate change high 2 train +6150 AR6_WGII 2068 22 Whether or not such events are related to long-term climate change remains unclear; however, it has been suggested that the influx may be related to strong Amazon discharge, enhanced West African upwelling, together with rising seawater temperatures in the Atlantic low 0 train +6151 AR6_WGII 2069 5 In small islands where the risk of loss to ecosystem services is high (Cross-Chapter Box DEEP in Chapter 17), many of these ecosystem services cannot be easily replaced medium 1 train +6152 AR6_WGII 2069 15 As corals, mangroves and seagrasses disappear, so do fish and other dependent organisms that directly benefit industries such as ecotourism and fisheries high 2 train +6153 AR6_WGII 2069 23 On high volcanic and granitic islands, freshwater ecosystems are often closely connected with coastal spaces, and changes in freshwater supply from river systems have direct implications for salinity and sediment loads high 2 train +6154 AR6_WGII 2069 24 Climate impacts on streamflow patterns in tropical islands also create shifts in water supply for downstream users and habitat conditions for organisms supporting a wide range of ecosystem services high 2 train +6155 AR6_WGII 2069 25 Projected changes in aridity are expected to impose freshwater stress on many small islands, especially SIDS high 2 train +6156 AR6_WGII 2069 26 In the Mediterranean region, freshwater resources will decline by 10–30% medium 1 train +6157 AR6_WGII 2071 7 Such changes in SLR could increase salinity in estuarine and aquifer water, affecting ground and surface water resources for drinking and irrigation water (Mycoo, 2018a) across the region high 2 train +6158 AR6_WGII 2071 14 This is likely to hinder the adaptation response of terrestrial biota–increasing the risk of biodiversity loss and, in turn, impairing the resilience capacity of ecosystem functioning and services high 2 train +6159 AR6_WGII 2071 24 This is due to projected strong shifts, reductions or even complete losses of climatic niches resulting from inadequate geographic space for species to track suitable climate envelopes high 2 train +6160 AR6_WGII 2072 3 Indirect impacts of SLR may potentially result in equal or more biodiversity loss than direct impacts medium 1 train +6161 AR6_WGII 2072 6 Tropical island natural habitats/systems are highly vulnerable to extreme weather events such as TCs, due to their small size, unique ecological systems and often low socioeconomic capacity high 2 test +6162 AR6_WGII 2072 15 There are too few studies available to suggest potential future response trends of these ecosystems to this increased intensity; however, it seems plausible that present resilience capacities may be adversely impacted medium 1 train +6163 AR6_WGII 2072 21 These may ultimately increase the risk of multiple extinctions, negatively impacting upon global biodiversity levels high 2 train +6164 AR6_WGII 2072 25 These are likely to enhance IAS impacts on islands including: restructuring of ecological communities leading to declines and extinctions/extirpations in flora and fauna, habitat degradation, declining ecosystem functioning, services and resilience and, in extreme cases, potential community homogenisation high 2 train +6165 AR6_WGII 2072 27 Compared to continents, terrestrial IAS are disproportionately prevalent on islands (almost three quarters of global species currently threatened by IAS and disease are found on islands) and also generate stronger impacts (e.g., within alpine ecosystems of high islands) than on continents high 2 train +6166 AR6_WGII 2074 6 The main settlements of small islands are located along the coast and with decades of high-density coastal urban development, their population, buildings and infrastructure are currently exposed to multiple climate change-related hazards (Kumar and Taylor, 2015; Mycoo, 2017) and face key risks high 2 train +6167 AR6_WGII 2077 9 These authors concluded that nine of 17 Pacific Island entities (Cook Islands, Federated States of Micronesia, Guam, Kiribati, Marshall Islands, Niue, Papua New Guinea, Solomon Islands, and Tuvalu) could experience ≥50% declines in maximum catch potential by 2100 relative to 1980–2000 under both an RCP2.6 and RCP8.5 scenario medium 1 train +6168 AR6_WGII 2077 12 The small islands that show the largest anticipated decrease in the maximum catch potential of fisheries by the end of the century (according to an RCP4.5 and RCP8.5 scenario) include the Federated States of Micronesia, Kiribati, Nauru, Palau, Tokelau, Tuvalu, São Tomé and Príncipe, whereas some other small islands such as Bermuda, Easter Island (Chile), and Pitcairn Islands (UK), might actually witness increases in fish catch potential medium 1 train +6169 AR6_WGII 2080 25 These KRs include loss of marine and coastal biodiversity and ecosystem services (high confidence) (KR1; for details on KR coverage, see Section 15.3.3.1); submergence of reef islands (low confidence) (KR2; Section 15.3.3.1.1); loss of terrestrial biodiversity and ecosystem services (high confidence) (KR3; Section 15.3.3.3); water insecurity (medium- high confidence) (KR4; Section 15.3.4.3); destruction of settlements and infrastructure (high confidence) (KR5; Section 15.3.4.1); degradation of human health and well-being (low confidence) (KR6; Section 15.3.4.2); economic decline and livelihood failure (high confidence) (KR7; Sections 15.3.4.4 and 15.3.4.5); and loss of cultural resources and heritage low 0 train +6170 AR6_WGII 2084 14 Risk accumulation and amplification through cascading effects from ecosystems and ecosystem services to human systems will likely cause reduced habitability of some small islands high 2 train +6171 AR6_WGII 2136 3 The geographic ranges of the animal and plant species assessed have shifted from low to high latitudes in response to climate warming on land and in the ocean very high 3 train +6172 AR6_WGII 2136 4 On land, climate change-induced shifts towards higher elevations are also common in biodiversity hotspots high 2 train +6173 AR6_WGII 2136 5 In the ocean, abrupt mortality of habitat- forming species on coral reefs and kelp forests, especially following heatwaves, are increasing in frequency in biodiversity hotspots high 2 train +6174 AR6_WGII 2136 9 All of these reduce climate resilience very high 3 train +6175 AR6_WGII 2136 11 In spite of the lower climate velocities inside terrestrial hotspots, these areas are not projected to serve as effective climate refugia from the effects of global warming, especially for endemic species (unique to a hotspot) medium 1 train +6176 AR6_WGII 2136 12 The greater climate velocities inside marine hotspots exposes their species to greater climate-induced pressures inside than outside hotspots high 2 train +6177 AR6_WGII 2136 13 The differences between temperatures inside and outside of hotspots narrow with increasing warming medium 1 train +6178 AR6_WGII 2136 15 Of the 6116 projections for more than 2700 species assessed in biodiversity hotspots, ~44% were found to be at high extinction risk, and ~24% at very high extinction risk due to climate change medium 1 test +6179 AR6_WGII 2136 18 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.endemic species than other native species high 2 train +6180 AR6_WGII 2136 19 For these endemic species, considering all scenarios and time periods evaluated, ~100% on islands, ~84% on mountains, ~12% on continents (medium confidence) and ~54% in the ocean (notably the Mediterranean) low 0 train +6181 AR6_WGII 2136 20 With further warming, increasingly high risks of local and global extinctions are projected in biodiversity hotspots from climate-related stressors high 2 train +6182 AR6_WGII 2136 23 Adaptation options for biodiversity (e.g., expanding fully protected areas, restoration and sustainable use practices) are as applicable inside biodiversity hotspots as outside high 2 train +6183 AR6_WGII 2137 19 Thus, terrestrial biodiversity hotspots have been warming slightly less, and marine hotspots considerably more than non-hotspots medium 1 train +6184 AR6_WGII 2137 25 This suggests that, if these areas are subject to increased velocities, they will lose species that are not able to disperse fast enough to cope with the pace of climate change medium 1 train +6185 AR6_WGII 2137 26 Climate velocities are 47% (Myers), 29% (G200, terrestrial) and 10% (G200, freshwater) lower inside biodiversity hotspots than outside, respectively (Kocsis et al., 2021), but are 69% higher inside marine hotspots than outside medium 1 train +6186 AR6_WGII 2142 2 Marine species tend to follow climate velocities more closely than terrestrial species high 2 train +6187 AR6_WGII 2142 5 The subtropics are then source areas of species that shift to temperate latitudes and so forth, such that observed impacts in marine biodiversity hotspots are largely attributable to species range shifts high 2 train +6188 AR6_WGII 2142 6 Because marine climate velocities are significantly greater within than outside hotspots, marine hotspots are especially prone to species redistributions medium 1 train +6189 AR6_WGII 2142 8 Thus, as already observed in the oceans around the equator, the loss of species in low latitudes will continue with future climate warming high 2 train +6190 AR6_WGII 2143 7 Since the 1950s, marine species richness has shifted poleward in the Northern Hemisphere, increased in mid-latitudes and declined at the equator in concert with ocean warming medium 1 train +6191 AR6_WGII 2143 8 Climate- driven altitudinal shifts are common on land (high confidence) (Lenoir and Svenning, 2015; Steinbauer et al., 2018), and depth shifts in the ocean may occur but are little studied low 0 train +6192 AR6_WGII 2143 9 While climate-induced range expansions can be viewed as opportunities for increasing regional biodiversity, range contractions adversely affect biodiversity through regional extirpations high 2 train +6193 AR6_WGII 2143 15 Modelled temperatures are projected to continue to be the highest in the tropics, indicating where there are more thermally stressful conditions for more species high 2 train +6194 AR6_WGII 2143 16 By the end of this century, all terrestrial biodiversity hotspots in Central and South America, Africa, India and southern and eastern Asia (including the Indo–West Pacific islands) are projected to experience climates unprecedented in their species’ evolutionary history medium 1 train +6195 AR6_WGII 2143 17 Based on WGI Interactive Atlas data (Gutiérrez et al., 2021), global warming is projected to affect terrestrial hotspots less than non- hotspot areas: 80% less for Myers and 95–96% less for G200 terrestrial and freshwater hotspots at global warming of 1.5°C–3°C medium 1 train +6196 AR6_WGII 2143 18 In contrast, warming is projected to be 12–13% greater inside than outside marine hotspots medium 1 train +6197 AR6_WGII 2143 19 Precipitation is generally projected to increase more in terrestrial and freshwater biodiversity hotspots compared to outside them low 0 train +6198 AR6_WGII 2143 20 The exception is Myers hotspots, which are projected to have, on average, ~28% less precipitation at 1.5°C warming, but ~33% more at 2°C and ~65% more at 3°C low 0 train +6199 AR6_WGII 2143 22 CCP1.2.1.2.2 Projected impacts on biodiversity Biodiversity hotspots are expected to be especially vulnerable to climate change because their endemic species have smaller geographic ranges high 2 train +6200 AR6_WGII 2143 26 All measures of biodiversity were found to be negatively impacted by projected climate change, namely, species abundance, diversity, area, physiology and fisheries catch potential medium 1 train +6201 AR6_WGII 2143 27 However, introduced species’ responses were neutral to positive medium 1 train +6202 AR6_WGII 2143 29 Land plants, insects, birds, reptiles and mammals were all projected to be negatively affected (medium confidence), as well as fish, coral reef, benthic, planktonic and other marine species medium 1 train +6203 AR6_WGII 2143 30 Of the 6116 projections for more than 2,700 species assessed in biodiversity hotspots, ~44% were found to be at high extinction risk, and ~24% at very high extinction risk due to climate change (Manes et al., 2021) medium 1 train +6204 AR6_WGII 2144 0 CCP1 2133Biodiversity Hotspots Cross-Chapter Paper 1 and ocean medium 1 train +6205 AR6_WGII 2144 5 This and previous assessments indicate that, while climate change varies spatially and taxa may respond differently, a loss of biodiversity is projected across all terrestrial hotspots high 2 train +6206 AR6_WGII 2145 3 CCP1.2.1.3 Compounding and Cascading Effects All biodiversity hotspots are already impacted, to differing degrees, by human activities high 2 train +6207 AR6_WGII 2145 5 Thus, climate change impacts on biodiversity hotspots are compounded by other anthropogenic impacts, increasing the vulnerability and reducing the resilience of biodiversity to climate change very high 3 train +6208 AR6_WGII 2146 8 Moreover, the present rates of species loss due to human activities are 130 times greater than those projected under future climate change medium 1 train +6209 AR6_WGII 2146 9 Marine systems are also vulnerable to cumulative human impacts, which can be direct (e.g., pollution, overfishing) and indirect (altered food webs) very high 3 train +6210 AR6_WGII 2147 1 Although there is a strong overlap of non-climatic and climatic impacts in marine ecosystems (Blowes et al., 2019; Bowler et al., 2020), the effects suggest that climate change impacts are most severe in tropical and northern high-latitude seas high 2 train +6211 AR6_WGII 2147 2 Temperature-driven range shifts and range expansions are projected to also lead to cascading effects on marine biodiversity through ecological interactions high 2 test +6212 AR6_WGII 2147 14 In Central and South America, observed impacts within Mesoamerica (H15, 16) and the Tropical Andes hotspots (H26, 27, 28, 32, 33) comprise upward altitudinal range shifts of birds, frogs, beetles and butterflies (Narins and Meenderink, 2014; Molina-Martínez et al., 2016; Moret et al., 2016; Freeman et al., 2018) medium 1 train +6213 AR6_WGII 2147 15 A shift of the Guianan-Amazon mangroves (H37) to higher grounds inland was attributed to the effects of observed sea level rise low 0 train +6214 AR6_WGII 2147 16 Range shifts in birds have been observed at higher elevations medium 1 train +6215 AR6_WGII 2147 20 Warming and drying trends have historically been shown to reduce the range of the Ethiopian wolf (Canis simensis), and they interact with land use pressures in the Ethiopian hotspot (H68) (Sintayehu, 2018) and plant species richness in the Cape Fynbos (H65) of southern Africa to reduce post-wildfire recruitment low 0 train +6216 AR6_WGII 2147 21 Observed impacts in Asia were mostly restricted to the Himalaya (H95, 98, 99), Sundaland (H109, 110, 111, 112, 117, 118) and Indo- Burma (H105, 106, 107, 114, 115) hotspots, showing negative impacts through increased invasion by exotic plants, decreased suitable area for endemic species and significant changes in phenology medium 1 train +6217 AR6_WGII 2147 24 In Australia, climate change has been implicated in: drought-induced canopy dieback across a range of forest and woodland types due to decades of declining rainfall in the southwestern hotspot (H133); fires in the palaeo-endemic pencil pine forests (Tasmania H142); declines in vertebrates in the Australian Wet Tropics World Heritage Area, which overlaps with the eastern part of the northern Australia hotspot (H131), related to warming and increased length of the dry season; and declines in grass and increases in shrubs in the Bogong High Plains high 2 train +6218 AR6_WGII 2147 25 The Australian Alps have seen increased species diversity following retreat of the snow line (Slatyer, 2010), replacement of long-lived trees by short-lived shrubs following multiple wildfires (Zylstra, 2018), and changing ecological interactions due to climate-related snow loss, drought and fires high 2 train +6219 AR6_WGII 2148 1 One positive observation was the high resilience to recovery of intact forest ecosystems to tropical cyclones within Caribbean (H20) and Pacific islands medium 1 train +6220 AR6_WGII 2148 9 About 85% of projections for assessed species showed a negative impact of climate change high 2 train +6221 AR6_WGII 2148 10 Projected impacts include contraction or loss of species’ geographic range, loss of diversity and high species turnover high 2 train +6222 AR6_WGII 2148 15 Climate change may also benefit invasive plant species in terms of range expansion (Wang et al., 2017) and physiology (de Faria et al., 2018) in the region.In European biodiversity hotspots, about 75% of projections for assessed species showed a negative impact of climate change, with ~30% at very high risk of extinction medium 1 train +6223 AR6_WGII 2148 17 Increased wildfire size and frequency is projected to have a strong effect on the Mediterranean basin (H216) ecosystems medium 1 train +6224 AR6_WGII 2148 18 Range reductions have been projected for endemic plants (Pérez-García et al., 2013; Casazza et al., 2014), reptiles (Ahmadi et al., 2019), birds (Abolafya et al., 2013) and insects (Sánchez-Guillén et al., 2013) medium 1 train +6225 AR6_WGII 2148 19 In African biodiversity hotspots, about 80% of projections for assessed species showed a negative impact of climate change, with ~10% at very high risk of extinction, especially of endemic species including birds, plants, bees across several taxa and hotspots if warming exceeds 2°C high 2 train +6226 AR6_WGII 2148 21 About ~70% of projections for assessed species showed a negative impact of climate change, with ~30% at very high risk of extinction medium 1 train +6227 AR6_WGII 2148 22 Impacts include species’ range changes, habitat loss for endemic plants, expansion of invasive species, decreased connectivity and overall species richness decline high 2 train +6228 AR6_WGII 2148 24 The few positive impacts of climate change were projected as increases in suitable habitat and distribution range for a few endangered plants and mammals medium 1 train +6229 AR6_WGII 2148 26 All projections for assessed species in Australia and New Zealand terrestrial biodiversity hotspots showed a negative impact of climate change, with half at very high risk of extinction low 0 train +6230 AR6_WGII 2149 0 While forest growth is projected to potentially increase due to carbon dioxide fertilization, this may be compromised by drought low 0 test +6231 AR6_WGII 2149 1 Seed production in native New Zealand beech forests is projected to increase due to climate warming, fuelling the abundance of invasive rats and stoats, which then predate native species and lead to loss of endemic fauna and flora medium 1 train +6232 AR6_WGII 2149 3 About 80% of projections for assessed terrestrial species within insular biodiversity hotspots showed a negative impact of climate change, with ~50% at very high risk of extinction, including 100% of endemic species medium 1 train +6233 AR6_WGII 2149 4 In addition to habitat loss and species range reductions, changes in precipitation are projected to be a major driver impacting tropical and subtropical island species medium 1 train +6234 AR6_WGII 2149 5 Compared to continents, island species are projected to undergo greater impacts from changing climate, especially birds and amphibians high 2 train +6235 AR6_WGII 2149 6 Of all biodiversity hotspots, island species face the highest proportion of extirpation risk at high elevations due to decreasing habitat area (e.g., Brown et al., 2015) and at low elevations from sea level rise, habitat loss and introduced species medium 1 train +6236 AR6_WGII 2154 0 CCP1 2143Biodiversity Hotspots Cross-Chapter Paper 1 CCP1.2.3.1 Observed Impacts An analysis of trends in 190 river basins in Australia found that stream- flows have been declining, including in the Central Australian (H194) and Kimberley (H191) hotspots, due to greater terrestrial plant uptake of water in response to climate-related increases in carbon dioxide low 0 train +6237 AR6_WGII 2154 3 CCP1.2.3.2 Projected Impacts Cold-water species are projected to lose habitat in Canada and this may apply in the Alaskan river (H143) and Russian Far East Lake Inle (H181) hotspots medium 1 train +6238 AR6_WGII 2154 5 In South America, in the Brazilian Amazon hotspot (H153, 154, 157), half the assessed fish species were considered sensitive to increased temperatures and reduced oxygen due to climate change low 0 train +6239 AR6_WGII 2154 8 In Europe, including the Mediterranean freshwater hotspots, climate change is projected to result in reduced river flow, low oxygen in summer, salinity incursions, further eutrophication and spread of invasive species, compromising the survival of native biodiversity medium 1 train +6240 AR6_WGII 2154 9 The longer growth season in the boreal and Arctic latitudes is projected to aid the invasion of exotic species, and increase lake stratification resulting in lower oxygen below the hypolimnion medium 1 train +6241 AR6_WGII 2154 11 An analysis of 1648 species of freshwater fish, amphibians, turtles, plants, molluscs, crayfish and dragonflies, projected ~6% of common and ~77% of rare species to lose 90% of their geographic range low 0 train +6242 AR6_WGII 2154 17 Thus, the areas where freshwater biodiversity is most threatened by climate change in Europe are in two of the three hotspots (high confidence).The African Rift Valley Lakes (H171), including Lakes Tanganyika and Turkana, are suffering from climate change influenced drought, po- tentially impacting freshwater biodiversity medium 1 train +6243 AR6_WGII 2154 18 Africa and Madagascar (H172) are projected to see a climate-driven 10% reduction in freshwater flow that is projected to threaten the survival of ~9% of freshwater-dependent fish and birds low 0 train +6244 AR6_WGII 2154 19 Climate change is projected to increase the extinction vulnerability of most freshwater fish in the western South Africa Cape hotspot (H170) low 0 train +6245 AR6_WGII 2154 20 In Asia, although climate change impacts on the Yangtze (H183) and Mekong river (H186) biodiversity hotspots have not been reported, they are subject to the range of human impacts of over-exploitation, pollution, water abstraction, altered flow regimes, habitat loss and spread of invasive species, which makes them more vulnerable to climate effects medium 1 train +6246 AR6_WGII 2154 21 The release of water from shrinking glaciers in Asia to some extent protects downstream freshwaters against drought, but half of these glaciers are projected to disappear by 2100 medium 1 train +6247 AR6_WGII 2154 22 In Australia, the Murray-Darling river basin occupies much of the Eastern Rivers hotspot (H195) and climate-related drought exacerbated by water abstraction is projected to drive declines in freshwater birds, fish and invertebrates high 2 train +6248 AR6_WGII 2154 23 A national scale analysis projected climate change to cause freshwater species range shifts, but no losses of species in this hotspot low 0 test +6249 AR6_WGII 2154 28 Marine heatwaves have increased over the past century, causing mass mortalities in the hotspots of the Mediterranean (H216), Great Barrier Reef (H236), western and southern Australia (H227, 228), northwest Atlantic (H207) and northeast Pacific (H197) high 2 test +6250 AR6_WGII 2154 29 The shift of thousands of species from equatorial latitudes since the 1950s has been attributed to climate warming medium 1 train +6251 AR6_WGII 2154 30 Climate change-related hazards, particularly marine heat events, have caused widespread coral bleaching and mass mortalities as the time between consecutive bleaching events decreases high 2 train +6252 AR6_WGII 2154 31 Coral reefs in some Indian Ocean hotspots (H230, 234) already exhibit net loss of coral reefs low 0 test +6253 AR6_WGII 2156 0 CCP1 2145Biodiversity Hotspots Cross-Chapter Paper 1 visible symptom of heat stress, warming has also induced restructuring of associated fish and invertebrate communities in the Great Barrier Reef (H236) medium 1 train +6254 AR6_WGII 2156 1 Although the number of coral species that are both exposed and vulnerable to climate hazards is greatest in the central Indo-Pacific, the proportion of corals at risk is greater in the lower diversity Caribbean hotspots (H209) medium 1 train +6255 AR6_WGII 2156 2 Some reef corals are able to acclimate to heatwaves (low confidence) (DeCarlo et al., 2019), and some have expanded their latitudinal ranges polewards high 2 train +6256 AR6_WGII 2156 4 The Mediterranean Sea hotspot (H216) is negatively affected by climate change high 2 train +6257 AR6_WGII 2156 5 Species entering via the Suez Canal from the Red Sea (H220) are facilitated by warming and lead to profound community changes high 2 train +6258 AR6_WGII 2156 7 Kelp forests are in decline in mid-latitudes due to warming and associated increased herbivory medium 1 train +6259 AR6_WGII 2156 10 Australia’s Great Barrier Reef (H236), kelp forests, seagrass meadows and mangroves (due to drought), have suffered mortalities due to climate change medium 1 train +6260 AR6_WGII 2156 12 However, while climate change is having measurable effects on kelp, the dominant effects on kelp projected to 2025 are fishing, through its effects on herbivores and predators medium 1 train +6261 AR6_WGII 2156 22 The distribution of krill has already contracted with ocean warming in the Southern Ocean medium 1 train +6262 AR6_WGII 2156 24 Paleo evidence supports projections of tropical biodiversity loss under high global warming high 2 train +6263 AR6_WGII 2156 25 Warm-water coral reefs are expected to decline with 1.5°C warming (very high confidence) (King et al., 2017; Bindoff et al., 2019) leading to systems with reduced biodiversity and structural complexity high 2 train +6264 AR6_WGII 2156 27 While some corals are expected to survive in deep ‘mesophotic’ reefs (Laverick and Rogers, 2019), the shallow coral reefs of today will not last the century if climate warming continues without mitigation high 2 train +6265 AR6_WGII 2159 3 Around Antarctica (H213), almost half of all species are endemic (Costello et al., 2010), and warming during this century is projected to cause a reduction in suitable thermal environment for 79% of its species (RCP8.5) low 0 train +6266 AR6_WGII 2159 5 Species richness in the northern polar hotspots is expected to increase substantially high 2 train +6267 AR6_WGII 2159 9 Where land barriers and other geographical limits to range shifts occur, limited dispersal and habitat fragmentation may also limit the capacity of some species to track climate velocities, such as in the Baltic Sea (H215) (Jonsson et al., 2018), Mediterranean Sea (H216) (Burrows et al., 2014; Arafeh- Dalmau et al., 2021) and Antarctica (H213) medium 1 train +6268 AR6_WGII 2160 3 Many of these hotspots are now faced with widespread fragmentation and habitat degradation high 2 train +6269 AR6_WGII 2160 7 Although mitigation can sharply reduce extinction risk associated with climate change high 2 train +6270 AR6_WGII 2160 8 Thus, in addition to mitigation, the literature consistently calls for reducing current non-climate impacts (e.g., habitat conversion, over-exploitation, hunting, fishing, wildfire, pollution, human-introduced invasive species) in order to increase biodiversity resilience to climate change very high 3 train +6271 AR6_WGII 2160 9 The main strategies to increase resilience rely on the combination of well-planned protected areas, restoration of degraded areas and the sustainable use of biodiversity high 2 train +6272 AR6_WGII 2160 10 On land, creating corridors for species is key for facilitating species movements high 2 train +6273 AR6_WGII 2160 14 Healthier marine ecosystems are more resilient to additional stressors, such as storms and climate change high 2 train +6274 AR6_WGII 2160 15 Extinction risk is lower when populations are larger and more genetically diverse, individuals are larger and older, and seabed habitats (e.g., coral, kelp, seagrass) are flourishing, as occurs in marine reserves high 2 train +6275 AR6_WGII 2160 17 Thus, a network of reserves representative of global biodiversity, helps attenuate the effects of climate change medium 1 train +6276 AR6_WGII 2160 18 However, the impacts of marine heatwaves on corals across marine reserves illustrates that enhanced resilience is not enough to protect against extreme and future climate change conditions high 2 train +6277 AR6_WGII 2160 22 If coastal management permits the expansion of mangroves inland with rising sea level, this will increase carbon sequestration because mangroves capture and preserve more carbon in their sediments than other terrestrial and marine forests and biomes high 2 train +6278 AR6_WGII 2160 24 Thus, the protection of existing natural habitats coupled with the restoration of the surrounding non-protected habitat can increase the effectiveness of adaptation strategies in terrestrial and freshwater hotspots very high 3 train +6279 AR6_WGII 2161 7 Islands have disproportionately higher rates of endemism and threat when compared to continents, with 80% of historical extinctions (since 1500 CE) having occurred on islands high 2 train +6280 AR6_WGII 2161 8 Current climate change projections suggest that insular species are particularly sensitive and, even at mild warming levels, substantial losses are expected high 2 train +6281 AR6_WGII 2161 9 Given islands’ characteristic high endemicity, current high threat levels and the fact that islands host almost half of all species currently considered to be at risk of extinction, especially at higher warming levels high 2 train +6282 AR6_WGII 2161 13 Unlike continental environments, insular species often have limited opportunities for autonomous adaptation from not having enough geographic space to shift their ranges to track suitable climatic conditions high 2 train +6283 AR6_WGII 2161 16 Intact island forests, for example, have shown rapid recovery rates after tropical cyclones, despite high levels of initial damage, especially in the Caribbean medium 1 train +6284 AR6_WGII 2161 19 However, this climate resilience will not be sustained under climate change, especially when coupled with habitat degradation high 2 train +6285 AR6_WGII 2161 24 Widespread unavailability of such data constrains accurate simulations of climatic variation within the small-scale mountainous and coastal regions of islands, associated with climate refugia and high habitat heterogeneity high 2 train +6286 AR6_WGII 2161 25 This is a key element contributing to the continued delay in development of robust adaptation strategies towards not only biodiversity conservation but other important cross-sectoral issues medium 1 train +6287 AR6_WGII 2162 0 CCP1 2151Biodiversity Hotspots Cross-Chapter Paper 1 Due to islands’ limited size and isolation, conventional conservation measures focused on expanding protected areas, dispersal corridors and buffer zones are of limited effectiveness on islands high 2 train +6288 AR6_WGII 2162 5 These lend to private–public partnerships, increasing the potential of solutions reaching beyond protected areas boundaries and affecting socio-political change high 2 train +6289 AR6_WGII 2176 1 Much of the world’s population, economic activities and critical infrastructure are concentrated near the sea (high confidence), with nearly 11% of the global population, or 896 million people, already living on low-lying coasts directly exposed to interacting climatic and non-climatic coastal hazards very high 3 train +6290 AR6_WGII 2176 2 Low-lying cities and settlements (C&S) by the sea are experiencing adverse climate impacts that are superimposed on extensive and accelerating anthropogenic coastal change very high 3 train +6291 AR6_WGII 2176 3 Depending on coastal C&S characteristics, continuing existing patterns of coastal development will worsen exposure and vulnerability high 2 train +6292 AR6_WGII 2176 4 With accelerating sea level rise (SLR) and worsening climate-driven risks in a warming world, prospects for achieving the Sustainable Development Goals (SDGs) and charting climate resilient development (CRD) pathways are dismal high 2 train +6293 AR6_WGII 2176 5 However, coastal C&S are also the source of SDG and CRD solutions, because they are centres of innovation with long histories of place-based livelihoods, many of which are globally connected through maritime trade and exchange medium 1 train +6294 AR6_WGII 2176 6 Regardless of climate and socioeconomic scenarios, many C&S face severe disruption to coastal ecosystems and livelihoods by 2050—extending to all C&S by 2100 and beyond—caused by compound and cascading risks, including submergence of some low-lying island states very high 3 train +6295 AR6_WGII 2176 9 These risks are acute for C&S on subsiding and/or low-lying small islands, the Arctic, and open, estuarine and deltaic coasts high 2 train +6296 AR6_WGII 2176 10 By 2050, more than a billion people located in low-lying C&S will be at risk from coast-specific climate hazards, influenced by coastal geomorphology, geographical location and adaptation action high 2 train +6297 AR6_WGII 2176 11 Between USD 7 and 14 trillion of coastal infrastructure assets will be exposed by 2100, depending on warming levels and socioeconomic development trajectories medium 1 train +6298 AR6_WGII 2176 16 The coastal flood risk will rapidly increase during coming decades, possibly by 2–3 orders of magnitude by 2100 in the absence of effective adaptation and mitigation, with severe impacts on coast-dependent livelihoods and socioecological systems high 2 train +6299 AR6_WGII 2176 19 Severely accelerated SLR resulting from rapid continental ice mass loss would bring impacts forward by decades, and adaptation would need to occur much faster and on a much greater scale than ever performed in the past medium 1 train +6300 AR6_WGII 2176 21 An adaptation-pathways approach sets out near-term ‘low-regret’ actions that align with societal goals, facilitates implementation of a locally appropriate sequence of interventions in the face of uncertain climate and development futures, and enables necessary transformation high 2 train +6301 AR6_WGII 2176 22 A mix of infrastructural, nature-based, institutional and sociocultural interventions are needed to reduce the multifaceted risk facing C&S, including vulnerability-reducing measures, avoidance (i.e., disincentivising developments in high-risk areas), hard and soft protection, accommodation, advance (i.e., building up and out to sea) and retreat (i.e., landward movement of people and development) very high 3 train +6302 AR6_WGII 2176 23 Depending on the C&S archetype, technical limits for hard protection may be reached beyond 2100 under high-emission scenarios, with socioeconomic and governance barriers reached before then medium 1 train +6303 AR6_WGII 2176 24 However, hard protection can set up lock-in of assets and people to risks and, in some cases, may reach limits—due to technical and financial constraints—by 2100 or sooner depending on the scenario, local SLR effects and community tolerance thresholds medium 1 train +6304 AR6_WGII 2176 25 Where sufficient space and adequate habitats are available, nature-based solutions can help to reduce coastal hazard risks and provide other benefits, but biophysical limits may be reached before end-century medium 1 train +6305 AR6_WGII 2176 28 As SLR is relentless on human timescales, the solution space will shrink without adoption of an adaptation-pathways planning approach high 2 train +6306 AR6_WGII 2177 0 CCP2 2166Cross-Chapter Paper 2 Cities and Settlements by the Sea if this is underpinned by sustained and ambitious mitigation to slow greenhouse gas emission rates high 2 train +6307 AR6_WGII 2177 1 Individual and collective choices founded on public-centred values and norms, as well as pro-social behaviour, help to foster climate-resilient coastal development in C&S high 2 train +6308 AR6_WGII 2177 2 The effectiveness of different approaches (e.g., awareness and education, market-based and legal strategies) is mediated by how well they address contextual and psychosocial factors influencing adaptation choices in coastal C&S medium 1 train +6309 AR6_WGII 2177 4 Locally appropriate institutional capabilities, including regulatory provisions and finances dedicated to maintaining healthy coastal socioecological systems, build adaptive capacity in C&S by the sea high 2 train +6310 AR6_WGII 2177 5 Implementing integrated multi-level coastal zone governance, pre- emptive planning, enabling behavioural change and alignment of financial resources with a wide set of values will provide C&S with greater flexibility to open up the solution space to adapt to climate change high 2 train +6311 AR6_WGII 2177 6 Insufficient financial resources are a key constraint for coastal adaptation, particularly in the Global South high 2 train +6312 AR6_WGII 2177 7 Engaging the private sector in coastal adaptation action with a range of financial tools is crucial to address the coastal adaptation funding gap high 2 test +6313 AR6_WGII 2177 8 Considering the full range of economic and non-economic values will improve adaptation effectiveness and equity across C&S archetypes high 2 train +6314 AR6_WGII 2177 11 Realising global aspirations for CRD depends on the extent to which coastal C&S institutionalise key enabling conditions and chart place-based adaptation pathways to close the coastal adaptation gap, and on the extent to which they take urgent action to mitigate greenhouse gas emissions medium 1 train +6315 AR6_WGII 2177 12 Extensive adaptation planning has been undertaken since the IPCC Fifth Assessment Report (AR5), but there has not been widespread effective implementation, thus giving rise to a ‘coastal adaptation gap’ high 2 train +6316 AR6_WGII 2177 13 To date, most interventions have been reactive and often rely on protective works alone high 2 train +6317 AR6_WGII 2177 14 The effectiveness of alternative interventions differs among C&S archetypes, while their feasibility is influenced by geomorphology and socioeconomic conditions as well as cultural, political and institutional considerations very high 3 train +6318 AR6_WGII 2177 15 Mismatches between adaptation needs and patterns of physical development are commonplace in many coastal C&S, with particularly adverse impacts on poor and marginalised communities in the Global North and Global South high 2 train +6319 AR6_WGII 2177 16 Overcoming this gap is key to transitioning towards CRD medium 1 train +6320 AR6_WGII 2177 17 Under higher warming levels and higher SLR, increasingly dichotomous coastal futures will become more entrenched medium 1 train +6321 AR6_WGII 2177 22 Progress towards these ends depends on the extent to which C&S mobilise urgent and transformational changes to institutionalise enabling conditions, close the coastal adaptation gap by addressing the drivers and root causes of exposure and vulnerability to climate-compounded coastal hazard risks, and drastically reduce greenhouse gas emissions medium 1 train +6322 AR6_WGII 2180 17 The ocean and climate impact drivers influencing these risks are assessed in WGI (Ranasinghe et al., 2021), which include extreme heat, pluvial floods from increasing rainfall intensity, coastal erosion and coastal flood driven by increasing SLR, and tropical cyclone storm surges high 2 train +6323 AR6_WGII 2180 18 Further, Arctic coastal settlements are particularly exposed to climate change due to sea ice retreat as well as from permafrost melt high 2 train +6324 AR6_WGII 2180 20 Across these studies, by 2100, 158–510 million people and USD 7,919–12,739 billion assets under RCP4.5, and 176–880 million people and USD 8,813– 14,178 billion assets under RCP8.5 will be within the 1-in-100-year floodplain very high 3 train +6325 AR6_WGII 2182 4 There is emerging evidence low 0 train +6326 AR6_WGII 2182 6 However, risks facing coastal C&S are high across the globe, especially under higher SLR projections high 2 train +6327 AR6_WGII 2183 4 To date, interventions are typically implemented reactively in response to extreme events high 2 train +6328 AR6_WGII 2184 26 Advance has occurred in all archetypes high 2 train +6329 AR6_WGII 2188 22 Moreover, drawing from places as distinct as small communities in Fiji (Neef et al., 2018) and Belize (Karlsson and Hovelsrud, 2015), and megacities like New York City and Shanghai (Oppenheimer et al., 2019), BCR provides only a limited view and consideration of feasibility, effectiveness, efficiency, equity, culture, politics and power, and attachment to place has a greater chance of fostering CRD high 2 train +6330 AR6_WGII 2192 13 In sum, prospects for addressing climate risk in archetypal coastal C&S around the world depend on the extent to which societal choices— and associated governance processes and practices—address the drivers and root causes of exposure and social vulnerability very high 3 train +6331 AR6_WGII 2192 23 This cross-chapter paper shows that a range of adaptation solutions, hard and soft protection, nature-based measures, accommodate, advance, retreat and behavioural change will need to be implemented as an integrated and sequenced portfolio of responses if coastal C&S are to contain the adverse risks of climate change high 2 train +6332 AR6_WGII 2193 2 Coastal C&S are on the frontline of observed climate change impacts and future risk high 2 train +6333 AR6_WGII 2193 7 Given the risks assessed in coastal C&S, the scale of climate impacts globally will depend to a large extent on whether coastal settlements develop and implement pre- emptive and flexible adaptation pathways, and whether a significant and timely reduction in greenhouse gas emissions is achieved in C&S and globally high 2 train +6334 AR6_WGII 2208 4 These changes have had varying and location-specific impacts on biodiversity, and have altered ecosystem carbon balance, water availability and the provision of ecosystem services high 2 train +6335 AR6_WGII 2208 5 There is no evidence, however, of a global trend in dryland expansion based on analyses of vegetation patterns, precipitation and soil moisture, with overall, more greening than drying in drylands since the 1980s medium 1 train +6336 AR6_WGII 2208 22 The impacts of climate change have affected the ecosystem services that humans can harness from drylands, with largely negative implications for livelihoods, human health and well- being, particularly in deserts and semiarid areas with lower adaptive capacities high 2 train +6337 AR6_WGII 2208 23 Ecosystem degradation (Section 16.5.2.3.2) and desertification threaten the abilities of both natural and human systems to adapt to climate change high 2 train +6338 AR6_WGII 2208 25 These groups also often have lower capacities to adapt, particularly given structural limitations of some drylands where healthcare, sanitation, infrastructure and efficient markets are lacking, reinforcing existing inequalities high 2 train +6339 AR6_WGII 2208 26 In rural drylands in tropical and Mediterranean areas, human populations are steadily expanding with mixed implications for ecosystem services under climate change, while rapid urbanisation in new and existing dryland megacities puts additional pressure on water ecosystem services high 2 train +6340 AR6_WGII 2208 30 Risks and adaptation under warming pathways Some drylands will expand by 2100, while others will shrink high 2 train +6341 AR6_WGII 2208 32 Projections are nevertheless uncertain and not well supported by observed trends, while different methodological approaches and indices exhibit different strengths and weaknesses medium 1 train +6342 AR6_WGII 2209 3 Nevertheless, the utility of the AI in delineating dryland biomes is limited under an increasing CO 2 environment medium 1 train +6343 AR6_WGII 2209 5 The characteristics and speed of human responses and adaptations also affect future risks and impacts high 2 train +6344 AR6_WGII 2209 6 Increased temperature and rainfall variability will significantly change the interannual variability in the global carbon cycle, which is strongly influenced by the world’s drylands and the ways they are managed medium 1 train +6345 AR6_WGII 2209 7 Increased variability of precipitation would generally contribute to increased vulnerability for people in drylands, intensifying the challenges that people living in deserts and semiarid areas will face for their sustainable development medium 1 train +6346 AR6_WGII 2209 11 Key enablers include supportive policies, institutions and governance approaches that strengthen the adaptive capacities of dryland farmers, pastoralists and other dryland resource users high 2 train +6347 AR6_WGII 2209 15 Land-based adaptations can help manage dryland changes, including sand and dust storms and desertification high 2 train +6348 AR6_WGII 2211 6 However, there is no evidence of a global trend in dryland expansion based on vegetation patterns, precipitation and soil moisture, based on the satellite record from the 1980s to the present medium 1 train +6349 AR6_WGII 2212 5 These systems are highly sensitive to annual precipitation and temperature variations high 2 train +6350 AR6_WGII 2212 20 Tree regeneration by farmers has also increased woody cover, particularly next to villages high 2 train +6351 AR6_WGII 2212 23 CCP3.2.1.5 Tree Death and Woody Cover Decline Field measurements have also detected tree mortality and loss of mesic tree species at some Sahel sites during drought periods (Gonzalez et al., 2012; Kusserow, 2017; Brandt et al., 2018; Ibrahim et al., 2018; Trichon et al., 2018; Zwarts et al., 2018; Bernardino et al., 2020; Zida et al., 2020) and a reduction of mesic species in favour of drought-tolerant species high 2 train +6352 AR6_WGII 2217 9 The spread of invasive Bromus tectorum may be enhanced by altered precipitation and freeze–thaw cycles low 0 train +6353 AR6_WGII 2217 10 Arid grassland has expanded (between 10–100 km) into the eastern Karoo, South Africa high 2 train +6354 AR6_WGII 2218 3 CCP3.2.1.7 Sand and Dust Storms Soil dust emissions are highly sensitive to changing climate conditions but also to changing land use and management practices high 2 train +6355 AR6_WGII 2219 13 CCP3.2.2 Observed Impacts of Climate Change on Human Systems in Desert and Semiarid Areas Climate change and desertification, alongside other drivers of degra- dation, reduce dryland ecosystem services, leading to losses of bio- diversity, water, food and impacts on human health (Section CCP4.2.3) and well-being high 2 train +6356 AR6_WGII 2219 21 SDS negatively impact human health through various pathways, causing respiratory, cardiovascular diseases and facilitating infections high 2 train +6357 AR6_WGII 2222 20 AI projections indicate potentially severe aridification in the Amazon, Australia, Chile, the Mediterranean region, northern, southern and western Africa, southwestern USA and South America medium 1 train +6358 AR6_WGII 2225 5 Supportive policies, institutions and good governance approaches can strengthen the adaptive capacities of dryland farmers, pastoralists and other resource users high 2 train +6359 AR6_WGII 2228 8 Building capacity by improving the knowledge base and access to information, as well as to financial and other resources, encourages vulnerable economic sectors and people to adopt more self-reliant measures that promote more integrated and sustainable use of natural resources high 2 train +6360 AR6_WGII 2246 8 Trends in precipitation are variable across the basin low 0 train +6361 AR6_WGII 2246 9 Droughts have become more frequent and intense, especially in the north Mediterranean high 2 train +6362 AR6_WGII 2246 11 Sea level has risen by 1.4±0.2 mm yr-1 during the 20th century (2.8±0.1 mm yr-1 over 1993–2018) high 2 train +6363 AR6_WGII 2246 12 Ocean acidity is increasing medium 1 train +6364 AR6_WGII 2246 23 Air and sea temperature and their extremes (notably heat waves) are likely2 to continue to increase more than the global average high 2 train +6365 AR6_WGII 2246 24 The projected annual mean warming on land at the end of the century is in the range of 0.9–5.6°C compared to the last two decades of the 20th century, depending on the emission scenario high 2 train +6366 AR6_WGII 2246 25 Precipitation will likely decrease in most areas by 4–22%, depending on the emission scenario medium 1 train +6367 AR6_WGII 2246 26 Rainfall extremes will likely increase in the northern part of the region high 2 train +6368 AR6_WGII 2246 27 Droughts will become more prevalent in many areas high 2 train +6369 AR6_WGII 2246 29 Higher values cannot be excluded (low confidence) and the process is irreversible at the scale of centuries to millennia high 2 train +6370 AR6_WGII 2246 31 The number of people exposed to sea level rise is projected to increase up to 2050, especially in the southern and eastern Mediterranean region, and may reach up to 130% compared to present in 2100 medium 1 train +6371 AR6_WGII 2246 32 Coastal settlements, World Heritage sites and ecosystems are at longer-term risk from sustained sea level rise over at least the coming three centuries high 2 train +6372 AR6_WGII 2246 35 The low-lying areas are the most vulnerable areas for coastal climate-related risks (e.g., sea level rise, floods, erosion) and other consequent risks (e.g., saltwater intrusion and agriculture damage) high 2 train +6373 AR6_WGII 2246 36 Climate change threatens water availability, reducing river low flows and annual runoff by 5–70%, reducing hydropower capacity high 2 train +6374 AR6_WGII 2246 37 Yields of rain- fed crops may decrease by 64% in some locations high 2 train +6375 AR6_WGII 2246 38 Ocean warming and acidification will impact marine ecosystems, with uncertain consequences on fisheries low 0 train +6376 AR6_WGII 2246 39 Desertification will affect additional areas, notably in the south and southeast medium 1 train +6377 AR6_WGII 2246 41 Beyond 3°C, 13–30% of the Natura 2000 protected area and 15–23% of Natura 2000 sites could be lost due to climate-driven habitat change medium 1 train +6378 AR6_WGII 2247 0 CCP4 2236Cross-Chapter Paper 4 Mediterranean Region The adaptive capacity of ecosystems and human systems is expected to encounter hard limits due to the interacting, cumulative and cascading effects of droughts, heat waves, sea level rise, ocean warming and acidification high 2 train +6379 AR6_WGII 2247 1 Coastal protection can reduce risks from sea level rise in some regions, but the costs of such interventions and their consequences for coastal ecosystems are high medium 1 train +6380 AR6_WGII 2247 4 To equitably enhance regional adaptive capacity and sustainable development, while safeguarding the rights of the most vulnerable people, regional cooperation can be strengthened with a focus on the link between adaptation, costs and financial limitation, and climate justice high 2 train +6381 AR6_WGII 2247 5 Cooperative policies across various sectors, involving all user groups and considering all regional and sectorial differences may enhance sustainable resource use in the region high 2 train +6382 AR6_WGII 2248 8 With the changing climate, marine ecosystems have already undergone changes in structure, including the spread of tropical species from the Atlantic Ocean and the Red Sea high 2 train +6383 AR6_WGII 2248 14 Wetlands and mountain summits are hotspots for biodiversity loss and extinctions medium 1 train +6384 AR6_WGII 2248 24 Since the 1980s, Mediterranean atmospheric warming has exceeded global average rates high 2 train +6385 AR6_WGII 2248 27 Temperature extremes and heat waves have increased in intensity, number, and length during recent decades, particularly in summer, and are projected to continue increasing high 2 train +6386 AR6_WGII 2248 28 Sea surface temperatures have increased in recent decades high 2 train +6387 AR6_WGII 2249 0 CCP4 2238Cross-Chapter Paper 4 Mediterranean Region 21st century, ocean warming in the range 0.8°C–3.8°C is projected near the surface high 2 train +6388 AR6_WGII 2249 1 The duration and intensity of marine heat waves have increased high 2 train +6389 AR6_WGII 2249 3 Salinity is projected to increase, with anomalies from +0.48 to +0.89 psu by the end of the century medium 1 train +6390 AR6_WGII 2249 4 Observed trends in annual precipitation are significant only in some areas and some periods, and they are stationary over the long term throughout the region medium 1 test +6391 AR6_WGII 2249 6 Precipitation extremes have increased in some northern areas (medium confidence), and are projected to increase in the north (high confidence for global warming levels above 2°C), potentially accompanied by an increase in of flash floods (Llasat et al., 2016), with no change in the south low 0 train +6392 AR6_WGII 2249 9 Widespread increase of evaporative demand and some decrease of precipitation explain the drying of the Mediterranean region during recent decades high 2 test +6393 AR6_WGII 2249 10 Droughts are projected to become more severe, more frequent and longer under moderate emission scenarios, and strongly enhanced under severe emission scenarios high 2 train +6394 AR6_WGII 2250 5 Mediterranean waters have acidified since the pre-industrial period, more rapidly than the global ocean, due to faster ventilation times high 2 train +6395 AR6_WGII 2251 0 CCP4 2240Cross-Chapter Paper 4 Mediterranean Region Mediterranean mean sea level has risen by 1.4±0.2 mm yr−1 during the 20th century (Wöppelmann and Marcos, 2012) and accelerated to 2.4±0.5 mm yr−1 for 1993 to 2012 (Bonaduce et al., 2016) and 3.4 mm yr−1 for 1990 to 2009 in the northwest medium 1 train +6396 AR6_WGII 2251 2 For 2150, sea level is likely to reach 0.52 m [0.32–0.81] for SSP1- 1.9, to 1.22 [0.91–1.78] for SSP5-8.5 relative to 1996–2014 medium 1 train +6397 AR6_WGII 2251 6 CCP4.1.4 Detection and Attribution of Climate Change Impacts New evidence published since Working Group II Assessment Report 5 (WGII AR5) confirms that climate change is increasingly affecting many systems and sectors in the Mediterranean region high 2 train +6398 AR6_WGII 2251 9 Despite increasing wildfire hazard, forest fires are generally decreasing in the European part of the basin, due to more efficient risk management medium 1 train +6399 AR6_WGII 2251 14 While land use and fisheries are still major non-climatic drivers of changing hazards and biodiversity losses (Aguilera et al., 2015; Turco et al., 2016; IPBES, 2018a; 2018b; Tramblay et al., 2019; Vicente-Serrano et al., 2019), impacts of climate change are now being observed in all parts of the Mediterranean region high 2 train +6400 AR6_WGII 2255 8 Warm-water fish species are expected to move northwards, while cold-water species will decline, and invasions of thermal-tolerant tropical species will increase high 2 train +6401 AR6_WGII 2256 5 CCP4.3.2 Coastal Systems Sea level rise is the origin of multiple risks for low-lying areas in the Mediterranean Basin; for example, the further increase in flooding at high tide in some locations, such as Venice high 2 train +6402 AR6_WGII 2256 10 Overall, sea level rise is projected to increase the risk of coastal flooding despite the potential slight reductions of marine storms high 2 train +6403 AR6_WGII 2256 12 Impacts are projected to increase nonlinearly during the 21st century with higher sea level rise, because coastal flooding will progressively change from overtopping to overflow, high-tide flooding and ultimately permanent flooding and shoreline retreat high 2 train +6404 AR6_WGII 2256 20 Increasing heat waves, combined with drought and land use change, reduce fuel moisture, thereby increasing fire risk, extending the duration of fire seasons and increasing the likelihood of large, severe fires high 2 train +6405 AR6_WGII 2257 5 Climate change will likely reduce crop yields in many areas (Table CCP4.1), mainly due to higher temperatures affecting crop phenology and the shortening of the crop growing season high 2 train +6406 AR6_WGII 2258 1 CCP4.3.5 Human Health and Cultural Heritage Warming is projected to impact human health, mostly through increased intensity, frequency and duration of heat waves high 2 train +6407 AR6_WGII 2258 10 Extreme high temperatures, hot days and nights and consequently cooling degree days will likely increase high 2 train +6408 AR6_WGII 2258 16 Sea level rise will increase these risks high 2 train +6409 AR6_WGII 2258 24 Many engineering-based coastal adaptation imply large residual impacts on coastal ecosystems high 2 train +6410 AR6_WGII 2286 5 Observed changes include increasing temperatures, changing seasonal weather patterns, reductions in snow cover extent and duration at low elevation, loss of glacier mass, increased permafrost thaw and an increase in the number and size of glacier lakes high 2 train +6411 AR6_WGII 2286 7 Around two-thirds of treeline ecotones have also shifted upwards in recent decades, though these shifts are not ubiquitous and slower than expected based on rising temperatures high 2 train +6412 AR6_WGII 2286 8 Impacts on biological communities and animal species are also increasingly being reported, with species of lower elevations increasing in mountain regions, creating more homogeneous vegetation and increasing risks to mountain-top species medium 1 train +6413 AR6_WGII 2286 10 These changes have variable impacts on water availability for people and economies, contributing to increasing tensions or conflicts over water resources, especially in seasonally dry regions medium 1 train +6414 AR6_WGII 2286 14 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.Climate-change-driven changes in precipitation, river flow regimes and landslides affect the production and use of energy in mountain regions, in particular hydropower high 2 train +6415 AR6_WGII 2286 16 The combined effects of climate change, hydropower development and other human interventions have exacerbated water security problems and social injustice medium 1 test +6416 AR6_WGII 2286 18 Agriculture has been negatively affected through increased exposure to hazards such as droughts and floods, changes in the onset of seasons, the timing and availability of water, increasing pests and decreasing pollinator diversity, which in turn have negatively influenced overall food production, dietary diversity and the nutritional value of food medium 1 train +6417 AR6_WGII 2286 19 Negative climate impacts on pastoralism, such as drought- induced degradation of rangelands and pastures, have affected livestock productivity and the livelihood of pastoralists, while other non-climatic factors, such as land use change and management, also play a role medium 1 train +6418 AR6_WGII 2286 21 Exposure and vulnerability exacerbate the negative effects of climate impacts on livelihoods and intertwine with power imbalances and gender and other inequalities medium 1 train +6419 AR6_WGII 2286 23 For winter activities such as skiing, diminishing snow at lower elevations has challenged operating conditions (medium confidence), increasing the demand for and dependence on snow management measures such as snow-making high 2 train +6420 AR6_WGII 2286 24 Climate-induced hazards are negatively affecting some climbing, mountaineering and hiking routes medium 1 train +6421 AR6_WGII 2286 25 In some regions, options to change routes or shift seasons to reduce hazard exposure have been employed as adaptation strategies, with variable outcomes medium 1 train +6422 AR6_WGII 2286 26 In some cases, higher temperatures and extreme heat conditions at lower elevations have made some mountain destinations more appealing, increasing the potential for summer visitation demand medium 1 train +6423 AR6_WGII 2287 0 CCP5 2276Cross-Chapter Paper 5 Mountains further downstream high 2 train +6424 AR6_WGII 2287 1 As a result, the number of disasters has increased; however, there is limited evidence that this is due to changes in the underlying hazard processes, pointing mainly to increasing levels of exposure medium 1 train +6425 AR6_WGII 2287 3 Adaptation responses to climate-driven impacts in mountain regions vary significantly in terms of goals and priorities, scope, depth and speed of implementation, governance and modes of decision-making and the extent of financial and other resources to implement them high 2 train +6426 AR6_WGII 2287 7 Many low-elevation and small glaciers around the world will lose most of their total mass at a 1.5°C global warming level (GWL) high 2 train +6427 AR6_WGII 2287 12 However, nearly all mountain regions will face at least moderate and some regions even high risks at around 2°C GWL medium 1 train +6428 AR6_WGII 2287 16 This may enable some species to persist in a region, though it may pose a threat to cold-adapted species, including endemics, which may be outcompeted high 2 train +6429 AR6_WGII 2287 18 Intangible losses and loss of cultural values will become increasingly more widespread in mountain regions, mainly driven by a decline in snow and ice and an increase in intangible harm to people from hazards medium 1 train +6430 AR6_WGII 2287 22 Reducing climate risks will depend on addressing the root causes of vulnerability, which include poverty, marginalisation and inequitable gender dynamics high 2 train +6431 AR6_WGII 2287 24 Risk management strategies that better integrate the adaptation needs of all affected sectors, account for different risk perceptions and build on multiple and diverse knowledge systems, including Indigenous knowledge and local knowledge, are important enabling conditions to reduce risk severity medium 1 train +6432 AR6_WGII 2287 27 There are increasing calls for more ambitious climate action in mountains, providing impetus for stronger cooperation within and across mountain regions and downstream areas medium 1 train +6433 AR6_WGII 2291 1 Mountain river systems are especially sensitive to and affected by climate change and continuing anthropogenic disturbance, including water pollution, hydropower development, water withdrawals for agriculture and human consumption and biodiversity loss and ecosystem changes high 2 train +6434 AR6_WGII 2291 8 Many mountain regions have one or more cryosphere components (glaciers, permafrost and perennial or seasonal snow), and the mountain cryosphere is among the natural systems most sensitive to climate change worldwide high 2 train +6435 AR6_WGII 2291 9 The SROCC assessed a decline in all cryosphere components due to climate change over recent decades, i.e., for low-elevation snow cover (high confidence), permafrost (high confidence) and glaciers very high 3 train +6436 AR6_WGII 2291 17 Groundwater may provide some resilience to loss of melt water from glacier and snow decline, but in the longer term groundwater recharge and contribution to streamflow are expected to decrease with ongoing climate change medium 1 train +6437 AR6_WGII 2291 23 A majority of documented adaptation efforts in mountain regions address water-related aspects (precipitation variability and extremes, including drought, water availability, floods) high 2 train +6438 AR6_WGII 2292 23 Taken together, this evidence suggests that a significant risk exists of wildfire exacerbating other impacts of climate change on already vulnerable ecosystems in many mountain regions medium 1 train +6439 AR6_WGII 2294 3 The combined effects of climate change, hydropower development and further anthropogenic effects in upstream mountain basins have increased and are expected to further negatively affect several aspects of ecosystem functioning and water security (e.g., negative effects on river geometry, water chemistry, sediment transport, fish composition and migration) high 2 train +6440 AR6_WGII 2294 4 With respect to hydropower, mountains play a unique role in the production of renewable energy for large downstream populations, but it also comes with important trade-offs affecting mountain ecosystems and populations high 2 train +6441 AR6_WGII 2294 20 Medicinal and aromatic plants and their secondary metabolites are also observed to be affected by climate change medium 1 train +6442 AR6_WGII 2294 23 Experimental trials have shown that when medicinal plants are stressed by drought, phytochemical content increases, either by a decrease in biomass or by an increase in the actual production of metabolites medium 1 train +6443 AR6_WGII 2296 11 Energy provision in and around mountain settlements is another key sector affected by climate-related impacts (Hock et al., 2019; CCP5.2.2.2), which bears relevance for the adaptation prospects for urban mountain settlements medium 1 train +6444 AR6_WGII 2296 15 CCP5.2.6 Natural Hazards and Disasters Climate- and weather-related disasters in mountain regions have increased over the last three decades medium 1 train +6445 AR6_WGII 2296 16 Disaster frequency shows increasing trends in the HKH, the Andes and mountain regions in Africa, whereas no clear trends are observed for the European Alps and Central Asia medium 1 train +6446 AR6_WGII 2296 17 Floods, debris flows, landslides and avalanches are the most frequent hazards affecting the highest number of people in mountain regions medium 1 train +6447 AR6_WGII 2297 1 Overview of key observed impacts on mountain communities and adaptation responses References and relevant AR6 WGII sections Mountain livelihoods and poverty Impacts –In some mountain regions, the incidence of poverty can be higher compared to other areas, with observed impacts of climate change intensifying the deterioration of socioeconomic conditions that support livelihoods, thereby exacerbating already existing conditions of non-climate-related vulnerabilities and livelihood insecurity medium 1 train +6448 AR6_WGII 2297 27 Capacity-building interventions strengthen adaptation capacity and links to access institutional support medium 1 train +6449 AR6_WGII 2298 1 Overview of key observed impacts and adaptation on select livelihood activities and economic sectorsReferences and relevant AR6 WGII sections Mountain agriculture and pastoralism Impacts –Changes in temperature and seasonal precipitation patterns affect the timing and availability of water for agricultural activities high 2 train +6450 AR6_WGII 2299 0 CCP5 2288Cross-Chapter Paper 5 Mountains The characteristics of natural hazards in mountain areas have been widely explored, and evidence suggests that conditions favouring cascading impacts are a common feature high 2 train +6451 AR6_WGII 2299 6 Evidence from different mountain regions shows that adaptation and risk reduction efforts are less successful if they focus on hazards or risks without considering diverse risk and value perceptions of the affected people medium 1 train +6452 AR6_WGII 2299 7 Previous experience and local social contexts of exposure to climate-related disasters affect people’s perceptions and influence the patterns associated with disaster risk management and associated coping strategies high 2 train +6453 AR6_WGII 2299 8 Important synergies exist between disaster risk reduction, climate change adaptation and sustainable development in mountain regions medium 1 train +6454 AR6_WGII 2300 10 Furthermore, the vast majority of detected impacts imply negative impacts on natural and human systems high 2 train +6455 AR6_WGII 2300 12 Since IPCC AR5, the evidence for meaningful climate change impacts being reported using local knowledge sources has increased substantially high 2 train +6456 AR6_WGII 2300 20 Adaptation measures most commonly found include farming- related changes (e.g., resilient or drought-tolerant crop varieties, irrigation techniques, crop storage and livestock insurance schemes), infrastructure development, Indigenous knowledge, community- based capacity-building and ecosystem-based adaptation high 2 train +6457 AR6_WGII 2300 26 Several lines of evidence converge, indicating that most observed adaptations in mountains are incremental in nature and not transformative high 2 train +6458 AR6_WGII 2300 29 Limits to adaptation are found in a majority (>80%) of the assessed adaptation studies; around half of the studies reported soft limits, and less than a third identified both hard and soft limits to adaptation high 2 train +6459 AR6_WGII 2301 5 Overall, adaptation in mountain regions is taking place in various ways, in different sectors, scales, levels, quality, and effectiveness high 2 train +6460 AR6_WGII 2302 12 Climate change impacts on food, fibre and ecosystem products will be highly variable across mountain regions medium 1 train +6461 AR6_WGII 2302 18 Water resources in mountains and dependent lowlands will continue to be strongly impacted by climate change throughout the 21st century high 2 train +6462 AR6_WGII 2302 19 The difference in impacts will be particularly strong in regions that greatly depend on glacier and snowmelt and, in pronounced dry seasons high 2 train +6463 AR6_WGII 2302 21 Many low- elevation and small glaciers around the world will lose most of their total mass at 1.5°C GWL high 2 train +6464 AR6_WGII 2302 22 For tropical and mid-latitude mountains, around half of the current ice mass can be preserved under low-emission scenarios, while between two-thirds and up to more than 90% will be lost under high emission scenarios compared to the 2000s medium 1 train +6465 AR6_WGII 2302 26 The dependence of lowland populations on mountain water resources will grow by mid-century across several climate and socioeconomic scenarios, and several seasonally dry or semiarid mountain regions (e.g., parts of South Asia, North America) are projected to be highly dependent on such resources medium 1 train +6466 AR6_WGII 2303 1 However, water supply from glacier melt will decrease once source glaciers pass peak discharge (Huss and Hock, 2018), and the areas with available sediment will grow as glaciers shrink, posing potential risks to downstream populations and assets high 2 train +6467 AR6_WGII 2303 5 Higher-latitude and high-elevation locations generally exhibit delayed declines in snow reliability compared to lower-latitude and lower-elevation locations high 2 train +6468 AR6_WGII 2303 6 In general, climate change impacts and risks to ski tourism are found to be spatially heterogeneous, within and across local and international markets, with potential for significant disruptions to related socioeconomic sectors due to a growing mismatch between ski area supply and skier demand in the coming decades high 2 train +6469 AR6_WGII 2303 8 Severe damage and disruptions to people and infrastructure from floods are projected to increase in Northwestern South America (NWS), South Asia (SAS), Tibetan Plateau (TIB) and Central Asia (WCA) between 1.5°C and 3°C GWL, mainly driven by river floods and an increase in the number of glacial lakes with high potential for outburst high 2 train +6470 AR6_WGII 2303 10 Projected changes in ice and snowmelt, as well as seasonal increases in extreme rainfall and permafrost thaw, will favour chain reactions and cascading processes, which can have devastating downstream effects well beyond the site of the original event high 2 train +6471 AR6_WGII 2303 11 The incidence of disasters is projected to increase in the future because some hazards will become more pervasive, with an increase in the exposure of people and infrastructure with future environmental and socioeconomic changes either contributing to reduce or enhance these disaster risks medium 1 train +6472 AR6_WGII 2303 15 Extreme precipitation in major mountain regions is projected to increase, leading to consequences such as floods and landslides medium 1 train +6473 AR6_WGII 2303 23 Comparatively, more severe risk consequences are expected under SSP3 and/or SSP4 given the high population projections in certain regions compared to SSP1 medium 1 train +6474 AR6_WGII 2304 1 Consequences that are considered severe refer to the magnitude (number of people and economic activities affected), timing (increase of water stress as early as mid-century in several regions) and likelihood (severe risk consequences are more likely where high population density is projected) high 2 train +6475 AR6_WGII 2304 3 Hotspot regions are those with large lowland populations depending on essential mountain water resource contributions and include river catchments such as the Ganges, Brahmaputra, Meghna, Yangtze, Nile, Niger, Indus, Euphrates-Tigris or Pearl high 2 train +6476 AR6_WGII 2304 4 Limited governance and integrated management of water resources, power and gender inequalities and level of disruption of local community practices also contribute to making risks more severe medium 1 train +6477 AR6_WGII 2306 12 However, the extent of adaptation in terms of time (i.e., speed), scale of change (i.e., scope) and depth of change (i.e., degree to which a change is substantial) is low in mountain regions, with the level of agreement across studies varying from one region to the other medium 1 train +6478 AR6_WGII 2306 14 Adaptation responses in mountains are mainly incremental changes from existing practices high 2 train +6479 AR6_WGII 2306 17 The portfolio of adaptation options to mitigate risks from changing water resources (KR2) is large but challenging and includes integrated catchment management, implementation of multiple use of water strategies, improved water governance (including community-based and participatory water governance), overcoming power inequalities among users and sectors and balancing economic pressure and sustainable development high 2 train +6480 AR6_WGII 2306 24 However, repeated and recurrent disturbances that increase recovery times can reduce the effectiveness of EbA medium 1 train +6481 AR6_WGII 2306 25 Adaptation in mountain areas is currently constrained predominantly by soft limits related to existing social, economic and political conditions high 2 train +6482 AR6_WGII 2306 26 Progress in overcoming soft limits is currently minimal due to insufficient engagement with socioeconomic and political issues in existing adaptation medium 1 train +6483 AR6_WGII 2307 1 Furthermore, varied and dynamic biophysical characteristics as well as high sociocultural diversity preclude one- size-fits-all responses; adaptation planning and action in mountains rooted in context-specific socioecological and climatic realities are more effective high 2 train +6484 AR6_WGII 2307 5 Autonomous responses can be appropriate when local resilience is high (Mishra et al., 2019; Ford et al., 2020); however, many mountain communities continue to face socioeconomic challenges that constrain their adaptive capacity high 2 train +6485 AR6_WGII 2307 8 Such limits are already observed and are likely to become more widespread as climatic stressors move beyond historical experience high 2 train +6486 AR6_WGII 2307 10 The potential for adaptation to contribute to sustainable development and transformative change in mountains is also becoming increasingly evident medium 1 train +6487 AR6_WGII 2307 11 To better achieve the adaptation potential in mountains, adaptation finance and private-sector inclusion and contribution are key enablers high 2 train +6488 AR6_WGII 2308 5 Progress in addressing climate risks requires targeting the root causes of vulnerability, which are often socioeconomic in origin and can include poverty, marginalisation and inequitable gender dynamics high 2 train +6489 AR6_WGII 2308 10 Furthermore, there is increasing evidence that key conventions related to mountains, such as the Alpine Climate Board (SROCC section 2.4 (Hock et al., 2019)), provide opportunities for accelerating adaptation efforts through mainstreaming responses into other policies aimed at addressing climate-related risks medium 1 train +6490 AR6_WGII 2308 15 More inclusive adaptation approaches, engagement with the root causes of vulnerability, improved coordination and monitoring activities and upscaling of support for adaptation are key enablers and are indicative of a substantial solution space for adaptation in mountain regions high 2 train +6491 AR6_WGII 2308 18 CCP5.4.3 Climate-Resilient and Sustainable Development in Mountains With accelerating warming and compounding risks increasing above 1.5°C warming, the need for climate resilient development in mountains is evident and intricately linked to achieving the SDGs and equity high 2 train +6492 AR6_WGII 2308 22 NbS can be pursued in mountains that will mitigate climate change and its impacts while at the same time contributing to improving livelihoods, social and economic well-being and sustainable environmental management high 2 train +6493 AR6_WGII 2309 3 Indigenous knowledge and local knowledge reinforce community adaptive capacity, yet governance structures and processes, including the deliberate design and implementation of climate policy, can constrain that capacity from being realised high 2 train +6494 AR6_WGII 2332 0 CCP6 2321Polar Regions Cross-Chapter Paper 6 Executive Summary Observed Impacts and Future Risks Climate change impacts and cascading impacts in polar regions, particularly the Arctic, are already occurring at a magnitude and pace unprecedented in recent history very high 3 train +6495 AR6_WGII 2332 1 The polar regions, notably the Arctic and maritime Antarctic, are experiencing impacts from climate change at magnitudes and rates that are among the highest in the world, and will become profoundly different in the near-term future (by 2050) under all warming scenarios high 2 train +6496 AR6_WGII 2332 2 In the Arctic, accelerated sea ice loss (particularly during summer), increased permafrost thaw and extreme high temperatures have substantially impacted marine, freshwater and terrestrial sociological–ecological systems very high 3 train +6497 AR6_WGII 2332 3 Multiple physical, ecological and societal elements of polar regions are approaching a level of change potentially irreversible for hundreds of years, if not millennia high 2 train +6498 AR6_WGII 2332 4 Evidence of borealisation of terrestrial and marine systems is emerging high 2 train +6499 AR6_WGII 2332 5 Loss of multi-year sea ice and the occurrence of a seasonally ice-free Arctic Ocean by the middle of this century will result in substantial range contraction, if not the disappearance of several Arctic fish, crab, bird and marine mammal species, including possible extinction of seals and polar bears in certain regions high 2 train +6500 AR6_WGII 2332 6 In the Arctic, permafrost thaw and snowfall decrease lead to profound hydrological changes, an overall greening of the tundra and regional browning of tundra and boreal forests high 2 train +6501 AR6_WGII 2332 8 Climate change has induced food web changes resulting in population declines in polar sea birds, including penguins, and marine and terrestrial mammals high 2 train +6502 AR6_WGII 2332 9 Globally and regionally important harvested fish and invertebrate species are also contracting ranges and declining productivity, including Pacific cod, salmon, snow and king crab in the Arctic and krill in the Antarctic (medium confidence), with implications for global food systems high 2 train +6503 AR6_WGII 2332 13 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.shipping, tourism (very high confidence) and Arctic maritime trade and resource extraction medium 1 train +6504 AR6_WGII 2332 14 Navigational risks have grown due to increasingly mobile multi-year ice, poor hydrographic charting in newly open areas, and limited weather, water, ice and climate data and services high 2 train +6505 AR6_WGII 2332 15 Cascading risks from polar shipping growth include increased air emissions, underwater noise pollution, disruption to subsistence hunting and cultural activities in the Arctic (high confidence) and potential for invasive marine species and geopolitical tensions medium 1 train +6506 AR6_WGII 2332 17 Arctic permafrost thaw is projected to impact most infrastructure by the middle of this century, impacting millions of people and their economies, and costing billions in damages high 2 train +6507 AR6_WGII 2332 19 It has negatively impacted mental health and increased risks of injury, food insecurity and foodborne and waterborne disease, with risks amplified for those reliant on the environment for subsistence, livelihoods and identity high 2 train +6508 AR6_WGII 2332 20 Permafrost thaw, sea level rise and reduced sea ice protection have already damaged or destroyed many cultural heritage sites in some Arctic regions (very high confidence) and are projected to continue across all Arctic regions very high 3 train +6509 AR6_WGII 2332 22 Polar zones will continue to contract and diminish in extent under climate change, and local adaptations will be insufficient to achieve long-term resilience of polar systems medium 1 train +6510 AR6_WGII 2332 23 The pace and extent of change in polar regions is challenging the ability of social and natural systems to adapt medium 1 train +6511 AR6_WGII 2333 0 CCP6 2322Cross-Chapter Paper 6 Polar Regions biodiversity medium 1 train +6512 AR6_WGII 2333 2 Governance around climate change planning, preparation and response has been limited in scope, and has often not considered interacting effects of climate change with other risks high 2 train +6513 AR6_WGII 2333 3 Reactive management strategies will not succeed in reducing risks in polar regions given the rapid change and increasing potential for extreme events high 2 train +6514 AR6_WGII 2333 4 Greater inclusivity of stakeholders and communities, along with using diverse sources of information, including Indigenous knowledge and local knowledge, can benefit robust planning and decision making, and uptake of adaptations high 2 train +6515 AR6_WGII 2333 5 Effectiveness in preparing for and adapting to climate risks can benefit from improved climate, weather and ice forecasting services, tools for integrating climate change data and different types of knowledge into management processes and enhanced polar search, rescue and emergency response capabilities high 2 train +6516 AR6_WGII 2333 7 Development of robust pathways for climate resilience in the Arctic can be accelerated by adaptation strategies and governance that reflect local conditions, cultures and adaptive capacities of communities and sectors high 2 train +6517 AR6_WGII 2333 8 Effectiveness of adaptation strategies will be enhanced by accounting for the geographic, climatic, ecological and cultural uniqueness of the polar regions medium 1 train +6518 AR6_WGII 2333 9 Colonialism can inhibit the development of robust climate adaptation strategies, and exacerbate climate risks very high 3 train +6519 AR6_WGII 2333 10 Inclusive decision making in establishing climate adaptations can foster resilience, reflect the unique environmental, cultural and economic imperatives of the region and support both market-based and sharing economies high 2 train +6520 AR6_WGII 2333 12 Arctic Indigenous self-determination in decision making can establish robust climate resilience, especially in Indigenous communities, incorporating locally derived definitions of social and economic success, culturally legitimate institutions of government, strategic visioning and thinking and public-spirited, nation-building leadership very high 3 train +6521 AR6_WGII 2334 2 These changes are causing a suite of direct and cascading risks for all polar ecosystems with larger effects to date in the Arctic than the Antarctic high 2 train +6522 AR6_WGII 2335 3 Driver Region Observed changes Projected changes Marine and sea ice Sea level (relative) ArcticNo consistent trend (increase in northwest America, decrease in northeast America, stable in Greenland and Arctic Russia) (WG1-12)Rise in all polar regions (except areas of substantial land uplift in northeast Canada, the west coast of Greenland) (high confidence); Increase of extreme sea levels in Russian Arctic and northwest America (high confidence) Greenland/Iceland and northeast America (given glacial isostatic adjustment) (medium confidence, WG1-12) AntarcticRise in all polar regions (except areas of substantial land uplift in west Antarctica) (high confidence, WG1-12) Sea surface temperature ArcticIncrease of ~0.5°C per decade during 1982–2017 in ice-free regions in summer (high confidence, SROCC-3)Further increases (high confidence, WG1-12) AntarcticWarmed in northern areas of Southern Ocean but cooled in its southernmost regions since the 1980s (high confidence, SROCC-3)Circumpolar increases (high confidence, WG1-12) Sea ice cover ArcticLoss (particularly of multi-year sea ice) accelerated since 2001 (very likely, WG1-9)Will become sea ice free (< 1 × 106 km2) during summer before 2050, irrespective of global warming level (likely, WG1-9) AntarcticNo significant circumpolar trend from 1979–2018 (very high confidence), but decrease off the Antarctic Peninsula high 2 train +6523 AR6_WGII 2336 6 Warming and wetting have persisted as key climatic impact drivers in polar regions very high 3 train +6524 AR6_WGII 2336 20 CCP6.2 Observed Impacts and Future Risks CCP6.2.1 Marine and Coastal Ecosystems CCP6.2.1.1 Warming and sea ice retreat cause shifts in distribution ranges of species In Arctic seas, warming and other climate impact drivers, primarily sea ice retreat, have led to range contractions of Arctic marine and ice-associated species and poleward expansions of boreal species very high 3 train +6525 AR6_WGII 2336 21 Altered conditions allow more microorganisms to move poleward and provide opportunities for invasive species (Cavicchioli et al., 2019; Nielsen et al., 2020; Driver Region Observed changes Projected changes Permafrost ArcticRising permafrost temperatures over past three to four decades (high confidence, WG1-9); decreases in permafrost active layer thickness very high 3 train +6526 AR6_WGII 2336 22 Submarine permafrost warming (medium confidence, WG1-9)Increases in temperature and active layer thickness (WG1-9); near-surface terrestrial permafrost extent will reduce under all scenarios by 2100 (virtually certain, WG1-9) AntarcticRising permafrost temperatures over past three to four decades (high confidence, WG1-9) Lake, river ice ArcticDeclines in seasonal lake ice cover thickness and duration over most Arctic lakes; declines in cold-season river ice extent (high confidence, WG1-12)Many lakes will lose >1 month lake ice cover by 2050 medium 1 train +6527 AR6_WGII 2338 1 Affected systemHazard *Cascading effectObserved impacts, future risks and natural adaptations identified in SROCC (confidence level) Arctic marine ecosystems Primary producers (PP-1)Sea ice loss * Freshening * StratificationImpact: timing (earlier and later blooms), distribution and magnitude (>30% increase in annual net primary production since 1998) (high confidence) Acidification Adaptation: phytoplankton may compensate for decrease in pH Zooplankton * PP-1 Impact: changing production and community composition (medium confidence) Benthos * PP-1 Impact: changing production and biodiversity medium 1 train +6528 AR6_WGII 2339 4 Numerous mammals and sea birds respond to changes in the distribution of their preferred habitats and prey by shifting their range, altering the timing or pathways for migration or switching prey very high 3 train +6529 AR6_WGII 2339 5 Ice-breeding seals (e.g., harp seals – Pagophilus groenlandicus) often have little scope to shift distribution, leading to increases in strandings and pup mortality in years with little ice cover medium 1 train +6530 AR6_WGII 2339 6 Recent studies confirm that polar bears (Ursus maritimus) are negatively affected by changing ice and snow conditions with decreases in denning, foraging, reproduction, genetic diversity and survival rates very high 3 train +6531 AR6_WGII 2339 8 Such shifts have so far only been detected for Antarctic krill (Euphausia superba), with a poleward contraction of the highest densities of krill in the Atlantic sector medium 1 train +6532 AR6_WGII 2339 9 Ocean warming is expected to put pressure on Antarctic phytoplankton (Pinkerton et al., 2021) and fish species unable to move further south in shelf areas, including waters off sub-Antarctic islands low 0 train +6533 AR6_WGII 2339 10 Off the Antarctic Peninsula and sub-Antarctic islands, invasive benthic invertebrates and macroalgae have already been detected medium 1 train +6534 AR6_WGII 2339 11 On a local to regional scale, the benthic recolonisation of the newly exposed seabed after the disintegration of ice shelves shows typical succession patterns, with mass occurrences of few pioneer species followed by gradual shifts to a more diverse typical shelf community, driven by increasing pelagic primary production upon ice-shelf collapse and strengthening of the pelagic–benthic coupling high 2 train +6535 AR6_WGII 2339 12 Range changes of Antarctic birds and marine mammals have been observed, which vary among sub- regions and are mostly attributable to changes in sea ice extent and food availability high 2 train +6536 AR6_WGII 2339 13 With projected sea ice retreat and associated change in prey distribution (Henley et al., 2020), foraging areas of sub-Antarctic sea birds and marine mammals will shift southwards, leading to elevated pressure on populations due to higher foraging costs during the breeding season medium 1 train +6537 AR6_WGII 2339 14 These changes are particularly impacting emperor penguins (Aptenodytes forsteri) (Table CCP6.2), with the projected population declining close to extinction by 2100 under Business-As-Usual climate scenarios (medium confidence) (Jenouvrier et al., 2020; Trathan et al., 2020; Jenouvrier et al., 2021), whereas population decline is halted by 2060 under the 1.5°C climate scenario low 0 train +6538 AR6_WGII 2339 15 CCP6.2.1.2 Ocean warming and sea ice changes affect marine primary productivity In the central Arctic Ocean, primary productivity remains low medium 1 train +6539 AR6_WGII 2339 16 In inflowing (Barents and Chukchi Sea) and interior shelf regions (Laptev, Kara, and Siberian Sea), changes in sea ice extent, thickness and seasonal timing have altered light and mixing regimes, causing increasing overall productivity in open-water and under-ice habitats, and in leads high 2 train +6540 AR6_WGII 2339 17 Productivity changes are associated with the earlier- onset phytoplankton spring blooms and the increasing occurrence of autumn blooms, particularly at lower latitudes of the Arctic high 2 train +6541 AR6_WGII 2339 18 Ice algal communities are expected to change in productivity and species composition in response to the transition from a predominantly multi-year to a seasonal sea ice pack high 2 train +6542 AR6_WGII 2340 0 CCP6 2329Polar Regions Cross-Chapter Paper 6 in Greenland medium 1 train +6543 AR6_WGII 2340 1 Macroalgae and seagrass are generally expanding in the Arctic medium 1 test +6544 AR6_WGII 2340 2 In the future Arctic Ocean, higher light availability in response to further sea ice decline and reduced deep mixing is projected to generally increase primary productivity medium 1 train +6545 AR6_WGII 2340 3 However, productivity may increase less than predicted and eventually even decrease once nutrient limitation outweighs the benefits of higher light availability low 0 train +6546 AR6_WGII 2340 4 Despite large-scale environmental changes in the Southern Ocean, such as the deepening of the summer mixed layer (medium confidence) (Panassa et al., 2018; Sallée et al., 2021), and the expected impacts via altered nutrient entrainment, light availability and grazer encounter rates (Chapter 3) (Behrenfeld and Boss, 2014; Llort et al., 2019), assessments indicated no consistent changes in primary production at the circumpolar scale, as sectors and regions show different trends medium 1 test +6547 AR6_WGII 2340 6 Primary productivity has increased in the Pacific sector and decreased in the Atlantic sector and the Ross Sea low 0 test +6548 AR6_WGII 2340 7 Higher productivity has also been observed in regions where rapid environmental changes occurred, such as in the vicinity of retreating IS and declining sea ice cover off the Antarctic Peninsula medium 1 train +6549 AR6_WGII 2340 10 Such an increase in Southern Ocean productivity will lead to a decline in global ocean productivity medium 1 train +6550 AR6_WGII 2340 11 CCP6.2.1.3 Impacts of ocean acidification vary spatially and among biotas In Arctic seas, areas with acidification levels corrosive to organisms forming CaCO 3 shells or skeletons expanded between the 1990s and 2010 high 2 train +6551 AR6_WGII 2340 12 Key species of diatom and picoeukaryote phytoplankton species yet appear relatively resilient to decreasing pH levels over a range of temperature and light conditions medium 1 train +6552 AR6_WGII 2340 13 In contrast, there is evidence for species- and stage-specific sensitivities of zooplankton, pteropods and fishes high 2 train +6553 AR6_WGII 2340 14 Warming, rising river-sediment discharge and coastal erosion in Arctic shelf regions are expected to increase the input of labile, often permafrost-derived organic carbon, the remineralisation of which further increases acidification rates medium 1 train +6554 AR6_WGII 2340 16 In the Southern Ocean, calcifying organisms are also most vulnerable to ocean acidification high 2 train +6555 AR6_WGII 2340 17 Calcifying species with low- magnesium calcite or mechanisms to protect their skeletons are less vulnerable to the corrosive effects of acidification than those using aragonite or high-magnesium calcite high 2 train +6556 AR6_WGII 2340 18 In diatom-dominated communities, silicification diminishes with reduced pH levels, albeit with rates differing among taxa low 0 train +6557 AR6_WGII 2340 19 Species-specific responses exist regarding growth and primary production, which are further strongly modulated by iron and light availability high 2 train +6558 AR6_WGII 2340 23 CCP6.2.1.4 Climate change alters food web dynamics Climate change has transformed Arctic marine ecosystems from sea ice- associated to open-water production regimes, with profound impacts on trophic energy transfer efficiencies and pathways (high confidence) (Behrenfeld et al., 2017; Meredith et al., 2019; Huntington et al., 2020) as well as benthic–pelagic coupling medium 1 train +6559 AR6_WGII 2340 24 Shifts in bloom phenology favour small phytoplankton and smaller zooplankton over large lipid-rich macro-zooplankton, leading to longer, less efficient food chains medium 1 train +6560 AR6_WGII 2341 1 Species range shifts have restructured higher trophic levels in Arctic food webs high 2 train +6561 AR6_WGII 2341 5 Climate impacts on Arctic marine food webs will be profound and intensify with GWL high 2 train +6562 AR6_WGII 2341 6 However, the exact nature of these impacts remains unclear due to attenuating and amplifying dynamics of both top-down and bottom-up processes in polar food webs and the management of fisheries high 2 train +6563 AR6_WGII 2341 8 Warming is expected to reduce the quantity and quality of lipid-rich copepod prey (high confidence) (Aarflot et al., 2018; Kimmel et al., 2018; Bouchard and Fortier, 2020; Møller and Nielsen, 2020; Mueter et al., 2020), leading to declines in survival and growth of multiple upper-trophic level fish species; these impacts are amplified over time under low mitigation scenarios (RCP8.5) high 2 train +6564 AR6_WGII 2341 9 Marine mammals and sea birds will continue to attenuate climate change impacts by shifting their diets and behaviour medium 1 train +6565 AR6_WGII 2341 10 However, sea birds generally have low temperature-mediated plasticity of reproductive timing, making them vulnerable to mismatches with their prey and limiting long-term adaptation medium 1 train +6566 AR6_WGII 2341 15 The optimum habitat for Antarctic krill is expected to decline with a shortening of suitable season for krill growth and reproduction, particularly in the northern Scotia and Bellingshausen Seas medium 1 train +6567 AR6_WGII 2341 18 Although salps have long been considered to be competitors of Antarctic krill (Suprenand and Ainsworth, 2017; Rogers et al., 2020), they provide a third energy pathway in pelagic food webs and, given the changing ocean conditions and their preference for smaller phytoplankton, may increase in importance for copepods low 0 train +6568 AR6_WGII 2341 20 CCP6.2.2 Terrestrial and Freshwater Ecosystems Since the publication of AR5 (IPCC, 2014) and SROCC (IPCC, 2019) and their findings (Table CCP6.2), more studies confirm rapid changes in Arctic terrestrial and freshwater systems including increased permafrost thaw, changes to tundra hydrology and vegetation (overall greening of the tundra, regional browning of tundra and boreal forests), coastal and riverbank erosion high 2 train +6569 AR6_WGII 2341 23 Further evidence shows that warming and changes to the Arctic hydrologic cycle increase the risk of wildfire medium 1 train +6570 AR6_WGII 2341 24 Both the frequency of and the area burned by wildfires during recent years are unprecedented compared with the last 10,000 years high 2 train +6571 AR6_WGII 2342 0 CCP6 2331Polar Regions Cross-Chapter Paper 6 most tundra and boreal regions, and interactions between climate and shifting vegetation (Song et al., 2018) will influence future fire intensity and frequency medium 1 train +6572 AR6_WGII 2342 2 Even though the overall regional water cycle will intensify, including increased precipitation, evapotranspiration and river discharge to the Arctic Ocean (Table CCP6.1), snow and permafrost decline may lead to further soil drying medium 1 train +6573 AR6_WGII 2342 5 Soil temperatures along the Antarctic Peninsula are now sufficient for germination of non-native plants; invasions by non- endemic species are expected to increase with rising temperatures (high confidence) (Bokhorst et al., 2021), posing a risk to endemic polar species medium 1 train +6574 AR6_WGII 2342 6 Vegetation responses to warming are contingent on water availability and local temperature medium 1 train +6575 AR6_WGII 2342 8 West Antarctica is showing evidence of greening in the dominant cryptogrammic vegetation, with greater growth in mosses high 2 train +6576 AR6_WGII 2342 9 Peatland ecosystems may increase on the west Antarctic Peninsula with future warming low 0 train +6577 AR6_WGII 2342 10 In contrast, some parts of East Antarctica and the subantarctic islands to the north have been experiencing a drying climate, with declining health of mosses and other vegetation high 2 train +6578 AR6_WGII 2342 15 However, this trend is not as great in southern colder locations (medium confidence) (e.g., Kim et al., 2018; Newsham et al., 2019), as the microbial community structure is affected by vegetation cover and water availability high 2 train +6579 AR6_WGII 2342 16 Antarctic terrestrial invertebrate communities on the West Antarctic Peninsula may be controlled more by vegetation and water availability than by air temperature medium 1 train +6580 AR6_WGII 2342 17 Evidence from laboratory studies, field programmes and sedimentary records indicate that Antarctic freshwater ecosystems may become more productive under climate warming scenarios medium 1 train +6581 AR6_WGII 2342 19 Since SROCC, there is further evidence that climate change alterations of polar ecosystems increasingly challenge production of, and access to, sufficient, healthy and nutritious food, posing risks to future food and nutritional security within and beyond polar regions high 2 train +6582 AR6_WGII 2342 21 Climate change has impacted Indigenous subsistence resources across the Arctic (very high confidence) (SMCCP6.2), and future food systems and ecological connections are at risk from future climate change hazards interacting with non-climate pressures, some of which are mediated or amplified by novel conditions and opportunities in Arctic regions high 2 train +6583 AR6_WGII 2342 22 Increasing heatwaves, wildfires, extreme precipitation, permafrost loss and rapid seasonal snow and ice thaw events will further threaten terrestrial subsistence food resources across the Arctic high 2 train +6584 AR6_WGII 2344 3 Shifting spatial distributions of fish stocks have led to transboundary management challenges in the Atlantic, Bering Sea and Arctic areas previously inaccessible due to sea ice (Table CCP6.6) (Gullestad et al., 2020).Cascading and interacting effects of climate change impacts in polar regions (Table CCP6.1) will reduce access to, and productivity of, future fisheries, and pose significant risks to regional and global food and nutritional security that increase with atmospheric carbon levels and declines in sea ice high 2 train +6585 AR6_WGII 2344 5 Some global- scale models project increases in potential fishery yields in Arctic Canada (Cheung, 2018; Bindoff et al., 2019; Tai et al., 2019), whereas many observational studies and high-resolution regional projections suggest overall declines in biomass, productivity and yield associated with warming and loss of sea ice in multiple regions such as the Bering Sea medium 1 test +6586 AR6_WGII 2344 6 Reduced production of macronutrients and protein by polar marine sources will disproportionately impact people already experiencing food and nutritional scarcity (Myers et al., 2017), marine-dependent communities within and beyond polar regions, and women and children who require higher quantities of macronutrients high 2 train +6587 AR6_WGII 2345 0 CCP6 2334Cross-Chapter Paper 6 Polar Regions Large-scale commercial fisheries are expected to continue to operate in polar regions (high confidence) (Barange et al., 2018; Cavanagh et al., 2021; Grant et al., 2021), and will shift poleward (high confidence) toward geopolitical and management boundaries high 2 train +6588 AR6_WGII 2345 2 Increased distances from ports to redistributed fishing grounds as well as increased frequency of storms and other extreme events are expected to increase risks and costs for fishery operations medium 1 train +6589 AR6_WGII 2345 4 There will be increased demand for new port infrastructure across the Arctic high 2 train +6590 AR6_WGII 2346 2 Coupling adaptation measures with global carbon mitigation strategies substantially decreases climate change risks to polar fisheries very high 3 train +6591 AR6_WGII 2346 3 CCP6.2.4 Economic Activities Climate change presents significant risks to economic activities in the polar regions (very high confidence) and simultaneously enables development possibilities for fisheries (CCP6.2.3.3), agriculture (CCP6.2.3.2), the sharing and subsistence economy (CCP6.2.3.1) (SMCCP6.2) (high confidence), maritime trade (Box CCP6.1), natural resource development (CCP6.2.4.1) medium 1 train +6592 AR6_WGII 2346 5 CCP6.2.4.1 Changing access to natural resources with consequences for safety, economic development and climate mitigation Climate change is improving access to natural resources in the Arctic with consequences for human safety (high confidence), economic development (very high confidence) and global mitigation efforts medium 1 train +6593 AR6_WGII 2346 13 Climate change has increased risks to, and demand for, polar tourism experiences with the development of a ‘last chance tourism market’ medium 1 test +6594 AR6_WGII 2346 21 Climate hazards create risks to transportation sectors with consequences for economic development high 2 test +6595 AR6_WGII 2347 2 Fog (low confidence) and an increase in precipitation falling as ice pellets or hail high 2 train +6596 AR6_WGII 2347 4 Polar settlements are at significant risk from climate change through shoreline erosion, permafrost thaw and flooding high 2 train +6597 AR6_WGII 2347 6 Degradation of ice-rich permafrost can threaten the structural stability and functional capacities of community-based infrastructure (i.e., airports and roads; CCP6.2.5) and can have implications for local economies with coupled impacts for local livelihoods, health and well- being (CCP6.2.5, CCP6.2.6) high 2 train +6598 AR6_WGII 2350 9 Permafrost thaw, sea-level rise and reduced sea ice protection also presents the potential for the introduction and propagation of invasive species (Chan et al., 2019; Rosenhaim et al., 2019), and sovereignty tensions with implications for global geopolitics (Drewniak et al., 2018) medium 1 test +6599 AR6_WGII 2351 4 CCP6.2.6 Human Health and Wellness in the Arctic Climate change continues to have wide-ranging physical human health risks in the Arctic, particularly for Indigenous Peoples high 2 train +6600 AR6_WGII 2351 14 Climate change has negative, widespread and cumulative impacts on mental health in the Arctic, particularly for Indigenous Peoples very high 3 train +6601 AR6_WGII 2351 17 The negative mental health impacts from climate change are amplified among those most reliant on the environment for subsistence and livelihoods, those who already face chronic physical or mental health issues, and those facing socioeconomic inequities and marginalisation, particularly for Indigenous Peoples high 2 train +6602 AR6_WGII 2351 19 Climate change will increase mental health risks in the Arctic in the future medium 1 train +6603 AR6_WGII 2353 1 Therefore, effective responses to climate change risks for Indigenous Peoples are self-determined and underpinned by Indigenous knowledge (IK) very high 3 train +6604 AR6_WGII 2358 3 Sector Direct and cascading risksEnabling principles of cli- mate resilience pathwaysAnticipated future condi- tions/level of certaintyCompounding risks (non-cli- matic factors) Coastal settlements (CCP6.2.5)Change in extent of sea ice with more storm surges, thawing of permafrost, SLR and coastal erosionLocal leadership and community-led initiatives to initiate and drive processes, responsive agencies, established processes for assessments and planning, geographic optionsIncreasing number of communities needing relocation (medium confidence), rising costs for mitigating erosion (high confidence)Limitations of government budgets, other disasters that may take priority, policies deficiencies for addressing mitigation and relocation Human health (CCP6.2.6)Increased food insecurity, waterborne disease, emerging pathogens, injury and death, and negative mental health outcomesResources to support public programmes; Indigenous self-determination; access to technology; supporting IK systems; interdisciplinary and integrated decision makingThe intersection of social determinants of health will modify or mediate climate change impacts on health (very high confidence)Underlying health conditions, advances in diagnosis and treatment, and other health system shocks (e.g., COVID) Transportation (aviation, rail, road, ice roads) (CCP6.2.4.3)Permafrost thaw, sea ice change, storm surge, coastal erosion, changing precipitation patterns (ice pellets, hail) and extreme events create risks to transportation infrastructure with consequences to navigation, economics, safety and securityFinancial and human resources for: climate-resilient infrastructure research, development and implementation; improved weather, water, ice and climate forecasting at appropriate scales; improved communications infrastructure; local search and rescueLimits to adaptation exist (high confidence), but strategic investments in technologically innovative infrastructure that offers mitigation co-benefits will greatly enhance adaptation effectiveness (very high confidence)Level of local, regional and national infrastructure development, commitment of national and state level government to sustainable development pathways, global economic and political trends, commodity prices, unforeseen system shocks Shipping (Box CCP6.1; FAQ CCP2)Sea ice reduction leading to increased shipping related to trade, tourism, fisheries, resource development and re-supply with cascading risks from ships such as: increased under-water noise, potential introduction of invasive species, fuel spill risks, release of black carbon and air emissions, impacts to cultural resources, implications for subsistence hunting and food security, increased accidents and incidentsFinancial support for ship-building technologies (e.g., low-emission fuels, propulsion technologies, hull strength); development of robust multi-national agreements (in addition to existing agreements); inclusion of Indigenous Peoples in decision making; investment in multi-national and longitudinal research on shipping impacts; and enhancing modern digital maritime chartingShip traffic will continue to grow in polar regions (high confidence), with Arctic trade routes becoming increasing accessible (very high confidence) albeit with more challenging navigation due to increases in mobile ice in the near-term compared with late century when ice is expected to diminish completely during the shipping season high 2 train +6605 AR6_WGII 2359 0 CCP6 2348Cross-Chapter Paper 6 Polar Regions Sector Direct and cascading risksEnabling principles of cli- mate resilience pathwaysAnticipated future condi- tions/level of certaintyCompounding risks (non-cli- matic factors) Non-renewable resource extraction (Arctic only) (CCP6.2.4.1)Reduced sea ice improves access to non-renewable resources in remote Arctic regions, while warming temperature and thawing permafrost affect production levels, quality, and reliability and season length of ice roads, leading to increased operational costsInvestment in climate-resilient infrastructure and low-emission transportation (shipping) and investment in solar powered ships and low-impact modular mining camp infrastructureIncrease in mining in newly accessible marine regions (medium confidence), frequent false starts (i.e., due to climatic and non-climatic factors) (high confidence) and high levels of operational uncertainty (i.e., commodity prices, economic trends, climate risks) (very high confidence)Commodity prices; global economic trends and shocks; Indigenous rights and decisions; changing regulatory environments, geopolitics, global demand for resources Tourism (CCP6.2.4.2)Increased demand for polar tourism activities including development of ‘last chance tourism’ market; increased tourism improves economic conditions but leads to increased environmental and cultural impactsFinancial resources for service and infrastructure development; Indigenous self-determination and development of co-management approaches for natural and cultural attractions; development of multi-stakeholder/rightsholder tourism task teamsPolar tourism demand will continue to increase, especially for cruise and yacht experiences (high confidence) and enhance risks related to ship groundings, accidents and incidents (medium confidence)Limited search and rescue capacity, poor infrastructure, aging expedition cruise ship fleet, uncharted waters, geological and sovereignty debates, global economic trends, unforeseen events (i.e., severe acute respiratory syndrome (SARS), COVID-19) altering tourism demand patterns Reindeer herding (CCP6.2.3.1; 6.2.3.2; 6.2.5; Box CCP6.2)Rain-on-snow events causing high mortality of herds, especially in the autumn season; shrubification of tundra pasture lowering forage qualityFlexibility in movement to respond to changes in pastures, secure land use rights; adaptive management; continued economic viability and cultural tradition; self-determination in decision making; adequate support for communication and technological services; Indigenous rights upheld and protectedIncreased frequency of extreme events and changing forage quality adding to vulnerabilities of reindeer and herders high 2 train +6606 AR6_WGII 2359 1 Declines in catch impact livelihoods, coastal communities, and pose a risk to regional and global food and nutritional security very high 3 train +6607 AR6_WGII 2359 3 Adaptation at the local, individual, and household level under low mitigation scenarios will be costly and possibly undermined by the scale and pace of change, including climate shocks and extreme events medium 1 train +6608 AR6_WGII 2360 2 Efforts to minimise and prevent extinctions; preservation of ecosystem processes and habitats during critical life stages; coordinated governance; measures and planning that consider dynamic interactions within and among social and ecological systems are more effectiveWithout institutional investment in sustaining climate resilience in ecosystems across sectors there is a high risk of failure high 2 train +6609 AR6_WGII 2361 2 Human health is generally under- represented in adaptation initiatives, along with adaptations being developed within larger Arctic settlements (Ford et al., 2014; Canosa et al., 2020), and in many sectors decisions continue to be made without explicit inclusion of climate change impacts and risk in planning and design high 2 train +6610 AR6_WGII 2362 15 In the Arctic, large landscapes, dispersed population centres, limited resources and settler colonialism are structural barriers to effective planning, emergency response, and relief and recovery from climate impacts medium 1 train +6611 AR6_WGII 2362 18 Large ‘responsiveness gaps’ between impacts and implementation, approaches that fail to consider dynamic responses within social and ecological systems (which amplify or attenuate climate impacts), and a paucity of a priori planning can contribute to maladaptation high 2 train +6612 AR6_WGII 2362 21 Inclusive and participatory decision making underpins long-term resilience to climate change medium 1 train +6613 AR6_WGII 2362 23 The rapid pace of change, such as sea ice loss, can outpace ecological processes and induce substantial ecological shifts (CCP6.2) medium 1 train +6614 AR6_WGII 2362 24 The speed of climate change in the Arctic limits options for adaptation in communities who rely on a narrow resource base, when adaptation involves loss of culture and livelihoods, and when the costs of adaptation make it infeasible medium 1 train +6615 AR6_WGII 2363 5 Notably, terrestrial areas of greatest interest for infrastructure development, agriculture and visitation potential are often also the same areas that have been identified as culturally and ecologically significant (PEW, 2016; Eliasson et al., 2017; Grant et al., 2021) high 2 train +6616 AR6_WGII 2363 13 For the Antarctic, the governance for managing climate impacts is emerging, particularly for terrestrial and nearshore habitats high 2 train +6617 AR6_WGII 2363 17 CCP6.4.2 Inclusive, Integrated Co-management Inclusive, low-cost participatory decision making can deliver equitable responses to climate change high 2 train +6618 AR6_WGII 2363 19 The capacity of governance systems in some Arctic regions to respond to climate change has strengthened recently high 2 train +6619 AR6_WGII 2364 3 For residents in the Arctic, a sustainable development pathway has been found to be highly effective if a self-determined development model is employed very high 3 train +6620 AR6_WGII 2382 0 CCP7 2371Tropical Forests Cross-Chapter Paper 7 Executive Summary Over 420 million ha of forest were lost to deforestation from 1990 to 2020; more than 90% of that loss took place in tropical areas high 2 train +6621 AR6_WGII 2382 4 Climate change is altering the structure and species composition of tropical tree communities (high confidence), including transitions from moist to drier forest in regions such as the Amazon (high confidence), and movement of species from lower to higher elevations high 2 train +6622 AR6_WGII 2382 5 Despite CO 2 fertilisation, ongoing climate change has weakened the carbon sink potential of tropical forests in Amazonia and, to a lesser extent, in Africa and Asia medium 1 train +6623 AR6_WGII 2382 7 Deforestation generally reduces rainfall and enhances temperatures, with effects depending on scales (high confidence), while often increasing surface runoff medium 1 train +6624 AR6_WGII 2382 8 Continued deforestation-driven landscape drying and fragmentation will aggravate fire risk and reduce forest resilience, leading to degradation or savannisation of the tropical forest biomes, in particular in combination with climate change high 2 train +6625 AR6_WGII 2382 11 Forest restoration using a diverse mix of native species can help rebuild the climate resilience of tropical forests, but is best implemented alongside other sustainable forest management strategies and adaptation interventions high 2 train +6626 AR6_WGII 2382 14 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.Community-based adaptation, built on Indigenous knowledge and local knowledge (IK and LK) over centuries or millennia, is often identified as an effective adaptation strategy to climate change high 2 train +6627 AR6_WGII 2382 15 For successful adaptation of tropical forest communities, it is vital to consider IK and LK in addition to modern scientific approaches, together with consideration of non-climatic vulnerabilities (e.g., poverty, gender inequality and power asymmetries) high 2 train +6628 AR6_WGII 2382 17 Transformative and sustainable practices are required for effective management of tropical forests high 2 train +6629 AR6_WGII 2382 19 Tropical deforestation is largely driven by agriculture, both from subsistence farming and industrial agriculture (e.g., oil palm, timber plantations, soybeans, livestock) high 2 train +6630 AR6_WGII 2382 20 While poverty and population growth combined with poor governance often fuel subsistence agriculture (high confidence), industrial agriculture is often driven by international market forces for commodities and large- scale land acquisitions high 2 train +6631 AR6_WGII 2382 22 Transformative levers towards improving environmental governance and resilience of tropical forests include: incentivising and building capacity for environmental responsibility and discontinuing harmful subsidies and disincentives; reforming segmented decision-making to promote integration across sectors and jurisdictions; pursuing pre-emptive and precautionary actions; managing for resilient social and ecological systems in the face of uncertainty and complexity; strengthening environmental laws and policies and their implementation; acknowledging land tenure and rights; and inclusive stakeholder participation medium 1 train +6632 AR6_WGII 2383 0 CCP7 2372Cross-Chapter Paper 7 Tropical Forests CCP7.1 Introduction Climate change is already impacting tropical forests around the world, including through distributional shifts of forest biomes, changes in species composition, biomass, pests and diseases, and increases in forest fires high 2 train +6633 AR6_WGII 2383 4 Reducing direct and indirect drivers of deforestation and forest degradation is therefore critical to building, maintaining or enhancing the resilience of tropical forests against climate and non-climate drivers alike high 2 train +6634 AR6_WGII 2383 6 To the degree to which forests are affected by climate change and other drivers, their resilience against these stressors is diminishing leading to a reduction in the regulating, supporting, provisioning and cultural ecosystem services they provide (Alroy, 2017; Cadman et al., 2017; Pörtner et al., 2021) (Chapter 2) high 2 train +6635 AR6_WGII 2383 8 While strong mitigation efforts are fundamental to minimising future climate impacts on forests, forest management can be improved in many places in support of enhancing the resilience of tropical forests, often with significant co-benefits for carbon storage, biodiversity, food security and ecosystem services high 2 test +6636 AR6_WGII 2383 10 While there are numerous approaches to managing forests and forest landscapes sustainably, an element that appears to be critical is property rights and tenure arrangements allowing stewards of the land, including Indigenous Peoples, securing long- term access and utilisation of forest resources medium 1 train +6637 AR6_WGII 2385 14 Overall, the net loss rate has slightly decreased (−4%) since 1990 high 2 train +6638 AR6_WGII 2386 1 CCP7.2.3 Drivers of Deforestation and Forest Degradation Deforestation and forest degradation both affect carbon stocks, biodiversity loss and the provision of ecosystem services, leading to a reduction in resilience to climate change and exacerbating forest landscape vulnerability even in the absence of direct anthropogenic action high 2 train +6639 AR6_WGII 2386 2 There is also clear evidence of deforestation influencing temperatures and the hydrological cycle at local to regional scales resulting in reduced precipitation and evaporation and increased runoff relative to unaffected areas high 2 train +6640 AR6_WGII 2386 4 Therefore, besides greenhouse gas (GHG) mitigation, reducing the driving forces leading to deforestation and forest degradation is of the utmost importance for forest resilience, biodiversity protection, avoiding regional climatic changes and the provision of critical ecosystem services, and communities whose livelihoods depend on forests high 2 train +6641 AR6_WGII 2386 15 Though not as visible as direct drivers, indirect or underlying causes can greatly influence direct drivers, and must be addressed to reduce pressures on forests high 2 train +6642 AR6_WGII 2387 0 CCP7 2376Cross-Chapter Paper 7 Tropical Forests International and market forces, particularly commodity markets and, increasingly, large-scale land acquisitions are also key underlying drivers high 2 train +6643 AR6_WGII 2389 0 CCP7 2378Cross-Chapter Paper 7 Tropical Forests CCP7.3.3 Fire Risks from Climate Change in Tropical Forests Temperature rise and prolonged droughts increase the danger of fires in drained peatlands and tropical forests in Southeast Asia and the Amazon (da Silva et al., 2018; Pan et al., 2018; Sullivan Martin et al., 2020), resulting in large carbon emissions, which reached 11.3 Tg CO 2 day−1 during September–October 2015 (Huijnen et al., 2016; Yin et al., 2020) and changes in forest composition and biodiversity (Asner et al., 2000; Hoffmann et al., 2003) high 2 train +6644 AR6_WGII 2389 1 In many cases, tree mortality due to fire is poorly recorded in the literature, but the available data suggest that fire-induced mortality has increased in recent years (Figure CCP7.2) (Malhi et al., 2014; Brando et al., 2019) high 2 train +6645 AR6_WGII 2390 19 Observational data show that structurally intact old-growth tropical forests have been net sinks of atmospheric carbon in recent decades, but there is evidence that the capacity of such intact tropical forests to build up carbon stock may be limited as biomass peaked during the 1990s and has since weakened by 30% in the Amazon since the 1990s high 2 train +6646 AR6_WGII 2391 1 The IPCC Special Report on Climate Change and Land (Jia et al., 2019) and IPCC AR6 WGI (Douville et al., 2021) both describe significant climate-related changes resulting from tropical deforestation high 2 train +6647 AR6_WGII 2391 2 Deforestation generally reduces rainfall and enhances temperatures and landscape dryness; effects that increase with the scale of forest loss, whereas reforestation and afforestation generally reverses these effects high 2 train +6648 AR6_WGII 2391 5 Modelling studies estimate that large-scale tropical deforestation will contribute to average warming of the deforested areas with +0.61 ± 0.48°C and will lead to large changes in diurnal temperature ranges owing to a reduction of nocturnal cooling medium 1 train +6649 AR6_WGII 2391 6 Large-scale deforestation will also strongly decrease average regional precipitation and evapotranspiration and further delay the onset of the wet season, enhancing the chance of dry spells and intensifying dry seasons, but the magnitude of the decline depends on the scale and type of land-cover change high 2 train +6650 AR6_WGII 2391 7 Continued forest landscape drying and fragmentation in connection with deforestation may also enhance surface flow variability (Farinosi et al., 2019; Souza et al., 2019) and will aggravate the risk of forest dieback (Zemp et al., 2017), elevate forest flammability (Alencar et al., 2015) and increase fire incidence high 2 train +6651 AR6_WGII 2391 14 The dearth of data for tropical forest regions other than the Amazon makes assessments of deforestation-related changes in temperature, precipitation and streamflow difficult (high confidence), and hampers estimates of tropical forest ecosystem health, biodiversity loss and vulnerability to current and future climatic and other pressures high 2 train +6652 AR6_WGII 2391 19 Livelihoods, gender, land-use change and dependency on forest resources for food, fuel, housing and other needs have been identified as key elements of vulnerability in Indigenous Peoples and rural communities in Africa and South America high 2 train +6653 AR6_WGII 2392 3 Globalised trade relations, agricultural expansion, illegal activities and violent conflicts have been identified as important non-climatic drivers of forest degradation high 2 train +6654 AR6_WGII 2392 23 Actions to protect the extent or reduce the disturbance pressure on forest systems contribute to the capacity of these systems to respond to climate change (increasing resistance and resilience) high 2 train +6655 AR6_WGII 2393 1 Indigenous knowledge systems, embedded in social and cultural structures, are integral to climate resilience and adaptation high 2 train +6656 AR6_WGII 2393 15 Role of IK and LK and CBA for Climate Change Adaptation in Tropical Forests Local forest and Indigenous forest management systems have developed over long time periods, generating social practices and institutions that have supported livelihoods and cultures for generations high 2 train +6657 AR6_WGII 2393 16 Archaeological evidence shows that humans have manipulated tropical forests for at least 45,000 years high 2 train +6658 AR6_WGII 2393 23 Integration of IK and LK Systems, CBA and Modern Scientific Systems Several authors have highlighted the need to foster a respectful dialogue between Indigenous knowledge (IK) and local knowledge (LK) and modern science towards a holistic research model high 2 train +6659 AR6_WGII 2394 1 Forest management strategies must take into account the wider picture if they seek to be not just temporally effective (at best), but transformative and sustainable over time high 2 train +6660 AR6_WGII 2395 2 Maximising tropical forest extent has well-described benefits in mitigating CO2 emissions and in the role of forests regulating global climate high 2 test +6661 AR6_WGII 2395 7 Ensuring forested areas are large and/or interconnected including the use of specific climate refugia and climate corridors is recommended for climate adaptation high 2 train +6662 AR6_WGII 2398 18 CCP7.6 Governance of Tropical Forests for Resilience and Adaptation to Climate Change Deforestation and forest degradation in tropical forests has grown in prominence as priorities for environmental governance in the face of climate change, given the large share of forest and land use GHG emissions in the national profiles of tropical forest countries high 2 train +6663 AR6_WGII 2398 24 Policy responses towards conservation and ecosystem resilience are found to be insufficient to stem the direct and indirect drivers of nature deterioration high 2 train +6664 AR6_WGII 2398 25 For governance measures to be effective, it is necessary to alter the direct and underlying drivers that are leading to forest destruction or impeding the implementation of sustainable forest management practices and actions to restore degraded forests high 2 train +6665 AR6_WGII 2425 4 There is also stronger evidence for impacts of long-term climate change on ecosystems, including the observed widespread mortality of warm water corals, far reaching shifts in phenology in marine and terrestrial ecosystems and the expansion of tropical species into the ranges of temperate species, and boreal species moving into Arctic regions high 2 train +6666 AR6_WGII 2425 8 A negative impact of long-term climate change on crop yields has been identified in some regions (e.g., wheat yields in Europe) medium 1 train +6667 AR6_WGII 2425 12 Overall, there is extensive observational evidence that extreme ambient temperatures increase human mortality (high confidence) and that the occurrence of water- and vector-borne diseases is sensitive to weather conditions high 2 train +6668 AR6_WGII 2425 16 Climate variability and extremes are associated with increased prevalence of conflict, with more consistent evidence for low- intensity organised violence than for major armed conflict. Compared with other socioeconomic drivers, the link is relatively weak and conditional on high population size, low socioeconomic development, high political marginalisation and high agricultural dependence medium 1 test +6669 AR6_WGII 2425 17 Literature suggests a larger climate-related influence on the dynamics of conflict than on the likelihood of initial conflict outbreak low 0 test +6670 AR6_WGII 2426 16 Evidence focuses on constraints that may lead to limits at some point of the adaptation process, with less information on how limits may be related to different levels of socioeconomic or climatic change high 2 train +6671 AR6_WGII 2426 18 Beginning at 1.5°C, autonomous and evolutionary adaptation responses by terrestrial and aquatic species and ecosystems face hard limits, resulting in biodiversity decline, species extinction and loss of related livelihoods high 2 train +6672 AR6_WGII 2426 19 Beginning at 3°C, hard limits are projected for water management measures, leading to decreased water quality and availability, negative impacts on health and well-being, economic losses in water and energy-dependent sectors and potential migration of communities medium 1 train +6673 AR6_WGII 2426 20 Adaptation to address risks of heat stress, heat mortality and reduced capacities for outdoor work for humans face soft and hard limits across regions beginning at 1.5°C, and are particularly relevant for regions with warm climates high 2 train +6674 AR6_WGII 2426 22 As sea levels rise and extreme events intensify, coastal communities face soft limits due to financial, institutional and socioeconomic constraints reducing the efficacy of coastal protection and accommodation approaches and resulting in loss of life and economic damages medium 1 train +6675 AR6_WGII 2426 24 Soft and hard limits for agricultural production are related to water availability and the uptake and effectiveness of climate-resilient crops which are constrained by socioeconomic and political challenges medium 1 train +6676 AR6_WGII 2426 27 While the rate, extent and timing of climate hazards largely determine hard limits of biophysical systems, these factors appear to be less influential in determining soft limits for human systems medium 1 train +6677 AR6_WGII 2426 29 Impacts of climate change may increase financial constraints (high confidence) and contribute to soft limits to adaptation being reached medium 1 train +6678 AR6_WGII 2426 32 These key risk are represented in eight so-called Representative Key Risks (RKRs) clusters of key risks relating to low-lying coastal systems; terrestrial and ocean ecosystems; critical physical infrastructure, networks and services; living standards; human health; food security; water security; and peace and human mobility high 2 train +6679 AR6_WGII 2427 2 For most Representative Key Risks (RKRs), potentially global and systemically pervasive risks become severe in the case of high warming, combined with high exposure/vulnerability, low adaptation, or both high 2 train +6680 AR6_WGII 2427 3 Under these conditions, there would be severe and pervasive risks to critical infrastructure (high confidence) and to human health from heat-related mortality, to low-lying coastal areas, aggregate economic output, and livelihoods (all medium confidence), of armed conflict low 0 train +6681 AR6_WGII 2427 6 Under these conditions, there would be severe and pervasive risks associated with water scarcity and water-related disasters high 2 train +6682 AR6_WGII 2427 10 Tropical and polar low-lying coastal human communities are experiencing severe impacts today high 2 train +6683 AR6_WGII 2427 11 Some systems will experience severe risks before the end of the century (medium confidence), for example critical infrastructure affected by extreme events medium 1 train +6684 AR6_WGII 2427 12 Food security for millions of people, particularly low-income populations, also faces significant risks with moderate to high warming or high vulnerability, with a growing challenge by 2050 in terms of providing nutritious and affordable diets high 2 train +6685 AR6_WGII 2427 21 Priority areas for regions are indicated by the intersection of hazards, risks and challenges, where, in the near term, challenges to SDGs indicate probable systemic vulnerabilities and issues in responding to climatic hazards high 2 train +6686 AR6_WGII 2427 23 Solar radiation modification (SRM) approaches have potential to offset warming and ameliorate other climate hazards, but their potential to reduce risk or introduce novel risks to people and ecosystems is not well understood high 2 train +6687 AR6_WGII 2427 24 SRM effects on climate hazards are highly dependent on deployment scenarios, and substantial residual climate change or overcompensating change would occur at regional scales and seasonal time scales high 2 train +6688 AR6_WGII 2427 26 Large negative impacts are projected from rapid warming for a sudden and sustained termination of SRM in a high-CO 2 scenario high 2 train +6689 AR6_WGII 2427 27 SRM would not stop CO 2 from increasing in the atmosphere or reduce resulting ocean acidification under continued anthropogenic emissions high 2 train +6690 AR6_WGII 2428 3 The wide variation across disparate methodologies does not allow a robust range of damage estimates to be identified with confidence, though the spread of estimates increases with warming in all methodologies, indicating higher risk (in terms of economic costs) at higher temperatures high 2 train +6691 AR6_WGII 2428 7 Compared with AR5 and SR15, risks increase to high and very high levels at lower global warming levels for all five RFCs high 2 train +6692 AR6_WGII 2428 8 Transitions from high to very high risk emerge in all five RFCs, compared with just two RFCs in AR5 high 2 train +6693 AR6_WGII 2428 10 Risks are already (at current warming of 1.1°C) in the transition from moderate to high very high 3 train +6694 AR6_WGII 2428 11 The transition to very high risk occurs between 1.2°C and 2.0°C warming high 2 train +6695 AR6_WGII 2428 14 The transition to very high risk occurs at between 2.0°C and 3.5°C warming medium 1 train +6696 AR6_WGII 2428 18 Remaining below 2°C warming (but above 1.5°C) would imply that risk for RFC3 through 5 would be transitioning to high, and risk for RFC1 and RFC2 would be transitioning to very high high 2 train +6697 AR6_WGII 2428 19 By 2.5°C warming, RFC1 will be in very high risk high 2 train +6698 AR6_WGII 2428 21 Once such risks materialise, as is expected at very high risk levels, the impacts would persist even if global temperatures would subsequently decline to levels associated with lower levels of risk in an ‘overshooting’ scenario high 2 train +6699 AR6_WGII 2432 28 This general reduction in poverty across the world is accompanied by a decrease in vulnerability to many types of climate change impacts medium 1 train +6700 AR6_WGII 2435 20 Climate change has started to induce such tipping points, with the first examples including mass mortality in coral reef ecosystems (e.g., Donner et al., 2017; Hughes et al., 2018; Hughes et al., 2019) high 2 train +6701 AR6_WGII 2435 23 Many terrestrial ecosystems on all continents show evidence of significant structural transformation, including woody thickening and ‘greening’ in more water-limited ecosystems, with a significant role played by rising atmospheric CO 2 fertilisation in these trends high 2 train +6702 AR6_WGII 2436 8 Although anthropogenic climate forcing has increased droughts’ intensity or probability in many regions of the world medium 1 train +6703 AR6_WGII 2436 12 However, anthropogenic climate forcing has become the dominant driver of sea level rise high 2 train +6704 AR6_WGII 2436 13 In addition, anthropogenic climate forcing has increased the amount of rainfall associated with tropical cyclones high 2 train +6705 AR6_WGII 2436 20 However, meanwhile, stagnation or decline in yields is also observed in parts of the harvested areas high 2 train +6706 AR6_WGII 2437 2 Results are relatively convergent that climate change has been an important driver of the recent declines in wheat yields in Europe medium 1 train +6707 AR6_WGII 2438 19 Going beyond extreme weather events, economic production depends nonlinearly on temperature fluctuations: below a certain threshold temperature, economic production increases with temperature, whereas it decreases above a certain threshold temperature high 2 train +6708 AR6_WGII 2438 21 In addition, the empirical findings on the sensitivity of macroeconomic development to weather fluctuations and extreme weather events have been used to estimate the cumulative effect of historical warming on long-term economic development (see ‘Other societal impacts—Macroeconomic output’, Table SM16.22): anthropogenic climate change is estimated to have reduced gross domestic product (GDP) growth over the last 50 years, with substantially larger negative effects on developing countries and in some cases positive effects on colder industrialised countries low 0 train +6709 AR6_WGII 2438 23 Climate change is estimated to have substantially slowed down this trend, that is, increased inequality compared with a counterfactual no-climate- change baseline low 0 train +6710 AR6_WGII 2442 19 While many species are increasingly responding to climate change, there is limited evidence that these responses will be fully adaptive, and for many species the rate of response appears insufficient to keep pace with the rate of climate change under mid- and high-range emissions scenarios medium 1 train +6711 AR6_WGII 2445 11 Some but not all adaptation-related responses are engaging vulnerable populations in planning or implementation high 2 train +6712 AR6_WGII 2445 12 Consideration of vulnerable populations is most frequently focused on low-income populations and women through the inclusion of informal or formal institutions or representatives in adaptation planning, or through targeted adaptations to reduce risk in these populations high 2 train +6713 AR6_WGII 2445 13 Consideration of vulnerable groups in adaptation responses is more frequently reported in the Global South medium 1 train +6714 AR6_WGII 2445 14 Engagement in adaptation planning of vulnerable elderly, migrants, and ethnic minorities remains low across all global regions medium 1 train +6715 AR6_WGII 2445 15 There is negligible literature on consideration of disabled peoples in planning and implementation of adaptation-related responses medium 1 train +6716 AR6_WGII 2446 15 Global adaptation is predominantly slow, siloed and incremental with little evidence of transformative adaptation high 2 train +6717 AR6_WGII 2446 18 Based on the literature, the overall transformative nature of adaptation across most global regions and sectors is low high 2 train +6718 AR6_WGII 2446 19 Documented adaptations tend to involve minor modifications to usual practices taken to address extreme weather conditions high 2 train +6719 AR6_WGII 2446 20 For example, changing crop variety or timing of crop planting to address floods or droughts, new types of irrigation, pursuing supplementary livelihoods, and home elevations are widely reported but typically do not reflect radical or novel shifts in practice or values and are therefore considered low depth high 2 train +6720 AR6_WGII 2446 21 Adaptations documented in the literature are also frequently focused on a single sector or small geographic area high 2 train +6721 AR6_WGII 2447 2 The speed of adaptation is rarely noted explicitly, but the average speed documented in the literature is slow medium 1 train +6722 AR6_WGII 2447 3 Adaptation efforts frequently encounter either soft or hard limits (see Section 16.4), but there is limited evidence to suggest these limits are being challenged or overcome medium 1 train +6723 AR6_WGII 2447 4 Few documented responses are simultaneously widespread, rapid and novel high 2 train +6724 AR6_WGII 2447 6 In general, adaptations that are broad in scope tend to be slow medium 1 train +6725 AR6_WGII 2448 38 Evidence on realised co-benefits of implemented adaptation responses with other priorities in the SDGs is emerging among the areas of poverty reduction, food security, health and well-being, terrestrial and freshwater ecosystem services, sustainable cities and communities, energy security, work and economic growth, and mitigation (Table 16.2) high 2 train +6726 AR6_WGII 2451 7 Van Bavel, for example, found that the involvement of local and diverse knowledge can improve the detection (medium confidence) and attribution (medium confidence) of health impacts, and improve the action high 2 train +6727 AR6_WGII 2452 13 Cross-Chapter Box INTEREG | Inter-regional Flows of Risks and Responses to Risk Authors: Birgit Bednar-Friedl (Austria, Chapter 13), Christopher Trisos (South Africa, Chapter 9), Laura Astigarraga (Uruguay, Chapter 12), Magnus Benzie (Sweden/UK), Aditi Mukherji (India, Chapter 4), Maarten Van Aalst (the Netherlands, Chapter 16) Introduction Our world today is characterised by a high degree of interconnectedness and globalisation which establish pathways for the transmission of climate-related risks across sectors and borders high 2 train +6728 AR6_WGII 2452 22 When climatic events like flooding or heat affect the location of these extraction and production activities, economies are not only disrupted locally but also across borders and in distant countries high 2 train +6729 AR6_WGII 2452 23 For many industrialised countries like the UK, Japan, the USA and the European Union, there is increasing evidence that the trade impacts of climate change are significant and can have substantial domestic impacts medium 1 train +6730 AR6_WGII 2454 3 The export of major food crops like wheat, maize and soybeans from many of the world’s water-scarce area—the Middle East, North Africa, parts of South Asia, North China Plains, southwest USA, Australia—to relatively water-abundant parts of the world carries a high virtual water content (the net volume of water embedded in trade) high 2 train +6731 AR6_WGII 2454 5 Climate change is projected to exacerbate risk and add new vulnerabilities for risk transmission medium 1 train +6732 AR6_WGII 2458 14 However, much of the available evidence focuses on constraints that may lead to limits at some point with little detailed information on how limits may be related to different levels of socioeconomic or environmental change high 2 train +6733 AR6_WGII 2458 16 Small islands and Central and South America show most evidence of constraints being linked to adaptation limits across sectors, while ocean and coastal ecosystems and health, well-being and communities show most evidence of constraints being linked to limits across regions medium 1 train +6734 AR6_WGII 2458 18 Evidence on limits to adaptation is largely focused on terrestrial and aquatic species and ecosystems, coastal communities, water security, agricultural production, and human health and heat high 2 train +6735 AR6_WGII 2458 19 Beginning at 1.5°C, autonomous and evolutionary adaptation responses by terrestrial and aquatic species and ecosystems face hard limits, resulting in biodiversity decline, species extinction and loss of related livelihoods high 2 train +6736 AR6_WGII 2458 20 Interventionist adaptation strategies to reduce risks for species and ecosystems face soft limits due to governance, financial and knowledge constraints medium 1 train +6737 AR6_WGII 2458 21 As sea levels rise and extreme events intensify, coastal communities face soft limits due to financial, institutional and socioeconomic constraints reducing the efficacy of coastal protection and accommodation approaches and resulting in loss of life and economic damages medium 1 train +6738 AR6_WGII 2458 22 Hard limits for coastal communities reliant on nature-based coastal protection will be experienced beginning at 1.5°C medium 1 train +6739 AR6_WGII 2458 23 Beginning at 3°C, hard limits are projected for water management measures, leading to decreased water quality and availability, negative impacts on health and well-being, economic losses in water and energy dependent sectors and potential migration of communities medium 1 train +6740 AR6_WGII 2458 24 Soft and hard limits for agricultural production are related to water availability and the uptake and effectiveness of climate-resilient crops, which is constrained by socioeconomic and political challenges medium 1 train +6741 AR6_WGII 2458 25 Adaptation measures to address risks of heat stress, heat mortality and reduced capacities for outdoor work for humans face soft and hard limits across regions beginning at 1.5°C and are particularly relevant for regions with warm climates high 2 train +6742 AR6_WGII 2459 1 While individual constraints may appear straightforward to address, the combination of constraints interacting with each other leads to soft limits that are difficult to overcome high 2 train +6743 AR6_WGII 2459 4 Impacts of climate change may increase financial constraints (high confidence) and contribute to soft limits to adaptation being reached medium 1 train +6744 AR6_WGII 2459 6 At the national level, negative macroeconomic responses to climate change may limit the availability of financial resources, impede access to financial markets and stunt economic growth high 2 train +6745 AR6_WGII 2461 7 Sectoral studies indicate that climate impacts will result in higher levels of losses and damages and decreases in income, thereby increasing financial constraints medium 1 train +6746 AR6_WGII 2461 13 At the household or community level, climate impacts may increase financial constraints high 2 train +6747 AR6_WGII 2466 7 In some Arctic communities and in communities reliant on warm- water coral reefs, even 1.5–2°C warming will lead to severe risks from loss of ecosystem services (Section 3.4.2.2; Cross-Chapter Paper 6) high 2 train +6748 AR6_WGII 2466 17 Warm-water coral reefs are at risk of widespread loss of structural complexity and reef accretion by 2050 under 1.5°C global warming (Section 3.4.2.1) high 2 train +6749 AR6_WGII 2468 3 Much evidence shows increased risk of the loss of 10% or more of terrestrial biodiversity with increasing anthropogenic climate change (Urban, 2015; Smith et al., 2018) medium 1 train +6750 AR6_WGII 2468 5 Mass population-level mortality (>50% of individuals or colonies killed) and resulting abrupt ecological changes can be caused by simple or compound climate extreme events, such as exceedance of upper thermal limits by vulnerable terrestrial species (Fey et al., 2015), who also note reduced mass mortality trends due to extreme low thermal events; marine heatwaves that can cause mortality, enhance invasive alien species establishment, and damage coastal ecological communities and small-scale fisheries high 2 train +6751 AR6_WGII 2468 7 Freshwater ecosystems and their biodiversity are at high risk of biodiversity loss and turnover due to climate change (precipitation change and warming, including warming of water bodies), due to high sensitivity of processes and life histories to thermal conditions and water quality (Chapter 2) high 2 test +6752 AR6_WGII 2468 8 In marine systems, heatwaves cause damages in coastal systems, including extensive coral bleaching and mortality (very high confidence) (Section 3.4.2.1), mass mortality of invertebrate species (low to high confidence, depending on system) (Sections 3.4.2.2, Section 3.4.2.5, Section 3.4.4.1), and abrupt mortality of kelp-forest (high confidence) (Section 3.4.2.3) and seagrass-meadow habitat high 2 train +6753 AR6_WGII 2468 9 The biodiversity of polar seas shows strong impacts of climate change on phenological timing of plankton activity, Arctic fish species range contractions and species community change (Table SM16.22) high 2 train +6754 AR6_WGII 2468 10 Extreme weather events and storm surges exacerbated by climate change have severe and sudden adverse impacts on coastal systems, including loss of seagrass meadows and mangrove forests high 2 train +6755 AR6_WGII 2468 15 Landscape- and larger-scale shifts in ecosystem structure and function (order-of-magnitude increases or abrupt decreases in cover and/or biomass of novel growth forms or functional types) are occurring in non- equilibrium ecosystems (systems which exist in multiple states, often disturbance-controlled) in response to changing disturbance regime, climate and rising CO 2 (high confidence) Woody plant encroachment has been occurring in multiple ecosystems, including subtropical and tropical fire driven grassland and savanna systems, upland grassland systems, arid grasslands and shrublands (high confidence), leading to large-scale biodiversity changes, albedo changes, and impacts on water delivery, grazing services and human livelihoods medium 1 train +6756 AR6_WGII 2468 17 In tropical forests, repeated droughts and recurrence of large-scale anthropogenic fires increase forest degradation, loss of biodiversity and ecosystem functioning high 2 train +6757 AR6_WGII 2468 20 The loss of a substantial fraction of biodiversity globally, abrupt impacts such as significant local biodiversity loss and mass population mortality events, and ecological disruption due to novel species interactions have been observed or are projected at global warming levels below 2°C (Chapter 2 Table SM2.5, Cross Chapter Box: EXTREMES in Chapter 2, Section 2.4.4.3.1, Section 2.4.2.3.3) medium 1 train +6758 AR6_WGII 2468 21 Simple and compound impacts of extreme climate events are already causing significant losses and damages in vulnerable ecosystems, including through the facilitation of important global change drivers of ecological disruption and homogenisation like invasive species high 2 train +6759 AR6_WGII 2468 24 Biodiversity conservation efforts may be hampered due to climate change impacts on the effectiveness of protected areas, with high sensitivity of effectiveness to forcing scenario medium 1 train +6760 AR6_WGII 2468 25 In addition, climate-related risks to ecosystems pose challenges to ecosystem-based adaptation responses (‘nature-based solutions’) (Section 2.1.3) medium 1 train +6761 AR6_WGII 2469 3 The major hazards driving such risks are acute extreme events such as cyclones, floods, droughts or fires high 2 train +6762 AR6_WGII 2469 11 Transport and energy infrastructure in coasts and polar systems and along rivers are projected to face a particularly steep rise in risk, resulting in severe risk even under medium warming high 2 train +6763 AR6_WGII 2469 12 Risk in relation to the increasing intensity and frequency of extreme events might become severe before the middle of the century medium 1 train +6764 AR6_WGII 2470 3 Already today, climate-related impacts on transport and energy infrastructure reach far beyond the direct impacts on physical infrastructure, triggering indirect impacts on, for example, health and income medium 1 train +6765 AR6_WGII 2470 19 Annual economic output losses in developing countries could exceed the worst country-level losses during historical economic recessions medium 1 train +6766 AR6_WGII 2471 3 More widespread severe risks would occur at high levels of warming (with high exposure/ vulnerability and low adaptation) where there is additional potential for one or more social or ecological tipping points to be triggered (Cai et al., 2015; Cai et al., 2016b; Kopp et al., 2016; Steffen et al., 2018; Lenton et al., 2019), and for severe impacts on livelihoods to cascade from relatively more climate-sensitive to relatively less climate-sensitive sectors and regions medium 1 train +6767 AR6_WGII 2472 1 Without effective adaptation measures, regions with high dependence on climate-sensitive livelihoods—particularly agriculture and fisheries in the tropics and coastal regions—would be severely impacted even at low levels of warming high 2 test +6768 AR6_WGII 2472 5 Within populations, the poor, women, children, the elderly and Indigenous populations are especially vulnerable due to a combination of factors, including gendered divisions of paid and/or unpaid labour, as well as barriers in access to information, skills, services or resources (Bose, 2017; Thomas et al., 2019b; Anderson and Singh, 2020; Adzawla and Baumüller, 2021) high 2 train +6769 AR6_WGII 2472 8 Climate change also could increase income inequality between countries high 2 train +6770 AR6_WGII 2473 23 Climate change will pose severe risks in terms of increasing the number of undernourished people, affecting tens to hundreds of million people under High vulnerability and High warming, particularly among low- income populations in developing countries high 2 train +6771 AR6_WGII 2473 24 Extreme weather events will increase risks of undernutrition even on a regional scale, via spikes in food price and reduced income high 2 train +6772 AR6_WGII 2474 5 Climate change risks of micronutrient deficiency will become severe in high-vulnerability development pathways and in the absence of societal adaptation, leading to hundreds of millions of additional people lacking key nutrients for atmospheric CO 2 levels above 500 ppm high 2 train +6773 AR6_WGII 2476 10 Regular human mobility will continue regardless of climate change, but mobility-related risks will increase with warming, notably in densely populated hazard-prone regions, in small islands and low-lying coastal zones, and among populations with limited coping capacity (RKR-A; Section CCP2.2.2; Chapter 7) high 2 train +6774 AR6_WGII 2476 14 Although climate-driven human mobility generally does not increase risks to peace (medium confidence), armed conflict is a major driver of forced displacement high 2 train +6775 AR6_WGII 2476 16 Expert elicitation estimates that 4°C warming above pre-industrial levels will have severe and widespread effects on armed conflict with 26% probability, assuming no change from present levels in non- climatic drivers (Mach et al., 2019). That judgement refers to impacts that exceed the threshold for severity considered here, suggesting that global warming of 4°C would produce severe risks to peace under present societal conditions low 0 test +6776 AR6_WGII 2476 19 A larger empirical literature offers indirect evidence that climate change may produce severe risks to peace within this century by demonstrating how climate variability and extremes affect contemporary conflict dynamics, especially in contexts marked by low economic development, high economic dependence on climate-sensitive activities, high or increasing social marginalisation, and fragile governance medium 1 train +6777 AR6_WGII 2476 27 There is high agreement that even moderate levels of future SLR will severely amplify involuntary migration and displacement in small islands and densely populated low-lying coastal areas in the absence of appropriate adaptive responses high 2 train +6778 AR6_WGII 2479 20 A wide range of climate change impacts depend strongly on development pathway high 2 train +6779 AR6_WGII 2479 24 Studies are increasingly going beyond exposure to incorporate future vulnerability, finding that it is often the dominant determinant of risk high 2 train +6780 AR6_WGII 2481 2 Future risk in these areas is heavily driven by climate change but also greatly depends on past, current and future socioeconomic changes which influence future trends in exposure, vulnerability and adaptive capacity of natural and human systems high 2 train +6781 AR6_WGII 2481 3 From a risk perspective, trends over the past decades have been unfavourable for many deltas, as most of them have experienced a simultaneous intensification of hazards, rise in exposure and stagnation or only limited reduction in vulnerability, particularly in low-income countries high 2 train +6782 AR6_WGII 2481 14 The combined effects of local subsidence and GMSL rise result in a significant increase in the potential for inundation of low-lying deltas across all RCPs, with some variation according to regional sea level change rates, without significant further adaptation measures very high 3 train +6783 AR6_WGII 2482 2 However, case study research from individual deltas suggests that delta populations, particularly those with agriculture-based livelihoods, have seen more limited vulnerability reduction due in particular to the impacts of environmental hazards, stress and disasters high 2 train +6784 AR6_WGII 2488 5 Cooling caused by SRM would increase the global land and ocean CO 2 sinks (medium confidence), but this would not stop CO 2 from increasing in the atmosphere or affect the resulting ocean acidification under continued anthropogenic emissions high 2 train +6785 AR6_WGII 2488 26 A gradual phase-out of SRM combined with emission reduction and CDR could avoid these termination effects medium 1 train +6786 AR6_WGII 2490 3 Aligning climate change adaptation to the SDGs could bring potential co-benefits and increased efficiency in funding, and reduce the gap between adaptation planning and implementation very high 3 train +6787 AR6_WGII 2490 14 From Figure 16.12, aside from SDG13 (climate action), the strongest connections and risk challenges are with zero hunger (SDG2), sustainable cities and communities (SDG11), life below water (SDG14), decent work and economic growth (SDG8), no poverty (SDG1), clean water and sanitation (SDG6) and good health and well-being (SDG3) high 2 train +6788 AR6_WGII 2490 15 Other SDGs have strong linkages with specific RKRs, for example, terrestrial and marine ecosystems with life on land (SDG15); infrastructure (RKR-C) with industry, innovation and infrastructure (SDG9) and affordable and clean energy (SDG7); living standards (RKR-D) with gender equality (SDG5); and peace and human mobility (RKR-H) with peace, justice and strong institutions (SDG 16) high 2 train +6789 AR6_WGII 2490 17 The greatest linkages and effects on the SDGs will be due to risks to water (RKR-G), living standards (RKR-D), coastal socio-ecological systems (RKR-A) and peace and human mobility (RKR-H) high 2 train +6790 AR6_WGII 2490 18 In particular, coastal socio-ecological systems (RKR-A), living standards (RKR-D), food security (RKR-F), water security (RKR-G) and peace and human mobility (RKR-H), have strong linkages with SDG 2 (zero hunger), for which there are significant to major challenges for all regions high 2 train +6791 AR6_WGII 2490 19 Almost all the RKRs are strongly linked to SDGs 8 (decent work and economic growth) and 11 (sustainable cities and communities) high 2 train +6792 AR6_WGII 2490 20 All regions also face major to significant challenges affecting SDGs 14 (life below water) and 15 (life on land), which relate to terrestrial and ocean ecosystems (RKR-B) high 2 train +6793 AR6_WGII 2492 1 They also will be indirectly linked to, and thus affect, the SDGs overall, due to the interactions between the key risks (Section 16.5) and between the SDGs themselves very high 3 train +6794 AR6_WGII 2492 5 In the near term, the strength of connection between the RKRs and the SDGs, with respect to existing SDG challenges, indicate probable systemic vulnerabilities and issues in responding to climatic hazards (UN-IATFFD, 2019; Leal Filho et al., 2020; Weaver et al., 2020; Tiedemann et al., 2021) high 2 train +6795 AR6_WGII 2495 17 Species extinction is irreversible, and Chapter 2 assesses that, at ~1.6°C, >10% of species are projected to become endangered as compared with >20% at ~2.1°C (median), representing high and very high biodiversity risk, respectively medium 1 train +6796 AR6_WGII 2498 9 Based on observed and modelled impacts to unique and threatened systems, including in particular coral reefs, sea-ice-dependent systems and biodiversity hotspots, AR6 assesses that the transition to high risks for RFC1 have already occurred at a median level of 0.9°C, with a lower bound at 0.7°C and an upper bound at the present-day level of global warming of 1.2°C (WMO, 2020) very high 3 train +6797 AR6_WGII 2498 21 Small island communities are confronted by socio-ecological limits to adaptation well before 2100, especially those reliant on coral reef systems for their livelihoods, even for a low-emissions pathway (Chapter 3) high 2 train +6798 AR6_WGII 2498 22 At warming levels beyond 1.5°C, the potential to reach biophysical limits to adaptation due to limited water resources are reported for small islands (medium confidence) and unique systems dependent on glaciers and snowmelt (Chapter 4) medium 1 train +6799 AR6_WGII 2498 25 In AR6, risks are considered to start to transition from high to very high risks above 1.2°C warming (present day, WMO, 2020), with a median value of 1.5°C, owing in particular to the observation of a present-day onset of ecosystem degradation in coral reefs, which are projected in the SR15 report ‘to decline by a further 70–90% at 1.5°C very high 3 train +6800 AR6_WGII 2499 2 Warming of 1.5°C is also assessed (Chapter 3) to reduce the habitability of small islands, due to the combined impacts of several key risks high 2 train +6801 AR6_WGII 2499 32 This includes increases in hot extremes over most land regions (virtually certain), increases in heavy precipitation at the global scale and over most regions with sufficient observations (high confidence), and increases in agricultural and ecological droughts in some regions medium 1 train +6802 AR6_WGII 2500 2 Furthermore, food security and livelihoods are being affected by short-term food shortages caused by climate extremes (Section 5.12.1; Chapter 16, Food Security RKR) which have affected the productivity of all agricultural and fishery sectors high 2 train +6803 AR6_WGII 2500 4 Droughts, floods and marine heatwaves contribute to reduced food availability and increased food prices, threatening food security, nutrition and livelihoods of millions high 2 test +6804 AR6_WGII 2500 6 Droughts induced by the 2015–2016 El Niño, partially attributable to human influences (medium confidence), caused acute food insecurity in various regions, including eastern and southern Africa and the dry corridor of Central America high 2 train +6805 AR6_WGII 2500 9 In the northeast Pacific, a recent 5-year warm period impacted the migration, distribution and abundance of key fish resources high 2 train +6806 AR6_WGII 2500 15 In particular, heavy precipitation and associated flooding are projected to intensify and be more frequent in most regions in Africa and Asia (high confidence), North America (medium to high confidence depending on the region) and Europe medium 1 train +6807 AR6_WGII 2500 16 Also, more frequent and/or severe agricultural and ecological droughts are projected in a few regions in all continents except Asia, compared with 1850–1900 (medium confidence); increases in meteorological droughts are also projected in a few regions medium 1 train +6808 AR6_WGII 2503 30 In SR15 Section 3.5.2.4 (Hoegh-Guldberg et al., 2018b), economic literature on potential socioeconomic threshold events and empirical studies of global economic damages, combined with new evidence on biome shifts, extinction risk, species range loss (especially noting the integral role of insects in ecosystem function) and ecosystem degradation, were assessed, and the upper bound of the transition to moderate risk was lowered to 1.5°C warming above pre-industrial levels, and the transition from moderate and high risk was lowered to between 1.5°C and 2.5°C medium 1 train +6809 AR6_WGII 2504 2 Cross-Working Group Box ECONOMIC in this chapter includes a more thorough discussion of advancements and limitations of global economic impact estimates and methodologies, finding significant variation in estimates that increases with warming, indicating higher risk in terms of economic costs at higher temperatures high 2 train +6810 AR6_WGII 2504 24 Higher atmospheric concentrations of carbon dioxide reduce the nutritional quality of wheat, rice and other major crops, potentially affecting millions of people at a doubling of carbon dioxide relative to pre-industrial very high 3 train +6811 AR6_WGII 2504 27 Section 16.5.2.3.2 (RKR-B, risks to terrestrial and marine ecosystems) finds that substantial biodiversity loss globally, abrupt local ecosystem mortality impacts, and ecological species disruption are all projected at global warming levels below 3°C, with insular systems and biodiversity hotspots at risk below 2°C medium 1 train +6812 AR6_WGII 2505 9 Under a 4°C warming scenario, models project global annual damages associated with SLR of $31,000 billion yr–1 in 2100 (Brown et al., 2021) In terms of global economic impact, while an emerging economic literature is addressing many gaps and critiques of previous damage estimates for high warming (e.g., Jensen and Traeger, 2014; Burke et al., 2015; Lontzek et al., 2015; Moore and Diaz, 2015; Lemoine and Traeger, 2016; Moore et al., 2017a; Cai and Lontzek; Takakura et al., 2019, discussed further in Cross-Working Group Box ECONOMIC; Carleton et al., 2020; Méjean et al., 2020; Rode et al., 2021), there remains wide variation across disparate methodologies, though the spread of estimates increases with warming in all methodologies, indicating higher risk in terms of economic costs at higher temperatures high 2 train +6813 AR6_WGII 2505 10 Risks to aggregate economic output would become severe at the global scale at high warming (~4.4°C) and minimal adaptation medium 1 test +6814 AR6_WGII 2505 12 Chapter 4 finds that, at 4°C, 4 billion people are projected to be exposed to physical water scarcity medium 1 train +6815 AR6_WGII 2509 6 The wide range of estimates, and the lack of comparability between methodologies, does not allow for identification of a robust range of estimates with confidence high 2 train +6816 AR6_WGII 2509 7 Evaluating and reconciling differences in methodologies is a research priority for facilitating use of the different lines of evidence high 2 train +6817 AR6_WGII 2510 13 Chapter 2 has assessed ecosystem carbon loss from tipping points in tropical forest and loss of Arctic permafrost, and finds a transition to moderate risk over the range 0.6–0.9°C medium 1 train +6818 AR6_WGII 2510 18 Technical issues with past and current modelling (e.g., Pezzey, 2019; Pindyck, 2019; EPRI, 2021) and the challenge of comparability across methodologies imply that many estimates are not robust high 2 train +6819 AR6_WGII 2510 20 Better representation of uncertainty in methods can improve robustness, while detailed methodology assessment and comparison will help define the relative biases of methods high 2 train +6820 AR6_WGII 2511 9 Therefore, we assess an increasing link between MISI, WAIS collapse and Antarctic mass loss, for increasing temperature levels high 2 train +6821 AR6_WGII 2511 19 AR6 assesses that it is virtually certain that sea level was higher than today at that time, likely by 5–10 m medium 1 train +6822 AR6_WGII 2512 5 Furthermore, a transition from high to very high risk is provided for the first time in this AR6 assessment, between 2.0°C and 3.5°C warming medium 1 train +6823 AR6_WGII 2512 10 In summary, risks to unique and threatened systems (RFC1) are higher at recent and projected levels of warming than assessed previously (very high confidence); risks associated with extreme weather events (RFC2) are assessed comparably to AR5 and SR15 at recent and low levels of warming, but notably much higher at projected warming above 1.8°C (medium confidence); risks associated with distribution of impacts (RFC3) and global aggregate impacts (RFC4) are similar to SR15 and higher than AR5 above 2°C (medium confidence); and those associated with large-scale singular events (RFC5) are similar to SR15 and higher at both recent and projected warming than AR5 medium 1 train +6824 AR6_WGII 2512 11 Limiting global warming to 1.5°C would ensure risk levels remain moderate for RFC3, RFC4 and RFC5 (medium confidence), but risk for RFC2 would have transitioned to a high risk at 1.5°C and RFC1 would be well into the transition to very high risk high 2 train +6825 AR6_WGII 2512 12 Remaining below 2°C warming (but above 1.5°C) would imply that risk for RFC3 through 5 would be transitioning to high, and risk for RFC1 and RFC2 would be transitioning to very high high 2 train +6826 AR6_WGII 2512 13 By 2.5°C warming, RFC1 will be in very high risk high 2 train +6827 AR6_WGII 2512 19 However, in some circumstances, expanded global adaptation could slow some of these transitions low 0 train +6828 AR6_WGII 2553 4 Many options are widely applicable and could be scaled up to reduce vulnerability or exposure for the majority of the world’s population and the ecosystems they depend on high 2 train +6829 AR6_WGII 2553 5 These include nature restoration (high confidence), changing diets and reducing food waste (high confidence), infrastructure retrofitting (high confidence), building codes (medium confidence), disaster early warning (high confidence) and cooperative governance medium 1 train +6830 AR6_WGII 2553 6 The portfolio of adaptation options that could be successfully implemented varies across locations, with resource-limited and conflict-affected contexts bearing large amounts of residual risk high 2 train +6831 AR6_WGII 2553 7 The majority of climate risk management and adaptation currently being planned and implemented is incremental high 2 train +6832 AR6_WGII 2553 8 Transformational adaptation will become increasingly necessary at higher global warming levels (medium confidence) but can be associated with significant and inequitable trade-offs medium 1 train +6833 AR6_WGII 2553 9 Adaptations with some of the highest transformative potential include migration (high confidence), spatial planning (medium confidence), governance cooperation (medium confidence), universal access to health care (medium confidence) and changing food systems medium 1 train +6834 AR6_WGII 2553 10 Options that tend to modify existing systems incrementally include early-warning systems (high confidence), insurance (medium confidence) and improved water use efficiency high 2 train +6835 AR6_WGII 2553 11 Governance, especially when inclusive and context sensitive, is an important enabling condition for climate risk management and adaptation very high 3 train +6836 AR6_WGII 2553 14 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.arrangements of public, private and community actors, is being increasingly recognised as important across many decision- making settings high 2 train +6837 AR6_WGII 2553 15 Public governance leadership has the largest role for social safety nets, spatial planning and building codes high 2 train +6838 AR6_WGII 2553 16 Private sector governance is important for insurance and for minimising the stressors that can negatively impact ecosystems and their functions, especially in the absence of public regulations or enforcement medium 1 train +6839 AR6_WGII 2553 17 Communities and individuals play the largest role in governance of adaptations to farming and fishery practices and ecosystem-based adaptations medium 1 train +6840 AR6_WGII 2553 18 Informal or individual-led decision-making is more common in food security and livelihood-related adaptations, such as changes to diets, livelihood diversification and seasonal migration high 2 test +6841 AR6_WGII 2553 19 People who have experienced climate shocks are more likely to take on informal adaptation measures, and in places where people are more exposed to extreme events, autonomous adaptation is more common high 2 train +6842 AR6_WGII 2553 20 National and international legal and policy frameworks and instruments support the planning and implementation of adaptation and climate risk management across scales, especially when combined with guidelines for action medium 1 train +6843 AR6_WGII 2553 21 Nationally Determined Contributions (NDCs) have been drivers of national adaptation planning, with cascading effects on sectors and sub-national action, especially in developing countries high 2 train +6844 AR6_WGII 2553 22 Nearly all developing countries (particularly Small Island Developing States [SIDS]) that included an adaptation component in their NDCs consider adaptation the most urgent aspect of their national climate change response high 2 train +6845 AR6_WGII 2553 23 A steady increase in national and sub-national laws, policies and regulations that mandate reporting and risk disclosure has promoted adaptation response across public agencies, private firms and community organisations high 2 train +6846 AR6_WGII 2553 24 Greater adaptation is present where national climate laws and policies require adaptation action from lower levels of government and include guidelines on how to do so medium 1 train +6847 AR6_WGII 2553 25 Recognition of the critical role of financing for adaptation and resilience as an important enabler for climate risk management has strengthened high 2 train +6848 AR6_WGII 2553 26 Yet, since the Intergovernmental Panel on Climate Change (IPCC) 5th Assessment Report (AR5), the gap between the estimated costs of adaptation and the documented (tracked) finance allocated to adaptation has widened high 2 train +6849 AR6_WGII 2553 27 Estimated global and regional costs of adaptation vary widely due to differences in assumptions, methods and data; the majority of more recent estimates are higher than the figures presented in AR5 high 2 train +6850 AR6_WGII 2553 28 A high proportion of developing country NDC adaptation contributions are conditional on external financial support, emphasising the crucial role of international finance to achieving adaptation efforts commensurate with climate risks high 2 train +6851 AR6_WGII 2554 2 Private sector financing for adaptation has been increasingly promoted as a response to realised adaptation finance needs high 2 train +6852 AR6_WGII 2554 3 However, private sector financing of adaptation has been limited, especially in developing countries high 2 train +6853 AR6_WGII 2554 4 Tracked private sector finance for climate change action has grown substantially since 2015, but the proportion directed towards adaptation has remained small high 2 train +6854 AR6_WGII 2554 5 A key challenge for private sector financing of adaptation is demonstrating financial return on investment, as many benefits of adaptation arise as avoided damages or public goods, rather than direct revenue streams medium 1 train +6855 AR6_WGII 2554 6 Leveraging private finance in developing countries is often more difficult because of risk (perceived and real) to investors, reducing the pool of potential investors and/or raising the cost (interest) of investment medium 1 train +6856 AR6_WGII 2554 7 Information and knowledge on climate risk and adaptation options, derived from different knowledge systems, can support risk management and adaptation decisions high 2 train +6857 AR6_WGII 2554 8 Processes, such as co-production, that link scientific, Indigenous, local, practitioner and other forms of knowledge can make climate risk management processes and outcomes more effective and sustainable high 2 train +6858 AR6_WGII 2554 9 Climate services that provide reliable, relevant and usable climate information for the short or long term are increasingly being produced and used in climate risk management high 2 train +6859 AR6_WGII 2554 10 In many regions and sectors, the utility of climate services is strengthened by sustained engagement between stakeholders and experts and by co-production medium 1 train +6860 AR6_WGII 2554 12 Catalysing conditions and windows of opportunity can drive shifts in motivation and adaptation effort, stimulating more rapid uptake of existing and new adaptation options medium 1 train +6861 AR6_WGII 2554 14 Disaster events or shocks such as wildfires, tropical cyclones, heatwaves or coral bleaching have catalysing characteristics high 2 train +6862 AR6_WGII 2554 15 Additional types of catalysing conditions include climate litigation and the presence of individuals and organisations that act as policy and decision innovators, including government and business innovators in cities (medium confidence), stimulating action within and beyond their immediate contexts medium 1 train +6863 AR6_WGII 2554 16 Litigation on failure of government and business to adapt is becoming more frequent and is expected to increase as climate impact attribution science matures further (high confidence) {Cross-Chapter Box LOSS in this Chapter; 17.4.5.2, 17.4.5.3}.Urgency can stimulate prompt climate risk management high 2 train +6864 AR6_WGII 2554 17 A moderate level of urgency contributes to enhanced climate action, while both high and low levels of urgency can impede response high 2 train +6865 AR6_WGII 2554 19 As conditions approach a crisis state, however, urgency can weaken decision-making rather than support it medium 1 train +6866 AR6_WGII 2554 20 Decision support tools and decision-analytic methods are available and are being applied for managing climate risks in varied contexts, including where deep uncertainty is present high 2 train +6867 AR6_WGII 2554 21 These tools and methods have been shown to support deliberative processes where stakeholders jointly consider factors such as the rate and magnitude of change and their uncertainties, associated impacts and timescales of adaptation needed along multiple pathways and scenarios of future risks high 2 train +6868 AR6_WGII 2554 22 However, comparative evidence on the relative utility of different analytical methods in their use by decision makers for managing climate risks is an important gap medium 1 train +6869 AR6_WGII 2554 23 Nevertheless, robust decision-making, using pathway analyses to determine ‘no regrets’ options among trade-offs, has been shown to be a useful starting point under deep uncertainty medium 1 train +6870 AR6_WGII 2554 24 Methods for analysing options differ across geo- political scales, with modelling studies being a particularly prominent method across scales from community and urban to regional and national high 2 train +6871 AR6_WGII 2554 25 Successful adaptation and maladaptation form the opposite poles of a continuum medium 1 train +6872 AR6_WGII 2554 26 The evaluation of an adaptation option and its location on this continuum are context specific and vary across time, place and evaluation perspectives high 2 train +6873 AR6_WGII 2554 27 Despite knowledge gaps, adaptation options can be assessed according to several criteria, such as benefits to humans, benefits to ecosystem services, benefits to equity (marginalised ethnic groups, gender, low-income populations), transformational potential and contribution to greenhouse gas emission reduction medium 1 train +6874 AR6_WGII 2554 28 These factors can aid evaluation of co-benefits and trade- offs within and between adaptation responses (high confidence) facilitating successful adaptation and reducing the likelihood of maladaptation medium 1 train +6875 AR6_WGII 2554 29 Adaptation options across a range of climate risk settings (Representative Key Risks) have potential for some degree of maladaptation alongside varied potential for success very high 3 train +6876 AR6_WGII 2554 30 Maladaptation can result from unaccounted trade- offs with low-income groups and the transformational potential of adaptation medium 1 train +6877 AR6_WGII 2554 31 Success is greatest when adaptation enhances gender equity (medium confidence) and supports ecosystem function and services medium 1 train +6878 AR6_WGII 2554 33 Examples of options with high potential for successful adaptation are nature restoration (medium confidence), social safety nets (medium confidence) and adaptations relating to changes of diets and reducing food waste medium 1 train +6879 AR6_WGII 2555 1 M&E application has increased since AR5 at the local, project and national level, but is still at an early stage in most countries (high confidence) and underutilised as a way to assess adaptation outcomes at longer time frames high 2 train +6880 AR6_WGII 2555 2 About one-third of countries have undertaken steps to develop national adaptation M&E systems, but fewer than half of these are reporting on implementation medium 1 train +6881 AR6_WGII 2555 3 M&E, as well as tracking global progress on adaptation, are confronted with a number of challenges high 2 train +6882 AR6_WGII 2555 5 Understanding of residual impacts and risks in vulnerable regions and implications for Loss and Damage (L&D) has become increasingly relevant as the limits to adaptation are projected to be reached in natural and human systems high 2 train +6883 AR6_WGII 2555 8 Yet, the policy space and concrete remit for L&D has remained vague, which renders policy formulation complex high 2 train +6884 AR6_WGII 2555 9 Effective management of climate risks is dependent on systematically integrating adaptations across interacting climate risks, ensuring that measures of success include factors important to climate resilient development, and accounting for the dynamic nature of climate risks over time very high 3 train +6885 AR6_WGII 2555 11 Climate risks can emerge at different rates and time horizons, and the interactions between risks vary from region to region very high 3 test +6886 AR6_WGII 2555 12 The need to manage these risks in an integrated manner is demonstrated by the diverse and interacting impacts of climate risks on ecosystems, cities, health, and poverty and livelihoods, such as in the water–energy– food nexus high 2 train +6887 AR6_WGII 2555 13 Expertise and resources for integrated risk management vary between the developed and developing countries high 2 train +6888 AR6_WGII 2555 14 Integrated pathways for managing climate risks will be most suitable when ‘low regrets’ anticipatory options are established jointly across sectors in a timely manner, path dependencies are avoided in order to not limit future options for climate resilient development, and maladaptations across sectors are avoided high 2 train +6889 AR6_WGII 2561 3 Of this list of adaptation options, many focus on reducing vulnerability to climate change high 2 train +6890 AR6_WGII 2561 10 Next to vulnerability reduction, a large number of adaptation options focus on reducing exposure to climate change high 2 train +6891 AR6_WGII 2561 16 Adaptation efforts can have negative impacts on ecosystems and vulnerable groups high 2 train +6892 AR6_WGII 2561 27 When it comes to decision-making, most of these 24 adaptations rely strongly on formal decision-making high 2 train +6893 AR6_WGII 2561 29 In contrast, informal or individual-led decision-making is more common in several food security-related and livelihood-related adaptations, such as changes to diets, livelihood diversification and seasonal migration high 2 train +6894 AR6_WGII 2561 31 All adaptation options can occur under a range of governance arrangements high 2 train +6895 AR6_WGII 2561 33 This is particularly true for social safety nets and spatial planning, where governments are often required to lead adaptation efforts high 2 train +6896 AR6_WGII 2563 2 There are a number of adaptation options that tend to be governed by communities and individuals, including adaptations to farming and fishery practices and ecosystem-based adaptations high 2 train +6897 AR6_WGII 2563 11 Many of the more generalisable adaptations have also been shown to have benefits to ecosystem services, such as nature restoration and changes to diets/food waste medium 1 train +6898 AR6_WGII 2564 11 Early-warning systems are an adaptation that can benefit more than 5 billion people high 2 train +6899 AR6_WGII 2565 8 While insurance is increasingly accepted as an adaptation option (Linnerooth-Bayer and Hochrainer-Stigler, 2015), positive outcomes are not guaranteed high 2 train +6900 AR6_WGII 2565 11 To avoid this, people simultaneously invest in insurance and adaptations that reduce vulnerability/exposure medium 1 train +6901 AR6_WGII 2565 14 Ex-post risk management relies on national assistance, social safety nets (Section 7.4.2.1.3; Béné et al., 2012; Elmi and Minja, 2019) and support from social networks as well as lending from international institutions high 2 train +6902 AR6_WGII 2565 18 Over the years, regional cooperation, such as through the regional sovereign insurance pools in the Caribbean, the Pacific and Africa, but also transboundary risk management elsewhere have become more important medium 1 train +6903 AR6_WGII 2566 15 Risk financing, especially insurance, is also common in higher-income countries with well-developed insurance markets and higher levels of insurance penetration than in lower-income countries, illustrated by the purple bar in Figure 17.5 high 2 train +6904 AR6_WGII 2566 20 Moderate investment focused on adaptive capacity In contrast to the ‘extensive protection’ scenario, many regions of the world bear greater resemblance to the second typology in Figure 17.5 ‘moderate investment focused on adaptive capacity’ medium 1 train +6905 AR6_WGII 2567 17 In conflict-affected areas similar to the third category of ‘little adaptation investment’, a combination of high vulnerability and relatively less support for adaptation means that there is a large amount of ‘residual risk’, in which residents cope with the impacts of extreme events on a regular basis high 2 train +6906 AR6_WGII 2567 27 While adaptation is rarely focused on taking advantage of opportunities presented by a changed climate, there are numerous co-benefits of adaptation opportunities, from health to reduced emissions to ecosystem services high 2 train +6907 AR6_WGII 2568 22 In contrast to a broadening literature on conceptualisation and policy proposal, there has been little evidence reported in the literature on transformational adaptation and risk management at scale of implementation high 2 train +6908 AR6_WGII 2569 11 For example, several adaptations related to the RKR on risks to peace and migration, namely permanent migration, and cooperative governance, require moderate to high levels of transformation high 2 train +6909 AR6_WGII 2569 12 Some behavioural adaptations, such as changing diets and reducing food waste, can also require large transformations in land use and food culture medium 1 train +6910 AR6_WGII 2569 13 Spatial planning, including urban zoning, also tends to be more transformational medium 1 train +6911 AR6_WGII 2569 14 On the other end of the spectrum, disaster early-warning systems tend to be incremental rather than transformational high 2 train +6912 AR6_WGII 2569 16 For example, improvements in agricultural and fishing practices can be done with moderate transformation to systems medium 1 train +6913 AR6_WGII 2569 17 Similarly, insurance tends to require less transformation, as it can allow people to maintain existing systems while being more resilient to climate-related shocks medium 1 train +6914 AR6_WGII 2569 18 None of the 24 adaptation options are consistently beneficial for vulnerable and marginalised groups high 2 train +6915 AR6_WGII 2569 30 Both RKRs and systems are facing substantial (residual) risk, characterised by adaptation limits and sharing heatwaves as the hazard, for which climate change has been considered the major driver of increasing intensity and frequency high 2 test +6916 AR6_WGII 2571 7 The discussion shows that heat has become a significant health risk globally, incurring severe mortality and morbidity in all world regions with annual heat-related deaths estimated around 300,000 with millions affected high 2 train +6917 AR6_WGII 2571 13 Adaptation, if upgraded to also consider transformational interventions, will thus help to reduce heat risks (medium to high confidence, limited evidence), albeit with reduced effectiveness at higher levels of warming, particularly in regions (Africa, Asia) where lethal heatwaves are projected to occur almost annually towards later in the 21st medium 1 train +6918 AR6_WGII 2571 18 Yet, across all regions there is limited evidence on proposed transformational adaptation and very little evidence regarding implementation high 2 train +6919 AR6_WGII 2571 22 Coral reefs across the tropics have recently seen massive bleaching events (such as for the Great Barrier Reefs) very high 3 train +6920 AR6_WGII 2573 2 Evidence suggests that already at further warming of 1.5°C coral reefs are put at high risk very high 3 train +6921 AR6_WGII 2573 5 Modelling has shown, however, that the effectiveness of such high-risk interventions declines beyond 2°C of global warming medium 1 test +6922 AR6_WGII 2573 6 Already for limited warming beyond 1.5°C for mid-century with increasing intensity and frequency of marine heatwaves, hard limits are projected to become manifest in terms of widespread decline and loss of structural integrity very high 3 train +6923 AR6_WGII 2573 9 Transformational adaptation, while requiring difficult choices to be made, is being discussed to help overcome soft limits through livelihood diversification for alternative income sources, assisted migration and planned relocation of communities dependent on the services provided by the reef ecosystem medium 1 train +6924 AR6_WGII 2573 16 The pathway to a decision may not be linear, depending on when and in what detail the decision-making or consultative group may need to be understanding the climate risk and its real-world context (sense-making, modelling), has sufficient background to analyse and explore options for ameliorating the risk (analysis, exploration), or is ready for interpreting the analyses and deciding on the requirements and strategies for implementing a chosen strategy (interpretation–implementation) high 2 train +6925 AR6_WGII 2573 17 The development of decision-support tools for climate risk management (Palutikof et al., 2019a; Palutikof et al., 2019b) and more generally (Papathanasiou et al., 2016), along with archives of experiences from practitioners (Watkiss and Hunt, 2013; Section 17.5; Bowyer et al., 2014; French, 2020), means that some aspects of the decision-making process can be circumvented or at least streamlined as that experience is re-used high 2 train +6926 AR6_WGII 2573 18 No single approach to decision-making best suits an individual climate risk across any adaptation context (Richards et al., 2013), although there is now a greater awareness of the methods and approaches that are available and their requirements for best practice (Hurlbert et al., 2019) high 2 train +6927 AR6_WGII 2574 7 Sections 8.4.5.6, 16.4 and 17.2 corroborate these findings concluding that, depending on mitigation and adaptation pathways, residual risks in key systems in many regions will create potential for negative impacts beyond adaptation limits medium 1 train +6928 AR6_WGII 2574 9 This ambiguity has persisted, and a policy space for L&D has not clearly been delimited high 2 train +6929 AR6_WGII 2574 14 There is evidence that, without strong risk management and adaptation, losses and damages will continue to affect the poorest vulnerable populations, potentially creating poverty traps high 2 train +6930 AR6_WGII 2575 11 The existential dimension There has been less and often implicit discussion on the existential dimension of climate-related risk as pertaining to L&D medium 1 train +6931 AR6_WGII 2576 4 As national and donor-related funding for impacts and risk management remains limited (Schäfer and Künzel, 2019; 17.2; Serdeczny, 2019) even at current global warming, many highly exposed developing countries remain financially constrained in their capacity to attend to residual impacts and risk management needs (Linnerooth-Bayer and Hochrainer-Stigler, 2015; Roberts et al., 2017; UNEP , 2021a) high 2 train +6932 AR6_WGII 2576 8 Liability and compensation, implying legally defined reimbursement of losses and damages attributable to climate change, remain contentious in L&D dialogue high 2 train +6933 AR6_WGII 2576 15 Overall, the L&D dialogue under the WIM supported by an increasing body of research has made important advances with regard to the two functions of knowledge generation and coordination, yet less so on action and support medium 1 train +6934 AR6_WGII 2576 16 Resolution on the last item will need additional attention as, despite the coalescence of themes, the L&D dialogue continues to proceed across interlinked yet contested discussion strands.Cross-Chapter Box LOSS (continued) Processes and methods to facilitate decision-making, from problem recognition to implementing a solution, have evolved in many contexts, disciplines and applications over the last century high 2 train +6935 AR6_WGII 2577 4 It focuses on decision-analytic methods, noting that decision-support tools will underpin many of these methods by organising information (Bourne et al., 2016; Papathanasiou et al., 2016; Ceccato et al., 2018; Haße and Kind, 2018) or support modelling (Papathanasiou et al., 2016; Kwakkel, 2017; Gardiner et al., 2018), sometimes with a particular decision-analytic process in mind (Hadka et al., 2015; Torresan et al., 2016; Tonmoy et al., 2018).17.3.1.1 Factors to Consider in Selecting Methods to Facilitate Decision-Making The choice of methods and approaches to decision-making for climate risks (next section) will depend on (i) the cognitive needs of the deliberations, otherwise considered to be the phase in developing a decision, (ii) the types of models and modelling available to facilitate the deliberations, (iii) the degree of uncertainty surrounding the choices and (iv) the context of a choice high 2 train +6936 AR6_WGII 2578 2 Sensitivity and robustness analyses can be useful if conditions are favourable to supplement the decision analysis, setting bounds on some of the residual uncertainty high 2 train +6937 AR6_WGII 2578 3 Validation of models and verification of data (Tittensor et al., 2018) are becoming highlighted as important steps in this phase or in the sense- making phase, particularly in their capacity to understand and test decision makers and stakeholders’ perceptions medium 1 train +6938 AR6_WGII 2578 4 Randomisation methods, Bayesian methods, interval methods, multi- criteria decision analysis (MCDA), decision-making under deep uncertainty (DMDU) and economic and financial approaches (e.g., Real Options Analysis) are tools of choice in this phase high 2 train +6939 AR6_WGII 2578 5 Decision-support tools in the provision of data and/or modelling methods are regularly used in this and the sense-making phase high 2 train +6940 AR6_WGII 2578 7 Decision process management tools and methods for communicating choices, outcomes and implementation are expected to be used to provide support in this phase, particularly for understanding whether the advice is fit for purpose, and the efficacy of choices are clear low 0 train +6941 AR6_WGII 2578 19 Decision makers will be better able to choose decision-analytic methods when they have an understanding of the types, scale and breadth of uncertainties around the climate risk high 2 train +6942 AR6_WGII 2578 20 The Cynefin framework (Snowden, 2002; French, 2013) is a policy-driven framework that broadly categorises the decision context of uncertainty within which decision makers and policy analysts may find themselves medium 1 train +6943 AR6_WGII 2579 1 Some analyses partially ignore uncertainties relating to the former in order to focus on conflicts in the values held by different stakeholders and help structure debate (Korhonen and Wallenius, 2020; French, 2020), while others build very sophisticated models of the external world to predict potential consequences, but in doing so lose transparency and risk becoming untrustworthy black boxes to many stakeholders low 0 train +6944 AR6_WGII 2579 10 Elicitation methods help reduce these uncertainties high 2 train +6945 AR6_WGII 2579 11 In addition, informal decision processes can assist in developing consensus in approaches and outcomes (Orlove et al., 2020).17.3.1.2 Decision-Analytic Methods Used in Decision-Making and Climate Risk Management Entities making decisions (countries, regions, organisations and individuals) select methods that best suit them in their context (Fünfgeld et al., 2018; Shi et al., 2019; French, 2020) high 2 train +6946 AR6_WGII 2586 7 Well-designed decision processes recognise the informal and seek to gain information from it without introducing bias medium 1 train +6947 AR6_WGII 2587 3 A higher degree of public participation can lead to more transformational adaptation as well as to higher ambition for local mitigation medium 1 train +6948 AR6_WGII 2587 4 Challenges to stakeholder participation are access to state-of-the-art science, capacity to recognise and respond to non-reliable or false climate science information, and the removal of cognitive and other biases high 2 train +6949 AR6_WGII 2587 5 Participatory and elicitation approaches, where the concerns and involvement of a broader range of interest groups and stakeholders are taken into account, can improve the effectiveness of decision-making medium 1 train +6950 AR6_WGII 2587 7 Specifically, for climate change adaptation, these decision- making strategies can incorporate expert, Indigenous and local knowledge high 2 train +6951 AR6_WGII 2587 10 Bayesian methods are increasingly used in advancing approaches for decision-making and support in climate adaptation (Sperotto et al., 2017), by being able to include stakeholder and decision-maker perceptions and biases (Dias et al., 2018; Engler et al., 2019; Phan et al., 2019; Fulton, 2021) in a transparent modelling environment, thereby facilitating consensus and impartiality medium 1 train +6952 AR6_WGII 2587 11 Increasing computational efficiency means that these methods can enable different approaches to be addressed and different descriptive and prescriptive models to be included within a single probabilistic environment, which also can be updated in iterative processes high 2 train +6953 AR6_WGII 2587 16 Scenarios and pathways, combined with elicitation methods, are becoming widely used to assess adaptation and resilience strategies high 2 train +6954 AR6_WGII 2587 17 They can support the consideration of a wide range of alternative possible futures (Catenacci and Giupponi, 2013; Jäger et al., 2018), enabling identification of potential path dependencies caused by adaptation options high 2 train +6955 AR6_WGII 2587 21 Trade-offs are often directly compared in cost–benefit analyses which require rigorous estimation of the monetised costs and benefits, where monetisation is feasible and values uncontested (such as for infrastructure) high 2 train +6956 AR6_WGII 2588 1 Reducing risk and building resilience under the context of these types of wicked problems require asking ‘what if’ questions about the future, remaining flexible in the face of uncertainty and seeking out policies that provide good outcomes no matter what the future climate might bring high 2 train +6957 AR6_WGII 2588 2 In these cases, trade- offs can be assessed and options can be prioritised through iterative decision-making processes, such as multi-criteria decision-making, robust decision-making and dynamic adaptation pathway planning high 2 train +6958 AR6_WGII 2588 8 Adaptive feedback management is important for managing climate risks that fall within the Cynefin context of chaos, relying on observations and indicators to learn about the system and to trigger actions medium 1 train +6959 AR6_WGII 2588 9 It has been a valued approach for managing wildfish fisheries in many oceans (high confidence) (Fulton et al., 2019; Hollowed et al., 2020; Bahri et al., 2021) and is important for responding to the challenges of climate change high 2 train +6960 AR6_WGII 2588 10 While the benefits of investment in data and assessments can outweigh the costs of implementation low 0 train +6961 AR6_WGII 2588 12 An emerging field in adapting fisheries to climate change is to embed the decision- making system in the scenario models in order to assess the capability of feedback management (decision-making, monitoring and capacity for adjustment of the options over time) to achieve satisfactory trade- offs among the objectives of the different stakeholders medium 1 train +6962 AR6_WGII 2588 17 Since AR5, recognition of the importance of using integrated adaptation to improve climate risk management across the nexus between many sectors and across regions has increased high 2 test +6963 AR6_WGII 2588 18 This was highlighted in the Special Report on Climate Change and Land (Hurlbert et al., 2019); advanced planning and integration of adaptation responses are needed over many levels medium 1 train +6964 AR6_WGII 2588 19 The complexity of managing this nexus may be compounded by the potential for antagonistic or synergistic effects among and between climate impacts, and changes arising from local sectoral activities and independent adaptation responses to those risks high 2 train +6965 AR6_WGII 2590 7 Literature assessed in sectoral and regional chapters of this report present several examples of potential risks to achieving development goals under climate change, at global as well as national and local levels high 2 train +6966 AR6_WGII 2590 13 Deep uncertainty approaches enhance the value of monitoring to detect signals of change in a timely manner medium 1 train +6967 AR6_WGII 2590 15 For early warning signals to be decision relevant, they need to have institutional connectivity to enable action (Haasnoot et al., 2018; Sections 1.4, 11.4, 11.7; Table 11.18) medium 1 train +6968 AR6_WGII 2591 2 Even at the lowest emissions scenarios, some local species will become extinct, but estimates of extinction risk are highly uncertain, typically varying by factors of two to three even for one species (Section 2.5) medium 1 train +6969 AR6_WGII 2591 8 In view of these multiple and diverse examples, it is evident that the application of deep uncertainty methods is enabling decisions to be made in a timely manner that avoid foreseeable and undesirable outcomes and take opportunities as they arise high 2 train +6970 AR6_WGII 2591 9 Prospects for adaptation decision-making Deep uncertainty is increasingly salient for decision-making as recognition of climate-related risks and related uncertainties has increased high 2 train +6971 AR6_WGII 2591 10 These risks can compound and cascade to become new risks, increasing the breadth, frequency and severity of climate change impacts and the consequently increasing scale and scope of adaptation high 2 train +6972 AR6_WGII 2591 13 Overall, there is growing evidence that effective implementation of strategies developed for deeply uncertain problems require adequate mandates and funding frameworks, preparedness and disaster response plans, and monitoring and evaluation of the strategy outcomes, against how the future unfolds medium 1 train +6973 AR6_WGII 2591 14 Collaborative and adaptive governance arrangements, and education and awareness raising, promote learning environments for community engagement, and are essential for the effective implementation of robust adaptation plans medium 1 train +6974 AR6_WGII 2592 5 The importance of supportive governance arrangements is re- iterated widely across regional and sectoral chapters in this report, in multiple different contexts very high 3 train +6975 AR6_WGII 2592 13 The successful implementation of national and sub-national climate change and related policies and strategies are often contingent upon the underlying legislative framework empowering, mandating or guiding their review, implementation and enforcement (Averchenkova and Matikainen, 2017; Scotford et al., 2017) medium 1 train +6976 AR6_WGII 2593 5 Substantial developments in adaptation policy have occurred since AR5 high 2 train +6977 AR6_WGII 2596 7 Capacity is a necessary enabling condition for knowledge to be put to use in adaptation activities high 2 train +6978 AR6_WGII 2597 5 However, there is wide agreement across these definitions that climate finance refers to financial resources devoted to addressing climate change, both mitigation and adaptation to current and projected climate change, and that these resources can come from both public and private sources high 2 train +6979 AR6_WGII 2597 7 Finance can be delivered through a range of instruments including grants, concessional and non-concessional debt, and internal budget reallocations high 2 train +6980 AR6_WGII 2597 16 Nonetheless, there remain differences of opinion on the types of finance that should count towards this goal, with several issues identified high 2 train +6981 AR6_WGII 2597 17 Progress towards the 100 Billion target has shown an upward trend over the last several years (high confidence), but will fall short in 2020, even when the most generous criteria are included high 2 train +6982 AR6_WGII 2598 1 The distribution between adaptation and mitigation has remained strongly weighted towards mitigation, although the proportion allocated to adaptation has increased from 17–25% in 2013/2014 to 19–30% in 2017/2018 high 2 train +6983 AR6_WGII 2598 3 Adaptation finance needs Estimates of global, regional or national finance needs for adaptation and resilience vary depending on both analysis approach, the level of climate change, and the geographic and sectoral scope of analysis high 2 train +6984 AR6_WGII 2599 9 Recent case studies and global level analyses continue to support the conclusion in IPCC AR5 WGII Chapter 17 (Chambwera et al., 2014) that the benefits of adaptation generally remain larger than the costs medium 1 train +6985 AR6_WGII 2599 12 Across these different sources, the main instruments used are grants, concessional debt, market debt, internal budget allocation, insurance, as well as personal savings in households high 2 train +6986 AR6_WGII 2601 19 The proportion of finance allocated to adaptation has remained small throughout, between 4% and 8% high 2 train +6987 AR6_WGII 2601 20 The large majority of tracked adaptation finance is from public sources high 2 train +6988 AR6_WGII 2604 22 Literature indicates the importance of effective governance for promoting integration of local and practitioner knowledge with scientific knowledge high 2 train +6989 AR6_WGII 2605 5 Co-production promotes iterative dialogue, experimentation, the tailoring of knowledge to context, needs and priorities, and learning, often promoting integration of Indigenous knowledge, local knowledge and practitioner knowledge with scientific knowledge high 2 train +6990 AR6_WGII 2605 20 Climate services are more effective and more widely used when they are tailored to specific decisions and decision makers high 2 train +6991 AR6_WGII 2605 21 Sustained iterative engagement between climate information users, producers and translators can improve the quality of the information and the decision-making and avoid maladaptation medium 1 train +6992 AR6_WGII 2606 18 It has been noted that some climate services, such as weather forecasts and early warnings, are an example of a public good, best provided by public agencies high 2 train +6993 AR6_WGII 2606 20 Evidence from many sectors (including water (Section 4.5.2), ocean and coastal ecosystems (Section 3.6.2), and agriculture (Section 5.4.2) and regions (including Africa [Section 9.8.4], Asia [Section 10.4.6] and North America [Section 10.4.5] shows that building capacity (e.g., adaptive capacity, institutional capacity, education/training in human capacity) can support adaptation and limited governance capacity can constrain it high 2 train +6994 AR6_WGII 2607 26 A moderate level of urgency serves as an important driver of climate action, but both high and low levels of urgency impede response high 2 train +6995 AR6_WGII 2608 3 First, greater levels of response to climate change- induced challenges can be motivated by communication strategies that move decision makers from low to moderate levels of urgency high 2 train +6996 AR6_WGII 2608 6 Second, very high levels of urgency are a barrier to effective action medium 1 train +6997 AR6_WGII 2609 20 Lawsuits against private entities contribute to articulating climate change as a legal and financial risk medium 1 train +6998 AR6_WGII 2611 25 The literature is converging to suggest that successful adaptation broadly refers to actions and policies that effectively and substantially reduce climate vulnerability, and exposure to and/or impacts of climate risk (Noble et al., 2014; Juhola et al., 2016), while creating synergies to other climate-related goals, increasing benefits to non-climate- related goals (such as current and future economic, societal and other environmental goals) and minimise trade-offs (Grafakos et al., 2019) across diverse objectives, perspectives, expectations and values (Eriksen et al., 2015; Gajjar et al., 2019a; Owen, 2020) high 2 train +6999 AR6_WGII 2613 15 According to an assessment (Figure 17.11; see SM 17.1 for full descriptions) of maladaptation-relevant outcome dimensions, here called criteria, that is, benefits to people, benefits to ecosystem services, benefits to equity (marginalised ethnic groups, gender, low-income populations), transformational potential and contribution to GHG emission reduction, no option is located at one or the other end of the adaptation-maladaptation continuum (Figure 17.11, right panel), showing that all options have some maladaptation potential, that is, trade-offs very high 3 train +7000 AR6_WGII 2613 19 The options most widely associated with successful adaptation are ‘nature restoration’, ‘social safety nets’, ‘change of farm/fishery practice’ and ‘change of diets/reducing food waste’ high 2 train +7001 AR6_WGII 2614 9 For example, if health care could be provided with low GHG emissions, it would move closer towards successful adaptation high 2 train +7002 AR6_WGII 2614 12 For example, particular attention should be paid to prioritising benefits to low-income groups and leveraging the transformational potential of adaptation (having the largest number of large circles), that is, many evaluated options become maladaptive by exacerbating the vulnerability of low-income groups and by fortifying the status quo medium 1 train +7003 AR6_WGII 2614 14 Through these criteria, a number of adaptation options contribute to a higher potential for successful adaptation high 2 train +7004 AR6_WGII 2616 14 Distributive equity and justice: Attention to distributional equity and justice aims to ensure that adaptation interventions do not exacerbate inequities (Atteridge and Remling, 2018) and that the benefits and burdens of interventions are distributed fairly (Tschakert et al., 2013; Reckien et al., 2017; Reckien et al., 2018b; Pelling and Garschagen, 2019).A global assessment of 1682 papers on adaptation (Araos et al., 2021) finds that about 60% of articles mentioned at least one vulnerable group being involved in the implementation of adaptation or targeted by it medium 1 train +7005 AR6_WGII 2617 7 Procedural justiceDifferential participation and power for more inclusive adaptation planning and implementation –Ensures that processes of representation and participation in adaptation planning, prioritisation and implementation are inclusive (Holland, 2017; Reckien et al., 2017; Reckien et al., 2018b) medium 1 train +7006 AR6_WGII 2618 14 The majority of adaptation M&E efforts have so far focused on processes and outputs rather than on achieved outcomes such as climate risks, vulnerability, well-being or development (Droesch et al., 2008; GIZ and Adelphi, 2017; UNDP Cambodia, 2014; Fawcett et al., 2017) high 2 train +7007 AR6_WGII 2618 17 Instead, they suggest that multiple complementary approaches combined with higher-frequency data collection produce a more elaborate picture of the effects of adaptation and resilience responses (Jones and d’Errico, 2019; Knippenberg et al., 2019; Singh et al., 2019; Jones, 2019a; see Cross-Chapter Box PROGRESS in this Chapter) medium 1 train +7008 AR6_WGII 2620 2 To understand adaptation progress, the assessment of implemented adaptation actions and their outcomes requires more attention very high 3 train +7009 AR6_WGII 2621 3 This Cross-Chapter Box responds to a growing demand for assessing global climate change adaptation progress, which currently faces the challenge of lacking consensus on how adaptation progress at this level can be tracked high 2 train +7010 AR6_WGII 2621 10 Moreover, climate risks are interconnected across scales, regions and sectors (Eakin et al., 2009; Challinor et al., 2017; Cross-Chapter Box INTERREG in Chapter 16; Hedlund et al., 2018) high 2 train +7011 AR6_WGII 2622 1 Yet very few scientific studies have addressed the adaptation-specific aspects of the Global Stocktake (Craft and Fisher, 2018; Tompkins et al., 2018), and there are different views and options on how assessing global progress could take place high 2 train +7012 AR6_WGII 2622 15 Due to the challenges inherent in measuring adaptation outcomes (Sections 16.3, 17.5.1 and 17.5.2.5), most global assessments to date have focused on outputs, such as whether countries have adopted adaptation plans (Berrang-Ford et al., 2021; UNEP , 2021a) high 2 train +7013 AR6_WGII 2622 17 Data Global assessments typically require global availability of consistent data, be they quantitative or qualitative, which has proven to be a constraining factor for attempts to assess global adaptation high 2 train +7014 AR6_WGII 2622 22 Adaptation is hence faced with a dilemma between globally available yet generic data and regionally or locally more detailed yet patchy data high 2 train +7015 AR6_WGII 2622 23 Assessment of existing approaches to assess adaptation progress at the global level Only few global assessments of adaptation progress across sectors have been undertaken to date high 2 train +7016 AR6_WGII 2623 1 The application of differing approaches shows that there is no single ‘best’ approach or data source to assess global progress on adaptation high 2 train +7017 AR6_WGII 2623 3 However, they do not provide comprehensive and robust answers so far on whether climate risk and vulnerability have been reduced (Berrang-Ford et al., 2021) high 2 train +7018 AR6_WGII 2623 13 Important considerations for a robust assessment framework (e.g., consistency), as well as the associated scientific challenges (e.g., aggregation, externalities, breadth versus depth of data) and the role of underlying objectives (e.g., on the contested issue of comparability) are increasingly understood high 2 train +7019 AR6_WGII 2624 2 Each approach and source of information can contribute additional knowledge, but also demonstrates limitations, so that there is no single ‘best’ approach high 2 train +7020 AR6_WGII 2624 4 Triangulated assessments have only rarely been applied high 2 train +7021 AR6_WGII 2625 5 Importantly, adaptations need to be designed to not only combat current and future climate risks but also ensure that they do not lock in undesirable pathways in the future as risks develop and change very high 3 train +7022 AR6_WGII 2625 6 Effective management of climate risks will therefore be dependent on satisfactorily managing current climate risks (Boxes 17.1, 17.2, 17.5), coupled with assessing prognoses for future climate risks, and developing responses in advance for reducing those risks to tolerable residual levels very high 3 train +7023 AR6_WGII 2625 7 The dynamic nature of risk (Viner et al., 2019; Simpson et al., 2021; Sections 16.3, 16.6) also means that the contribution of current adaptations to ameliorating future risks needs to be regularly reviewed high 2 train +7024 AR6_WGII 2625 13 These examples demonstrate that the emergence of climate risks can be at different rates and different time horizons, and the interactions between risks vary from region to region very high 3 train +7025 AR6_WGII 2625 17 Different decision-making approaches can be complementary (high confidence) (Section 17.3.1; Kwakkel et al., 2016), and multiple approaches will likely be necessary in managing the risks across sectors, over different spatial scales, and over short to long time scales medium 1 train +7026 AR6_WGII 2625 18 Deciding on which adaptations to adopt when managing climate risks inevitably needs examination of trade-offs in outcomes very high 3 train +7027 AR6_WGII 2625 20 For climate resilient development, dimensions of poverty, equity, justice and health need to be factored into analyses (Boxes 17.1, 17.5), many of which are difficult to quantify high 2 train +7028 AR6_WGII 2625 21 Moreover, uncertainties on the interactions within and between sectors can make trade-off analyses uneven in their precision across sectors and uncertain as to the outcome of an implemented adaptation medium 1 train +7029 AR6_WGII 2625 22 Expertise and resources for using tools and approaches for integrated risk management vary between the developed and developing countries high 2 train +7030 AR6_WGII 2625 23 Exploration of adaptation scenarios can be derived from Earth System Models high 2 train +7031 AR6_WGII 2626 1 The current levels of uncertainty surrounding the effectiveness of many adaptation options (Section 17.5.2; Cross- Chapter Box PROGRESS in this Chapter) means that decision- making approaches applicable to deep uncertainty (Cross-Chapter Box DEEP in this Chapter; Section 17.3.1) will apply in many if not most cases medium 1 train +7032 AR6_WGII 2626 2 An early step in identifying suitable integrated pathways for managing climate risks, establishing ‘no regrets’ anticipatory options in a timely manner, and avoiding path dependencies is to jointly map the steps for adapting to sectoral risks and determine suitable ways to avoid maladaptations arising high 2 train +7033 AR6_WGII 2626 7 Since AR5, the challenges facing the management of climate risks have been articulated (Adger et al., 2018; Balasubramanian, 2018), and greater clarity on the steps that could be taken to better mainstream adaptation has been developed high 2 train +7034 AR6_WGII 2626 11 Integrated risk assessments and adaptation processes are being developed but with much less experience evident in their implementation high 2 train +7035 AR6_WGII 2626 13 They found that the plans were more oriented at the strategic level or at the level of specific projects rather than identifying methods for resolving cross-sectoral or cross-jurisdictional interactions or issues medium 1 train +7036 AR6_WGII 2626 14 A key recommendation from their review and supported by other studies (e.g., Abutaleb et al., 2018) is that plans would be improved greatly by having inputs from multiple government agencies and multiple sectors medium 1 train +7037 AR6_WGII 2626 16 Hence, satisfactorily managing intersecting climate risks in different settings, of which RKRs provide examples, is central to achieving sustainable development (high confidence) (Section 16.6.4), requiring integrated risk management within and across regions, jurisdictions, sectors and ecosystems high 2 train +7038 AR6_WGII 2626 17 Iterative processes will enable measuring progress and updating adaptation at a satisfactory rate, to account for the different needs within regions and across sectors at different times high 2 train +7039 AR6_WGII 2626 18 The degree to which equity and justice will be achieved will be determined by the participatory processes in deciding on suitable adaptation options, the investment in the adaptation processes and the coordination and collaboration built among institutions and people across regions high 2 train +7040 AR6_WGII 2668 5 Similarly, universal water and energy access can help to reduce poverty and improve well-being while making populations less vulnerable and more resilient to adverse climate impacts very high 3 train +7041 AR6_WGII 2668 8 Severe risks to natural and human systems are already observed in some places (high confidence) and could occur in many more systems worldwide before mid-century (medium confidence) and by the end of the century at all scales, from the local to the global, and at all latitudes and altitudes high 2 train +7042 AR6_WGII 2668 10 Various global trends, including rising income inequality, continued growth in greenhouse gas emissions, land use change, food and water insecurity, human displacement and reversals of long-term increasing life expectancy trends in some nations, run counter to the SDGs very high 3 train +7043 AR6_WGII 2668 15 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.exacerbate injustices, as well as constrain the implementation of CRD for all very high 3 train +7044 AR6_WGII 2668 16 Climate change intensifies existing vulnerability and inequality, with adverse impacts of climate change on the most vulnerable groups, including women and children in low-income households, Indigenous or other minority groups, small-scale producers and fishing communities, and low-income countries high 2 train +7045 AR6_WGII 2668 17 Most vulnerable regions and population groups, such as in East, Central and West Africa, South Asia, Micronesia and Melanesia, and Central America, present the most urgent need for adaptation high 2 train +7046 AR6_WGII 2668 27 Moving towards different pathways involves confronting complex synergies and trade-offs between development pathways, and the options, contested values and interests that underpin climate mitigation and adaptation choices very high 3 train +7047 AR6_WGII 2669 5 Prevailing development pathways are not advancing CRD very high 3 train +7048 AR6_WGII 2669 7 Some low-emissions pathways and climate outcomes are unlikely2 to be realised very high 3 train +7049 AR6_WGII 2669 10 Increasing mitigation effort across multiple sectors exhibits opportunities for synergies with sustainable development, but also trade-offs that increase with mitigation efforts, that need to be balanced and managed high 2 train +7050 AR6_WGII 2669 13 Developing and transitional economies are estimated as low-cost mitigation opportunities but are often at high risk from climate change due to their regional and development context high 2 train +7051 AR6_WGII 2669 20 This Report also uses the term ‘likely range’ to indicate that the assessed likelihood of an outcome lies within the 17–83% probability range.Systems transitions can enable CRD when accompanied by appropriate enabling conditions and inclusive arenas of engagement very high 3 train +7052 AR6_WGII 2669 30 Past choices have already eliminated some development pathways, but other pathways for CRD remain very high 3 train +7053 AR6_WGII 2669 38 Prospects for transformation towards CRD increase when key governance actors work together in inclusive and constructive ways to create a set of appropriate enabling conditions {Section 18.4.2} high 2 train +7054 AR6_WGII 2670 4 CRDPs are determined through engagement in different arenas, the degree to which the emergent pathways foster just and CRD depends on how contending societal interests, values and worldviews are reconciled through inclusive and participatory interactions between governance actors in these arenas of engagement {Section 18.4.3} high 2 train +7055 AR6_WGII 2670 9 Economic sectors and global regions are exposed to different opportunities and challenges in facilitating CRD, suggesting adaptation and mitigation options should be aligned to local and regional context and development pathways very high 3 train +7056 AR6_WGII 2670 17 People, acting through enabling social, economic and political institutions are the agents of system and societal transformations that facilitate CRD founded on the principles of inclusion, equity, climate justice, ecosystem health and human well-being very high 3 train +7057 AR6_WGII 2670 20 For example, grounding adaptation actions in local realities could help to ensure that adaptive actions do not worsen existing gender and other inequities within society (e.g., leading to maladaptation practices) high 2 train +7058 AR6_WGII 2670 24 Pursuing CRD involves considering a broader range of sustainable development priorities, policies and practices, as well as enabling societal choices to accelerate and deepen their implementation very high 3 train +7059 AR6_WGII 2671 5 The literature presenting research findings on climate resilient development (CRD) and pathways and processes for successfully achieving CRD has expanded significantly in the several years since the AR5 very high 3 train +7060 AR6_WGII 2671 7 Furthermore, the literature describing the role of system transitions and societal transformation in enabling climate action (Box 18.1, Section 18.3), compliance with the Paris Agreement (Sections 18.1.3, 18.2.1) and achievement of the SDGs (Section 18.1.3; Box 18.4) has expanded significantly very high 3 train +7061 AR6_WGII 2671 13 However, studies that explicitly refer to CRD as a concept or a guide for policy and practice remain modest very high 3 train +7062 AR6_WGII 2672 3 Moreover, current commitments to reduce greenhouse gas emissions are not yet consistent with limiting changes in global mean temperature elevation to well-below 2°C or 1.5°C very high 3 train +7063 AR6_WGII 2672 5 Exceeding these boundaries poses increased risk of large-scale abrupt or irreversible environmental changes that would threaten human and ecological well-being very high 3 train +7064 AR6_WGII 2672 31 Furthermore, while a given pathway might lead to a set of desired outcomes for one region or set of actors, the process of getting there may come at high environmental, socio- and economic cost to others very high 3 train +7065 AR6_WGII 2677 11 Since the AR5, the volume of research at the nexus of climate action and sustainable development has changed markedly very high 3 train +7066 AR6_WGII 2677 13 Nevertheless, the concept of resilience generally, and CRD specifically, has come under increasing criticism in recent years very high 3 train +7067 AR6_WGII 2677 22 For example, the COVID-19 pandemic is estimated to have reversed gains over the past several years in terms of global poverty reduction very high 3 train +7068 AR6_WGII 2677 26 For example, recent literature on CRD emphasises the urgency of climate action that achieve significant reduction in greenhouse gas emissions, as well as the implementation of adaptation options that result in significant gains in human and natural system resilience very high 3 train +7069 AR6_WGII 2678 8 Second, CRD is dependent on achieving transitions in key systems including energy, land and ecosystem, urban and infrastructure, and industrial systems very high 3 train +7070 AR6_WGII 2678 20 Fifth, CRD involves processes involving diverse actors, at different scales operating within an environmental, developmental, socioeconomic, cultural and political context, as typified in the SDG and the Paris Agreement negotiations very high 3 train +7071 AR6_WGII 2679 9 Systems transitions alone are insufficient to achieve the rapid, fundamental and comprehensive changes required for humanity and planetary health in the face of climate change high 2 train +7072 AR6_WGII 2679 10 Transformative action is increasingly urgent across all sectors, systems and scales to avert dangerous climate change and meet the SDGs (Pelling et al., 2015; IPCC, 2018a; IPCC, 2021b; Shi and Moser, 2021; Vogel and O’Brien, 2021) high 2 train +7073 AR6_WGII 2679 13 Transformative actions aimed at ‘deliberately and fundamentally changing systems to achieve more just and equitable outcomes’, (Shi and Moser, 2021: 2) shift pathways towards climate resilient development (CRD) high 2 train +7074 AR6_WGII 2679 15 Climate actions that support CRD are embedded in these dimensions of development; for example, social cohesion and equity, individual and collective agency, and democratising knowledge processes have been identified as steps to transform practices and governance systems for increased resilience (Ziervogel et al., 2016b; Nightingale et al., 2020; Colloff et al., 2021; Vogel and O’Brien, 2021) high 2 train +7075 AR6_WGII 2679 20 Transformation may require actions that disrupt moral or social boundaries and structures that are perpetuating unsustainable systems and pathways (Vogel and O’Brien, 2021) high 2 train +7076 AR6_WGII 2679 21 Extreme events and long-term climatic changes can trigger a realigning of practices, politics and knowledge (Carr, 2019; Schipper et al., 2020b) high 2 train +7077 AR6_WGII 2679 23 Climate shocks, when managed within socio-political systems in ways that safeguard rather than alter practices and structures, can also reinforce rather than shift the status quo (Mosberg et al., 2017; Carr, 2019; Marmot and Allen, 2020; Arifeen and Nyborg, 2021) high 2 train +7078 AR6_WGII 2680 1 There has been an increase in transformative actions taking place through city-level resilience building aimed at shifting inequitable relations and opening up space for a plurality of actors (Rosenzweig and Solecki, 2018; Ziervogel et al., 2021) high 2 train +7079 AR6_WGII 2680 4 Transformative actions meet resistance by precisely the political, social, knowledge and technical systems and structures they are attempting to transform (Blythe et al., 2018; Shi and Moser, 2021) high 2 train +7080 AR6_WGII 2680 5 There is expanding evidence that many adaptation efforts have failed to be transformative, but instead entrenched inequities, exacerbated power imbalances and reinforced vulnerability among marginalised groups and that, instead, marginalised groups and future trends in vulnerability need to be placed at the centre of adaptation planning (Atteridge and Remling, 2018; Mikulewicz, 2019; Owen, 2020; Eriksen et al., 2021a; Eriksen et al., 2021b; Garschagen et al., 2021) high 2 train +7081 AR6_WGII 2682 1 Others may experience acute challenges with adaptation due to existing vulnerability associated with poverty and social inequality very high 3 train +7082 AR6_WGII 2682 3 While demonstrable progress has been made towards the SDGs and improving human well-being, globally and in specific nations, some observed patterns of development are inconsistent with sustainable development and the principles of CRD very high 3 train +7083 AR6_WGII 2682 14 While there are many possible successful pathways to future development in the context of climate change, history shows that pathways positive for the vast majority of people typically induce significant impacts and costs, especially on marginal and vulnerable people high 2 test +7084 AR6_WGII 2683 28 As empirical studies in the Global South have demonstrated (Lele et al., 2018), developing countries face multiple stressors, climate change being just one among them, and there are multiple normative concerns in developing country contexts, such as equity and justice, and not merely resilience very high 3 train +7085 AR6_WGII 2685 10 For example, pathways that lead to poverty reduction can have synergies with food security, water, gender, terrestrial and ocean ecosystems that support climate risk management, but also poverty alleviation projects with unintended negative consequences that increase vulnerability (e.g., Ley, 2017; Ley et al., 2020).While the scenarios literature is useful for characterising the potential climate risk implications of different global societal futures, important limitations impact their use in climate risk management planning very high 3 train +7086 AR6_WGII 2686 4 Importantly, the scenarios literature overall has found trade-offs such that none of the global societal projections achieve all the SDGs very high 3 train +7087 AR6_WGII 2686 14 For all climate projections and variables, there is significant regional heterogeneity and uncertainty in projected climate change very high 3 train +7088 AR6_WGII 2687 5 Thus, adaptation interventions and pathways can further the achievement of development goals such as food security (Campbell et al., 2016; Douxchamps et al., 2016; Richardson et al., 2018; Bezner Kerr et al., 2019) and improvements in human health (Watts et al., 2019) including in systems where animals and humans live in close proximity very high 3 train +7089 AR6_WGII 2687 7 Since AR5, the scientific community has deepened its understanding of the relationship between adaptation and sustainable development very high 3 train +7090 AR6_WGII 2687 19 The critical literature on socio-ecological resilience, which has grown substantially since the last AR very high 3 train +7091 AR6_WGII 2690 3 Adaptation gaps or deficits arise from a lack of adequate technological, financial, social, and institutional capacities to adapt effectively to climate change and extreme weather events, which are in turn linked to development very high 3 train +7092 AR6_WGII 2690 7 Despite the challenges of measurement associated with adaptation gaps, available evidence from smaller scales across several regions, communities and businesses suggest that significant adaptation gaps have existed in historical contexts of climate change, while expectations of extreme heat, increasing storm intensity and rising sea levels will create the context for the emergence of new gaps very high 3 train +7093 AR6_WGII 2690 10 A higher level of adaptation finance is critical to enhance adaptation planning and implementation and reduce adaptation gaps, particularly in developing countries very high 3 train +7094 AR6_WGII 2694 1 There are numerous individual and system mitigation options throughout the economy and within human and natural systems very high 3 train +7095 AR6_WGII 2694 2 Limiting global average warming has been found to reduce climate risks (IPCC, 2018a; IPCC, 2019b), and limiting global average warming to any temperature level has also been found to be associated with broad ranges of potential global emissions pathways that represent future uncertainty in the evolution of socioeconomic, technological, market and physical systems very high 3 train +7096 AR6_WGII 2695 16 In general, mitigation alters development opportunities by constraining the emissions future society can produce, which affects markets, resource allocation, economic structure, income distribution, consumers and the environment (besides climate) very high 3 train +7097 AR6_WGII 2698 21 Nonlinearities have been estimated in global and regional mitigation costs and potential economic damages from climate change very high 3 train +7098 AR6_WGII 2698 28 Not only do trade-offs vary by climate level, as do synergies, but they increase at an increasing rate and their relative importance can shift across climate levels very high 3 train +7099 AR6_WGII 2698 31 For instance, additional mitigation options and more economically efficient policy designs have been shown to reduce mitigation costs and the nonlinearities in mitigation costs very high 3 train +7100 AR6_WGII 2699 1 Central to this framing is a shift away from viewing adaptation as discrete sets of options that are selected and implemented to manage risk, to thinking about adaptation as a social process that evolves over time, includes multiple decision points, and requires dynamic adjustments in response to new information about climate risk, socioeconomic conditions and the value of potential adaptation responses very high 3 train +7101 AR6_WGII 2699 3 While ensuring development and adaptation produce synergies that allow for the achievement of sustainable development is challenging, modelling exercises suggest that there are pathways where synergies among the SDGs are realised very high 3 train +7102 AR6_WGII 2699 22 Synergies between adaptation, mitigation and sustainable development might be promoted by prioritising those CRD strategies most likely to generate synergies very high 3 test +7103 AR6_WGII 2699 24 Similarly, trade- offs might be managed by prioritising strategies such as disqualifying mitigation options thought to have negative social implications (Section 18.2.5.3.1), internalising externalities, such as placing a fee or constraint on a negative externality or related activity (Dubash et al., 2022) (Bistline and Rose, 2018), or using complementary policies, such as transfer payments to offset negative mitigation, adaptation or sustainable development strategy implications very high 3 train +7104 AR6_WGII 2700 9 Globally, low climate change projections, versus higher climate change projections, imply greater mitigation, lower climate risks and less adaptation. This implies greater mitigation trade-offs in terms of overall economic development, food crop prices, energy prices and overall household consumption, but lower climate risk, with sustainable development synergies such as human health and lower adaptation trade-offs, and an uneven distribution of effects very high 3 test +7105 AR6_WGII 2700 29 Prioritising sustainable development locally is also supported by the insight that the impacts on poverty depend at least as much or more on development than on the level of climate change very high 3 train +7106 AR6_WGII 2701 2 In the climate change solution space, system transitions represent an important mechanism for linking and enabling mitigation, adaptation and sustainable development options and actions very high 3 train +7107 AR6_WGII 2701 12 This growth in demand, however, has been moderated by improvements in energy efficiency in industry, buildings and transportation sectors very high 3 train +7108 AR6_WGII 2701 14 Features of sustainable development, such as enhanced energy access, energy security, reductions in air pollution and economic growth, continue to be the dominant influence on the evolution of energy systems and decision making regarding energy investments and portfolios very high 3 train +7109 AR6_WGII 2701 16 Yet there are examples at the local, regional and national level of policy incentivising rapid changes in energy systems very high 3 train +7110 AR6_WGII 2701 17 Many sustainable development priorities have co-benefits in terms of climate mitigation, such as air pollution and conservation policies reducing short-lived climate forcers and sequestering carbon respectively, as well adaptation benefits, such as improved energy access and environmental quality enhancing adaptive capacity very high 3 train +7111 AR6_WGII 2701 24 Energy systems have been a historical driver of climate change, but are also adversely affected by climate change impacts, including short- term shocks and stressors from extreme weather as well as long-term shifts in climatic conditions very high 3 train +7112 AR6_WGII 2702 1 Available literature indicates that greenhouse gas emissions reductions have been achieved in response to climate actions including financial incentives to promote renewable energy, carbon taxes and emissions trading, removal of fossil fuel subsidies, and promotion of energy efficiency standards very high 3 train +7113 AR6_WGII 2703 1 Rather, urban areas are increasingly conceptualised as complex socio-ecological or socio- technical systems very high 3 train +7114 AR6_WGII 2703 3 Urban transitions will be associated with synergies as well as trade-offs with respect to sustainable development very high 3 test +7115 AR6_WGII 2709 25 Accordingly, significant acceleration in the pace of system transitions is necessary to enable the implementation of mitigation, adaptation and sustainable development initiatives consistent with CRD very high 3 train +7116 AR6_WGII 2711 1 Contributing Authors: Seema Arora-Jonsson (Sweden/India), Emily Baker (USA), Graeme Dean (Ireland), Emily Hillenbrand (USA), Alison Irvine (Canada), Farjana Islam (Bangladesh/ UK), Katriona McGlade (UK/Germany), Hanson Nyantakyi-Frimpong (Ghana), Nitya Rao (UK/India), Federica Ravera (Italy), Emilia Reyes (Mexico), Diana Hinge Salili (Fiji), Corinne Schuster-Wallace (Canada), Alcade C. Segnon (Benin), Divya Solomon (India), Shreya Some (India), Indrakshi Tandon (India), Sumit Vij (India), Katharine Vincent (UK/South Africa), Margreet Zwarteveen (the Netherlands) Key Messages • Gender and other social inequities (e.g., racial, ethnic, age, income, geographic location) compound vulnerability to climate change impacts high 2 train +7117 AR6_WGII 2711 8 Efforts are needed to change unequal power dynamics and to foster inclusive decision making for climate adaptation to have a positive impact for gender equality high 2 train +7118 AR6_WGII 2712 8 Climate change is reducing the quantity and quality of safe water available in many regions of the world and increasing domestic water management responsibilities high 2 train +7119 AR6_WGII 2712 10 Water insecurity and the lack of water, sanitation and hygiene (WASH) infrastructure have resulted in psychosocial distress and gender-based violence, as well as poor maternal and child health and nutrition (Collins et al., 2019a; Wilson et al., 2019; Geere and Hunter, 2020; Islam et al., 2020; Mainali et al., 2020) (Sections 4.3.3 and 4.6.4.4) high 2 train +7120 AR6_WGII 2712 15 Increased domestic responsibilities of women and youth, due to migration of men, can increase their vulnerability due to their reduced capacity for investment in off-farm activities and reduced access to information (Sugden et al., 2014; O’Neil et al., 2017) (Sections 4.3, 4.6) high 2 train +7121 AR6_WGII 2712 20 In most regions where there are climate change policies that consider gender, they inadequately address structural inequalities resulting from climate change impacts, or how gender and other social inequalities can compound risk high 2 train +7122 AR6_WGII 2713 6 However, attention to the following has the potential to bring about change: Creation of new, deliberative policymaking spaces that support inclusive decision making processes and opportunities to (re)negotiate pervasive gender and other social inequalities in the context of climate change for transformation (Tschakert et al., 2016; Harris et al., 2018; Ziervogel, 2019; Garcia et al., 2020) high 2 train +7123 AR6_WGII 2713 7 Increased access to reproductive health and family planning services, which contributes to climate change resilience and socioeconomic development through improved health and well-being of women and their children, including increased access to education, gender equity and economic status (Onarheim et al., 2016; Starbird et al., 2016; Lopez-Carr, 2017; Hardee et al., 2018) (Section 7.4) high 2 train +7124 AR6_WGII 2713 14 Strengthened adaptive capacity and resilience through integrated approaches to adaptation that include social protection measures, disaster risk management and ecosystem-based climate change adaptation high 2 train +7125 AR6_WGII 2713 15 For example, gender-transformative and nutrition-sensitive agroecological approaches strengthen adaptive capacities and enable more resilient food systems by increasing leadership for women and their participation in decision making and a gender-equitable domestic work high 2 train +7126 AR6_WGII 2716 13 The agency of people to act on CRD is grounded in their worldviews, beliefs, values and consciousness (Woiwode, 2020), and is shaped through social and political processes including how policies and decision making recognise the voices, knowledges and rights of particular actors over others very high 3 train +7127 AR6_WGII 2717 10 Given these insights, CRD can be understood as the sum of complex multi-dimensional processes consisting of large numbers of actions and societal choices made by multiple actors from government, the private sector and civil society, with important influences by science and the media very high 3 train +7128 AR6_WGII 2718 7 However, the pursuit of a given NDC within a specific country will likely necessitate a range of other policy interventions that have more immediate impact on technologies and behaviour, implicating transitions in energy, industry, land and infrastructure very high 3 train +7129 AR6_WGII 2718 10 As reflected by the SDGs (and SDG 13 specifically), the mainstreaming of climate change concerns into development policies is one mechanism for pursuing sustainable development and CRD very high 3 train +7130 AR6_WGII 2718 18 As a consequence, the characteristics of economic systems will play an important role in determining their resilience very high 3 train +7131 AR6_WGII 2718 23 Addressing climate impacts in isolation is unlikely to achieve equitable, efficient or effective adaptation outcomes very high 3 test +7132 AR6_WGII 2718 26 Although higher levels of poverty, corruption, and economic and social inequalities can increase the risk of negative outcomes, these potential negative effects would be mitigated if inequality impacts were taken into consideration in all stages of policy making very high 3 test +7133 AR6_WGII 2719 4 These gains can provide more resource-efficient production technologies and positively affect economic competitiveness very high 3 train +7134 AR6_WGII 2719 21 Financial actors increasingly recognise that the generation of long-term, sustainable financial returns is dependent on stable, well-functioning and well-governed social, environmental and economic systems very high 3 train +7135 AR6_WGII 2719 29 Stable and predictable carbon-pricing regimes would significantly contribute to fostering financial innovation that can help further accelerate the decarbonisation of the global economy, even in jurisdictions which are more lenient in implementing climate mitigation actions very high 3 test +7136 AR6_WGII 2720 8 As a consequence, they and can become a significant barrier to change, whether incremental or more transformational very high 3 train +7137 AR6_WGII 2720 12 There is agreement in this literature that such an approach allows for the effective integration of climate challenges into existing policy and planning processes very high 3 train +7138 AR6_WGII 2720 21 However, access to, and the benefits of, that innovation have not been evenly distributed among global regions and communities, and continued innovation is needed to facilitate climate action and sustainable development very high 3 train +7139 AR6_WGII 2721 35 The need to assess resilience and adaptation against a background of evolving climate hazards, and to link resilience and adaptation with development outcomes, present further methodological challenges (very high confidence) (Brooks et al., 2014).Currently, the ability to monitor different components of CRD are in various stages of maturity very high 3 train +7140 AR6_WGII 2722 3 This allows measurement of key indicators that are proxies for resilience at regular intervals, even in the absence of significant climate hazards and associated disruptions very high 3 train +7141 AR6_WGII 2722 28 Alternatively, they can also serve as significant barriers to system transitions and transformation, based on anthropocentric, mechanistic and materialistic worldviews and the utilitarian, individualist or skeptical values and attitudes they often promote very high 3 train +7142 AR6_WGII 2723 7 On the other hand, the tendency for certain worldviews to dominate the policy discourse has the potential to exacerbate social, economic and political inequities as well as ontological, epistemic and procedural injustices very high 3 train +7143 AR6_WGII 2724 4 Yet these knowledge systems represent a range of cultural practices, wisdom, traditions and ways of knowing the world that provide accurate and useful climate change information, observations and solutions very high 3 train +7144 AR6_WGII 2724 20 Instead, Indigenous knowledge and local knowledge can shape how climate change risk is understood and experienced, the possibility of developing climate change solutions grounded in place-based experiences, and the development of governance systems that match the expectations of different Indigenous knowledge and local knowledge holders very high 3 train +7145 AR6_WGII 2725 5 Supporting Indigenous Peoples’ leadership and rights in climate adaptation options at the local, regional, national and international levels is an effective way to ensure that such options are adapted to their living conditions and do not pose additional detrimental impacts to their lives very high 3 train +7146 AR6_WGII 2725 8 Climate-resilient futures will depend on recognising the socioeconomic, political and health inequities that often affect Indigenous Peoples (Mapfumo et al., 2016; Ludwig and Poliseli, 2018) very high 3 test +7147 AR6_WGII 2725 16 Adaptation efforts have benefited from the inclusion of Indigenous knowledge and local knowledge (IPCC, 2019e) very high 3 train +7148 AR6_WGII 2725 17 Moreover, it has been recognised that including Indigenous knowledge and local knowledge in IPCC reports can contribute to overcoming the combined challenges of climate change, food security, biodiversity conservation, and combating desertification and land degradation (IPCC, 2019c) high 2 train +7149 AR6_WGII 2727 17 More integration of government policy and interventions across scales, accompanied by capacity building to accelerate adaptation is needed very high 3 test +7150 AR6_WGII 2727 26 However, the resource limitations of local governments as well as their small geographic sphere of influence suggests the need for more funding for this from higher levels of government, particularly national governments, to address adaptation gaps very high 3 train +7151 AR6_WGII 2728 10 Wealthy nations of the Global North, including for example the USA, Great Britain, Iceland and Japan, have had success over the past decade in reducing their GHG emissions while growing their economies very high 3 train +7152 AR6_WGII 2728 15 While GDP growth can drive growth in income, it can also drive growth in inequality which can undermine poverty reduction efforts very high 3 test +7153 AR6_WGII 2729 1 Nevertheless, consistent with earlier assessment of enabling conditions for system transitions (Section 18.4.2.1), implementation of Blue Growth initiatives is contingent upon the successful achievement of social innovation as well as creating an inclusive and cooperative governance structure very high 3 train +7154 AR6_WGII 2729 13 However, while scientific and technology knowledge may be useful, in some cases, they remain subordinate to political agendas, or are controlled by actors in positions of power and thus not equitably distributed very high 3 train +7155 AR6_WGII 2729 20 Realising the benefits of STI, however, may be contingent on building broader STI capacity and bolstering nations’ systems of innovation very high 3 train +7156 AR6_WGII 2731 11 Since AR5, both formal and informal setting are increasingly arenas of debate and contestation regarding development choices and pathways very high 3 train +7157 AR6_WGII 2731 24 Social movements are demanding radical action as the only option to achieve the mobilisation necessary for deep societal transformation very high 3 train +7158 AR6_WGII 2732 2 Research shows that new climate movements have increased public awareness, and also stimulated unprecedented public engagement with climate change very high 3 train +7159 AR6_WGII 2732 26 The WGI data also indicate that increases in globally averaged temperatures will have different consequences for regional climate change (Table 18.4), including variation in the magnitude and, for precipitation, even the direction of change very high 3 train +7160 AR6_WGII 2733 3 Given outcomes in many systems including public health, agriculture, ecosystems and biodiversity, and infrastructure are often associated with biophysical thresholds (e.g., physiological or design thresholds), those regions where such thresholds are increasingly exceeded due to climate change may experience disproportionately higher impacts very high 3 train +7161 AR6_WGII 2733 11 Nevertheless, such higher SSTs have implications not only for ocean ecosystems and the distribution of marine species, but also for weather patterns, such as formation and intensity of tropical cyclones very high 3 train +7162 AR6_WGII 2733 15 If that demand is not met, then the adaptation gap will be larger, with greater risk of loss and damage very high 3 train +7163 AR6_WGII 2733 19 Common indicators of development reflect the significant diversity that exists across different global regions with respect to their development context very high 3 test +7164 AR6_WGII 2733 25 However, nations and regions with high PPAHDI values also tend to have higher per capita CO 2-e emissions production, indicating that economic development based on fossil fuel use undermines both efforts on climate action as well as the SDGs very high 3 train +7165 AR6_WGII 2733 30 In addition to development indicators, the literature assessed in the WGII regional chapters indicates that different regions experience a range of development challenges and opportunities that affect the pursuit of CRD very high 3 train +7166 AR6_WGII 2733 32 For example, significant challenges exist within regions with respect to managing debt and the ability to fund or finance climate action and sustainable development interventions very high 3 train +7167 AR6_WGII 2733 33 On the other hand, a broad range of opportunities exist to pursue CRD including challenges with debt and financing of adaptation competing policy objectives, social protection programmes, economic diversification, investing in education and human capital development, and expanding disaster risk reduction efforts very high 3 train +7168 AR6_WGII 2733 34 There are a wide variety of more focused options for climate action and sustainable development very high 3 train +7169 AR6_WGII 2735 2 For example, in all regions, existing vulnerability and inequality exacerbate climate risk and therefore pose challenges to CRD very high 3 train +7170 AR6_WGII 2735 3 Furthermore, low prioritisation of sustainability and climate action in government decision making, low perceptions of climate risk, and path dependence in governance systems and decision-making processes all pose barriers to system transitions, transformation and CRD very high 3 train +7171 AR6_WGII 2735 8 Issues associated with natural resource dependency, access to information for decision making, access to human and financial capital, and path dependence of institutions represent barriers that must be overcome if sectors are to support transitions that enable CRD. These challenges are more acute within vulnerable communities or nations where capacity to innovate and invest are constrained and social inequities reinforce the status quo very high 3 test +7172 AR6_WGII 2735 15 Moreover, progress across multiple sectors simultaneously creates opportunities for synergies for achieving the SDGs, but also enhances the risk of potential trade-offs very high 3 train +7173 AR6_WGII 2743 8 These include enhancing understanding of mainstreaming of climate change into institutional decision making, managing risk under conditions of uncertainty, catalysing system transitions and transformation, and processes for enhancing participation, equity and accountability in sustainable development very high 3 train +7174 AR6_WGII 2743 16 This includes consideration for risk and science communication; decision analysis and decision support systems; and mechanisms for knowledge co-production between scientists and public policy actors very high 3 train +7175 AR6_WGII 2780 3 For energy systems transitions, the adaptation options of infrastructure resilience, efficient water use and water management, and reliable power systems enable energy systems to work during disasters with reduced costs, demonstrating the synergistic relationships between mitigation and adaptation high 2 train +7176 AR6_WGII 2780 5 New evidence has focused on both options for peri-urban and rural areas through distributed generation and isolated renewable energy systems, which also provide multiple social co-benefits medium 1 train +7177 AR6_WGII 2780 6 For efficient water use and management, the synergistic potential with mitigation can make processes more efficient and cost effective high 2 train +7178 AR6_WGII 2780 7 With regards to adaptation feasibility, efficient water use is especially useful in drought-stricken areas and provides better water management for multiple uses high 2 train +7179 AR6_WGII 2780 9 Forest- and biodiversity-based adaptation options are generally promoted on the basis of their positive impacts on adaptive and ecological capacities, increased provision of ecosystem services and goods, with a particularly strong contribution to carbon sequestration high 2 train +7180 AR6_WGII 2780 10 However, large afforestation projects and the introduction of non-native and fast-growing vegetation reduce water availability, impoverish habitats for wildlife and reduce overall ecological resilience, threatening the achievement of some Sustainable Development Goals (SDGs), and potentially leading to maladaptation (high confidence).Over-reliance on forest-based solutions may increase the susceptibility to wildfires, with detrimental consequences both for mitigation and adaptation medium 1 train +7181 AR6_WGII 2780 11 Over the last decade, forest- and biodiversity-based solutions have gained considerable political traction and social acceptability (high confidence), but in countries with economies highly dependent on the export of agricultural commodities, opportunity costs continue to hinder the expansion of these alternatives, particularly against more profitable land uses high 2 train +7182 AR6_WGII 2780 13 Agro-forestry solutions have strong ecological and adaptive co-benefits high 2 train +7183 AR6_WGII 2780 16 Successful implementation requires a strong socioeconomic framework and can offer diverse social, ecological and economic benefits, as well as sequestering carbon high 2 train +7184 AR6_WGII 2780 17 There is extensive experience with hard coastal defence structures (e.g., sea walls), which can be cost-effective in economic terms, depending on the location (medium confidence); however, they are considered maladaptive and unsustainable in some contexts (medium confidence) due to their lack of flexibility or robustness in response to a changing climate, as well as their carbon-intensiveness and potential ecological impacts medium 1 train +7185 AR6_WGII 2781 1 There are financial barriers to implementing sustainable aquaculture and fisheries, even though they can improve employment opportunities, especially for local communities medium 1 train +7186 AR6_WGII 2781 2 Technical resource availability is still lacking and could represent a barrier to implementing sustainable aquaculture and fisheries medium 1 train +7187 AR6_WGII 2781 3 Robust institutional and legal frameworks are needed to guarantee effective adaptation high 2 train +7188 AR6_WGII 2781 4 Sustainable aquaculture and fisheries are highly dependent on healthy and resilient ecosystems high 2 train +7189 AR6_WGII 2781 5 They can provide diverse ecosystem services and support coastal ecosystems restoration medium 1 train +7190 AR6_WGII 2781 7 This suite of strategies has strong feasibility to build resilience while improving incomes (medium confidence) and providing mitigation co-benefits high 2 train +7191 AR6_WGII 2781 8 While technological and ecological feasibility is high, institutional, market and socio-political acceptability remain significant barriers medium 1 train +7192 AR6_WGII 2781 9 Improving water use efficiency and water resource management under land and ecosystem transitions has high technological feasibility high 2 train +7193 AR6_WGII 2781 10 However, economic and institutional barriers remain and are based on type, scale and location of interventions medium 1 train +7194 AR6_WGII 2781 11 Notably, inadequate institutional capacities to prepare for changing water availability, especially in the long term, unsustainable and unequal water use and sharing practices, and fragmented water resource management approaches remain critical barriers to feasibility high 2 train +7195 AR6_WGII 2781 13 These strategies have high economic and environmental feasibility (high confidence) and substantial mitigation co-benefits medium 1 train +7196 AR6_WGII 2781 14 However, high costs, inadequate information and technical know-how, delays between actions and tangible benefits, lack of comprehensive policies, fragmentation across different sectors, inadequate access to credit, and unequal access to resources constrain technological, institutional and socio-cultural feasibility medium 1 train +7197 AR6_WGII 2781 15 For urban and infrastructure system transitions, sustainable urban planning can support both adaptation and decarbonisation by mainstreaming climate concerns, including effective land use into urban policies, by promoting resilient and low-carbon infrastructure, and by protecting and integrating carbon-reducing biodiversity and ecosystem services into city planning medium 1 train +7198 AR6_WGII 2781 16 Urban green infrastructure and ecosystem services have high feasibility to support climate adaptation and mitigation efforts in cities, for example to reduce flood exposure and attenuate the urban heat island high 2 train +7199 AR6_WGII 2781 17 While green infrastructure options are cost-effective and provide co-benefits in terms of ecosystem services such as improved air quality or other health benefits high 2 train +7200 AR6_WGII 2781 18 Governments across scales can support urban sustainable water management by undertaking projects to recycle wastewater and runoff through green infrastructure; enabling greater coherence between urban water and riverine basin management; decentralising water systems; supporting networks for sharing best practices in water supply and storm runoff treatment to scale sustainable management; and foregrounding equity and justice concerns, especially through participation involving informal settlement residents medium 1 train +7201 AR6_WGII 2781 19 Strong and equitable health systems can protect the health of populations in the face of known and unexpected stressors medium 1 train +7202 AR6_WGII 2781 20 Health and health systems adaptation is feasible where capacity is well developed, and where options align with national priorities and engage local and international communities medium 1 train +7203 AR6_WGII 2781 21 Socio-cultural acceptability of health and health systems adaptation is high and there is significant potential for risk-mitigation and social co-benefits where adaptation addresses the needs of vulnerable regions and populations medium 1 train +7204 AR6_WGII 2781 22 Microeconomic feasibility and socioeconomic vulnerability reduction potentials are also high (high confidence), although economic feasibility may pose a significant challenge in low-income settings medium 1 train +7205 AR6_WGII 2781 23 However, inadequate institutional capacity and resource availability represent major barriers, particularly for health systems struggling to manage current health risks high 2 train +7206 AR6_WGII 2781 24 There is strong evidence that disaster risk management (DRM) is highly feasible when supported by strong institutions, good governance, local engagement and trust across actors medium 1 train +7207 AR6_WGII 2781 25 DRM is constrained by lack of capacity, inadequate institutions, limited coordination across levels of government (high confidence), lack of transparency and accountability, and poor communication medium 1 train +7208 AR6_WGII 2781 26 There is a preference for top-down DRM processes, which can undermine local institutions and perpetuate uneven power relationships medium 1 train +7209 AR6_WGII 2782 1 Moves towards community-based and ecosystem-based DRM are promising but uneven and may increase vulnerability if they fail to address underlying and structural determinants of vulnerability high 2 train +7210 AR6_WGII 2782 2 Climate services that are demand-driven and context-specific (e.g., to a particular crop or agricultural system) build adaptation capacity and enable short- and longer-term risk management decisions high 2 train +7211 AR6_WGII 2782 3 Metrics to assess the economic outcomes of climate services remain insufficient to capture longer-term benefits of interventions medium 1 train +7212 AR6_WGII 2782 4 While technological capacity and political acceptance is high medium 1 train +7213 AR6_WGII 2782 5 Risk insurance can be a feasible tool to adapt to climate risks and support sustainable development high 2 train +7214 AR6_WGII 2782 7 Insurance mechanisms enjoy wide legal and regulatory acceptability among policymakers and are institutionally feasible high 2 train +7215 AR6_WGII 2782 8 However, socio-cultural and financial barriers make insurance spatially and temporally challenging to implement (high confidence), even though it can improve the health and well-being of populations medium 1 train +7216 AR6_WGII 2782 9 The risk of generating maladaptive outcomes can further limit the uptake of insurance, as it can provide disincentives for reducing risk over the long term medium 1 train +7217 AR6_WGII 2782 11 Ensuring equitable access to and benefits from innovative financial products (e.g., loans) is needed to guarantee successful uptake of insurance across all the population high 2 train +7218 AR6_WGII 2782 12 Migration has been used by millions around the world to maintain and improve their well-being in the face of changed circumstances, often as part of labour or livelihood diversification very high 3 train +7219 AR6_WGII 2782 13 Properly supported and, where levels of agency and assets are high, migration as a climate response can reduce exposure and socioeconomic vulnerability medium 1 train +7220 AR6_WGII 2782 15 These households can undertake distress migration, which results in negative adaptive and resilience outcomes high 2 train +7221 AR6_WGII 2782 16 Outcomes can be improved through a systematic examination of the political economy of local and regional sectors that employ precarious communities and by addressing vulnerabilities that pose barriers to in situ adaptation and livelihood strategies medium 1 train +7222 AR6_WGII 2782 17 Migrants and their sending and receiving communities can be supported through temporary labour-migration schemes, improving discourses on migration, and matching existing migration agreements with development objectives medium 1 train +7223 AR6_WGII 2782 18 Planned relocation and resettlement have low feasibility as climate responses medium 1 train +7224 AR6_WGII 2782 19 Previous disaster- and development-related relocation has been expensive, contentious, posed multiple challenges for governments and amplified existing, and generated new, vulnerabilities for the people involved high 2 train +7225 AR6_WGII 2782 20 Planned relocation will be increasingly required as climate change undermines habitability, especially for coastal areas medium 1 train +7226 AR6_WGII 2782 21 Full participation of those affected, ensuring human rights-based approaches, preserving cultural, emotional and spiritual bonds to place, and dedicated governance structures and associated funding are associated with improved outcomes high 2 train +7227 AR6_WGII 2782 22 Improving the feasibility of planned relocation and resettlement is a high priority for managing climate risks high 2 train +7228 AR6_WGII 2784 15 Since SR1.5, there has not been significant change in the feasibility of the first two options as they continue to be implemented successfully, allowing for power generation to maintain or increase its reliability during extreme weather events high 2 train +7229 AR6_WGII 2784 17 The main difference from SR1.5 is that resilient power infrastructure now includes distributed generation utilities, such as microgrids, as there is increasing evidence of its role in reducing vulnerability, especially within underserved populations high 2 train +7230 AR6_WGII 2785 3 As with previous options, the technological means exist to create redundancy in power generation, transmission and distribution systems and their implementation ensures the continuous functionality of emergency services, such as communications, health and water pumping, amongst others, in urban, peri-urban and rural landscapes high 2 train +7231 AR6_WGII 2791 9 Moreover, while once neglected, rapidly increasing attention has been paid to the equity and justice dimensions of planning and implementing green infrastructure initiatives, such as inclusion of citizens in decision making or the allocation of benefits and impacts of projects (Anguelovski et al., 2019b; Buijs et al., 2019; Langemeyer et al., 2020; Venter et al., 2020) Institutional barriers constrain the feasibility of urban green infrastructure medium 1 train +7232 AR6_WGII 2840 1 The major drivers of the increase are climate hazards including droughts, floods and storms.Over the past 30 years, major crop yields decreased by 4–10% globally due to climate change high 2 train +7233 AR6_WGIII 18 23 By 2019, the largest growth in absolute emissions occurred in CO2 from fossil fuels and industry followed by CH4, whereas the highest relative growth occurred in fluorinated gases, starting from low levels in 1990 high 2 train +7234 AR6_WGIII 18 24 Net anthropogenic CO2 emissions from land use, land-use change and forestry (CO2-LULUCF) are subject to large uncertainties and high annual variability, with low confidence even in the direction of the long-term trend.9 (Figure SPM.1) {Figure 2.2, Figure 2.5, 2.2, Figure TS.2} B.1.3 Historical cumulative net CO2 emissions from 1850 to 2019 were 2400 ± 240 GtCO2 high 2 train +7235 AR6_WGIII 19 2 The annual average CO2-FFI emissions reduction in 2020 relative to 2019 was about 5.8% [5.1–6.3%], or 2.2 [1.9–2.4] GtCO2 high 2 train +7236 AR6_WGIII 20 6 Emissions growth in AFOLU, comprising emissions from agriculture (mainly CH4 and N2O) and forestry and other land use (mainly CO2) is more uncertain than in other sectors due to the high share and uncertainty of CO2-LULUCF emissions medium 1 train +7237 AR6_WGIII 20 7 About half of total net AFOLU emissions are from CO2-LULUCF, predominantly from deforestation14 medium 1 train +7238 AR6_WGIII 20 13 For comparison, the carbon intensity of primary energy is projected to decrease globally by about 3.5% yr–1 between 2020 and 2050 in modelled scenarios that limit warming to 2°C (>67%), and by about 7.7% yr–1 globally in scenarios that limit warming to 1.5°C (>50%) with no or limited overshoot.16 high 2 train +7239 AR6_WGIII 21 6 Least developed countries (LDCs) and Small Island Developing States (SIDS) have much lower per capita emissions (1.7 tCO2-eq and 4.6 tCO2-eq, respectively) than the global average (6.9 tCO2-eq), excluding CO2-LULUCF.18 high 2 train +7240 AR6_WGIII 23 26 Innovation has provided opportunities to lower emissions and reduce emission growth and created social and environmental co-benefits high 2 train +7241 AR6_WGIII 23 31 For example, sensors, internet of things, robotics, and artificial intelligence can improve energy management in all sectors, increase energy efficiency, and promote the adoption of many low-emission technologies, including decentralised renewable energy, while creating economic opportunities high 2 train +7242 AR6_WGIII 23 32 However, some of these climate change mitigation gains can be reduced or counterbalanced by growth in demand for goods and services due to the use of digital devices high 2 train +7243 AR6_WGIII 23 34 Digital technology supports decarbonisation only if appropriately governed high 2 train +7244 AR6_WGIII 25 5 At least 18 countries that had Kyoto targets for the first commitment period have had sustained absolute emission reductions for at least a decade from 2005, of which two were countries with economies in transition very high 3 test +7245 AR6_WGIII 25 6 The Paris Agreement, with near universal participation, has led to policy development and target-setting at national and sub-national levels, in particular in relation to mitigation, as well as enhanced transparency of climate action and support medium 1 train +7246 AR6_WGIII 25 8 By 2020, over 20% of global GHG emissions were covered by carbon taxes or emissions trading systems, although coverage and prices have been insufficient to achieve deep reductions medium 1 train +7247 AR6_WGIII 25 9 By 2020, there were ‘direct’ climate laws focused primarily on GHG reductions in 56 countries covering 53% of global emissions medium 1 train +7248 AR6_WGIII 25 10 Policy coverage remains limited for emissions from agriculture and the production of industrial materials and feedstocks high 2 train +7249 AR6_WGIII 25 12 Multiple lines of evidence suggest that mitigation policies have led to avoided global emissions of several GtCO2-eq yr–1 medium 1 train +7250 AR6_WGIII 25 16 These financial flows remained heavily focused on mitigation, are uneven, and have developed heterogeneously across regions and sectors high 2 train +7251 AR6_WGIII 25 17 In 2018, public and publicly mobilised private climate finance flows from developed to developing countries were below the collective goal under the UNFCCC and Paris Agreement to mobilise USD100 billion per year by 2020 in the context of meaningful mitigation action and transparency on implementation medium 1 train +7252 AR6_WGIII 25 18 Public and private finance flows for fossil fuels are still greater than those for climate adaptation and mitigation high 2 train +7253 AR6_WGIII 26 3 A gap remains between global GHG emissions in 2030 associated with the implementation of NDCs announced prior to COP26 and those associated with modelled mitigation pathways assuming immediate action (for quantification see Table SPM.1).26 The magnitude of the emissions gap depends on the global warming level considered and whether only unconditional or also conditional elements of NDCs27 are considered.28 (high confidence) {3.5, 4.2, Cross-Chapter Box 4 in Chapter 4} B.6.2 Global emissions in 2030 associated with the implementation of NDCs announced prior to COP26 are lower than the emissions implied by the original NDCs29 high 2 train +7254 AR6_WGIII 26 4 The original emissions gap has fallen by about 20% to one-third relative to pathways that limit warming to 2°C (>67%) with immediate action (category C3a in Table SPM.2), and by about 15–20% relative to pathways limiting warming to 1.5°C (>50%) with no or limited overshoot (category C1 in Table SPM.2) medium 1 train +7255 AR6_WGIII 27 1 Continued investments in unabated high-emitting infrastructure and limited development and deployment of low-emitting alternatives prior to 2030 would act as barriers to this acceleration and increase feasibility risks high 2 train +7256 AR6_WGIII 29 15 Without a strengthening of policies beyond those that are implemented by the end of 2020, GHG emissions are projected to rise beyond 2025, leading to a median global warming of 3.2 [2.2 to 3.5] °C by 210038, 39 medium 1 train +7257 AR6_WGIII 29 17 This compares with reductions of 43% [34–60%] by 2030 and 84% [73–98%] by 2050 in pathways that limit warming to 1.5°C (>50%) with no or limited overshoot (C1, Table SPM.2) (high confidence).41 In modelled pathways that return warming to 1.5°C (>50%) after a high overshoot,42 GHG emissions are reduced by 23% [0–44%] in 2030 and by 75% [62–91%] in 2050 (C2, Table SPM.2) high 2 train +7258 AR6_WGIII 29 18 Modelled pathways that are consistent with NDCs announced prior to COP26 until 2030 and assume no increase in ambition thereafter have higher emissions, leading to a median global warming of 2.8 [2.1–3.4] °C by 2100 medium 1 test +7259 AR6_WGIII 29 20 There are similar reductions of non-CO2 emissions by 2050 in both types of pathways: CH4 is reduced by 45% [25–70%]; N2O is reduced by 20% [–5 to +55%]; and F-gases are reduced by 85% [20–90%].43 Across most modelled pathways, this is the maximum technical potential for anthropogenic CH4 reductions in the underlying models high 2 train +7260 AR6_WGIII 29 21 Further emissions reductions, as illustrated by the IMP-SP pathway, may be achieved through changes in activity levels and/or technological innovations beyond those represented in the majority of the pathways medium 1 train +7261 AR6_WGIII 29 24 Pathways that exceed warming of >4°C (≥50%) (C8, SSP5-8.5, Table SPM.2) would imply a reversal of current technology and/or mitigation policy trends medium 1 train +7262 AR6_WGIII 29 25 Such warming could occur in emission pathways consistent with policies implemented by the end of 2020 if climate sensitivity is higher than central estimates high 2 train +7263 AR6_WGIII 36 12 In modelled pathways that limit warming to 2°C (>67%) (C3 category), there is no significant difference in warming by 2100 between those pathways that reach net zero GHGs (around 30%) and those that do not high 2 train +7264 AR6_WGIII 36 13 In pathways that limit warming to 2°C (>67%) or lower and that do reach net zero GHG, net zero GHG occurs around 10–40 years later than net zero CO2 emissions medium 1 train +7265 AR6_WGIII 37 1 As indicated by the ranges, choices in one sector can be compensated for by choices in another while being consistent with assessed warming levels.52 high 2 train +7266 AR6_WGIII 40 15 Depending on its availability, CCS could allow fossil fuels to be used longer, reducing stranded assets high 2 train +7267 AR6_WGIII 40 16 The combined global discounted value of the unburned fossil fuels and stranded fossil fuel infrastructure has been projected to be around USD1–4 trillion from 2015 to 2050 to limit global warming to approximately 2°C, and it will be higher if global warming is limited to approximately 1.5°C medium 1 train +7268 AR6_WGIII 40 20 About 50–80% of CH4 emissions from these fossil fuels could be avoided with currently available technologies at less than USD50 tCO2-eq–1 medium 1 train +7269 AR6_WGIII 42 4 With ambitious and immediate mitigation efforts, including high levels of electrification and improved energy and material efficiency, global consumption-based urban CO2 and CH4 emissions could be reduced to 3 GtCO2-eq in 2050 in the modelled scenario with very low GHG emissions (SSP1-1.9).56 (medium confidence) {8.3} C.6.2 The potential and sequencing of mitigation strategies to reduce GHG emissions will vary depending on a city’s land use, spatial form, development level, and state of urbanisation high 2 train +7270 AR6_WGIII 42 7 New and emerging cities will have significant infrastructure development needs to achieve high quality of life, which can be met through energy efficient infrastructures and services, and people-centred urban design high 2 train +7271 AR6_WGIII 42 8 For cities, three broad mitigation strategies have been found to be effective when implemented concurrently: (i) reducing or changing energy and material use towards more sustainable production and consumption; (ii) electrification in combination with switching to low-emission energy sources; and (iii) enhancing carbon uptake and storage in the urban environment, for example through bio-based building materials, permeable surfaces, green roofs, trees, green spaces, rivers, ponds and lakes. very high 3 test +7272 AR6_WGIII 44 1 Demand-focused interventions can reduce demand for all transport services and support the shift to more energy efficient transport modes medium 1 train +7273 AR6_WGIII 44 2 Electric vehicles powered by low-emissions electricity offer the largest decarbonisation potential for land-based transport, on a life cycle basis high 2 train +7274 AR6_WGIII 44 3 Sustainable biofuels can offer additional mitigation benefits in land-based transport in the short and medium term medium 1 train +7275 AR6_WGIII 44 4 Sustainable biofuels, low-emissions hydrogen, and derivatives (including synthetic fuels) can support mitigation of CO2 emissions from shipping, aviation, and heavy-duty land transport but require production process improvements and cost reductions medium 1 train +7276 AR6_WGIII 44 5 Many mitigation strategies in the transport sector would have various co-benefits, including air quality improvements, health benefits, equitable access to transportation services, reduced congestion, and reduced material demand high 2 train +7277 AR6_WGIII 44 8 In both categories of scenarios, the transport sector likely does not reach zero CO2 emissions by 2100 so negative emissions are likely needed to counterbalance residual CO2 emissions from the sector high 2 train +7278 AR6_WGIII 44 10 Investments in public inter- and intra-city transport and active transport infrastructure (e.g., bicycle and pedestrian pathways) can further support the shift to less GHG-intensive transport modes high 2 train +7279 AR6_WGIII 44 11 Combinations of systemic changes, including teleworking, digitalisation, dematerialisation, supply chain management, and smart and shared mobility may reduce demand for passenger and freight services across land, air, and sea high 2 train +7280 AR6_WGIII 44 12 Some of these changes could lead to induced demand for transport and energy services, which may decrease their GHG emissions reduction potential medium 1 train +7281 AR6_WGIII 44 14 Costs of electrified vehicles, including automobiles, two- and three-wheelers, and buses, are decreasing and their adoption is accelerating, but they require continued investments in supporting infrastructure to increase scale of deployment high 2 train +7282 AR6_WGIII 44 15 Advances in battery technologies could facilitate the electrification of heavy-duty trucks and complement conventional electric rail systems medium 1 train +7283 AR6_WGIII 44 17 Material and supply diversification strategies, energy and material efficiency improvements, and circular material flows can reduce the environmental footprint and material supply risks for battery production medium 1 train +7284 AR6_WGIII 44 18 Sourced sustainably and with low-GHG emissions feedstocks, bio-based fuels, blended or unblended with fossil fuels, can provide mitigation benefits, particularly in the short and medium term medium 1 train +7285 AR6_WGIII 44 19 Low-GHG emissions hydrogen and hydrogen derivatives, including synthetic fuels, can offer mitigation potential in some contexts and land-based transport segments medium 1 train +7286 AR6_WGIII 44 21 For aviation, such technologies include high energy density biofuels (high confidence), and low-emission hydrogen and synthetic fuels medium 1 train +7287 AR6_WGIII 44 22 Alternative fuels for shipping include low-emission hydrogen, ammonia, biofuels, and other synthetic fuels medium 1 train +7288 AR6_WGIII 44 23 Electrification could play a niche role for aviation and shipping for short trips (medium confidence) and can reduce emissions from port and airport operations high 2 train +7289 AR6_WGIII 44 24 Improvements to national and international governance structures would further enable the decarbonisation of shipping and aviation medium 1 train +7290 AR6_WGIII 44 25 Such improvements could include, for example, the implementation of stricter efficiency and carbon intensity standards for the sectors medium 1 train +7291 AR6_WGIII 44 28 Integrated transport and energy infrastructure planning and operations can enable sectoral synergies and reduce the environmental, social, and economic impacts of decarbonising the transport and energy sectors high 2 train +7292 AR6_WGIII 44 29 Technology transfer and financing can support developing countries leapfrogging or transitioning to low-emissions transport systems thereby providing multiple co-benefits high 2 train +7293 AR6_WGIII 45 13 Assisting countries to overcome barriers will help to achieve significant mitigation medium 1 train +7294 AR6_WGIII 46 18 SM.2, 8.4, 9.5, 10.2, 11.3, 11.4, Table 11.6, Box TS.12} C.10.4 Choice architecture62 can help end-users adopt, as relevant to consumers, culture and country contexts, low-GHG-intensive options such as balanced, sustainable healthy diets61 acknowledging nutritional needs; food waste reduction; adaptive heating and cooling choices for thermal comfort; building-integrated renewable energy; and electric light-duty vehicles, and shifts to walking, cycling, shared pooled and public transit; and sustainable consumption by intensive use of longer-lived repairable products high 2 train +7295 AR6_WGIII 46 19 Addressing inequality and many forms of status consumption63 and focusing on wellbeing supports climate change mitigation efforts high 2 train +7296 AR6_WGIII 48 4 CDR methods vary in terms of their maturity, removal process, time scale of carbon storage, storage medium, mitigation potential, cost, co-benefits, impacts and risks, and governance requirements high 2 train +7297 AR6_WGIII 48 5 Specifically, maturity ranges from lower maturity (e.g., ocean alkalinisation) to higher maturity (e.g., reforestation); removal and storage potential ranges from lower potential (<1 GtCO2 yr–1, e.g., blue carbon management) to higher potential (>3 GtCO2 yr–1, e.g., agroforestry); costs range from lower cost (e.g., USD-45–100 per tCO2 for soil carbon sequestration) to higher cost (e.g., USD100–300 per tCO2 for DACCS) medium 1 train +7298 AR6_WGIII 48 6 Estimated storage time scales vary from decades to centuries for methods that store carbon in vegetation and through soil carbon management, to 10,000 years or more for methods that store carbon in geological formations high 2 train +7299 AR6_WGIII 48 8 Afforestation, reforestation, improved forest management, agroforestry and soil carbon sequestration are currently the only widely practiced CDR methods high 2 train +7300 AR6_WGIII 48 10 Reforestation, improved forest management, soil carbon sequestration, peatland restoration and blue carbon management are examples of methods that can enhance biodiversity and ecosystem functions, employment and local livelihoods, depending on context high 2 train +7301 AR6_WGIII 48 11 In contrast, afforestation or production of biomass crops for BECCS or biochar, when poorly implemented, can have adverse socio-economic and environmental impacts, including on biodiversity, food and water security, local livelihoods and on the rights of Indigenous Peoples, especially if implemented at large scales and where land tenure is insecure high 2 train +7302 AR6_WGIII 48 12 Ocean fertilisation, if implemented, could lead to nutrient redistribution, restructuring of ecosystems, enhanced oxygen consumption and acidification in deeper waters medium 1 train +7303 AR6_WGIII 49 2 The global economic benefit of limiting warming to 2°C is reported to exceed the cost of mitigation in most of the assessed literature medium 1 train +7304 AR6_WGIII 49 12 The corresponding reductions in average annual global GDP growth over 2020–2050, in percentage points, are as follows: 0.09–0.14 (C1), 0.05–0.09 (C2), 0.03–0.07 (C4), 0.02–0.04 (C5).66 There are large variations in the modelled effects of mitigation on GDP across regions, depending notably on economic structure, regional emissions reductions, policy design and level of international cooperation67 high 2 train +7305 AR6_WGIII 49 13 Country-level studies also show large variations in the effect of mitigation on GDP depending notably on the level of mitigation and on the way it is achieved high 2 train +7306 AR6_WGIII 49 14 Macroeconomic implications of mitigation co-benefits and trade-offs are not quantified comprehensively across the above scenarios and depend strongly on mitigation strategies high 2 train +7307 AR6_WGIII 49 16 Models that incorporate the economic damages from climate change find that the global cost of limiting warming to 2°C over the 21st century is lower than the global economic benefits of reducing warming, unless: (i) climate damages are towards the low end of the range; or, (ii) future damages are discounted at high rates (medium confidence).68 Modelled pathways with a peak in global emissions between now and 2025 at the latest, compared to modelled pathways with a later peak in global emissions, entail more rapid near-term transitions and higher up-front investments, but bring long-term gains for the economy, as well as earlier benefits of avoided climate change impacts high 2 train +7308 AR6_WGIII 52 14 Electrification combined with low-GHG energy, and shifts to public transport can enhance health, employment, and can elicit energy security and deliver equity high 2 train +7309 AR6_WGIII 52 16 However, some industrial options could impose high costs medium 1 train +7310 AR6_WGIII 54 12 Limited economic, social and institutional resources often result in high vulnerability and low adaptive capacity, especially in developing countries medium 1 train +7311 AR6_WGIII 54 14 However, land and aquatic ecosystems can be adversely affected by some mitigation actions, depending on their implementation medium 1 train +7312 AR6_WGIII 54 15 Coordinated cross-sectoral policies and planning can maximise synergies and avoid or reduce trade-offs between mitigation and adaptation high 2 train +7313 AR6_WGIII 54 17 These options can also reduce flood risks, pressure on urban sewer systems, urban heat island effects, and can deliver health benefits from reduced air pollution high 2 train +7314 AR6_WGIII 54 19 For example, increasing urban density to reduce travel demand, could imply high vulnerability to heat waves and flooding high 2 train +7315 AR6_WGIII 54 28 Absence or limited resources in social and institutional systems can lead to poorly coordinated responses, thus reducing the potential for maximising mitigation and adaptation benefits, and increasing risk high 2 train +7316 AR6_WGIII 55 12 Applying just transition principles and implementing them through collective and participatory decision-making processes is an effective way of integrating equity principles into policies at all scales, in different ways depending on national circumstances medium 1 train +7317 AR6_WGIII 55 14 A multitude of actors, networks, and movements are engaged high 2 train +7318 AR6_WGIII 55 16 The consideration of ethics and equity can help address the uneven distribution of adverse impacts associated with 1.5°C and higher levels of global warming, in all societies high 2 train +7319 AR6_WGIII 55 17 Consideration of climate justice can help to facilitate shifting development pathways towards sustainability, including through equitable sharing of benefits and burdens of mitigation, increasing resilience to the impacts of climate change, especially for vulnerable countries and communities, and equitably supporting those in need high 2 train +7320 AR6_WGIII 56 13 This enables deployment in many regions high 2 train +7321 AR6_WGIII 56 14 While many mitigation options have environmental co-benefits, including improved air quality and reducing toxic waste, many also have adverse environmental impacts, such as reduced biodiversity, when applied at very large scale, for example very large scale bioenergy or large scale use of battery storage, that would have to be managed medium 1 train +7322 AR6_WGIII 56 15 Almost all mitigation options face institutional barriers that need to be addressed to enable their application at scale medium 1 train +7323 AR6_WGIII 56 22 Strengthened and coordinated near-term actions in cost-effective modelled global pathways that limit warming to 2°C (>67%) or lower, reduce the overall risks to the feasibility of the system transitions, compared to modelled pathways with relatively delayed or uncoordinated action.73 high 2 train +7324 AR6_WGIII 57 1 Policies that shift development pathways towards sustainability can broaden the portfolio of available mitigation responses, and enable the pursuit of synergies with development objectives medium 1 train +7325 AR6_WGIII 57 2 Actions can be taken now to shift development pathways and accelerate mitigation and transitions across systems high 2 train +7326 AR6_WGIII 57 4 Choices made by policymakers, citizens, the private sector and other stakeholders influence societies’ development pathways high 2 train +7327 AR6_WGIII 57 5 Actions that steer, for example, energy and land systems transitions, economy-wide structural change, and behaviour change, can shift development pathways towards sustainability74 medium 1 train +7328 AR6_WGIII 57 7 It can also facilitate the combination of mitigation and other development goals high 2 train +7329 AR6_WGIII 57 8 For example, measures promoting walkable urban areas combined with electrification and renewable energy can create health co-benefits from cleaner air and benefits from enhanced mobility high 2 train +7330 AR6_WGIII 57 9 Coordinated housing policies that broaden relocation options can make mitigation measures in transport more effective medium 1 train +7331 AR6_WGIII 57 13 In some situations, such as with innovation in technology at an early stage of development and some changes in behaviour towards low emissions, because the enabling conditions may take time to be established, action in the near term can yield accelerated mitigation in the mid-term medium 1 train +7332 AR6_WGIII 57 14 In other situations, the enabling conditions can be put in place and yield results in a relatively short time frame, for example the provision of energy related information, advice and feedback to promote energy saving behaviour high 2 train +7333 AR6_WGIII 57 16 Climate governance is most effective when it integrates across multiple policy domains, helps realise synergies and minimise trade-offs, and connects national and sub-national policymaking levels high 2 train +7334 AR6_WGIII 57 17 Effective and equitable climate governance builds on engagement with civil society actors, political actors, businesses, youth, labour, media, Indigenous Peoples and local communities medium 1 train +7335 AR6_WGIII 57 19 These functions can be promoted by climate-relevant laws, which are growing in number, or climate strategies, among others, based on national and sub-national context medium 1 train +7336 AR6_WGIII 57 20 Framework laws set an overarching legal basis, either operating through a target and implementation approach, or a sectoral mainstreaming approach, or both, depending on national circumstance medium 1 train +7337 AR6_WGIII 57 21 Direct national and sub-national laws that explicitly target mitigation and indirect laws that impact emissions through mitigation-related policy domains have both been shown to be relevant to mitigation outcomes medium 1 train +7338 AR6_WGIII 58 2 Complementary sub-national institutions tailor mitigation actions to local context and enable experimentation but can be limited by inequities and resource and capacity constraints high 2 train +7339 AR6_WGIII 58 3 Effective governance requires adequate institutional capacity at all levels high 2 train +7340 AR6_WGIII 58 10 These instruments could support deep emissions reductions and stimulate innovation if scaled up and applied more widely high 2 train +7341 AR6_WGIII 58 11 Policy packages that enable innovation and build capacity are better able to support a shift towards equitable low-emission futures than are individual policies high 2 train +7342 AR6_WGIII 58 12 Economy-wide packages, consistent with national circumstances, can meet short-term economic goals while reducing emissions and shifting development pathways towards sustainability medium 1 train +7343 AR6_WGIII 58 14 These instruments, and broad-based approaches including relevant economic instruments,75 are complementary high 2 train +7344 AR6_WGIII 58 15 Regulatory instruments that are designed to be implemented with flexibility mechanisms can reduce costs medium 1 train +7345 AR6_WGIII 58 16 Scaling up and enhancing the use of regulatory instruments, consistent with national circumstances, could improve mitigation outcomes in sectoral applications, including but not limited to renewable energy, land use and zoning, building codes, vehicle and energy efficiency, fuel standards, and low-emissions industrial processes and materials high 2 train +7346 AR6_WGIII 58 18 Where implemented, carbon pricing instruments have incentivised low-cost emissions reduction measures, but have been less effective, on their own and at prevailing prices during the assessment period, in promoting the higher-cost measures necessary for further reductions medium 1 train +7347 AR6_WGIII 58 19 Equity and distributional impacts of such carbon pricing instruments can be addressed by using revenue from carbon taxes or emissions trading to support low-income households, among other approaches high 2 train +7348 AR6_WGIII 58 20 Practical experience has informed instrument design and helped to improve predictability, environmental effectiveness, economic efficiency, distributional goals and social acceptance high 2 train +7349 AR6_WGIII 58 21 Removing fossil fuel subsidies would reduce emissions, improve public revenue and macroeconomic performance, and yield other environmental and sustainable development benefits; subsidy removal may have adverse distributional impacts especially on the most economically vulnerable groups which, in some cases can be mitigated by measures such as redistributing revenue saved, all of which depend on national circumstances (high confidence); fossil fuel subsidy removal is projected by various studies to reduce global CO2 emissions by 1–4%, and GHG emissions by up to 10% by 2030, varying across regions medium 1 train +7350 AR6_WGIII 59 4 They can meet short-term economic goals while reducing emissions and shifting development pathways towards sustainability medium 1 train +7351 AR6_WGIII 59 5 Infrastructure investments can be designed to promote low-emissions futures that meet development needs medium 1 train +7352 AR6_WGIII 59 7 There is no consistent evidence that current emission trading systems have led to significant emissions leakage, which can be attributed to design features aimed at minimising competitiveness effects, among other reasons medium 1 train +7353 AR6_WGIII 59 10 Scaling up mitigation financial flows can be supported by clear policy choices and signals from governments and the international community high 2 train +7354 AR6_WGIII 59 11 Accelerated international financial cooperation is a critical enabler of low-GHG and just transitions, and can address inequities in access to finance and the costs of, and vulnerability to, the impacts of climate change high 2 train +7355 AR6_WGIII 59 13 Mitigation investment gaps are wide for all sectors, and widest for the AFOLU sector in relative terms and for developing countries76 high 2 train +7356 AR6_WGIII 59 14 Financing and investment requirements for adaptation, reduction of losses and damages, general infrastructure, regulatory environment and capacity building, and climate-responsive social protection further exacerbate the magnitude of the challenges for developing countries to attract financing high 2 train +7357 AR6_WGIII 59 16 Barriers to the deployment of commercial finance from within the financial sector as well as macroeconomic considerations include: inadequate assessment of climate-related risks and investment opportunities; regional mismatch between available capital and investment needs; home bias factors; country indebtedness levels; economic vulnerability; and limited institutional capacities high 2 train +7358 AR6_WGIII 59 19 Scaled-up public grants for mitigation and adaptation funding for vulnerable regions, especially in Sub-Saharan Africa, would be cost-effective and have high social returns in terms of access to basic energy high 2 train +7359 AR6_WGIII 60 1 A coordinated effort to make the post-pandemic recovery sustainable and increased flows of financing over the next decade can accelerate climate action, including in developing regions and countries facing high debt costs, debt distress and macroeconomic uncertainty high 2 train +7360 AR6_WGIII 60 11 Challenges in and opportunities to enhance innovation cooperation exist, including in the implementation of elements of the UNFCCC and the Paris Agreement as per the literature assessed, such as in relation to technology development and transfer, and finance high 2 train +7361 AR6_WGIII 60 12 International cooperation on innovation works best when tailored to specific institutional and capability contexts, when it benefits local value chains, when partners collaborate equitably and on voluntary and mutually agreed terms, when all relevant voices are heard, and when capacity building is an integral part of the effort medium 1 train +7362 AR6_WGIII 60 13 Support to strengthen technological innovation systems and innovation capabilities, including through financial support in developing countries would enhance engagement in and improve international cooperation on innovation high 2 train +7363 AR6_WGIII 60 16 Transnational networks of city governments are leading to enhanced ambition and policy development and a growing exchange of experience and best practices medium 1 train +7364 AR6_WGIII 60 18 Agreements addressing ozone depletion and transboundary air pollution are contributing to mitigation, and in other areas, such as atmospheric emissions of mercury, may contribute to mitigation high 2 train +7365 AR6_WGIII 60 19 Trade rules have the potential to stimulate international adoption of mitigation technologies and policies, but may also limit countries’ ability to adopt trade-related climate policies medium 1 train +7366 AR6_WGIII 60 20 Current sectoral levels of ambition vary, with emission reduction aspirations in international aviation and shipping lower than in many other sectors medium 1 train +7367 AR6_WGIII 70 23 Original INDCs and NDCs refer to those submitted to the UNFCCC in 2015 and 2016.GHG emissions continued to rise to 2019, although the growth of global GHG emissions has slowed over the past decade high 2 train +7368 AR6_WGIII 71 1 Since 2010, GHG emissions have continued to grow reaching 59 ± 6.6 GtCO 2-eq in 2019,4 but the average annual growth in the last decade (1.3%, 2010–2019) was lower than in the previous decade (2.1%, 2000–2009) high 2 train +7369 AR6_WGIII 71 2 Average annual GHG emissions were 56 GtCO 2-eq yr –1 for 2010–2019 (the highest decadal average on record) growing by about 9.1 GtCO 2-eq yr –1 from the previous decade (2000–2009) high 2 train +7370 AR6_WGIII 71 8 The average annual emission levels of the last decade (2010–2019) were higher than in any previous decade for each group of greenhouse gases high 2 train +7371 AR6_WGIII 72 1 F-gases excluded from GHG emissions inventories such as chlorofluorocarbons and hydrochlorofluorocarbons are about the same size as those included high 2 train +7372 AR6_WGIII 75 10 Furthermore, for mitigation pathways that limit global warming to 2°C (>67%) or lower, using GWP100 to inform cost-effective abatement choices between gases would achieve such long-term temperature goals at close to least global cost within a few percent high 2 train +7373 AR6_WGIII 75 15 IFrom a mitigation perspective, this makes these metrics well-suited in principle to estimate the effect on the remaining carbon budget from more, or less, ambitious SLCF mitigation over multiple decades compared to a given reference scenario high 2 train +7374 AR6_WGIII 77 16 The lifestyle consumption emissions of the middle income and poorest citizens in emerging economies are between five and 50 times below their counterparts in high-income countries medium 1 train +7375 AR6_WGIII 77 17 Increasing inequality within a country can exacerbate dilemmas of 11 Bookkeeping models and dynamic global vegetation models.redistribution and social cohesion, and affect the willingness of the rich and poor to accept policies to protect the environment, and to accept and afford lifestyle changes that favour mitigation medium 1 train +7376 AR6_WGIII 77 21 Average annual GHG emissions growth during 2010–2019 slowed compared to the previous decade in energy supply (from 2.3% to 1.0%) and industry (from 3.4% to 1.4%, direct emissions only), but remained roughly constant at about 2% yr –1 in the transport sector high 2 train +7377 AR6_WGIII 77 22 Emission growth in AFOLU is more uncertain due to the high share of CO 2-LULUCF emissions medium 1 train +7378 AR6_WGIII 77 24 Accounting for this discrepancy would assist in assessing collective progress in a global stocktake high 2 train +7379 AR6_WGIII 77 29 The accounting method used will affect the assessment of collective progress in a global stocktake medium 1 train +7380 AR6_WGIII 77 34 This slowing of growth is attributable to further improvements in energy efficiency and reductions in the carbon intensity of energy supply driven by fuel switching from coal to gas, reduced expansion of coal capacity, particularly in Eastern Asia, and the increased use of renewables medium 1 train +7381 AR6_WGIII 78 12 Important drivers include technology transfer and cooperation, international policy and financial support, and harnessing synergies among technologies within a sustainable energy system perspective medium 1 train +7382 AR6_WGIII 78 13 A fast global low-carbon energy transition enabled by finance to facilitate low-carbon technology adoption in developing and particularly in least developed countries can facilitate achieving climate stabilisation targets high 2 train +7383 AR6_WGIII 79 1 The rapid deployment and unit cost decrease of modular technologies like solar, wind, and batteries have occurred much faster than anticipated by experts and modelled in previous mitigation scenarios, as shown in Figure TS.7 high 2 train +7384 AR6_WGIII 79 4 Emerging evidence since AR5 indicates that small-scale technologies (e.g., solar, batteries) tend to improve faster and be adopted more quickly than large-scale technologies (nuclear, CCS) medium 1 train +7385 AR6_WGIII 80 2 This compares to overall cumulative net CO 2 emissions until reaching net zero CO 2 of 510 (330–710) GtCO 2 in pathways that limit warming to 1.5°C (>50%) with no or limited overshoot, and 890 (640–1160) GtCO 2 in pathways that limit warming to 2°C (>67%) high 2 train +7386 AR6_WGIII 80 3 While most future CO 2 emissions from existing and currently planned fossil fuel infrastructure are situated in the power sector, most remaining fossil fuel CO 2 emissions in pathways that limit warming to 2°C (>67%) and below are from non-electric energy – most importantly from the industry and transportation sectors high 2 train +7387 AR6_WGIII 80 4 Decommissioning and reduced utilisation of existing fossil fuel installations in the power sector as well as cancellation of new installations are required to align future CO 2 emissions from the power sector with projections in these pathways high 2 train +7388 AR6_WGIII 81 2 While there is considerable literature on country-level mitigation pathways, including but not limited to NDCs, the country distribution of this literature is very unequal high 2 train +7389 AR6_WGIII 82 1 NDCs with unconditional and conditional elements14 lead to 53 (50–57) and 50 (47–55) GtCO 2-eq, respectively medium 1 train +7390 AR6_WGIII 82 9 This closes the emission gaps by about one third to 2°C and about 20% to 1.5°C compared to the original NDCs submitted in 2015/16 medium 1 train +7391 AR6_WGIII 82 10 An implementation gap also exists between the projected emissions with ‘current policies’ and the projected emissions resulting from the implementation of the unconditional and conditional elements of NDCs; this is estimated to be around 4 and 7 GtCO 2-eq in 2030, respectively medium 1 train +7392 AR6_WGIII 82 21 The literature does not adequately include demand-side options and systems analysis, and captures the impact from non-CO 2 GHGs medium 1 train +7393 AR6_WGIII 82 25 The way countries develop determines their capacity to accelerate mitigation and achieve other sustainable development objectives simultaneously medium 1 train +7394 AR6_WGIII 86 2 Policies can shift development pathways. There are examples of policies implemented in the pursuit of overall societal development objectives, such as job creation, macroeconomic stability, economic growth, and public health and welfare. In some countries, such policies are framed as part of a Just Transition, however, they can have major influence on mitigative capacity, and hence can be seen as tools to broaden mitigation options medium 1 test +7395 AR6_WGIII 86 14 Detailed design of mitigation policies is critical for distributional impacts and avoiding lock-in high 2 train +7396 AR6_WGIII 86 17 To this aim, mobilising a range of policies is preferable to single policy instruments high 2 train +7397 AR6_WGIII 86 22 Accelerated mitigation and shifting development pathways necessitates both redirecting existing financial flows from high- to low-emissions technologies and systems, and providing additional resources to overcome current financial barriers high 2 train +7398 AR6_WGIII 86 24 At the national level, public finance for actions promoting sustainable development helps broaden the scope of mitigation medium 1 train +7399 AR6_WGIII 86 25 Changes in behaviour and lifestyles are important to move beyond mitigation as incremental change, and when supporting shifts to more sustainable development pathways will broaden the scope of mitigation medium 1 train +7400 AR6_WGIII 86 31 Equity can be an important enabler, increasing the level of ambition for accelerated mitigation high 2 train +7401 AR6_WGIII 86 33 Transition pathways have distributional consequences such as large changes in employment and economic structure high 2 train +7402 AR6_WGIII 89 1 Differences between pathways typically represent choices that can steer the system in alternative directions through the selection of different combinations of response options high 2 train +7403 AR6_WGIII 89 4 The likelihood of limiting warming to 1.5°C with no or limited overshoot has dropped in AR6 WGIII compared to AR6 SR1.5 because global GHG emissions have risen since 2017, leading to higher near-term emissions (2030) and higher cumulative CO 2 emissions until the time of net zero medium 1 train +7404 AR6_WGIII 89 7 Pathways limiting warming to 2°C (>67%) reach 50% reductions in the 2040s and net zero CO 2 by the 2070s medium 1 train +7405 AR6_WGIII 90 8 Final energy demand in the absence of any new climate policies is projected to grow to around 480 to 750 EJ yr –1 in 2050 (compared to around 390 EJ yr –1 in 2015) medium 1 train +7406 AR6_WGIII 90 9 The highest emissions scenarios in the literature result in global warming of >5°C by 2100, based on assumptions of rapid economic growth and pervasive climate policy failures high 2 train +7407 AR6_WGIII 94 23 Only 30% of the pathways limiting warming to 2°C (>67%) or below reach net zero GHG emissions in the 21st century high 2 train +7408 AR6_WGIII 94 24 In those pathways reaching net zero GHGs, net zero GHGs is achieved around 10–20 years later than net zero CO 2 is achieved medium 1 train +7409 AR6_WGIII 94 26 Reaching and sustaining global net zero GHG emissions – when emissions are measured and reported in terms of GWP100 – results in a gradual decline in temperature high 2 train +7410 AR6_WGIII 94 28 Pathways that limit warming to 1.5°C (>50%) with no or limited overshoot entail CO 2 emissions reductions between 2019 and 2050 of around 77% (31–96%) for energy demand, around 115% (90–167%) for energy supply, and around 148% (94–387%) for AFOLU.16 In pathways that limit warming to 2°C (>67%), projected CO 2 emissions are reduced between 2019 and 2050 by around 49% for energy demand, 97% for energy supply, and 136% for AFOLU medium 1 train +7411 AR6_WGIII 94 32 In cost-effective mitigation pathways, the energy supply sector typically reaches net zero CO 2 before the economy as a whole, while the demand sectors reach net zero CO 2 later, if ever high 2 train +7412 AR6_WGIII 97 2 The use of bioenergy can lead to either increased or reduced emissions, depending on the scale of deployment, conversion technology, fuel displaced, and how, and where, the biomass is produced high 2 train +7413 AR6_WGIII 97 4 CDR deployment in pathways serves multiple purposes: accelerating the pace of emissions reductions, offsetting residual emissions, and creating the option for net negative CO 2 emissions in case temperature reductions need to be achieved in the long term high 2 train +7414 AR6_WGIII 97 6 CDR through some measures in AFOLU can be maintained for decades but not over the very long term because these sinks will ultimately saturate high 2 train +7415 AR6_WGIII 97 18 These estimates do not account for the economic benefits of avoided climate change impacts medium 1 train +7416 AR6_WGIII 97 24 This holds true even without accounting for benefits in other sustainable development dimensions or non-market damages from climate change medium 1 train +7417 AR6_WGIII 97 28 When aggregate economic benefits from avoided climate change impacts are accounted for, mitigation is a welfare-enhancing strategy high 2 test +7418 AR6_WGIII 98 1 Delayed global cooperation increases policy costs across regions, especially in those that are relatively carbon intensive at present high 2 train +7419 AR6_WGIII 98 2 Pathways with uniform carbon values show higher mitigation costs in more carbon-intensive regions, in fossil fuel-exporting regions, and in poorer regions high 2 train +7420 AR6_WGIII 98 7 Mitigation at the speed and scale required to limit warming to 2°C (>67%) or below implies deep economic and structural changes, thereby raising multiple types of distributional concerns across regions, income classes, and sectors high 2 train +7421 AR6_WGIII 99 12 Furthermore, avoided impacts for poorer households and poorer countries represent a smaller share in aggregate quantifications expressed in GDP terms or monetary terms, compared to their influence on well-being and welfare high 2 train +7422 AR6_WGIII 100 1 However, emerging evidence suggests that, even without accounting for co-benefits of mitigation on other sustainable development dimensions, the global benefits of pathways limiting warming to 2°C (>67%) outweigh global mitigation costs over the 21st century medium 1 train +7423 AR6_WGIII 101 3 TS.5.1 Energy A broad-based approach to deploying energy-sector mitigation options can reduce emissions over the next ten years and set the stage for still deeper reductions beyond 2030 high 2 train +7424 AR6_WGIII 101 8 In scenarios limiting warming to 1.5°C with no or limited overshoot (likely below 2°C), net electricity sector CO 2 emissions reach zero globally between 2045 and 2055 (2050 and 2080) high 2 train +7425 AR6_WGIII 101 10 This includes reduced fossil fuel consumption, increased production from low- and zero-carbon energy sources, and increased use of electricity and alternative energy carriers high 2 train +7426 AR6_WGIII 102 24 The impacts, however, are uncertain, particularly at the regional scale high 2 train +7427 AR6_WGIII 104 9 Multiple energy supply options are available to reduce emissions over the next decade high 2 train +7428 AR6_WGIII 104 14 It will not be possible to widely deploy all of these and other options without efforts to address the geophysical, environmental-ecological, economic, technological, socio-cultural, and institutional factors that can facilitate or hinder their implementation high 2 train +7429 AR6_WGIII 104 17 Energy systems can be integrated across district, regional, national, and international scales high 2 train +7430 AR6_WGIII 106 3 Energy-sector mitigation and efforts to achieve SDGs generally support one another, though there are important region- specific exceptions high 2 train +7431 AR6_WGIII 106 7 Advances in low-carbon energy resources and carriers such as next-generation biofuels, hydrogen produced from electrolysis, synthetic fuels, and carbon-neutral ammonia would substantially improve the economics of net zero energy systems medium 1 train +7432 AR6_WGIII 106 10 For most regions, per-capita urban emissions are lower than per-capita national emissions (excluding aviation, shipping and biogenic sources) very high 3 test +7433 AR6_WGIII 106 15 However, urbanisation can result in increased global GHG emissions through emissions outside the city’s boundaries very high 3 train +7434 AR6_WGIII 106 19 In 2015, urban emissions were estimated to be 25GtCO 2-eq (about 62% of the global share) and in 2020 were 29 GtCO 2-eq (67–72% of the global share).21 Around 100 of the highest-emitting urban areas account for approximately 18% of the global carbon footprint high 2 train +7435 AR6_WGIII 106 22 For 2000 to 2015, the urban emissions share increased from 28% to 38% in Africa, from 46% to 54% in Asia and Pacific, from 62% to 72% in Developed Countries, from 57% to 62% in Eastern Europe and West Central Asia, from 55% to 66% in Latin America and Caribbean, and from 68% to 69% in the Middle East high 2 train +7436 AR6_WGIII 108 4 Under a scenario with aggressive but not immediate urban mitigation policies to limit global warming to 2°C (>67%) (low emissions, SSP1-2.6), urban emissions could reach 17 GtCO 2-eq in 2050.23 (Figure TS.13) {8.3.4} Urban land areas could triple between 2015 and 2050, with significant implications for future carbon lock-in medium 1 train +7437 AR6_WGIII 109 11 Three broad mitigation strategies have been found to be effective in reducing emissions when implemented concurrently: (i) reducing or changing urban energy and material use towards more sustainable production and consumption across all sectors, including through compact and efficient urban forms and supporting infrastructure; (ii) electrification and switching to low-carbon energy sources; and (iii) enhancing carbon uptake and storage in the urban environment high 2 train +7438 AR6_WGIII 109 24 New and emerging cities will have significant infrastructure development needs to achieve high quality of life, which can be met through energy-efficient infrastructures and services, and people-centred urban design high 2 train +7439 AR6_WGIII 110 11 Transport-related emissions in developing regions of the world have increased more rapidly than in Europe or North America, a trend that is expected to continue in coming decades high 2 train +7440 AR6_WGIII 110 32 These same technologies and expanded use of available electric rail systems can support rail decarbonisation medium 1 train +7441 AR6_WGIII 111 3 Increased capacity for low-carbon hydrogen production would also be essential for hydrogen-based fuels to serve as an emissions reduction strategy high 2 train +7442 AR6_WGIII 111 5 Increased efficiency has been insufficient to limit the emissions from shipping and aviation, and natural gas-based fuels are expected to be inadequate to meet stringent decarbonisation goals for these segments high 2 train +7443 AR6_WGIII 111 7 Advanced biofuels could provide low-carbon jet fuel medium 1 test +7444 AR6_WGIII 111 8 The production of synthetic fuels using low-carbon hydrogen with CO2 captured through DACCS/BECCS could provide jet and marine fuels but these options still require demonstration at scale low 0 test +7445 AR6_WGIII 111 9 Ammonia produced with low-carbon hydrogen could also serve as a marine fuel medium 1 train +7446 AR6_WGIII 111 12 The scenarios literature projects continued growth in demand for freight and passenger services, particularly in developing countries in Africa and Asia high 2 train +7447 AR6_WGIII 111 15 While many global scenarios place greater reliance on emissions reduction in sectors other than transport, a quarter of the 1.5°C scenarios describe transport-related CO 2 emissions reductions in excess of 68% (relative to modelled 2020 levels) medium 1 train +7448 AR6_WGIII 111 33 Given the high degree of potential recyclability of lithium-ion batteries, a nearly closed-loop system in the future could mitigate concerns about critical mineral issues medium 1 train +7449 AR6_WGIII 112 1 Of this, 57% (6.8 GtCO 2-eq) were indirect emissions from off-site generation of electricity and heat, 24% (2.9 GtCO 2-eq) were direct emissions produced on-site and 18% (2.2 GtCO 2-eq) were embodied emissions from the production of cement and steel used in buildings high 2 train +7450 AR6_WGIII 113 1 Sufficiency measures tackle the causes of GHG emissions by limiting the demand for energy and materials over the lifecycle of buildings and appliances high 2 train +7451 AR6_WGIII 113 6 Density, compacity, bioclimatic design to optimise the use of nature-based solutions, multi-functionality of space through shared space and to allow for adjusting the size of buildings to the evolving needs of households, circular use of materials and repurposing unused existing buildings to avoid using virgin materials, optimisation of the use of buildings through lifestyle changes, use of the thermal mass of buildings to reduce thermal needs, and moving from ownership to usership of appliances, are among the sufficiency interventions implemented in leading municipalities high 2 train +7452 AR6_WGIII 113 7 At a global level, up to 17% of the mitigation potential in the buildings sector could be captured by 2050 through sufficiency interventions medium 1 train +7453 AR6_WGIII 113 9 The construction of high-performance buildings is expected to become a business-as- usual technology by 2050 with costs below USD20 tCO 2–1 in developed countries and below USD100 tCO 2–1 in developing countries medium 1 train +7454 AR6_WGIII 113 11 However, for the whole building stock they tend to be in cost intervals of USD–200 tCO 2–1 and >USD200 tCO 2–1 medium 1 train +7455 AR6_WGIII 113 12 Literature emphasises the critical role of the 2020–2030 decade in accelerating the learning of know-how and skills to reduce the costs and remove feasibility constraints for achieving high-efficiency buildings at scale and set the sector on the pathway to realise its full potential high 2 train +7456 AR6_WGIII 113 15 The complementarity and interdependency of measures leads to cost reductions, while optimising the mitigation potential achieved and avoiding the lock-in-effect medium 1 train +7457 AR6_WGIII 115 1 The lack of institutional capacity, especially in developing countries, and appropriate governance structures slow down the decarbonisation of the global building stock medium 1 train +7458 AR6_WGIII 115 6 However, this is not enough by far to close the investment gap high 2 train +7459 AR6_WGIII 115 8 Building energy codes represent the main regulatory instrument to reduce emissions from both new and existing buildings high 2 train +7460 AR6_WGIII 115 10 Building energy codes have proven to be effective if compulsory and combined with other regulatory instruments such as minimum energy performance standard for appliances and equipment, if the performance level is set at the level of the best available technologies in the market high 2 test +7461 AR6_WGIII 115 11 Market-based instruments such as carbon taxes with recycling of the revenues and personal or building carbon allowances could also contribute to fostering the decarbonisation of the building sector medium 1 train +7462 AR6_WGIII 115 13 Expected heatwaves will inevitably increase cooling needs to limit the health impacts of climate change medium 1 train +7463 AR6_WGIII 116 2 Mitigation actions in the building sector bring health gains through improved indoor air quality and thermal comfort, and have positive significant macro- and micro-economic effects, such as increased productivity of labour, job creation, reduced poverty, especially energy poverty, and improved energy security high 2 train +7464 AR6_WGIII 116 7 The Paris Agreement, the SDGs and the COVID-19 pandemic provide a new context for the evolution of industry and mitigation of industry greenhouse gas (GHG) emissions high 2 train +7465 AR6_WGIII 118 12 Producer, user, and regulator education, as well as innovation and commercialisation policy are needed medium 1 train +7466 AR6_WGIII 118 25 Pulp mills have access to biomass residues and by-products and in paper mills the use of process heat at low to medium temperatures allows for electrification high 2 train +7467 AR6_WGIII 119 18 In the energy sector, CO 2 emissions from biomass combustion for energy are recorded as an information item that is not included in the sectoral total emissions for the that sector.At the same time the capacity of the land to support these functions may be threatened by climate change high 2 train +7468 AR6_WGIII 119 20 At the same time managed and natural terrestrial ecosystems were a carbon sink, absorbing around one third of anthropogenic CO 2 emissions medium 1 train +7469 AR6_WGIII 119 24 If the responses of all managed and natural land to both anthropogenic environmental change and natural climate variability, estimated to be a gross sink of –12.5 ± 3.2 GtCO 2 yr –1 for the period 2010–2019, are added to land-use emissions, then land overall constituted a net sink of –6.6 ± 5.2 GtCO 2 yr –1 in terms of CO 2 emissions medium 1 train +7470 AR6_WGIII 119 26 The rate of deforestation, which accounts for 45% of total AFOLU emissions, has generally declined, while global tree cover and global forest-growing stock levels are likely increasing medium 1 train +7471 AR6_WGIII 119 30 Similarly, AFOLU N 2O emissions are increasing, dominated by agriculture, notably from manure application, nitrogen deposition, and nitrogen fertiliser use high 2 train +7472 AR6_WGIII 121 3 Climate change itself could reduce the mitigation potential from the AFOLU sector, although an increase in the capacity of natural sinks could occur despite changes in climate medium 1 train +7473 AR6_WGIII 121 4 The continued loss of biodiversity makes ecosystems less resilient to climate change extremes and this may further jeopardise the achievement of the AFOLU mitigation potentials indicated in this chapter high 2 train +7474 AR6_WGIII 121 7 While not as closely connected to the AFOLU sector as bioenergy, other renewable energy options can influence AFOLU activities in both synergistic and detrimental ways high 2 train +7475 AR6_WGIII 121 10 Deployment can be partly decoupled from additional land use, for example, use of organic waste and residues and integration of solar PV into buildings and other infrastructure high 2 train +7476 AR6_WGIII 121 11 Wind and solar power can coexist with agriculture in beneficial ways medium 1 train +7477 AR6_WGIII 121 21 The agriculture and forestry sectors can devise management approaches that enable biomass production and use for energy in conjunction with the production of food and timber, thereby reducing the conversion pressure on natural ecosystems medium 1 train +7478 AR6_WGIII 121 23 Integrated responses that contribute to mitigation, adaptation, and other land challenges will have greater likelihood of being successful high 2 train +7479 AR6_WGIII 121 31 This is well short of the more than USD400 billion yr –1 that is estimated to be necessary to deliver the up to 30% of global mitigation effort envisaged in deep mitigation scenarios medium 1 train +7480 AR6_WGIII 126 1 Some also include direct air CO 2 capture and storage (DACCS) high 2 train +7481 AR6_WGIII 129 5 Rapid and deep changes in demand make it easier for every sector to reduce GHG emissions in the near and mid-term high 2 train +7482 AR6_WGIII 133 14 Behavioural nudges promote easy behaviour change, for example, ‘Improve’ actions such as making investments in energy efficiency, but fail to motivate harder lifestyle changes high 2 train +7483 AR6_WGIII 137 3 TS.6.1 Policy and Institutions Long-term deep emission reductions, including the reduction of emissions to net zero, is best achieved through institutions and governance that nurture new mitigation policies, while at the same time reconsidering existing policies that support the continued emission of GHGs high 2 train +7484 AR6_WGIII 137 21 The awareness of co-benefits for the public increases support of climate policies high 2 train +7485 AR6_WGIII 137 27 They are able to experiment with climate solutions and can forge partnerships with the private sector and internationally to leverage enhanced climate action high 2 train +7486 AR6_WGIII 139 17 Subsidy removal may have adverse distributional impacts especially on the most economically vulnerable groups which, in some cases can be mitigated by measures such as redistributing revenue saved, all of which depend on national circumstances (high confidence); fossil fuel subsidy removal is projected by various studies (using alternative methodologies) to reduce global CO 2 emissions by 1–4%, and GHG emissions by up to 10% by 2030, varying across regions medium 1 train +7487 AR6_WGIII 139 28 There is no consistent evidence of significant emissions leakage or competitiveness effects between countries, including for emissions- intensive trade-exposed industries covered by emission-trading systems medium 1 train +7488 AR6_WGIII 143 2 AFOLU {Chapter 7}Regulation of land-use rights and practices have led to falling aggregate AFOLU-sector emissions.Regulation of land-use rights and practices, payments for ecosystem service, and offsets, have led to decreasing rates of deforestation medium 1 train +7489 AR6_WGIII 144 3 Other international agreements and institutions have led to avoided CO 2 emissions from land-use practices, as well as avoided emissions of some non-CO 2 greenhouse gases medium 1 train +7490 AR6_WGIII 144 5 Both new and pre-existing forms of cooperation are vital for achieving climate mitigation goals in the context of sustainable development high 2 train +7491 AR6_WGIII 144 6 While previous IPCC assessments have noted important synergies between the outcomes of climate mitigation and achieving sustainable development objectives, there now appear to be synergies between the two processes themselves medium 1 train +7492 AR6_WGIII 144 9 International cooperation helps countries achieve long-term mitigation targets when it supports development and diffusion of low-carbon technologies, often at the level of individual sectors, which can simultaneously lead to significant benefits in the areas of sustainable development and equity medium 1 train +7493 AR6_WGIII 144 20 The extent to which countries increase the ambition of their NDCs and ensure they are effectively implemented will depend in part on the successful implementation of the support mechanisms in the Paris Agreement, and in turn will determine whether the goals of the Paris Agreement are met high 2 train +7494 AR6_WGIII 144 24 Transnational partnerships and alliances involving non-state and sub-national actors are also playing a growing role in stimulating low-carbon technology diffusion and emissions reductions medium 1 train +7495 AR6_WGIII 145 1 Moreover, there are cases where international cooperation may be hindering mitigation efforts, namely evidence that trade and investment agreements, as well as agreements within the energy sector, impede national mitigation efforts medium 1 train +7496 AR6_WGIII 145 22 Complementarity in policies helps in the design of an optimal demand-side policy mix medium 1 test +7497 AR6_WGIII 145 25 Fundamental inequities in access to finance as well as finance terms and conditions, and countries’ exposure to physical impacts of climate change overall, result in a worsening outlook for a global Just Transition high 2 train +7498 AR6_WGIII 145 31 Other major challenges for commercial climate finance include: the mismatch between capital and investment needs, home bias31 considerations, differences in risk perceptions for regions, as well as limited institutional capacity to ensure safeguards are effective high 2 train +7499 AR6_WGIII 145 33 This increased awareness can support climate policy development and implementation high 2 train +7500 AR6_WGIII 145 41 This leaves high uncertainty, both near term (2021–30) and longer term (2021– 50), on the feasibility of an alignment of financial flows with the Paris Agreement goals high 2 train +7501 AR6_WGIII 146 5 When the perceived risks are too high, the misallocation of abundant savings persists and investors refrain from investing in infrastructure and industry in search of safer financial assets, even earning low or negative real returns high 2 train +7502 AR6_WGIII 147 2 To meet the needs for rapid deployment of mitigation options, global mitigation investments are expected to need to increase by the factor of three to six high 2 train +7503 AR6_WGIII 147 3 The gaps represent a major challenge for developing countries, especially Least-Developed Countries (LDCs), where flows have to increase by the factor of four to seven for specific sectors such as AFOLU, and for specific groups with limited access to, and high costs of, climate finance high 2 train +7504 AR6_WGIII 147 6 Soft costs for regulatory environment and institutional capacity, upstream funding needs as well as R&D and venture capital for development of new technologies and business models are often overlooked despite their critical role to facilitate the deployment of scaled-up climate finance high 2 train +7505 AR6_WGIII 147 9 This will particularly impact urban infrastructure and the energy and transport sectors high 2 train +7506 AR6_WGIII 147 10 A common understanding of debt sustainability and debt transparency, including negative implications of deferred climate investments on future GDP, and how stranded assets and resources may be compensated, has not yet been developed medium 1 train +7507 AR6_WGIII 147 13 A significant push for international climate finance access for vulnerable and poor countries is particularly important given these countries’ high costs of financing, debt stress and the impacts of ongoing climate change high 2 train +7508 AR6_WGIII 147 15 Approaches include de-risking investments, robust ‘green’ labelling and disclosure schemes, in addition to a regulatory focus on transparency and reforming international monetary system financial sector regulations medium 1 train +7509 AR6_WGIII 147 18 These relatively new labelled financial products will help by allowing a smooth integration into existing asset allocation models high 2 train +7510 AR6_WGIII 147 19 Green bond markets and markets for sustainable finance products have also increased significantly since AR5, but challenges nevertheless remain, in particular, there are concerns about ‘greenwashing’ and the limited application of these markets to developing countries high 2 train +7511 AR6_WGIII 147 25 Greater public-private cooperation can also encourage the private sector to increase and broaden investments, within a context of safeguards and standards, and this can be integrated into national climate change policies and plans high 2 train +7512 AR6_WGIII 147 27 It can also help address macroeconomic uncertainty and alleviate developing countries’ debt burden post-COVID-19 high 2 train +7513 AR6_WGIII 147 30 In addition to indirect and direct subsidies, the public sector’s role in addressing market failures, barriers, provision of information, and risk-sharing can encourage the efficient mobilisation of private sector finance high 2 train +7514 AR6_WGIII 147 34 Existing policy misalignments – for example, in fossil fuel subsidies – undermine the credibility of public commitments, reduce perceived transition risks and limit financial sector action high 2 train +7515 AR6_WGIII 148 3 This highlights the importance of trust in political leadership which, in turn, affects risk perception and ultimately financing costs high 2 train +7516 AR6_WGIII 148 7 A coordinated effort to green the post-pandemic recovery is also essential in countries facing much higher debt costs high 2 train +7517 AR6_WGIII 148 13 Technological innovation can also bring about new and improved ways of delivering services that are essential to human well-being high 2 test +7518 AR6_WGIII 148 15 Trade-offs include negative externalities’ – for instance, greater environmental pollution and social inequalities – rebound effects leading to lower net emission reductions or even increases in emissions, and increased dependency on foreign knowledge and providers high 2 train +7519 AR6_WGIII 148 18 This systemic view of innovation takes into account the role of actors, institutions, and their interactions, and can inform how innovation systems that vary across technologies, sectors and countries, can be strengthened high 2 train +7520 AR6_WGIII 150 1 The effectiveness of such international cooperation arrangements, however, depends on the way they are developed and implemented high 2 train +7521 AR6_WGIII 150 4 International diffusion of low-emission technologies is also facilitated by knowledge spillovers from regions engaged in clean R&D medium 1 train +7522 AR6_WGIII 150 6 The evidence on the role of intellectual property rights (IPR) in innovation is mixed. Some literature suggests that it is a barrier while other sources suggests that it is an enabler to the diffusion of climate-related technologies medium 1 test +7523 AR6_WGIII 150 11 These gaps could be filled by enhancing financial support for international technology cooperation, by strengthening cooperative approaches, and by helping build suitable capacity in developing countries across all technological innovation system functions high 2 train +7524 AR6_WGIII 150 13 For example, despite building a large market for mitigation technologies in developing countries, the lack of a systemic perspective in the implementation of the Clean Development Mechanism (CDM), operational since the mid-2000s, has only led to some technology transfer, especially to larger developing countries, but limited capacity building and minimal technology development medium 1 train +7525 AR6_WGIII 150 16 Addressing both sets of challenges simultaneously presents multiple and recurrent obstacles that systemic approaches to technological change could help resolve, provided they are well managed high 2 train +7526 AR6_WGIII 150 17 Obstacles include both entrenched power relations dominated by vested interests that control and benefit from existing technologies, and governance structures that continue to reproduce unsustainable patterns of production and consumption medium 1 train +7527 AR6_WGIII 150 23 Inspiration can be drawn from the global unit-cost reductions of solar PV, which were accelerated by a combination of factors interacting in a mutually reinforcing way across a limited group of countries high 2 train +7528 AR6_WGIII 152 5 At present, the understanding of both the direct and indirect impacts of digitalisation on energy use, carbon emissions and potential mitigation is limited medium 1 train +7529 AR6_WGIII 153 2 Looking at climate change from a justice perspective also means placing the emphasis on: (i) the protection of vulnerable populations from the impacts of climate change, (ii) mitigating the effects of low-carbon transformations, and (iii) ensuring an equitable decarbonised world high 2 train +7530 AR6_WGIII 153 4 Understanding the co-benefits and trade-offs associated with mitigation is key to understanding how societies prioritise among the various sectoral policy options medium 1 train +7531 AR6_WGIII 153 10 There are synergies and trade-offs between adaptation and mitigation as well as synergies and trade-offs with sustainable development high 2 train +7532 AR6_WGIII 153 27 There can be many synergies in urban areas between mitigation policies and the SDGs but capturing these depends on the overall planning of urban structures and on local integrated policies such as combining affordable housing and spatial planning with walkable urban areas, green electrification and clean renewable energy medium 1 test +7533 AR6_WGIII 158 1 The feasibility challenges associated with mitigation pathways are predominantly institutional and economic rather than technological and geophysical medium 1 train +7534 AR6_WGIII 159 8 A wide range of factors have been found to enable sustainability transitions, ranging from technological innovations to shifts in markets, and from policies and governance arrangements to shifts in belief systems and market forces high 2 train +7535 AR6_WGIII 159 10 Those same conditions that may serve to impede the transition (i.e., organisational structure, behaviour, technological lock-in) can also ‘flip’ to enable both the transition and the framing of sustainable development policies to create a stronger basis for policy support high 2 test +7536 AR6_WGIII 159 12 For example, rebuilding more sustainably after an extreme event, or renewed public debate about the drivers of social and economic vulnerability to multiple stressors medium 1 train +7537 AR6_WGIII 159 14 Climate change is the result of decades of unsustainable production and consumption patterns, as well as governance arrangements and political economic institutions that lock-in resource-intensive development patterns high 2 train +7538 AR6_WGIII 159 20 Strengthening different stakeholders’ ‘response capacities’ to mitigate and adapt to a changing climate will be critical for a sustainable transition high 2 train +7539 AR6_WGIII 168 9 It found this would require rapid and far-reaching transitions in energy, land, urban and infrastructure (including transport and buildings), and industrial systems high 2 train +7540 AR6_WGIII 168 13 Modelled direct mitigation costs of pathways to 1.5°C, with no/limited overshoot, span a wide range, but were typically three to four times higher than in pathways to 2°C high 2 test +7541 AR6_WGIII 229 1 Since 2010, GHG emissions have continued to grow, reaching 59 ± 6.6 GtCO 2-eq in 2019,1 but the average annual growth in the last decade (1.3%, 2010–2019) was lower than in the previous decade (2.1%, 2000–2009) high 2 test +7542 AR6_WGIII 229 2 Average annual GHG emissions were 56 ± 6.0 GtCO 2-eq yr –1 for the decade 2010–2019 growing by about 9.1 GtCO 2-eq yr –1 from the previous decade (2000–2009) – the highest decadal average on record high 2 train +7543 AR6_WGIII 229 4 The average annual emission levels of the last decade (2010–2019) were higher than in any previous decade for each group of GHGs high 2 train +7544 AR6_WGIII 229 8 F-gases excluded from GHG emissions inventories such as chlorofluorocarbons and hydrochlorofluorocarbons are about the same size as those included high 2 train +7545 AR6_WGIII 229 11 This growth outpaced the reduction in the use of energy per unit of GDP (–2% yr –1, globally) as well as improvements in the carbon intensity of energy (–0.3% yr –1) high 2 train +7546 AR6_WGIII 229 14 Emissions, however, have rebounded globally by the end of December 2020 medium 1 train +7547 AR6_WGIII 229 22 For comparison, the remaining carbon budget for keeping warming to 1.5°C with a 67% (50%) probability is about 400 (500) ± 220 GtCO 2 medium 1 train +7548 AR6_WGIII 230 10 Average annual GHG emissions growth during 2010 to 2019 slowed compared to the previous decade in energy supply (from 2.3% to 1.0%) and industry (from 3.4% to 1.4%, direct emissions only), but remained roughly constant at about 2% per year in the transport sector high 2 train +7549 AR6_WGIII 230 11 Emission growth in AFOLU is more uncertain due to the high share of CO 2-LULUCF emissions medium 1 train +7550 AR6_WGIII 230 14 Reductions in global carbon intensity by –0.2% yr –1 contributed further – reversing the trend during 2000 to 2009 (+0.2% yr –1) medium 1 train +7551 AR6_WGIII 231 9 Increasing inequality within a country can exacerbate dilemmas of redistribution and social cohesion, and affect the willingness of rich and poor to accept lifestyle changes for mitigation and policies to protect the environment (medium evidence, medium agreement) {2.6.1, 2.6.2, Figure 2.25} Estimates of future CO 2 emissions from existing fossil fuel infrastructures already exceed remaining cumulative net CO 2 emissions in pathways limiting warming to 1.5°C with no or limited overshoot high 2 train +7552 AR6_WGIII 231 11 This compares to overall cumulative net CO 2 emissions until reaching net zero CO 2 of 510 (330–710) Gt in pathways that limit warming to 1.5°C with no or limited overshoot, and 890 (640–1160) Gt in pathways that limit warming to 2°C (<67%) high 2 train +7553 AR6_WGIII 231 12 While most future CO 2 emissions from existing and currently planned fossil fuel infrastructure are situated in the power sector, most remaining fossil fuel CO 2 emissions in pathways that limit warming to 2°C (<67%) and below are from non-electric energy – most importantly from the industry and transportation sectors high 2 train +7554 AR6_WGIII 231 13 Decommissioning and reduced utilisation of existing fossil fuel installations in the power sector as well as cancellation of new installations are required to align future CO 2 emissions from the power sector with projections in these pathways high 2 train +7555 AR6_WGIII 231 16 Countries with a lower carbon pricing gap (higher carbon price) tend to be less carbon intensive medium 1 train +7556 AR6_WGIII 238 32 For methane, GWP100 implies a social discount rate of about 3–5% depending on the assumed damage function, whereas GWP20 implies a much higher discount rate, greater than 10% medium 1 train +7557 AR6_WGIII 239 5 These studies indicate that, for mitigation pathways that limit warming to 2°C (<67%) above pre-industrial levels or lower, using GWP100 to inform cost-effective abatement choices between gases would achieve such long-term temperature goals at close to least global cost within a few percent high 2 train +7558 AR6_WGIII 239 6 Using the dynamic GTP instead of GWP100 could reduce global mitigation costs by a few percent in theory (high confidence), but the ability to realise those cost savings depends on the temperature limit, policy foresight and flexibility in abatement choices as the weighting of SLCF emissions increases over time medium 1 train +7559 AR6_WGIII 239 35 The ability of these metrics to relate changes in emission rates of short-lived gases to cumulative CO 2 emissions makes them well-suited, in principle, to estimating the effect on the remaining carbon budget from more, or less, ambitious SLCF mitigation over multiple decades compared to a given reference scenario high 2 train +7560 AR6_WGIII 240 6 A fundamental change in GHG emission metrics used to monitor achievement of existing emission targets could therefore inadvertently change their intended climate outcomes or ambition, unless existing emission targets are re-evaluated at the same time very high 3 train +7561 AR6_WGIII 240 20 GHG emissions levels in 2019 were higher compared to 10 and 30 years earlier high 2 train +7562 AR6_WGIII 240 21 GHG emissions growth slowed compared to the previous decade high 2 train +7563 AR6_WGIII 242 6 The CO 2-FFI share in total CO 2-eq emissions has plateaued at about 65% in recent years and its growth has slowed considerably since AR5 high 2 train +7564 AR6_WGIII 242 18 Starting in the spring of 2020 a major break in global emissions trends was observed due to lockdown policies implemented in response to the COVID-19 pandemic high 2 train +7565 AR6_WGIII 242 28 Daily emissions, estimated based on activity and power-generation data, declined substantially compared to 2019 during periods of economic lockdown, particularly in April 2020 – as shown in Figure 2.6 – but rebounded by the end of 2020 medium 1 train +7566 AR6_WGIII 243 10 Cumulative CO 2 emissions since 1850 reached 2400 ± 240 GtCO 2 in 2019 high 2 train +7567 AR6_WGIII 243 14 Emissions in the last decade are about the same size as the remaining carbon budget of 400 ± 220 (500, 650) GtCO 2 for limiting global warming to 1.5°C and between one-third and half the 1150 ± 220 (1350, 1700) GtCO 2 for limiting global warming below 2°C with a 67% (50%, 33%) probability, respectively medium 1 train +7568 AR6_WGIII 245 38 GHG and CO 2-FFI levels diverge starkly between countries and regions high 2 train +7569 AR6_WGIII 262 11 Residential buildings accounted for the majority of this sector’s emissions (64%, 6.3 GtCO 2-eq, including both direct and indirect emissions), followed by non-residential buildings (35%, 3.5 GtCO 2-eq) high 2 train +7570 AR6_WGIII 263 31 Road transport passenger and freight emissions represented by far the largest component and source of this growth (6.1 GtCO 2-eq, 69% of all transport emissions in 2019) high 2 train +7571 AR6_WGIII 263 36 North America’s absolute and per capita transport emissions are the highest amongst world regions, but those of South, South-East and East Asia are growing the fastest high 2 train +7572 AR6_WGIII 267 48 Technological change has had a mitigating effect on emissions over the long term and is central to efforts to achieving climate goals high 2 train +7573 AR6_WGIII 267 49 Progress since AR5 shows that multiple low-carbon technologies are improving and falling in cost (high confidence); technology adoption is reaching substantial shares, and small-scale technologies are particularly promising on both medium 1 train +7574 AR6_WGIII 267 51 However, the historical pace of technological change is still insufficient to catalyse a complete and timely transition to a low-carbon energy system: technological change needs to accelerate high 2 train +7575 AR6_WGIII 268 31 Reasons that these exemplars could be applied more broadly in the future include: growing urgency on climate change, shifting motivation from price response to proactive resource scarcity, and an increase in the likelihood of technological breakthroughs medium 1 train +7576 AR6_WGIII 269 9 Technological change has been at the core of transitions, but is best understood as part of a system in which social aspects are crucial medium 1 train +7577 AR6_WGIII 269 19 The observed pace of these changes and the likelihood of their continuation support the arguments in the previous section that future energy transitions are likely to occur more quickly than in the past medium 1 train +7578 AR6_WGIII 269 20 Among the most notable are solar PV, wind power, and batteries high 2 train +7579 AR6_WGIII 269 23 The future potential for PV and batteries seems especially promising given that neither industry has yet begun to adopt alternative materials with attractive properties as the cost reductions and performance improvements associated with the current generation of each technology continue medium 1 train +7580 AR6_WGIII 269 29 Solar PV is by far the most dynamic technology, and its cost since AR5 has continued on its steep decline at about the same rate of change as before AR5, but now costs are well within the range of fossil fuels high 2 train +7581 AR6_WGIII 269 41 Smaller unit sizes, sometimes referred to as ‘granularity’, tend to be associated with faster learning rates medium 1 train +7582 AR6_WGIII 270 9 The large residual has motivated studies, which find that small-scale technologies provide opportunities for rapid change, but they do not make rapid change inevitable; a supportive context, including supportive policy and complementary technologies, can stimulate more favourable technology outcomes high 2 train +7583 AR6_WGIII 270 10 There is also evidence that small technologies not only learn but become adopted faster than large technologies medium 1 train +7584 AR6_WGIII 271 2 Cost reductions facilitate adoption, which generates opportunities for further cost reductions through a process of learning by doing medium 1 train +7585 AR6_WGIII 271 11 In contrast, IAMs indicate that they expect much lower rates of growth in future years for the technologies that have been growing fastest in recent years (wind and solar), without strong evidence for why this should occur.The overall pattern shows that IAMs expect growth in small-scale renewables to fall to less than half of their recent pace, and large- scale CCS to more than double from the limited deployment assessed high 2 train +7586 AR6_WGIII 279 26 Hence, cumulative net CO 2 emissions to limit warming to 2°C (<67%) or lower could already be exhausted by current and planned fossil fuel infrastructure high 2 train +7587 AR6_WGIII 281 2 Therefore, our overall assessment of these available lines of evidence strongly emphasises the importance of decommissioning, reduced utilisation of existing power sector infrastructure, as well as continued cancellation of new power sector infrastructures in order to limit warming to well below 2°C high 2 train +7588 AR6_WGIII 282 44 Despite such evidence, studies of carbon pricing find that additional policies are often needed to stimulate sufficient emissions reductions in transportation medium 1 train +7589 AR6_WGIII 283 18 Public transit can reduce vehicle travel and lower GHG emissions by reducing the number of trips taken by private vehicles and the length of those trips medium 1 train +7590 AR6_WGIII 283 27 Bike- and car-sharing programmes can reduce GHG emissions medium 1 train +7591 AR6_WGIII 351 7 Final energy demand in the absence of any new climate policies is projected to grow to around 480–750 EJ yr–1 in 2050 (compared to around 390 EJ in 2015) medium 1 train +7592 AR6_WGIII 351 8 The highest emissions scenarios in the literature result in global warming of >5°C by 2100, based on assumptions of rapid economic growth and pervasive climate policy failures high 2 train +7593 AR6_WGIII 351 10 The likelihood of limiting warming to 1.5°C with no or limited overshoot has dropped in AR6 compared to the Special Report on Global Warming of 1.5°C (SR1.5) because global GHG emissions have risen since the time SR1.5 was published, leading to higher near-term emissions (2030) and higher cumulative CO 2 emissions until the time of net zero medium 1 train +7594 AR6_WGIII 352 1 Pathways limiting warming to 2°C (>67%) reach 50% reductions in the 2040s and net zero CO 2 by 2070s medium 1 train +7595 AR6_WGIII 352 15 Only 30% of the pathways limiting warming to 2°C (>67%) or lower reach net zero GHG emissions in the 21st century high 2 train +7596 AR6_WGIII 352 16 In those pathways reaching net zero GHGs, it is achieved around 10 to 40 years later than for net zero CO 2 medium 1 train +7597 AR6_WGIII 352 18 Reaching and sustaining global net zero GHG emissions, measured in terms of GWP-100, results in a gradual decline of temperature high 2 train +7598 AR6_WGIII 352 21 In pathways limiting warming to 2°C (>67%), projected CO 2 emissions are reduced between 2019 and 2050 by around 49% for energy demand, 97% for energy supply, and 136% for AFOLU medium 1 train +7599 AR6_WGIII 352 25 In cost-effective mitigation pathways, the energy-supply sector typically reaches net zero CO 2 before the economy as a whole, while the demand sectors reach net zero CO 2 later, if ever high 2 train +7600 AR6_WGIII 352 33 The use of bioenergy can lead to either increased or reduced emissions, depending on the scale of deployment, conversion technology, fuel displaced, and how/ where the biomass is produced high 2 train +7601 AR6_WGIII 353 1 CDR deployment in pathways serves multiple purposes: accelerating the pace of emissions reductions, offsetting residual emissions, and creating the option for net negative CO 2 emissions in case temperature reductions need to be achieved in the long term high 2 train +7602 AR6_WGIII 353 3 CDR through some measures in AFOLU can be maintained for decades but not in the very long term because these sinks will ultimately saturate high 2 train +7603 AR6_WGIII 353 14 Yet, these estimates do not account for the economic benefits of avoided climate change impacts medium 1 train +7604 AR6_WGIII 353 16 This holds true even without accounting for benefits in other sustainable development dimensions or non- market damages from climate change medium 1 train +7605 AR6_WGIII 353 20 When aggregate economic benefits from avoided climate change impacts are accounted for, mitigation is a welfare-enhancing strategy high 2 train +7606 AR6_WGIII 353 24 Mitigation has implications for employment through multiple channels, each of which impacts geographies, sectors and skill categories differently medium 1 train +7607 AR6_WGIII 353 26 Delayed global cooperation increases policy costs across regions, especially in those that are relatively carbon intensive at present high 2 train +7608 AR6_WGIII 353 27 Pathways with uniform carbon values show higher mitigation costs in more carbon-intensive regions, in fossil fuel exporting regions and in poorer regions high 2 train +7609 AR6_WGIII 353 28 Aggregate quantifications expressed in GDP or monetary terms undervalue the economic effects on households in poorer countries; the actual effects on welfare and well-being are comparatively larger high 2 train +7610 AR6_WGIII 353 29 Mitigation at the speed and scale required to limit warming to 2°C (>67%) or lower implies deep economic and structural changes, thereby raising multiple types of distributional concerns across regions, income classes and sectors high 2 train +7611 AR6_WGIII 354 3 Dimensions with potential trade-offs include food, employment, water stress, and biodiversity, which come under pressure from large-scale CDR deployment, energy affordability/ access, and mineral-resource extraction high 2 train +7612 AR6_WGIII 354 10 Feasibility challenges are transient and concentrated in the next two to three decades high 2 train +7613 AR6_WGIII 390 8 However, delaying emissions reductions, or more limited emissions reductions in one sector or region, involves compensating reductions in other sectors or regions if warming is to be limited high 2 train +7614 AR6_WGIII 390 20 So, while pathways indicate some flexibility in emissions reductions across sectors, all pathways involve substantial CO 2 emissions reductions in all sectors and regions high 2 train +7615 AR6_WGIII 395 15 All pathways that limit warming to 2°C (>67%) or lower show substantial reductions in fossil fuel consumption and a near elimination of the use of coal without CCS high 2 train +7616 AR6_WGIII 395 34 Stringent emissions reductions at the level required to limit warming to 2°C (>67%) or 1.5°C are achieved through increased electrification of end use, resulting in increased electricity generation in all pathways high 2 train +7617 AR6_WGIII 400 1 Limiting warming to likely 2°C (>67%) or lower can result in large-scale transformation of the land surface high 2 train +7618 AR6_WGIII 400 26 Some AFOLU mitigation options can enhance vegetation and soil carbon stocks such as reforestation, restoration of degraded ecosystems, protection of ecosystems with high carbon stocks and changes to agricultural land management to increase soil carbon high 2 train +7619 AR6_WGIII 400 31 The time scales associated with these options indicate that carbon sinks in terrestrial vegetation and soil systems can be maintained or enhanced so as to contribute towards long-term mitigation high 2 train +7620 AR6_WGIII 400 34 In the very long term (the latter part of the century and beyond), it will become more challenging to continue to enhance vegetation and soil carbon stocks, so that the associated carbon sinks could diminish or even become sources high 2 train +7621 AR6_WGIII 401 8 Anthropogenic land CO 2 emissions and removals in IAM pathways cannot be directly compared with those reported in national GHG inventories high 2 train +7622 AR6_WGIII 402 2 The assessment in this section shows that if mitigation ambitions in NDCs announced prior to COP262,18 are followed until 2030, leading to estimated emissions of 47–57 GtCO 2-eq in 203019 (Section 4.2.2), it is no longer possible to limit warming to 1.5°C (>50%) with no or limited overshoot high 2 train +7623 AR6_WGIII 402 4 It would also strongly increase mitigation challenges to limit warming to 2°C (>67%) high 2 train +7624 AR6_WGIII 402 7 A lock-in to fossil fuel-intensive production systems (carbon lock-in) will increase the societal, economic and political strain of a rapid low-carbon transition after 2030 high 2 train +7625 AR6_WGIII 413 1 However, it is robust across modelling frameworks that the marginal abatement cost of carbon increases for lower temperature categories, with a higher increase in the short term than in the longer term (Figure 3.32, left panel) high 2 train +7626 AR6_WGIII 413 10 Pathways that correspond to NDCs announced prior to COP26 in 2030 and strengthen action after 2030 imply higher marginal abatement costs of carbon in the longer run than pathways with stronger immediate global mitigation action (Figure 3.32b) high 2 train +7627 AR6_WGIII 415 12 Mitigation pathways with early emissions reductions represent higher mitigation costs in the short-run but bring long-term gains for the economy compared to delayed transition pathways high 2 train +7628 AR6_WGIII 418 15 The AR6 WGII finds that variation in estimated global economic impacts increases with warming in all methodologies, indicating higher risk in terms of economic impacts at higher temperatures high 2 train +7629 AR6_WGIII 418 17 The AR6 WGII’s assessment finds that the lack of comparability between methodologies does not allow for identification of robust ranges of global economic impact estimates high 2 train +7630 AR6_WGIII 418 18 Further, AR6 WGII identifies evaluating and reconciling differences in methodologies as a research priority for facilitating use of the different lines of evidence high 2 train +7631 AR6_WGIII 422 34 Through avoiding impacts, mitigation thus reduces economic inequalities and poverty high 2 train +7632 AR6_WGIII 422 46 Targeted policy areas must include healthy nutrition, sustainable consumption and production, inequality and poverty alleviation, air quality and international collaboration high 2 train +7633 AR6_WGIII 422 47 Lower energy demand enables synergies between mitigation and sustainability, with lower reliance on CDR high 2 train +7634 AR6_WGIII 423 9 Reducing warming reduces the population impacted by all impact categories shown high 2 train +7635 AR6_WGIII 423 16 Areas of co-benefits include human health, ambient air pollution and other specific kinds of pollution, while areas of trade-off include food access, habitat loss and mineral resources medium 1 train +7636 AR6_WGIII 425 21 Recent research shows that mitigation is compatible with reductions in inequality and poverty (Box 3.6).Lower demand – for example, for energy and land-intensive consumption such as meat – represents a synergistic strategy for achieving ambitious climate mitigation without compromising Sustainable Development Goals high 2 train +7637 AR6_WGIII 426 7 Climate change will reduce crop yields, increase food insecurity, and negatively influence nutrition and mortality. Climate mitigation will thus reduce these impacts, and hence reduce food insecurity high 2 test +7638 AR6_WGIII 427 2 Mitigating climate change while ensuring that food security is not adversely affected requires a range of different strategies and interventions high 2 train +7639 AR6_WGIII 427 18 Taken together, climate changes will reduce crop yields, increase food insecurity and influence nutrition and mortality high 2 train +7640 AR6_WGIII 427 28 Climate change also affects the occurrence of and exposure to hydrological extremes high 2 train +7641 AR6_WGIII 427 32 Climate models project increases in drought risk very high 3 test +7642 AR6_WGIII 427 34 The effect of climate change on water availability and hydrological extremes varies by region high 2 train +7643 AR6_WGIII 427 44 Limiting warming could reduce water-related risks high 2 train +7644 AR6_WGIII 427 50 The effect of climate change on water depends on the climate model, the hydrological model, and the metric high 2 train +7645 AR6_WGIII 427 54 However, the effect of socio-economic development could be larger than the effect of climate change high 2 train +7646 AR6_WGIII 427 78 Bioenergy and BECCS can increase water withdrawals and water consumption high 2 train +7647 AR6_WGIII 428 14 The effect of mitigation on water quality depends on the mitigation option, its implementation, and the aspect of quality considered high 2 train +7648 AR6_WGIII 430 29 Marine and coastal ecosystems Marine ecosystems are being affected by climate change and growing non-climate pressures including temperature change, acidification, land-sourced pollution, sedimentation, resource extraction and habitat destruction high 2 train +7649 AR6_WGIII 430 32 The danger or warming and acidification to coral reefs, rocky shores and kelp forests is well established high 2 train +7650 AR6_WGIII 430 49 Stringent mitigation that includes reductions in demand for animal-based foods and food waste could also relieve pressures on land use and biodiversity high 2 train +7651 AR6_WGIII 479 25 Working Group II contribution to this Assessment finds that the overall extent of adaptation-related responses in human systems is low high 2 train +7652 AR6_WGIII 521 33 IPCC SR1.5 found that sustainable development pathways to 1.5°C broadly support and often enable transformations and that ‘sustainable development has the potential to significantly reduce systemic vulnerability, enhance adaptive capacity, and promote livelihood security for poor and disadvantaged populations high 2 train +7653 AR6_WGIII 566 5 Rapid and deep changes in demand make it easier for every sector to reduce greenhouse gas (GHG) emissions in the short and medium term high 2 train +7654 AR6_WGIII 567 7 Behavioural nudges promote easy behaviour change, for example ‘Improve’ actions such as making investments in energy efficiency, but fail to motivate harder lifestyle changes high 2 train +7655 AR6_WGIII 582 45 The study shows that of all demand- side option effects on well-being, 79% are positive, 18% are neutral (or not relevant or specified), and only 3% are negative high 2 train +7656 AR6_WGIII 582 50 Well-being improvements are most notable in health, air, and energy high 2 train +7657 AR6_WGIII 582 53 Food (medium confidence), mobility (high confidence), and water medium 1 train +7658 AR6_WGIII 582 57 An exception is economic stability, suggesting that demand-side options generate stable opportunities to participate in economic activities high 2 train +7659 AR6_WGIII 589 11 Around 931 million tonnes of food waste was generated in 2019 globally, 61% of which came from households, 26% from food service and 13% from retail.Demand-side mitigations are achieved through changing Socio- cultural factors, Infrastructure use and Technology adoption by various social actors in urban and other settlements, food choice and waste management high 2 train +7660 AR6_WGIII 589 17 The technical mitigation potential of food loss and waste reductions globally has been estimated at 0.1–5.8 GtCO 2-eq high 2 train +7661 AR6_WGIII 590 1 The estimated technical potential for GHG emissions reductions associated with shifts to sustainable healthy diets is 0.5–8 GtCO 2-eq high 2 train +7662 AR6_WGIII 616 3 They also offer potential levers to change normative ideas and social practices in order to achieve extensive emissions cuts high 2 train +7663 AR6_WGIII 618 7 Countermovement coalitions work to oppose climate mitigation high 2 train +7664 AR6_WGIII 622 25 Although there is variability across innovations, sectors, and countries, the transitions literature distinguishes four phases, characterised by generic core processes and challenges: (i) emergence, (ii) early adaptation, (i) diffusion, (iv) stabilisation high 2 train +7665 AR6_WGIII 622 29 These four phases do not imply that transitions are linear, teleological processes, because set-backs or reversals may occur as a result of learning processes, conflicts, or changing coalitions very high 3 train +7666 AR6_WGIII 622 32 In the first phase, radical innovations emerge in peripheral niches, where researchers, inventors, social movement organisations or community activists dedicate time and effort to their development high 2 train +7667 AR6_WGIII 737 13 Replacing old coal with new coal facilities is inconsistent with limiting warming to 2°C or below high 2 train +7668 AR6_WGIII 742 10 Global fossil fuel subsidies represent more than half of total energy subsidies with predominantly adverse environmental, economic, and social effects high 2 train +7669 AR6_WGIII 742 23 Fossil fuel subsidies most commonly pursue non-climate objectives, for example, enhanced access to energy sources high 2 test +7670 AR6_WGIII 743 10 Costs have declined by 62% since 2015 high 2 train +7671 AR6_WGIII 743 11 Key areas for continued improvement are grid integration and non- module costs for rooftop systems high 2 train +7672 AR6_WGIII 743 12 Most deployment is now utility-scale high 2 train +7673 AR6_WGIII 743 13 Global future potential is not limited by solar irradiation, but by grid integration needed to address its variability, as well as access to finance, particularly in developing countries high 2 train +7674 AR6_WGIII 743 14 The global technical potential of direct solar energy far exceeds that of any other renewable energy resource and is well beyond the total amount of energy needed to support ambitious mitigation over the current century high 2 train +7675 AR6_WGIII 743 25 In many parts of the world, the cost of electricity from PV is below the cost of electricity generated from fossil fuels; in some, it is below the operating costs of electricity generated from fossil fuels high 2 train +7676 AR6_WGIII 745 2 The share of total costs of PV-intensive electricity systems attributed to integration costs has been increasing but can be reduced by enhancing grid flexibility high 2 train +7677 AR6_WGIII 745 23 The most important ways to minimise PV’s impact on the environment lie in recycling materials at end of life and making smart land-use decisions medium 1 train +7678 AR6_WGIII 745 60 Material demand for PV will likely increase substantially to limit warming to well below 2°C, but PV materials are widely available, have possible substitutes, and can be recycled medium 1 train +7679 AR6_WGIII 746 11 Many alternative PV materials are improving in efficiency and stability, providing longer-term pathways for continued PV costs reductions and better performance high 2 train +7680 AR6_WGIII 746 68 Solar energy elicits favourable public responses in most countries high 2 train +7681 AR6_WGIII 747 1 Costs have declined by 18% and 40% on land and offshore since 2015 (high confidence), and further reductions can be expected by 2030 medium 1 train +7682 AR6_WGIII 747 2 Wind power is increasingly competitive with other forms of electricity generation and is the low-cost option in many applications, and critical areas for continued improvement are technology advancements and economies of scale high 2 test +7683 AR6_WGIII 747 3 Global future potential is primarily limited by onshore land availability in wind power-rich areas, lack of supporting infrastructure, grid integration, and access to finance (especially in developing countries) (high confidence).Energy from wind is abundant, and the estimated technical potentials surpass the total amount of energy needed to limit warming to well below 2°C high 2 train +7684 AR6_WGIII 749 5 Electricity from onshore wind is less expensive than electricity generated from fossil fuels in a growing number of markets high 2 train +7685 AR6_WGIII 750 6 Wind power plants pose relatively low environmental impact, but sometimes locally significant ecological effects high 2 train +7686 AR6_WGIII 750 23 Public support for onshore and particularly offshore wind energy is generally high, although people may oppose specific wind farm projects high 2 train +7687 AR6_WGIII 751 2 However, the future mitigation potential of hydropower depends on minimising environmental and social impacts during the planning stages, reducing the risks of dam failures, and modernising the ageing hydropower fleet to increase generation capacity and flexibility high 2 train +7688 AR6_WGIII 751 12 Hydropower is a mature technology with locally adapted solutions high 2 train +7689 AR6_WGIII 752 15 Hydroelectric power plants may pose serious environmental and societal impacts high 2 train +7690 AR6_WGIII 752 29 Yet, public support for hydro seems to differ for existing and new projects high 2 train +7691 AR6_WGIII 752 54 Doing so will require improvements in managing construction of reactor designs that hold the promise of lower costs and broader use medium 1 train +7692 AR6_WGIII 752 55 At the same time, nuclear power continues to be affected by cost overruns, high upfront investment needs, challenges with final disposal of radioactive waste, and varying public acceptance and political support levels high 2 train +7693 AR6_WGIII 752 56 There are sufficient resources for substantially increasing nuclear deployment medium 1 train +7694 AR6_WGIII 753 1 There are several possible nuclear technology options for the period from 2030 to 2050 medium 1 train +7695 AR6_WGIII 753 20 Nuclear power costs vary substantially across countries high 2 train +7696 AR6_WGIII 753 28 Cost-cutting opportunities, such as design standardisation and innovations in construction approaches, are expected to make SMRs competitive against large reactors by 2040 (Rubio and Tricot 2016) medium 1 train +7697 AR6_WGIII 753 34 However, new reactor designs with passive and enhanced safety systems reduce the risk of such accidents significantly high 2 train +7698 AR6_WGIII 753 55 Nuclear power continues to suffer from limited public and political support in some countries high 2 train +7699 AR6_WGIII 754 18 Most of the countries that might introduce nuclear power in the future for their climate change mitigation benefits do not envision developing their own full fuel cycle, significantly reducing any risks that might be linked to proliferation (IAEA 2014, 2019).6.4.2.5 Carbon Dioxide Capture, Utilisation and Storage Since AR5, there have been increased efforts to develop novel platforms that reduce the energy penalty associated with CO 2 capture, develop CO 2 utilisation pathways as a substitute to geologic storage, and establish global policies to support CCS high 2 train +7700 AR6_WGIII 754 21 The theoretical global geologic storage potential is about 10,000 GtCO 2, with more than 80% of this capacity existing in saline aquifers medium 1 train +7701 AR6_WGIII 754 42 CO2 utilisation (CCU) – instead of geologic storage – could present an alternative method of decarbonisation high 2 train +7702 AR6_WGIII 755 16 Existing post-combustion approaches relying on absorption are technologically ready for full-scale deployment high 2 train +7703 AR6_WGIII 755 31 CO2 capture costs present a key challenge, remaining higher than USD50 tCO 2–1 for most technologies and regions; novel technologies could help reduce some costs high 2 train +7704 AR6_WGIII 755 37 The potential for such reductions is limited in several regions due to low sink availability, but it could jump-start initial investments medium 1 train +7705 AR6_WGIII 756 17 Because CCS always adds cost, policy instruments are required for it to be widely deployed high 2 train +7706 AR6_WGIII 756 28 Its long-term role in low-carbon energy systems is therefore uncertain high 2 train +7707 AR6_WGIII 758 11 Several technological and institutional barriers exist for large-scale BECCS implementation, including large energy requirements for CCS, limit and cost of biomass supply and geologic sinks for CO 2 in several regions, and cost of CO 2 capture technologies high 2 train +7708 AR6_WGIII 759 59 Fossil fuel reserves have continued to rise because of advanced exploration and utilisation techniques high 2 train +7709 AR6_WGIII 759 60 A fraction of these available reserves can be used consistent with mitigation goals when paired with CCS opportunities in close geographical proximity high 2 train +7710 AR6_WGIII 760 36 The cost of producing electricity from fossil sources has remained roughly the same with some regional exceptions while the costs of producing transport fuels has gone down significantly high 2 train +7711 AR6_WGIII 760 49 Owing to climate constraints, these may become stranded, causing considerable economic impacts high 2 train +7712 AR6_WGIII 761 4 Oil and coal consistently rank among the least preferred energy sources in many countries high 2 train +7713 AR6_WGIII 761 26 The geophysical potential of geothermal resources is 1.3 to 13 times the global electricity demand in 2019 medium 1 train +7714 AR6_WGIII 762 16 Public awareness and knowledge of geothermal energy is relatively low high 2 train +7715 AR6_WGIII 762 17 Geothermal energy is evaluated as less acceptable than other renewable energy sources such as solar and wind, but is preferred over fossil and nuclear energy, and in some studies, over hydroelectric energy high 2 train +7716 AR6_WGIII 764 4 Rather than incremental planning, strategic energy system planning can help minimise long-term mitigation costs high 2 train +7717 AR6_WGIII 766 13 No single, sufficiently mature energy storage technology can provide all the required grid services – a portfolio of complementary technologies working together can provide the optimum solution high 2 train +7718 AR6_WGIII 768 47 TES can be much cheaper than batteries and has the unique ability to capture and reuse waste heat and cold, enabling the efficiency of many industrial, buildings, and domestic processes to be greatly improved high 2 train +7719 AR6_WGIII 770 17 At present hydrogen has limited applications – mainly being produced onsite for the creation of methanol and ammonia (IEA 2019c), as well as in refineries.Low- or zero-carbon produced hydrogen is not currently competitive for large-scale applications, but it is likely to have a significant role in future energy systems, due to its wide-range of applications high 2 train +7720 AR6_WGIII 774 21 Contextual factors, such as physical and climate conditions, infrastructure, available technology, regulations, institutions, culture, and financial conditions define the costs and benefits of mitigation options that enable or inhibit their adoption high 2 train +7721 AR6_WGIII 779 1 However, globally climate change impacts on electricity generation – including hydro, wind and solar power potentials – should not compromise climate mitigation strategies high 2 train +7722 AR6_WGIII 779 10 Areas with decreased runoff are anticipated to experience reduced hydropower production and increased water conflict among different economic activities high 2 train +7723 AR6_WGIII 780 14 Changing wind variability may have a small-to-modest impact on backup energy and storage needs low 0 test +7724 AR6_WGIII 781 13 Climate change will shift the suitable range for bioenergy towards higher latitudes, but the net change in the total suitable area is uncertain high 2 train +7725 AR6_WGIII 781 24 The effect of climate change on bioenergy crop yields will increase in high latitudes low 0 test +7726 AR6_WGIII 782 1 Peak load may increase more than energy consumption, and the changing spatial and temporal load patterns can impact transmission and needs for storage, demands- side management, and peak-generating capacity high 2 train +7727 AR6_WGIII 783 38 Their environmental impacts of renewable energy production are mostly confined to areas close to the production sources and have been shown to be trivial compared to the mitigation benefits of renewable energy high 2 train +7728 AR6_WGIII 785 36 High-fidelity models and analyses are needed to assess the economic and environmental characteristics and the feasibility of many aspects of net-zero or net-negative emissions energy systems high 2 train +7729 AR6_WGIII 785 43 Configurations of net-zero energy systems will vary by region but are likely to share several common characteristics high 2 train +7730 AR6_WGIII 785 46 The precise quantity of fossil fuels will largely depend on the relative costs of such fuels, electrification, alternative fuels, and CDR (Section 6.6.2.4) in the energy system high 2 train +7731 AR6_WGIII 785 60 There is considerable flexibility regarding the overall quantity of liquid and gaseous fuels that will be required in net-zero energy systems high 2 train +7732 AR6_WGIII 787 4 There are many possible configurations and technologies for zero- or net- negative-emissions electricity systems high 2 train +7733 AR6_WGIII 787 17 Based on their increasing economic competitiveness, VRE technologies, especially wind and solar power, will likely comprise large shares of many regional generation mixes high 2 train +7734 AR6_WGIII 787 42 Energy storage will be increasingly important in net-zero energy systems, especially in systems with shares of VRE high 2 train +7735 AR6_WGIII 788 11 It is technically feasible to use very high renewable shares (e.g., above 75% of annual regional generation) to meet hourly electricity demand under a range of conditions, especially when VRE options, notably wind and solar, are complemented by other resources high 2 train +7736 AR6_WGIII 789 15 Several end uses, such as passenger transportation (light-duty electric vehicles, two and three wheelers, buses, rail) as well as building energy uses (lighting, cooling) are likely to be electrified in net- zero energy systems high 2 train +7737 AR6_WGIII 789 24 Regions endowed with cheap and plentiful low-carbon electricity resources (wind, solar, hydropower) are likely to emphasise electrification, while those with substantial bioenergy resources or availability of other liquid fuels might put less emphasis on electrification, particularly in hard-to-electrify end uses medium 1 train +7738 AR6_WGIII 789 28 Electrification of most buildings services, with the possible exception of space heating in extreme climates, is expected in net-zero energy systems high 2 train +7739 AR6_WGIII 789 37 Energy systems that are 100% renewable (including all parts of the energy sector, and not only electricity generation) raise a range of technological, regulatory, market, and operational challenges that make their competitiveness uncertain high 2 train +7740 AR6_WGIII 790 7 A significant share of transportation, especially road transportation, is expected to be electrified in net-zero energy systems high 2 train +7741 AR6_WGIII 790 8 In road transportation, two- and three-wheelers, light-duty vehicles (LDVs), and buses, are especially amenable to electrification, with more than half of passenger LDVs expected to be electrified globally in net-zero energy systems medium 1 train +7742 AR6_WGIII 790 24 A non-trivial number of industry applications could be electrified as a part of a net-zero energy system, but direct electrification of heavy industry applications such as cement, primary steel manufacturing, and chemical feedstocks is expected to be challenging medium 1 train +7743 AR6_WGIII 790 67 Costs are the main barrier to synthesis of net-zero emissions fuels high 2 train +7744 AR6_WGIII 792 2 Net-zero energy systems will use energy more efficiently than those of today high 2 train +7745 AR6_WGIII 792 9 Characterising efficiency of net-zero energy systems is problematic due to measurement challenges high 2 train +7746 AR6_WGIII 793 5 Measurement issues notwithstanding, virtually all studies that address net-zero energy systems assume improved energy intensity in the future high 2 train +7747 AR6_WGIII 793 13 Net-zero energy systems will be characterised by greater efficiency and more efficient use of energy across all sectors high 2 train +7748 AR6_WGIII 793 34 Carbon-neutral energy systems are likely to be more interconnected than those of today high 2 train +7749 AR6_WGIII 798 10 Warming cannot be limited to well below 2°C without rapid and deep reductions in energy system GHG emissions high 2 train +7750 AR6_WGIII 800 10 To decarbonise most cost- effectively, global net CO 2 emissions from electricity generation will likely reach zero before the rest of the energy sector medium 1 train +7751 AR6_WGIII 801 10 Limiting warming to well below 2°C requires a rapid and dramatic increase in energy produced from low- or zero-carbon sources high 2 train +7752 AR6_WGIII 801 14 Low- and zero-carbon sources produce 97–99% of global electricity in 2050 in scenarios limiting warming to 1.5°C (>50%) with no or limited overshoot and 93–97% in scenarios limiting warming to 2°C (>67%) (Figure 6.29) medium 1 train +7753 AR6_WGIII 809 10 Strong path dependencies, even in early formative stages, can have lasting impacts on energy systems, producing inertia that cuts across technological, economic, institutional and political dimensions high 2 train +7754 AR6_WGIII 810 1 These current investments combined with emissions from proposed fossil infrastructure exceed the emissions required to limit warming to 1.5°C medium 1 train +7755 AR6_WGIII 810 39 Existing policies and the NDCs are insufficient to prevent an increase in fossil infrastructure and associated carbon lock-in high 2 train +7756 AR6_WGIII 810 58 Past and present energy sector investments have created technological, institutional, and behavioural path dependencies aligned towards coal, oil, and natural gas high 2 train +7757 AR6_WGIII 811 3 Box 6.13 | Stranded Assets Limiting warming to 2°C (>67%) or lower will result in stranded assets high 2 train +7758 AR6_WGIII 811 16 About 200 GW of fossil fuel electricity generation per year will likely need to be retired prematurely after 2030 to limit warming to 2°C, even if countries achieve their Nationally Determined Contributions (NDCs) medium 1 train +7759 AR6_WGIII 811 30 If warming is limited to 2°C, the discounted economic impacts of stranded assets, including unburned fossil reserves, could be as high as USD1–4 trillion from 2015 through 2050 (USD10–20 trillion in undiscounted terms) medium 1 train +7760 AR6_WGIII 811 55 Global coal consumption without CCS needs to be largely eliminated by 2040–2050 to limit warming to 1.5°C (>50%), and 2050–2060 to limit warming to 2°C (>67%) high 2 train +7761 AR6_WGIII 811 56 New investments in coal-fired electricity without CCS are inconsistent with limiting warming to 2°C (>67%) or lower high 2 train +7762 AR6_WGIII 812 13 Natural gas may remain part of energy systems through mid-century, both for electricity generation and use in industry and buildings, and particularly in developed economies, even if warming is limited to 2°C (>67%) or lower medium 1 train +7763 AR6_WGIII 813 23 While policy interventions are necessary to achieve low-carbon energy system transitions, appropriate governance frameworks are crucial to ensure policy implementation high 2 train +7764 AR6_WGIII 813 44 Well-designed policy mixes can support the pursuit of multiple policy goals, target effectively different types of imperfections and framework conditions and take into account the technological, economical, and societal situation high 2 train +7765 AR6_WGIII 814 18 However, comprehensive evaluation of policy mixes requires a broader set of criteria that reflect different considerations, such as broader goals (e.g., SDGs) and the feasibility of policies high 2 train +7766 AR6_WGIII 814 32 Potential future policies are difficult to evaluate due to methodological challenges high 2 train +7767 AR6_WGIII 816 26 Cost reductions in key technologies, particularly in electricity and light-duty transport, have increased the economic attractiveness of near-term low-carbon energy system transitions high 2 train +7768 AR6_WGIII 816 27 The near-term, economic outcomes of low-carbon energy system transitions in some sectors and regions may be on par with or superior to those of an emissions-intensive future high 2 train +7769 AR6_WGIII 816 47 The long-term economic characteristics of low-emissions energy system transitions are not well understood, and they depend on policy design and implementation along with future costs and availability of technologies in key sectors (e.g., process heat, long- distance transport), and the ease of electrification in end-use sectors high 2 train +7770 AR6_WGIII 817 1 Advances in low-carbon energy resources and carriers such as next-generation biofuels, hydrogen produced from electrolysis, synthetic fuels, and carbon-neutral ammonia would substantially improve the economics of net-zero energy systems high 2 train +7771 AR6_WGIII 817 8 Improving efficiency and energy conservation will promote sustainable consumption and production of energy and associated materials (SDG 12) high 2 train +7772 AR6_WGIII 818 18 Phasing out fossil fuels in favour of low-carbon sources is likely to have considerable SDG benefits, particularly if trade-offs such as unemployment to fossil fuel workers are minimised high 2 train +7773 AR6_WGIII 818 31 CDR and CCS can create significant land and water trade-offs high 2 train +7774 AR6_WGIII 819 11 Greater energy system integration (Sections 6.4.3 and 6.6.2) would enhance energy-SDG synergies while eliminating trade-offs associated with deploying mitigation options high 2 train +7775 AR6_WGIII 886 3 The rapid deployment of AFOLU measures is essential in all pathways staying within the limits of the remaining budget for a 1.5°C target high 2 train +7776 AR6_WGIII 886 6 At the same time the capacity of the land to support these functions may be threatened by climate change itself high 2 train +7777 AR6_WGIII 886 8 At the same time managed and natural terrestrial ecosystems were a carbon sink, absorbing around one third of anthropogenic CO 2 emissions medium 1 train +7778 AR6_WGIII 886 12 If the managed and natural responses of all land to both anthropogenic environmental change and natural climate variability, estimated to be a gross sink of –12.5 ± 3.2 GtCO 2 yr–1 for the period 2010–2019, are included with land use emissions, then land overall, constituted a net sink of –6.6 ± 5.2 GtCO 2 yr–1 in terms of CO 2 emissions medium 1 train +7779 AR6_WGIII 886 14 The rate of deforestation has generally declined, while global tree cover and global forest growing stock levels are likely increasing medium 1 train +7780 AR6_WGIII 886 22 AFOLU CH 4 emissions continue to increase (high confidence), the main source of which is enteric fermentation from ruminant animals high 2 train +7781 AR6_WGIII 886 23 Similarly, AFOLU N 2O emissions are increasing, dominated by agriculture, notably from manure application, nitrogen deposition, and nitrogen fertiliser use high 2 train +7782 AR6_WGIII 887 7 Assisting countries to overcome barriers will help to achieve significant short-term mitigation medium 1 train +7783 AR6_WGIII 887 10 Climate change could also emerge as a barrier to AFOLU mitigation, although the IPCC AR6 WGI contribution to AR6 indicated that an increase in the capacity of natural sinks may occur, despite changes in climate medium 1 train +7784 AR6_WGIII 887 11 The continued loss of biodiversity makes ecosystems less resilient to climate change extremes and this may further jeopardise the achievement of the AFOLU mitigation potentials indicated in this chapter (IPCC AR6 WGII and IPBES) high 2 train +7785 AR6_WGIII 887 15 Poorly planned deployment of biomass production and afforestation options for in-forest carbon sequestration may conflict with environmental and social dimensions of sustainability high 2 train +7786 AR6_WGIII 887 16 The global technical CDR potential of BECCS by 2050 (considering only the technical capture of CO 2 and storage underground) is estimated at 5.9 mean (0.5–11.3) GtCO 2 yr–1, of which 1.6 (0.8–3.5) GtCO 2 yr–1 is available at below USD100 tCO 2–1 medium 1 train +7787 AR6_WGIII 887 17 Bioenergy and other bio-based products provide additional mitigation through the substitution of fossil fuels fossil-based products high 2 train +7788 AR6_WGIII 887 20 The agriculture and forestry sectors can devise management approaches that enable biomass production and use for energy in conjunction with the production of food and timber, thereby reducing the conversion pressure on natural ecosystems medium 1 train +7789 AR6_WGIII 887 22 Such risks can best be managed if AFOLU mitigation is pursued in response to the needs and perspectives of multiple stakeholders to achieve outcomes that maximise synergies while limiting trade-offs medium 1 train +7790 AR6_WGIII 887 26 Integrated responses that contribute to mitigation, adaptation, and other land challenges will have greater likelihood of being successful high 2 train +7791 AR6_WGIII 887 28 Globally, the AFOLU sector has so far contributed modestly to net mitigation, as past policies have delivered about 0.65 GtCO 2 yr–1 of mitigation during 2010–2019 or 1.4% of global gross emissions high 2 train +7792 AR6_WGIII 887 29 The majority (>80%) of emission reduction resulted from forestry measures high 2 train +7793 AR6_WGIII 887 34 To date USD0.7 billion yr–1 is estimated to have been spent on AFOLU mitigation, which is well short of the more than USD400 billion yr–1 that is estimated to be necessary to deliver the up to 30% of global mitigation effort envisaged in deep mitigation scenarios medium 1 test +7794 AR6_WGIII 888 5 In addition to funding, these factors include governance, institutions, long-term consistent execution of measures, and the specific policy setting high 2 train +7795 AR6_WGIII 888 8 It would also assist in assessing collective progress in a global stocktake high 2 train +7796 AR6_WGIII 888 13 To enable a like-with- like comparison, the remaining cumulative global CO 2 emissions budget can be adjusted medium 1 train +7797 AR6_WGIII 888 19 These options could support more specific NDCs with 2 Bookkeeping models and dynamic global vegetation models.AFOLU measures that enable mitigation while also contributing to biodiversity conservation, ecosystem functioning, livelihoods for millions of farmers and foresters, and many other Sustainable Development Goals (SDGs) high 2 train +7798 AR6_WGIII 889 19 Since the IPCC Fifth Assessment Report (AR5), the share of AFOLU to anthropogenic GHG emissions had remained largely unchanged at 13–21% of total GHG emissions medium 1 train +7799 AR6_WGIII 894 1 While there is low agreement in the trend of global AFOLU CO 2 emissions over the past few decades (Section 7.2.2), they have remained relatively constant medium 1 train +7800 AR6_WGIII 894 7 This showed a removal of –12.5 ± 3.2 GtCO 2 yr–1 medium 1 train +7801 AR6_WGIII 902 40 Following changes in land conditions, CO 2, CH 4 and N 2O fluxes are quickly mixed into the atmosphere and dispersed, resulting in the biogeochemical effects being dominated by the biophysical effects at local scales high 2 train +7802 AR6_WGIII 911 12 If implemented at appropriate scales and in a sustainable manner, land-based mitigation practices have the capacity to reduce emissions and sequester billions of tonnes of carbon from the atmosphere over coming decades, while also preserving or enhancing biodiversity, water quality and supply, air quality, soil fertility, food and wood security, livelihoods, resilience to droughts, floods and other natural disasters, and positively contributing to ecosystem health and human well-being high 2 train +7803 AR6_WGIII 911 25 The SRCCL assessed the full range of technical, economic and sustainability mitigation potentials in AFOLU for the period 2030–2050 and identified reduced deforestation and forest degradation to have greatest potential for reducing supply-side emissions (0.4 to 5.8 GtCO 2-eq yr–1) (high confidence) followed by combined agriculture measures, 0.3 to 3.4 GtCO 2-eq yr–1 medium 1 train +7804 AR6_WGIII 911 27 For the demand-side estimates, shifting towards healthy, sustainable diets (0.7 to 8.0 GtCO 2-eq yr–1) (high confidence) had the highest potential, followed by reduced food loss and waste (0.8 to 4.5 GtCO 2-eq yr–1) high 2 train +7805 AR6_WGIII 911 28 Measures with greatest potential for CDR were afforestation/reforestation (0.5 to 10.1 GtCO 2-eq yr–1) (medium confidence), soil carbon sequestration in croplands and grasslands (0.4 to 8.6 GtCO 2-eq yr–1) (medium confidence) and BECCS (0.4 to 11.3 GtCO 2-eq yr–1) medium 1 train +7806 AR6_WGIII 913 30 The feasibility of implementing AFOLU mitigation measures, including those with multiple co-benefits, depends on varying economic, technological, institutional, socio- cultural, environmental and geophysical barriers high 2 train +7807 AR6_WGIII 916 1 Reducing deforestation and forest degradation represents one of the most effective options for climate change mitigation, with technical potential estimated at 0.4–5.8 GtCO 2 yr–1 by 2050 high 2 train +7808 AR6_WGIII 916 28 Based on studies since AR5, the technical mitigation potential for reducing deforestation and degradation is significant, providing 4.5 (2.3–7) GtCO 2 yr–1 globally by 2050, of which 3.4 (2.3–6.4) GtCO 2 yr–1 is available at below USD100 tCO 2–1 medium 1 train +7809 AR6_WGIII 916 59 The SRCCL remained with a reported wide range of mitigation potential for A/R of 0.5–10.1 GtCO 2 yr–1 by 2050 medium 1 train +7810 AR6_WGIII 917 74 In the SRCCL, forest management activities have the potential to mitigate 0.4–2.1 GtCO 2-eq yr–1 by 2050 medium 1 train +7811 AR6_WGIII 919 10 In the SRCCL, fire management is among the nine options that can deliver medium-to-large benefits across multiple land challenges (climate change mitigation, adaptation, desertification, land degradation, and food security) high 2 train +7812 AR6_WGIII 921 19 In the SRCCL (Chapters 2 and 6), it was estimated that avoided peat impacts could deliver 0.45–1.22 GtCO 2-eq yr–1 technical potential by 2030–2050 medium 1 train +7813 AR6_WGIII 922 22 According the SRCCL, peatland restoration could deliver technical mitigation potentials of 0.15 – 0.81GtCO 2-eq yr–1 by 2030–2050 low 0 train +7814 AR6_WGIII 922 27 Peatlands are highly sensitive to climate change high 2 train +7815 AR6_WGIII 934 16 Like AE, it is likely that RA can contribute to mitigation, the extent to which is currently unclear and by its case-specific design, will vary medium 1 train +7816 AR6_WGIII 934 32 There is evidence that CA can contribute to mitigation, but its contribution is depended on multiple factors including climate and residue returns high 2 train +7817 AR6_WGIII 936 43 The SRCCL reported a technical CDR potential for BECCS at 0.4–11.3 GtCO 2 yr–1 medium 1 train +7818 AR6_WGIII 938 16 Based on studies to date, the technical net CDR potential of BECCS (including LUC and other supply chain emissions, but excluding energy carrier substitution) by 2050 is 5.9 (0.5–11.3) GtCO 2 yr–1 globally, of which 1.6 (0.5–3.5) GtCO 2 yr–1 is available at below USD100 tCO 2–1 medium 1 train +7819 AR6_WGIII 939 40 Shifting to sustainable healthy diets has large potential to achieve global GHG mitigation targets as well as public health and environmental benefits high 2 train +7820 AR6_WGIII 939 43 A shift to more sustainable and healthy diets is generally feasible in many regions medium 1 train +7821 AR6_WGIII 940 9 In line with these SDG targets, it is estimated that reducing FLW can free up several million km2 of land high 2 train +7822 AR6_WGIII 940 44 Regionally, FLW reduction is feasible anywhere but its potential needs to be understood in a wider and changing socio-cultural context that determines nutrition high 2 train +7823 AR6_WGIII 941 16 The SRCCL (Chapters 2 and 6) finds that some studies indicate significant mitigation potentials for material substitution, but concludes that the global, technical mitigation potential for material substitution for construction applications ranges from 0.25–1 GtCO 2-eq yr–1 medium 1 train +7824 AR6_WGIII 949 12 The total amounts to 0.66 GtCO 2 yr–1 for the period 2010–2019, which is 1.2% of total global, and 5.5% of AFOLU emissions reported in Table 7.1, over the same time period high 2 train +7825 AR6_WGIII 951 3 In total, reported funding for AFOLU projects and programmes has been USD4.4 billion over the past decade, or about USD569 million yr–1 low 0 train +7826 AR6_WGIII 951 8 Despite widespread effort, AFOLU measures have thus far failed to achieve the large potential for climate mitigation described in earlier IPCC WG III reports high 2 train +7827 AR6_WGIII 951 9 The limited gains from AFOLU to date appear largely to result from lack of investment and other institutional and social barriers, rather than methodological concerns high 2 train +7828 AR6_WGIII 951 49 A large literature has investigated whether PES programmes have successfully protected habitats. Despite concerns, the many lessons learned from PES programme implementation provide critical information that will help policymakers refine future PES programmes to increase their effectiveness medium 1 test +7829 AR6_WGIII 952 28 When combined with increased non-CO 2 gas emissions, the emission intensity of US agriculture increased from 1.5 to 1.7 tCO 2 ha–1 between 2005 and 2018 high 2 train +7830 AR6_WGIII 952 46 The extent to which the combined influence of these regulations has enhanced carbon storage in ecosystems is not quantified although they are likely to explain some of the persistent carbon sink that has emerged in temperate forests of OECD countries high 2 train +7831 AR6_WGIII 953 12 Because protected areas limit not just land- use change, but also logging or harvesting non-timber forest products, they may be relatively costly approaches for forest conservation medium 1 train +7832 AR6_WGIII 954 8 Efforts to expand property rights, especially community forest management, have reduced carbon emissions from deforestation in tropical forests in the last two decades high 2 train +7833 AR6_WGIII 954 28 Supply chain management in the food sector encourages more widespread use of conservation measures in agriculture high 2 train +7834 AR6_WGIII 956 18 While these analyses establish that many projects to reduce deforestation have overcome hurdles related to additionality high 2 train +7835 AR6_WGIII 957 46 The protocols to quantify emission reductions in the agricultural sector are available and have been tested, and the main limitation appears to be the lack of available financing or the unwillingness to re-direct current subsidies medium 1 train +7836 AR6_WGIII 958 15 The presence of significant subsidy programmes intended to improve farmer welfare and rural livelihoods makes it more difficult to implement regulatory programmes aimed at reducing net emissions in agriculture, however, it may increase the potential to implement new subsidy programmes that encourage practices aimed at reducing net emissions medium 1 train +7837 AR6_WGIII 959 7 It is argued that a carbon tax on only fossil fuel derived emissions, may lead to massive deployment of bioenergy, although the effects of such a policy can be mitigated when combined with policies that encourage sustainable forest management and protection of forest carbon stocks as well as forest management certification high 2 train +7838 AR6_WGIII 959 17 However, if private markets emerge for biomass and BECCS on the scale suggested in the SR1.5, policy efforts must ramp up to substantially value, encourage, and protect terrestrial carbon stocks and ecosystems to avoid outcomes inconsistent with many SDGs high 2 train +7839 AR6_WGIII 960 1 The lack of resources thus far committed to implementing AFOLU mitigation, income and access to alternative sources of income in rural households that rely on agriculture or forests for their livelihoods remains a considerable barrier to adoption of AFOLU high 2 train +7840 AR6_WGIII 960 4 Without quickly ramping up spending, the lack of funding to implement projects remains a substantial barrier high 2 train +7841 AR6_WGIII 961 3 Barriers to adoption of AFOLU mitigation will be strongest where historical practices represent long-standing traditions high 2 train +7842 AR6_WGIII 961 5 AR6 presents new estimates of the mitigation potential for shifts in diets and reductions in food waste, but given long-standing dietary traditions within most cultures, some of the strongest barriers exist for efforts to change diets medium 1 train +7843 AR6_WGIII 961 50 Implementation of nature-based solution may have local or regionally important consequences for other ecosystem services, some of which may be negative high 2 train +7844 AR6_WGIII 962 9 Climate is expected to reduce crop yields, increase crop and livestock prices, and increase pressure on undisturbed forest land for food production creating new barriers and increasing costs for implementation of many nature-based mitigation techniques medium 1 train +7845 AR6_WGIII 962 27 Factors that reduce permanence or slow forest growth will drive up costs of forest mitigation measures, suggesting that climate change presents a formidable challenge to implementation of nature-based solutions beyond 2030 high 2 train +7846 AR6_WGIII 962 38 Forest management strategies based on biodiversity and ecosystems functioning interactions can augment the effectiveness of forests in reducing climate change impacts on ecosystem functioning high 2 train +7847 AR6_WGIII 964 22 There are synergies, trade-offs and co- benefits between ecosystem services and mitigation options with impacts on ecosystem services differing by scale and contexts high 2 train +7848 AR6_WGIII 965 6 Conservation of biodiversity and ecosystem services is critical to sustaining the well-being and livelihoods of poor and marginalised people, and indigenous communities who depend on natural resources high 2 train +7849 AR6_WGIII 966 7 Nature-based solutions (NBS) with safeguards has immense potential for cost-effective adaptation to climate change; but their impacts will vary by scale and contexts high 2 train +7850 AR6_WGIII 999 22 There is a large range in the forecasts of urban land expansion across scenarios and models, which highlights an opportunity to shape future urban development towards low- or net-zero GHG emissions and minimise the loss of carbon stocks and sequestration in the agriculture, forestry and other land use (AFOLU) sector due to urban land conversion medium 1 train +7851 AR6_WGIII 1000 17 New and emerging cities will have significant infrastructure development needs to achieve high quality of life, which can be met through energy- efficient infrastructures and services, and people-centred urban design high 2 train +7852 AR6_WGIII 1013 6 Urbanisation: A megatrend driving global climate risk and potential for low-carbon and resilient futures Severe weather events, exacerbated by anthropogenic emissions, are already having devastating impacts on people who live in urban areas, on the infrastructure that supports these communities, as well as people living in distant places high 2 train +7853 AR6_WGIII 1013 26 Urban settlements contribute to climate change, generating about 70% of global CO 2-eq emissions high 2 train +7854 AR6_WGIII 1014 1 As urbanisation unfolds, its legacy continues to be the locking-in of emissions and vulnerabilities high 2 train +7855 AR6_WGIII 1015 8 Enabling action Innovative governance and finance solutions are required to manage complex and interconnected risks across essential key infrastructures, networks, and services, as well as to meet basic human needs in urban areas medium 1 train +7856 AR6_WGIII 1015 16 Increasing the pace of investments will put pressure on governance capability, transparency, and accountability of decision-making medium 1 train +7857 AR6_WGIII 1021 14 Third, the urban share of regional GHG emissions increased between 2000 and 2015, with much inter-region variation in the magnitude of the increase high 2 train +7858 AR6_WGIII 1021 18 Fourth, the global average per capita urban GHG emissions increased between 2000 and 2015, with cities in the Developed Countries region producing nearly seven times more per capita than the lowest emitting region medium 1 train +7859 AR6_WGIII 1026 3 Future urban expansion will amplify the background warming caused by GHG emissions, with extreme warming most pronounced during night-time very high 3 test +7860 AR6_WGIII 1038 20 Over 30 years, wood-based construction can accumulate between 0.25 and 20 GtCO 2 and reduce cumulative emissions from 4 GtCO 2 (range of 7–20 GtCO 2) to 2 GtCO 2 (range of 0.3–10 GtCO 2) high 2 train +7861 AR6_WGIII 1042 5 This might suggest that in other geographies, similar adjacent non-urban forest types may store similar carbon stocks per unit area medium 1 train +7862 AR6_WGIII 1048 3 The involvement of governance at multiple levels is necessary to enable cities to plan and implement emissions reductions targets high 2 train +7863 AR6_WGIII 1048 5 Further, regional, national, and international climate goals are most impactful when local governments are involved alongside higher levels, rendering urban areas key foci of climate governance more broadly high 2 train +7864 AR6_WGIII 1051 28 To meet the multi-trillion-dollar annual investment needs in urban areas, cities in partnership with international institutions, national governments, and local stakeholders increasingly play a pivotal role in mobilising global climate finance resources for a range of low-carbon infrastructure projects and related urban land use and spatial planning programmes across key sectors high 2 train +7865 AR6_WGIII 1052 36 Moreover, the potential application of blockchain for land-based funding instruments is possibly associated with urban form attributes, such as density, compactness, and land-use mixture, to disincentivise urban expansion and emissions growth around city-fringe locations medium 1 train +7866 AR6_WGIII 1257 4 Transport- related emissions in developing regions of the world have increased more rapidly than in Europe or North America, a trend that is likely to continue in coming decades high 2 test +7867 AR6_WGIII 1257 12 Appropriate infrastructure, including protected pedestrian and bike pathways, can also support much greater localised active travel.1 Transport demand management incentives are expected to be necessary to support these systemic changes high 2 train +7868 AR6_WGIII 1257 22 The continued growth of electromobility for land transport would require investments in electric charging and related grid infrastructure high 2 train +7869 AR6_WGIII 1257 23 Electromobility powered by low-carbon electricity has the potential to rapidly reduce transport GHG and can be applied with multiple co-benefits in the developing world’s growing cities high 2 train +7870 AR6_WGIII 1257 25 These same technologies and expanded use of available electric rail systems can support rail decarbonisation medium 1 train +7871 AR6_WGIII 1257 26 Initial deployments of battery electric, hydrogen- and bio-based haulage are underway, and commercial operations of some of these technologies are considered feasible by 2030 medium 1 train +7872 AR6_WGIII 1257 29 Increased capacity for low-carbon hydrogen production would also be essential for hydrogen-based fuels to serve as an emissions reduction strategy high 2 train +7873 AR6_WGIII 1257 31 Increased efficiency has been insufficient to limit the emissions from shipping and aviation, and natural gas-based fuels are likely inadequate to meet stringent decarbonisation goals for these segments high 2 train +7874 AR6_WGIII 1257 33 Advanced biofuels could provide low-carbon jet fuel medium 1 train +7875 AR6_WGIII 1257 34 The production of synthetic fuels using low-carbon hydrogen with CO2 captured through direct air capture (DAC) or bioenergy with carbon capture and storage (BECCS) could provide jet and marine fuels but these options still require demonstration at scale low 0 test +7876 AR6_WGIII 1257 35 Ammonia produced with low-carbon hydrogen could also serve as a marine fuel medium 1 train +7877 AR6_WGIII 1258 1 The scenarios literature projects continued growth in demand for freight and passenger services, particularly in developing countries in Africa and Asia high 2 train +7878 AR6_WGIII 1258 4 While many global scenarios place greater reliance on emissions reduction in sectors other than transport, a quarter of the 1.5°C degree scenarios describe transport-related CO2 emissions reductions in excess of 68% (relative to modelled 2020 levels) medium 1 train +7879 AR6_WGIII 1258 8 In general terms, electrification tends to play the key role in land- based transport, but biofuels and hydrogen (and derivatives) could play a role in decarbonisation of freight in some contexts high 2 train +7880 AR6_WGIII 1258 9 Biofuels and hydrogen (and derivatives) are likely more prominent in shipping and aviation high 2 train +7881 AR6_WGIII 1258 10 The shifts towards these alternative fuels must occur alongside shifts towards clean technologies in other sectors high 2 train +7882 AR6_WGIII 1258 22 Given the high degree of potential recyclability of LIBs, a nearly closed-loop system in the future could mitigate concerns about critical mineral issues medium 1 train +7883 AR6_WGIII 1369 5 Producer, user, and regulator education, as well as innovation and commercialisation policy are needed medium 1 train +7884 AR6_WGIII 1369 18 Pulp mills have access to biomass residues and by-products and in paper mills the use of process heat at low to medium temperatures allows for electrification high 2 train +7885 AR6_WGIII 1453 1 Key to maximising benefits and managing trade-offs are sectoral integration, transparent governance, and stakeholder involvement high 2 train +7886 AR6_WGIII 1453 2 A sustainable bioeconomy relying on biomass resources will need to be supported by technology innovation and international cooperation and governance of global trade to disincentivise environmental and social externalities medium 1 train +7887 AR6_WGIII 1465 19 This is about, or below half, the most recent (2019) emissions value of 59 ± 6.6 GtCO 2-eq high 2 train +7888 AR6_WGIII 1471 13 Compared to other CDR methods, the primary barrier to upscaling DAC is its high cost and large energy requirement high 2 train +7889 AR6_WGIII 1471 27 Costs: As the process captures dilute CO 2 (~0.04%) from the ambient air, it is less efficient and more costly than conventional carbon capture applied to power plants and industrial installations (with a CO 2 concentration of ~10%) high 2 train +7890 AR6_WGIII 1471 53 Risks and impacts: DACCS requires a considerable amount of energy high 2 train +7891 AR6_WGIII 1502 19 The SRCCL concluded that conversion of land for A/R and bioenergy crops at the scale commonly found in pathways limiting warming to 1.5°C or 2°C is associated with multiple feasibility and sustainability constraints, including land carbon losses high 2 train +7892 AR6_WGIII 1512 1 Increased technology innovation, stakeholder integration and transparent governance structures and procedures at local to global scales are key to successful bioeconomy deployment maximising benefits and managing trade-offs high 2 train +7893 AR6_WGIII 1512 2 Limited global land and biomass resources accompanied by growing demands for food, feed, fibre, and fuels, together with prospects for a paradigm shift towards phasing out fossil fuels, set the frame for potentially fierce competition for land3 and biomass to meet burgeoning demands, even as climate change increasingly limits natural resource potentials high 2 train +7894 AR6_WGIII 1516 18 In summary, there is significant scope for optimising use of land resources to produce more biomass while reducing adverse effects high 2 train +7895 AR6_WGIII 1516 19 Context-specific prioritisation, technology innovation in bio-based production, integrative policies, coordinated institutions and improved governance mechanisms to enhance synergies and minimise trade-offs can mitigate the pressure on managed as well as natural and semi-natural ecosystems medium 1 train +7896 AR6_WGIII 1516 20 Yet, energy conservation and efficiency measures, and deployment of technologies and systems that do not rely on carbon-based energy and materials, are essential for mitigating biomass demand growth as countries pursue ambitious climate goals high 2 train +7897 AR6_WGIII 1672 2 It also built national capacity for greenhouse gas (GHG) accounting, catalysed the creation of GHG markets, and increased investments in low-carbon technologies medium 1 train +7898 AR6_WGIII 1672 3 Other international agreements and institutions have led to avoided carbon dioxide (CO 2) emissions from land use practices, as well as avoided emissions of some non- CO2 greenhouse gases medium 1 train +7899 AR6_WGIII 1672 5 Both new and pre-existing forms of cooperation are vital for achieving climate mitigation goals in the context of sustainable development high 2 train +7900 AR6_WGIII 1672 6 While previous IPCC assessments have noted important synergies between the outcomes of climate mitigation and achieving sustainable development objectives, there now appear to be synergies between the two processes themselves medium 1 train +7901 AR6_WGIII 1672 10 International cooperation helps countries achieve long-term mitigation targets when it supports development and diffusion of low-carbon technologies, often at the level of individual sectors, which can simultaneously lead to significant benefits in the areas of sustainable development and equity medium 1 train +7902 AR6_WGIII 1672 13 By contrast, the commitments under the Paris Agreement are primarily procedural, extend to all Parties, and are designed to trigger domestic policies and measures, enhance transparency, stimulate climate investments, particularly in developing countries, and to lead iteratively to rising levels of ambition across all countries high 2 train +7903 AR6_WGIII 1672 14 Issues of equity remain of central importance in the UN climate regime, notwithstanding shifts in the operationalisation of ‘common but differentiated responsibilities and respective capabilities’ from Kyoto to Paris high 2 train +7904 AR6_WGIII 1672 17 The recent proliferation of national mid-century net zero GHG targets can be attributed in part to the Paris Agreement medium 1 train +7905 AR6_WGIII 1672 18 Moreover, its processes and commitments will enhance countries’ abilities to achieve their stated level of ambition, particularly among developing countries medium 1 train +7906 AR6_WGIII 1672 19 Arguments against the Paris Agreement are that it lacks a mechanism to review the adequacy of individual Parties’ Nationally Determined Contributions (NDCs), that collectively current NDCs are inconsistent in their level of ambition with achieving the Paris Agreement’s temperature goal, that its processes will not lead to sufficiently rising levels of ambition in the NDCs, and that NDCs will not be achieved because the targets, policies and measures they contain are not legally binding at the international level medium 1 train +7907 AR6_WGIII 1672 20 To some extent, arguments on both sides are aligned with different analytic frameworks, including assumptions about the main barriers to mitigation that international cooperation can help overcome medium 1 train +7908 AR6_WGIII 1672 21 The extent to which countries increase the ambition of their NDCs and ensure they are effectively implemented will depend in part on the successful implementation of the support mechanisms in the Paris Agreement, and in turn will determine whether the goals of the Paris Agreement are met high 2 train +7909 AR6_WGIII 1672 23 Agreements addressing ozone depletion, transboundary air pollution, and release of mercury are all leading to reductions in the emissions of specific greenhouse gases high 2 train +7910 AR6_WGIII 1672 25 Transnational partnerships and alliances involving non-state and sub-national actors are also playing a growing role in stimulating low-carbon technology diffusion and emissions reductions medium 1 train +7911 AR6_WGIII 1672 27 Climate change is being addressed in a growing number of international agreements operating at sectoral levels, as well as within the practices of many multilateral organisations and institutions high 2 train +7912 AR6_WGIII 1672 28 Sub-global and regional cooperation, often described as climate clubs, can play an important role in accelerating mitigation, including the potential for reducing mitigation costs through linking national carbon markets, although actual examples of these remain limited high 2 train +7913 AR6_WGIII 1673 1 Many developing countries’ NDCs have components or additional actions that are conditional on receiving assistance with respect to finance, technology development and transfer, and capacity building, greater than what has been provided to date high 2 train +7914 AR6_WGIII 1673 3 In some cases, notably with respect to aviation and shipping, sectoral agreements have adopted climate mitigation goals that fall far short of what would be required to achieve the temperature goal of the Paris Agreement high 2 test +7915 AR6_WGIII 1673 4 Moreover, there are cases where international cooperation may be hindering mitigation efforts, namely evidence that trade and investment agreements, as well as agreements within the energy sector, impede national mitigation efforts medium 1 train +7916 AR6_WGIII 1673 5 International cooperation is emerging but so far fails to fully address transboundary issues associated with Solar Radiation Modification and CO 2 removal high 2 train +7917 AR6_WGIII 1711 18 Cooling caused by SRM would increase the global land and ocean CO 2 sinks (medium confidence), but this would not stop CO 2 from increasing in the atmosphere or affect the resulting ocean acidification under continued anthropogenic emissions high 2 train +7918 AR6_WGIII 1712 2 A gradual phase-out of SRM combined with emissions reduction and CDR could avoid these termination effects medium 1 train +7919 AR6_WGIII 1768 1 Fundamental inequities in access to finance as well as its terms and conditions, and countries’ exposure to physical impacts of climate change overall result in a worsening outlook for a global just transition high 2 train +7920 AR6_WGIII 1768 9 This increased awareness can support climate policy development and implementation high 2 train +7921 AR6_WGIII 1768 17 This leaves high uncertainty, both near-term (2021–30) and longer- term (2021–50), on the feasibility of an alignment of financial flows with the Paris Agreement high 2 train +7922 AR6_WGIII 1768 19 There is a climate financing gap which reflects a persistent misallocation of global capital high 2 train +7923 AR6_WGIII 1768 25 This reflects policy misalignment, the current perceived risk-return profile of fossil fuel-related investments, and political economy constraints high 2 train +7924 AR6_WGIII 1768 31 To meet the needs for rapid deployment of mitigation options, global mitigation investments are expected to need to increase by the factor of 3 to 6 medium 1 train +7925 AR6_WGIII 1768 32 The gaps are wide for all sectors and represent a major challenge for developing countries,3 especially Least-Developed Countries (LDCs), where flows have to increase by factor 4 to 7, for specific sectors like agriculture, forestry and other land use (AFOLU) in relative terms, and for specific groups with limited access to, and high costs of, climate finance high 2 train +7926 AR6_WGIII 1768 34 Financing needs for the creation and strengthening of regulatory environment and institutional capacity, upstream financing needs as well as R&D and venture capital for development of new technologies and business models are often overlooked despite their critical role to facilitate the deployment of scaled-up climate finance high 2 train +7927 AR6_WGIII 1769 1 This will particularly impact urban infrastructure and the energy and transport sectors high 2 train +7928 AR6_WGIII 1769 2 A common understanding of debt sustainability and debt transparency, including negative implications of deferred climate investments on future GDP, and how stranded assets and resources may be compensated, has not yet been developed medium 1 train +7929 AR6_WGIII 1769 6 This highlights the importance of trust in political leadership which, in turn, affects risk perception and ultimately financing costs high 2 train +7930 AR6_WGIII 1769 9 A significant push for international climate finance access for vulnerable and poor countries is particularly important given these countries’ high costs of financing, debt stress and the impacts of ongoing climate change high 2 train +7931 AR6_WGIII 1769 15 In addition to indirect and direct subsidies, the public sector’s role in addressing market failures, barriers, provision of information, and risk sharing (equity, various forms of public guarantees) can encourage the efficient mobilisation of private sector finance high 2 test +7932 AR6_WGIII 1769 19 Existing policy misalignments – for example in fossil fuel subsidies – undermine the credibility of public commitments, reduce perceived transition risks and limit financial sector action high 2 train +7933 AR6_WGIII 1769 21 Approaches include de-risking investments, robust ‘green’ labelling and disclosure schemes, in addition to a regulatory focus on transparency and reforming international monetary system financial sector regulations medium 1 train +7934 AR6_WGIII 1769 24 These relatively new labelled financial products will help by allowing a smooth integration into existing asset allocation models high 2 train +7935 AR6_WGIII 1769 33 The following policy options can have important long-term catalytic benefits high 2 train +7936 AR6_WGIII 1773 11 The AR5 concluded that published assessments of financial flows whose expected effect was to reduce net greenhouse gas (GHG) emissions and/or to enhance resilience to climate change aggregated USD343–385 billion4 yr–1 globally between 2010 and 2012 medium 1 train +7937 AR6_WGIII 1774 1 The resulting estimated gaps in annual mitigation financing during 2014 to 2017, using reporting of climate financing from published sources, was about 67% for 2015, and 76% for the energy sector alone in 2017 medium 1 train +7938 AR6_WGIII 1774 6 The gap is expected to be aggravated by COVID-19 high 2 train +7939 AR6_WGIII 1774 19 Four major events and macro trends mark the developments in climate finance in the previous five years and likely developments in the near term.• First, the 2015 Paris Agreement, with the engagement of the financial sector institutions in the climate agenda, has been followed by a series of related developments in financial regulation in relation to climate change and in particular to the disclosure of climate-related financial risk high 2 train +7940 AR6_WGIII 1774 22 At the same time, while it is still too early to draw positive conclusions, this crisis highlights opportunities in terms of political and policy feasibility and behavioural change in respect of realigning climate finance medium 1 train +7941 AR6_WGIII 1776 4 As each warmer year keeps producing more negative impacts – arising from greater and rising variability and intensity of rainfall, floods, droughts, forest fires and storms – the negative consequences have become more macro- economically significant, and worst for the most climate-vulnerable developing countries high 2 train +7942 AR6_WGIII 1776 27 The third headwind is rising financial and insurance sector risks and stresses (distinct from real ‘physical’ climate risks above) arising from the impacts of climate change, and systematically affecting both national and international financial institutions and raising their credit risks high 2 train +7943 AR6_WGIII 1783 9 Climate finance in developing countries remains heavily concentrated in a few large economies high 2 train +7944 AR6_WGIII 1783 10 Least-developed countries (LDCs), on the other hand, continue to represent less than 5% year-on-year medium 1 train +7945 AR6_WGIII 1783 20 Mitigation continues to represent the lion’s share of global climate finance (consistently above 90% between 2017 and 2020), and in particular renewable energy, followed by energy efficiency and transport high 2 train +7946 AR6_WGIII 1785 7 More generally, significant gaps remain to track climate finance comprehensively at a global level: • Available estimates are based on a good coverage of investments in renewable energy and, where available, energy efficiency and transport, while other sectors remain more difficult to track, such as industry, agriculture and land use high 2 train +7947 AR6_WGIII 1785 10 Data on private and commercial finance remain very patchy, particularly for corporate financing (including debt financing provided by commercial banks), for which it is difficult to establish a link with activities and projects on the ground high 2 train +7948 AR6_WGIII 1785 28 To avoid locking GHG emissions incompatible with remaining carbon budgets, this implies a rapid scaling down of new fossil fuel- related investments, combined with a scaling up of financing to allow energy and infrastructure systems to transition high 2 train +7949 AR6_WGIII 1796 13 Estimated mitigation financing needs as a percentage of mean 2017–2020 GDP in USD 2015 comes in at around 2–4% for developed countries, and around 4-9% for developing countries high 2 train +7950 AR6_WGIII 1796 15 This disparity is further exacerbated when considering adaptation, infrastructure and SDG-related investment needs high 2 train +7951 AR6_WGIII 1796 24 Notably, climate finance flows to African countries might have even decreased for mitigation technology deployment (stagnated for adaptation between 2017 and 2020), widening the finance gap in African countries in the recent years high 2 train +7952 AR6_WGIII 1798 29 Taking into account the inertia of the financial system as well as the magnitude of the challenge to align financial flows with the long-term global goals, fast action is required to ensure the readiness of the financial sector as an enabler of the transition high 2 train +7953 AR6_WGIII 1799 18 On the one hand, unmitigated climate change implies an increased potential for adverse socio-economic impacts especially in more exposed economic activities and areas high 2 train +7954 AR6_WGIII 1800 7 On the other hand, the mitigation of climate change, by means of a transition to a low- carbon economy, requires a transformation of the energy and production system at a pace and scale that implies adverse impacts on a range of economic activities, but also opportunities for some other activities high 2 train +7955 AR6_WGIII 1800 38 Significant cost increases have been observed related to increases in frequency and magnitude of extreme events high 2 train +7956 AR6_WGIII 1801 10 Fossil fuel reserve and resource estimates exceed in equivalent quantity of CO 2 with virtual certainty the carbon budget available to reach the 1.5°C and 2°C targets high 2 train +7957 AR6_WGIII 1801 33 One outstanding challenge for the analysis of investors’ exposure to climate risks is the difficulty of gathering granular and standardised information on the breakdown of non-financial firms’ revenues and CAPEX in terms of low-/high- carbon activities high 2 train +7958 AR6_WGIII 1802 10 One fundamental challenge is that climate-related financial risk is endogenous high 2 train +7959 AR6_WGIII 1802 24 The endogeneity of risk and its associated deep uncertainty implies that the standard approach to financial risk, consisting of computing expected values and risk based on historical values of market prices, is not adequate for climate risk high 2 train +7960 AR6_WGIII 1802 36 This development is key to mainstreaming the assessment of transition risk among financial institutions, but the following challenges emerge high 2 train +7961 AR6_WGIII 1803 3 Global macroeconomic changes that may affect asset prices are expected to take place as a result of a possible reduction in growth or contraction of fossil fuel demand, in scenarios in which climate targets are met according to carbon budgets, but also following ongoing energy efficiency changes high 2 train +7962 AR6_WGIII 1803 16 Due to the predominantly international nature of fossil fuel markets, assets may be at risk from regulatory and technological changes both domestically and in foreign countries medium 1 train +7963 AR6_WGIII 1803 24 Framing climate risk as a financial risk (not just as an ethical issue) is key for it to become an actionable criterion for investment decision among mainstream investors high 2 train +7964 AR6_WGIII 1805 6 The role of government is crucial for creating an enabling environment for climate (Clark 2018), and governments are critical in the launching and maintenance of this circle of trust by lowering the political, regulatory, macroeconomic and business risks high 2 train +7965 AR6_WGIII 1805 12 Transparency: Policy de-risking measures, such as robust policy design and better transparency, as well as financial de-risking measures, such as green bonds and guarantees, at both domestic and international levels, enhance the attractiveness of clean energy investments high 2 train +7966 AR6_WGIII 1805 14 However, risk disclosures alone would likely be insufficient as long as market failures that inhibit the emergence of low-carbon investment initiatives with positive risk-weighted returns high 2 train +7967 AR6_WGIII 1805 17 Central banks in all economies will likely have to play a critical role in supporting the financing of fiscal operations, particularly in a post-COVID-19 world high 2 train +7968 AR6_WGIII 1805 26 A green QE programme ‘would have the benefit of providing large amounts of additional liquidity to companies interested’ in green projects medium 1 train +7969 AR6_WGIII 1805 33 Additional monetary policies and macroprudential financial regulation may facilitate the expected role of carbon pricing on boosting low-carbon investments medium 1 train +7970 AR6_WGIII 1805 40 Despite increasing challenges to the theory (Sewell 2011), especially by repeated episodes of global financial crashes and crises, and other widely noted anomalies, a weaker form of the efficient markets hypothesis may still apply medium 1 train +7971 AR6_WGIII 1806 44 As a result, low-carbon investments do not take place to socially and economically optimal levels, and the correct market signals would involve setting carbon prices high enough or equivalent trading in reduced carbon emissions by regulatory action to induce sufficient and faster shift towards low-carbon investments high 2 train +7972 AR6_WGIII 1807 1 Carbon tax can be a simpler and easier way to implement carbon pricing, especially in developing countries, because countries can utilise the existing fiscal tools and do not need concrete enabling conditions as market-based frameworks high 2 train +7973 AR6_WGIII 1807 39 There is a growing awareness of the low leverage ratio of public to private capital in climate blended finance (Blended Finance Taskforce 2018b) and of a ‘glass ceiling’, caused by a mix of agencies’ inertia and perceived loss of control over the use of funds, on the use of public guarantees by MDBs to increase it high 2 train +7974 AR6_WGIII 1809 1 Limited prioritisation and agreement on prioritisation of sectors and/or project categories being ready and/or preferred for direct private sector involvement might become a challenge in the coming years high 2 train +7975 AR6_WGIII 1809 15 A missing alignment of public funding and investment activity with the Paris Agreement (and Sustainable Development Goals) would result in significant carbon lock-ins, stranded assets and thus increase transition risks and ultimately economic costs of the transition high 2 train +7976 AR6_WGIII 1809 21 Lessons from the global financial crises show that although deep economic crises create a sharp short-term emission drop, and green stimulus is argued to be the ideal response to tackle both the economic and the climate crises at once, disparities between regional strategies hinder the low-carbon transition high 2 train +7977 AR6_WGIII 1809 32 Boosting investment should propel job creation, increasing household income and therefore demand across economic sectors high 2 train +7978 AR6_WGIII 1809 35 First, only those countries and regions with highest credit-ratings (AAA or AA) with access to deep financial markets and excess savings will be able to mount such counter-cyclical climate investment paths, typically high-income developed economies high 2 train +7979 AR6_WGIII 1810 6 Box 15.6 | Macroeconomics and Finance of a Post-COVID-19 Green Stimulus Economic Recovery Path Financial history suggests that capital markets may be willing to accommodate extended public borrowing for transient spending spikes (Barro 1987) when macroeconomic conditions suggest excess savings relative to private investment opportunities (Summers 2015) and when public spending is seen as timely, effective and productive, with governments able to repay when conditions improve as economic crisis conditions abate high 2 train +7980 AR6_WGIII 1811 2 However, in terms of increase of supply of, in particular, public finance, often the debate is still driven by the question on affordability, considerations around financial debt sustainability and budgetary constraints against the background of macroeconomic headwinds – even more in the (post-)COVID-19 world high 2 train +7981 AR6_WGIII 1811 3 The level of climate alignment of debt is hardly considered in debt-related regulation and/ or debt sustainability agreements like the Maastricht Treaty ceilings (3% of GDP government deficit and 60% of GDP (gross) government debt) not considering economic costs of deferred climate action as well as economic benefits of the transformation.Robust studies on the economic costs and benefits in the short- to long-term of reaching the LTGG exist for only few countries and/or regions, primarily in the developed world high 2 train +7982 AR6_WGIII 1811 6 For many developing countries, the focus of debt sustainability discussions is on the negative effect of climate change on the future GDP and the uncertainty with regard to short-term effects of climate change and their economic implications high 2 train +7983 AR6_WGIII 1812 6 Debt levels globally but particularly in developing and vulnerable countries have significantly increased over the past years with current and expected climate change impacts further burdening debt sustainability high 2 train +7984 AR6_WGIII 1812 19 Definition of triggers is likely the most complex challenge in this context.The use of debt-for-nature and debt-for-climate-swaps is still very limited and not mainstreamed but offers significant potential if used correctly high 2 train +7985 AR6_WGIII 1812 31 The debate around stranded assets focuses strongly on the loss of value to financial assets for investors (Section 15.6.1), however, stranded assets and resources in the context of the transition towards a low-emission economy ‘are expected to become a major economic burden for states and hence the tax payers’ high 2 train +7986 AR6_WGIII 1813 2 Late government action can delay action and consequently strengthen the magnitude of action needed at a later point in time with implications for employment and economic development in impacted regions requiring higher level of fiscal burden high 2 train +7987 AR6_WGIII 1813 17 While there are questions about the sufficiency of insurance products to address the losses and damages of climate- related disasters, insurance can help to cover immediate needs directly, provide rapid response and transfer financial risk in times of extreme crisis high 2 train +7988 AR6_WGIII 1814 4 Risk pooling among countries and regions is relatively advantageous when compared to conventional insurance because of the effective subsidising of ‘affected regions’ using revenues from unaffected regions which involve pooling among a large subset of countries high 2 train +7989 AR6_WGIII 1814 8 At the same time, this approach has substantial basis risk (actual losses do not equal financial compensation) high 2 train +7990 AR6_WGIII 1814 17 Increasingly, climate risk insurance schemes are being blended into disaster risk management as part of a comprehensive risk management approach high 2 train +7991 AR6_WGIII 1814 27 There are also challenges with risk diversification, replication, and scalability high 2 train +7992 AR6_WGIII 1815 13 Other gaps and challenges flagged by Kreft and Schäfer (2017) include limited coverage of the full spectrum of contingency risks experienced by countries, inadequate role of risk management as a standard for all regional pools, though there are some emerging best practices in terms of data provision on weather-related risks, and incentivisation of risk reduction high 2 train +7993 AR6_WGIII 1815 40 It has transformative potential as a key enabler of inclusive urban economic development through the building of resilient communities high 2 train +7994 AR6_WGIII 1816 7 The effect of deficit in investment for global infrastructure towards the growing subnational-level debt also creates pressure on subnational finances and constrains future access to financing high 2 train +7995 AR6_WGIII 1816 22 PPPs are particularly important in cities with mature financial systems as the effectiveness of PPPs depends on appropriate investment architecture at scale and government capacity high 2 train +7996 AR6_WGIII 1816 25 National-level investment vehicles can provide leadership for subnational climate financing and crowd in private finance by providing early-stage market support to technologies or evidence related to asset performance and costs-benefits high 2 train +7997 AR6_WGIII 1816 27 Debt financing via subnational bonds and borrowing, including municipal bonds, is another potential tool for raising upfront capital, especially for rich cities high 2 train +7998 AR6_WGIII 1816 33 Across all types of cities, five key challenges constrain the flow of subnational climate finance high 2 train +7999 AR6_WGIII 1816 37 Access to capital markets has been one of the major sources for subnational financing and is generally limited to rich cities, and much of this occurs through loans high 2 train +8000 AR6_WGIII 1816 43 The depth and character of the local capital market also affect cities differently in generating bonds high 2 train +8001 AR6_WGIII 1817 4 There is insufficient evidence about which financing schemes contribute to climate change mitigation and adaptations at community level high 2 train +8002 AR6_WGIII 1817 15 Integrated urban capital investment planning is an option to develop cross-sectoral solutions that reduce investment needs, boost coordination capacity, and increase climate-smart impacts high 2 train +8003 AR6_WGIII 1817 23 Investor demand is driving developments in innovative financial products high 2 train +8004 AR6_WGIII 1817 26 The growth and diversity of the green bond market illustrates how innovative financial products can attract both public and private investors high 2 train +8005 AR6_WGIII 1818 3 Significant potential exists for continued growth in innovative financial products, though some challenges remain high 2 train +8006 AR6_WGIII 1818 18 Financial disclosure regulatory developments can help further align and specify definitions of green in the financial sector but are not a substitute for climate policy high 2 train +8007 AR6_WGIII 1818 35 Identifying green financial products can increase uptake and may result in a lower cost of capital in certain parts of the market high 2 train +8008 AR6_WGIII 1818 41 There is some evidence of a premium, or an acceptance of lower yields by the investor, for green bonds medium 1 train +8009 AR6_WGIII 1819 11 Financial flows via these instruments have limited measurable environmental impact to date, however they can support capacity building on climate risk and opportunities within institutions to realise future impacts high 2 train +8010 AR6_WGIII 1821 19 One challenge is unlocking unencumbered endowments to contribute to Paris and SDGs high 2 train +8011 AR6_WGIII 1822 8 Institutional domestic and international investors have growing assets estimated to exceed USD100 trillion high 2 train +8012 AR6_WGIII 1822 18 However, progress remains pilot, slow and piecemeal high 2 train +8013 AR6_WGIII 1822 24 Options to mobilise institutional investors currently remain small pilots, relative to Paris and SDG ambitions high 2 test +8014 AR6_WGIII 1822 37 Paris-aligned NDCs that integrate policies on COVID-19 pandemic recovery, climate action, sustainable development, just transition and equity can harness co-benefits including contribution to Invisible UN SDG 7 energy poverty sectors high 2 train +8015 AR6_WGIII 1823 16 Combining approaches: (i) developed countries meeting UNFCCC USD100 billion commitment on a grant-equivalent basis, (ii) stepped up technical assistance, (iii) infrastructure coordination, (iv) knowledge sharing by project preparation entities, and (iv) harnessing project risk facilities such as guarantees could be instrumental for scaling climate finance for Paris-SDGs high 2 train +8016 AR6_WGIII 1824 13 Existing project risk facilities including guarantees could benefit from coordination, scaling and better reporting frameworks high 2 train +8017 AR6_WGIII 1824 27 Investment vehicles into green infrastructure come in various forms high 2 train +8018 AR6_WGIII 1825 4 Developing government bond yield curves with different maturities can be an important policy objective high 2 train +8019 AR6_WGIII 1825 24 Developing countries are using fiscal incentives, grants, and guarantees to support nascent bond markets with most taxonomies under development high 2 train +8020 AR6_WGIII 1826 11 This makes a business case for deferred investments in grid infrastructure medium 1 train +8021 AR6_WGIII 1826 28 MaaS tends to deliver significant urban benefits (e.g., cleaner air) and brings in efficiency gains in the use of resources high 2 train +8022 AR6_WGIII 1827 2 If effectively implemented, nature-based solutions can be cost- effective measures and able to provide multiple benefits, such as enhanced climate resilience, enhanced climate change mitigation, biodiversity habitat, water filtration, soil health, and amenity values high 2 train +8023 AR6_WGIII 1827 6 Nature-based solutions have large potential to address climate change and other sustainable development issues high 2 train +8024 AR6_WGIII 1827 13 REDD+ can significantly contribute to climate change mitigation and also produce other co-benefits like climate change adaptation, biodiversity conservation, and poverty reduction, if well-implemented high 2 test +8025 AR6_WGIII 1827 45 The development of nature-based solutions faces barriers that relate to the value proposition, value delivery and value capture of nature- based solutions business models and sustainable sources of public/ private finance to tap into high 2 train +8026 AR6_WGIII 1828 7 Policies and frameworks to expand and enhance financial inclusion also extend to the area of climate finance high 2 train +8027 AR6_WGIII 1828 14 At the level of public multilateral climate funds, there have been significant improvements in integrating gender equality and women’s empowerment issues in the governance structures, policies, project approval and implementation processes of existing multilateral climate funds such as the UNFCCC’s funds managed by the Global Environment Facility, the Green Climate Fund and the World Bank’s CIFs high 2 train +8028 AR6_WGIII 1828 24 Overall, efforts to promote gender responsive/sensitive climate finance, at national and local levels, both in the public and private dimensions and more specifically in mitigation-oriented sectors such as clean and renewable energy, remain deficient high 2 train +8029 AR6_WGIII 1863 1 Innovation has also led to, and exacerbated, trade-offs in relation to sustainable development high 2 train +8030 AR6_WGIII 1863 4 The currently widespread implementation of solar photovoltaic (solar PV) and light-emitting diodes (LEDs), for instance, could not have happened without technological innovation high 2 train +8031 AR6_WGIII 1863 7 Trade-offs include negative externalities – for instance, greater environmental pollution and social inequalities – rebound effects leading to lower net emission reductions or even increases in emissions, and increased dependency on foreign knowledge and providers high 2 train +8032 AR6_WGIII 1863 10 This systemic view of innovation takes into account the role of actors, institutions and their interactions, and can inform how innovation systems that vary across technologies, sectors and countries, can be strengthened high 2 train +8033 AR6_WGIII 1863 19 A combination of scaled-up innovation investments with demand-pull interventions can achieve faster technology unit cost reductions and more rapid scale-up than either approach in isolation high 2 train +8034 AR6_WGIII 1863 28 This inhibition occurs more often in developing countries high 2 train +8035 AR6_WGIII 1863 30 Capabilities play a key role in these functions, the build-up of which can be enhanced by domestic measures, but also by international cooperation high 2 train +8036 AR6_WGIII 1863 34 The effectiveness of such international cooperation arrangements, however, depends on the way they are developed and implemented high 2 train +8037 AR6_WGIII 1863 37 International diffusion of low-emission technologies is also facilitated by knowledge spillovers from regions engaged in clean R&D medium 1 train +8038 AR6_WGIII 1863 39 Some literature suggests that it is a barrier, while other sources suggest that it is an enabler to the diffusion of climate-related technologies medium 1 train +8039 AR6_WGIII 1864 4 These gaps could be filled by enhancing financial support for international technology cooperation, by strengthening cooperative approaches, and by helping build suitable capacity in developing countries across all technological innovation system functions high 2 train +8040 AR6_WGIII 1864 6 For example, despite building a large market for mitigation technologies in developing countries, the lack of a systemic perspective in the implementation of the Clean Development Mechanism, operational since the mid- 2000s, has only led to some technology transfer, especially to larger developing countries, but limited capacity building and minimal technology development medium 1 train +8041 AR6_WGIII 1864 9 Addressing both sets of challenges simultaneously presents multiple and recurrent obstacles that systemic approaches to technological change could help resolve, provided they are well managed high 2 train +8042 AR6_WGIII 1864 10 Obstacles include both entrenched power relations dominated by vested interests that control and benefit from existing technologies, and governance structures that continue to reproduce unsustainable patterns of production and consumption medium 1 train +8043 AR6_WGIII 1864 16 System-level rebound effects may also occur high 2 train +8044 AR6_WGIII 1864 21 At present, the understanding of both the direct and indirect impacts of digitalisation on energy use, carbon emissions and potential mitigation, is limited medium 1 train +8045 AR6_WGIII 1864 24 Inspiration can be drawn from the global unit cost reductions of solar PV, which were accelerated by a combination of factors interacting in a mutually reinforcing way across a limited group of countries high 2 train +8046 AR6_WGIII 1870 40 Spillovers related to energy and low-carbon technologies have been documented by a number of empirical studies high 2 train +8047 AR6_WGIII 1870 44 The presence of spillovers can have both positive and negative impacts on climate change mitigation high 2 train +8048 AR6_WGIII 1870 55 In addition, the presence of spillovers implies that a unilateral effort to reduce emissions in one region could reduce emissions in other regions medium 1 train +8049 AR6_WGIII 1917 20 In conjunction with other enabling conditions, technological innovation can support system transitions to limit warming, help shift development pathways, and bring about new and improved ways of delivering goods and services that are essential to human well-being high 2 train +8050 AR6_WGIII 1917 22 Trade-offs include negative externalities, such as environmental impacts and social inequalities, rebound effects leading to lower net emission reductions or even increases in emissions, and increased dependency on foreign knowledge and providers high 2 train +8051 AR6_WGIII 1917 23 Digitalisation, for example, holds both opportunity for emission reduction and emission-saving behaviour change, but at the same time causes significant environmental, social and greenhouse gas (GHG) impacts high 2 train +8052 AR6_WGIII 1917 28 Like other enabling conditions, technological innovation plays a balancing role – by inhibiting change as innovation strengthens incumbent technologies and practices – and a reinforcing role, by allowing new technologies and practices to disrupt the existing socio-technical regimes high 2 train +8053 AR6_WGIII 1917 31 Against this backdrop, international cooperation on technological innovation is one of the enablers of climate action in developing countries on both mitigation and adaptation high 2 train +8054 AR6_WGIII 2011 1 Sustainable Development Goals SectorSectoral mitigation optionsSDG 1 End poverty SDG 2 Zero hungerSDG 3 Good health and wellbeingSDG 4 Quality education SDG 5 Gender equalitySDG 6 Clean water and sanitationSDG 7 Affordable and clean energy SDG 8 Decent work and economic growthSDG 9 Industry, innovation and infrastructure SDG 10 Reduced inequalitiesSDG 11 Sustainable cities and communitiesSDG 12 Responsible consumption and productionSDG 13 Climate action SDG 14 Life below water SDG 15 Life on landSDG 16 Peace, justice and strong institutions SDG 17 PartnershipLine of sight (section numbers, tables, figures, box)Remarks (context specificity/scale) Energy systemsWind energy+ Wind can provide low-cost electricity to several communities (high confidence)± Land use for wind energy needs to be coordinated based on local circumstances, otherwise can have negative implications on food security (medium confidence)+ Minimal air pollution, also integration with health sector frequently discussed (high confidence)+ Low consumption of water (high confidence)+ Low-cost and low-carbon electricity in several regions (high confidence)+ Large job creation per unit investment (medium confidence)+ Integration with offshore and other infrastructure (medium confidence)+ Could help through net metering (medium confidence)± Cater to sustain- able production, however has significant material consumption and disposal needs (medium confidence)+ Low-carbon emissions (high confidence)± Offshore wind could pose risk to marine life if not appropriately managed (high confidence)± Land use for wind energy needs to be coordinated, otherwise can have negative implications on biodiversity medium 1 train +8055 AR6_WGIII 2011 2 Will likely require significantly more critical minerals Key context would include availability of land that does not compromise biodiversity Solar energy+ Solar PV can provide low-cost electricity to several communities (high confidence)± Land use for solar energy needs to be coordinated based on local circum- stances, otherwise can have negative implications on food security (medium confidence)+ Minimal air pollution, also inte- gration with health sector frequently discussed (high confidence)± Low consumption of water for PV but higher for CSP (Concentrating Solar Power) (high confidence)+ Low-cost and low- carbon electricity in several regions (high confidence)+ Large job creation per unit investment (medium confidence)+ Solar heat may be used in industrial heating (medium confidence)+ Could help through net metering (medium confidence)± Cater to sustain- able production, how- ever has significant material consumption and disposal needs (medium confidence)+ Low-carbon emissions (high confidence)± Land use for solar energy needs to be coordinated, otherwise can have negative implications on biodiversity medium 1 train +8056 AR6_WGIII 2011 5 Large-scale bioenergy projects with CCS may be expensive medium 1 test +8057 AR6_WGIII 2011 6 The reason for the tradeoff is competition between food and bioenergy crops (medium confidence)± Depending on the scale and infrastructural efficacy, bioenergy may result in good or poor air quality (medium confidence)± Some bioenergy feedstocks may cause competition for water high 2 train +8058 AR6_WGIII 2011 7 High depen- dence of 2.2 billion people on traditional and non-sustainable biomass, with nega- tive impact on health and deforestation (high confidence)+ Potential to provide employ- ment, including to workers who may be transitioning from fossil sectors (high confidence)+ Considerable opportunities for integration with other industries such as wastewater treatment (high confidence)+ Could lead to low- carbon transport fuels (high confidence)+ Use of waste biomass could be useful (high confidence)+ Low-carbon emissions high 2 train +8059 AR6_WGIII 2011 8 At the same time, effluents from biofuel production can also cause negative impacts on marine ecosystems when effluent treat- ment is not meeting high standards (high confidence)± Land use needs to be coordinated, otherwise can have negative implications on biodiversity and food production (high confidence)Sections 6.4.2.6 and 12.5, Box 6.1The regional context in terms of the types of biomass/land being utilised is critical Has potential for development of low-carbon organic materials, chemicals and plastics that can be produced Hydropower± Could lead to fisheries damage if not properly managed (medium confidence)+ Minimal air pollution (high confidence)+ Coordination with water infrastructure (medium confidence)+ Low-cost and low-carbon electricity in several regions (high confidence)+ Low-carbon emissions (high confidence)± Could lead to fish- eries damage if not properly managed (medium confidence)± Land use needs to be coordinated, otherwise can have negative implications on biodiversity (medium confidence)Section 6.4.2Key context would include availability of land that does not compromise biodiversity Geothermal energy+ Potential to provide energy in several energy- scarce regions (low confidence)± Low air pollution but some water pollution risks (medium confidence)± Water flowback, water pollution and other issues (medium confidence)+ Low-cost and low-carbon electric- ity and heat in several regions (high confidence)+ Heat may be used in industrial heating (medium confidence)+ Potential for air conditioning and heating (medium confidence)+ Low-carbon emissions high 2 train +8060 AR6_WGIII 2011 9 Much Literature on both the health benefits as well as risks arising from such power plants (high confidence)– Significant water consumption (high confidence)± Increased use of nuclear power can provide stable baseload power supply and reduce price volatility but has nuclear waste management needs (medium confidence)+ Local employment and reduced price volatility medium 1 train +8061 AR6_WGIII 2011 11 In the latter case, a wide range of infrastructure issues need to be addressed, including facilities and equipment, as well as human and financial resources, and the legal and regulatory framework (medium confidence)± Cater to sustain- able production, low resource consumption, but has significant waste management needs (medium confidence)+ Low-carbon emissions high 2 train +8062 AR6_WGIII 2011 12 Long- term solutions for high-level radioactive waste are under development (medium confidence)± Low impacts to biodiversity but high impact in case of an accident (medium confidence)Section 6.4.2.4, Figure 6.18Depends on the type of power plants being displaced Nuclear requires significantly less minerals than other low-carbon technologies Carbon capture and storage (CCS)+ CCS infrastructure generally requires reduction of air pol- lutants for optimal operation (high confidence)– Water use generally increases significantly; significant water treatment needs may also arise for brines (high confidence)+ Potential to spur technological innova- tion; also could reduce inequity risks for fossil workers (high confidence)+ Could help decarbonise some hard- to-decarbonise sectors (high confidence)± Use of resources and chemicals could increase unless appropriately managed (medium confidence)+ Low-carbon emissions high 2 train +8063 AR6_WGIII 2012 2 Water management for reducing drought and adapting to climate change is important issue high 2 train +8064 AR6_WGIII 2012 3 Water management for reducing drought and adapting to climate change is important issue (high confidence)± Probably no direct impact (soil–human health nexus through nutritional transfer: may contribute to better nutrient security through quality and nutrient-rich products and better living if higher prof- its and diversified products) (medium confidence)+ Better landscape water balance by influencing the quality and availability of water supply (high confidence)+ Better soil management can lead to improved productivity and thus economic growth (medium confidence)± Low environment footprints, quality and healthy food production and economic and social viability (high confidence)+ Clear climate benefit (high confidence)+ Better sponge function to life in water, and less nutrients into the water (medium confidence)+ Proved beneficial for combating soil degradation and improving soil health and beneficial to biodiversity (high confidence)+ Securing local food production and higher and stable profits may reduce migration and prevent conflict and support peace and justice medium 1 train +8065 AR6_WGIII 2012 6 Trade-offs with other uses of the organic matter Reduce CH 4 and N2O emissions in agriculture± When part of improved agricul- ture, it may reduce hunger (low confidence)+ Cleaner air and soil-plant-herbivore nexus (high confidence)± Less use of water and less manure into water streams (low confidence)± Requires innova- tive food production (low confidence)+ Reduction of emissions (which is a part of responsible production) (medium confidence)+ Clear reduction of emissions (high confidence)+ Less impact on water (high confidence)+ Less impact on land (high confidence)Section 7.4Risks include mitigation persistence, ecological impacts associated with improving feed quality and supply, or potential toxicity and animal welfare issues concerning feed additives Reduced conver - sion of forests and other ecosystems (deforestation, loss and degrada- tion of peatlands and coastal wetlands)± Protecting huge areas may lead to poverty (medium confidence)– May lead to some competition for land (medium confi- dence)+ Cleaner air, greener environ- ment generally leads to better health (high confidence)+ Better landscape water balance (high confidence)± May lead to competition for land and less economic benefits (medium confidence)± When surrounding cities, it may lead to cooling (high confidence)+ Clear climate benefit (high confidence)+ Better sponge function to life in water, and less nutrients into the water (high confidence)+ Beneficial to biodiversity (high confidence)± May lead to more competition for resources and thus pressures between actors (medium confidence)± May lead to joint forces low 0 train +8066 AR6_WGIII 2012 7 It can compete for land and thus with food provision Ecosystem restoration, reforestation, afforestation+ If it provides income, food and wood products, then neutral to positive medium 1 train +8067 AR6_WGIII 2012 8 Reforestation and forest restoration can have co-benefits for food security, and cleaner air and greener environment generally leads to better health medium 1 test +8068 AR6_WGIII 2012 9 Afforestation (on naturally unforested land) can compound climate-related risks to water security (medium confidence)– May lead to competition for land and less economic benefits (medium confidence)± When done with involvement of locals it can reduce inequality (low confidence)+ When surrounding cities, it may lead to cooling (high confidence)+ Clear climate benefit (high confidence)+ Better sponge function to life in water when done in proper manner (high confidence)+ Beneficial to biodiversity when done in proper manner medium 1 train +8069 AR6_WGIII 2012 10 If car - ried out at massive scale, competition for food will arise; when carried out adapted to local circumstances, and for various needs, the trade-offs are small; may have high opportunity costs Improved sustainable forest management+ If it provides income and wood products, then neutral to positive (high confidence)± If it provides income and wood products, then neutral to positive (low confidence)+ Cleaner air, greener environ- ment generally leads to better health (low confidence)+ Better landscape water balance (medium confidence)± Can lead to improved wood chain, including bioenergy from residues (high confidence)+ If it provides income and wood products, then neutral to positive (medium confidence)+ Can lead to improved wood chain, and biobased product innovation (high confidence)± When done with involvement of locals it can reduce inequality (low confidence)± Can lead to cooling of cities and building with biobased products (medium confidence)+ Most likely climate benefit (medium confidence)+ Better sponge function to life in water, and less nutrients into the water (medium confidence)+ Beneficial to biodiversity high 2 train +8070 AR6_WGIII 2012 12 Reduce food loss and food waste+ Reduced food loss will reduce prices, and may lead to less poverty (high confidence)+ Reduced food loss will reduce prices, and may lead to more food available (high confidence)+ Reduced food loss will reduce prices, and may lead to more food available (high confidence)+ Less use of water (high confidence)+ Lead to less energy use (high confidence)+ Balanced food distribution globally may reduce prices for many (low confidence)+ Balanced food distribution globally may provide more sustainable societies (low confidence)+ Balanced food distribution and reduced losses is part of responsible consumption (high confidence)+ Clear reduction of emissions (high confidence)+ Less impact on water (medium confidence)+ Less impact on land (high confidence)+ Leads to less competition for resources and thus less pressures between actors (medium confidence)Section 7.5Occurs in all societies, there are no trade-offs Shift to balanced, sustainable healthy diets± Depends whether the healthier diet is cheaper, often not (medium confi- dence)+ Balanced food distribution globally may reduce prices for many (low confidence)+ Healthier diets for affluent populations (high confidence)+ Less use of water (high confidence)+ May lead to less energy use (high confidence)± More innovative food production (medium confidence)+ Balanced food globally may reduce prices for many (low confidence)+ Balanced food globally may provide more sustainable societies (low confidence)+ Clear reduction of emissions – Part of responsible production (high confidence)+ Clear reduction of emissions (high confidence)+ Less impact on water (medium confidence)+ Less impact on land (high confidence)Section 7.4Only for affluent societies; synergies occur because it may lead to freeing up land Renewables supply timber, biomass, agri feedstock)± Can lead to both positive and nega- tive outcomes for livelihoods and food security; if it pro- vides income, food and wood products, then neutral to positive (medium confidence)± May lead to competition when done at large scale and not taking into account local circumstances or needs (medium confidence)± Can lead to greener landscape, but can also result in large-scale undesirable changes (medium confidence)± Can lead to better landscape-level water balance when done in proper manner (medium confidence)± Can lead to better and more stable products and energy provision if done properly (medium confidence)+ If it provides income and wood products, then neutral to positive (medium confidence)+ Can lead to improved wood chain, and bio-based product innovation (high confidence)± Clear reduction of emissions up to certain scales and volumes and when negative land-use effects are avoided (high confidence)± Depending on type of land use, less impact on water (medium confidence)± Depending on type of land use, less impact on life on land medium 1 train +8071 AR6_WGIII 2013 1 Sustainable Development Goals SectorSectoral mitigation optionsSDG 1 End poverty SDG 2 Zero hungerSDG 3 Good health and wellbeingSDG 4 Quality education SDG 5 Gender equalitySDG 6 Clean water and sanitationSDG 7 Affordable and clean energy SDG 8 Decent work and economic growthSDG 9 Industry, innovation and infrastructure SDG 10 Reduced inequalitiesSDG 11 Sustainable cities and communitiesSDG 12 Responsible consumption and productionSDG 13 Climate action SDG 14 Life below waterSDG 15 Life on landSDG 16 Peace, justice and strong institutions SDG 17 PartnershipLine of sight (section numbers , tables, figures , box)Remarks (context specificity/ scale) Urban systemsUrban land use and spatial planning(+) Provides employ- ment density and supports productivity (high confidence) (+) Can reduce exposure and vulner - ability to climate change given policy integration (high confidence)(+) Better spatial planning will reduce pressures on land use change, including croplands (high confidence) (–) Growth in urban extent can still reduce cropland if not sufficiently managed (high confidence)(+) Improves access to health infrastruc- ture; improves air quality when coupled to shifting energy use; improves wellbeing with green and blue infrastructure (high confidence)(+) Better spatial planning increases educational oppor - tunities (medium confidence)(+) Can increase equal opportunities and effective participation of women, including urban governance (medium confidence)(+) Can improve water quality, water-use efficiency, water harvesting and wastewater treatment; efficient urbanisation can also reduce GHG emissions from water infrastructure (high confidence)(+) Can reduce energy use and enable access to modern energy infrastructure while urban infrastruc- ture for energy services varies (high confidence)(+) Provides employment density and supports productivity (high confidence)(+) Sustainable urban- isation and settlement planning requires development across all infrastructure sectors (high confidence)(+) Spatial inequalities within cities can be reduced; urban infra- structure gap between cities can be reduced (high confidence) (–) Unintended gentrification and spatial inequalities are still possible (medium confidence)(+) Supports capacity for participatory, inte- grated and sustainable human settlement plan- ning (Target 11.3) and protecting the poor and vulnerable (Target 11.5) (high confidence)(+) Urbanisation with lower material demands will support responsible consumption and production (high confidence) (–) Urban population growth contributes to increased demand for resources with differences in scenarios; increase in urban water demand can increase pressures on water scarcity; over-exploitation of groundwater needs to be avoided (medium confidence)(+) Contributes to both climate mitiga- tion and adaptation given integration in urban planning (high confidence)(+) Can reduce growth in urban expansion that can help protect coastal and marine eco- systems (medium confidence) (–) Urban develop- ment can still impact coastal and marine eco- systems (medium confidence)(+) Can reduce growth in urban expansion that can help protect biodiversity on land and terrestrial and inland freshwaters (high confidence) (–) Urban develop- ment can still impact biodiversity (medium confidence)(+) Has synergies with responsive, inclusive and participatory decision-making at all levels and transparent institu- tions medium 1 train +8072 AR6_WGIII 2013 6 Electrifica- tion of the urban energy system(+) Can address energy poverty that is linked to poverty; eradicating poverty is supported by access to modern energy services for all (medium confidence)(+) Electrification can support welfare; electric stoves can support nutritional food intake (medium confidence) (–) Can have trade-offs if food systems are coupled with electricity and bioenergy (medium confidence)(+) Improves air quality when coupled to shifting energy use as included in the option; Avoids air pollution from energy and transport infrastructure; Sup- ports energy services for quality health services in hospitals (high confidence)(+) Electrification and access to electricity supports quality education and educational attainment (high confidence)(+) Supports equal opportunities, also through electricity for internet access if previously lacking (medium confidence)(+) Renewable energy powered water treatment facilities can sup- port clean water and sanitation (medium confi- dence)(+) Supports renewable energy, energy efficiency and access to affordable, reliable and modern energy; renewable- energy generation technologies can enhance infrastruc- ture resilience (high confidence)(+) Supports tech- nological upgrading, innovation and decent job creation (high confidence)(+) Supports sustain- able and resilient infrastructure and can support domestic tech- nology development; renewable-energy generation tech- nologies can enhance infrastructure resilience high 2 train +8073 AR6_WGIII 2014 1 Sustainable Development Goals SectorSectoral mitigation optionsSDG 1 End poverty SDG 2 Zero hungerSDG 3 Good health and wellbeingSDG 4 Quality education SDG 5 Gender equalitySDG 6 Clean water and sanitationSDG 7 Affordable and clean energy SDG 8 Decent work and economic growthSDG 9 Industry, innovation and infrastructure SDG 10 Reduced inequalitiesSDG 11 Sustainable cities and communitiesSDG 12 Responsible consumption and productionSDG 13 Climate actionSDG 14 Life below waterSDG 15 Life on landSDG 16 Peace, justice and strong institutions SDG 17 PartnershipLine of sight (section numbers, tables, figures, box)Remarks (context specificity/scale) BuildingsDemand- side management + Reduce poverty due to less energy expenditures and other financial savings (high confidence)+ Result in avoiding the ‘heat or eat’ dillemma low 0 train +8074 AR6_WGIII 2014 2 Furthermore, smart controllers and wireless communications that are used for control- ling lighting, windows, HVAC equipment, water heaters, and other building equipment provide many other non-energy benefits such as improved security, access control, fire and other emergency detection and management, and on-time identification of maintenance issues (high confidence)+ Lower energy demand can lead to reduced water demand for thermal cooling at energy production facilities (medium confidence)+ Result in fuel poverty alleviation and improving the security of energy supply high 2 train +8075 AR6_WGIII 2014 5 Also, buildings with high energy efficiency and/or green features are sold/rented at higher prices than conventional, low-energy- efficient houses (high confidence)+ Result in reduced consumption of natural resources (medium confidence)+ Reduce emissions and increase resilience high 2 train +8076 AR6_WGIII 2014 7 Economic impacts (synergies and trade-offs) are associated with reduced energy demand, resulting in lower energy prices, energy efficiency investments, fostering innovation, and improvements in labour productivity Efficient heating, ventilation and air con- ditioning (HVAC)+ Reduce poverty due to less energy expenditures and other financial savings – The distributional costs of some mitigation policies supporting energy efficiency may reduce the dispos- able income of the poor (medium confidence)+ Result in avoiding the ‘heat or eat’ dillemma (low confidence)+ Improve health through better indoor air quality, fuel poverty alleviation, better ambient air quality, and reduction of the heat island effect (high confidence)+ Lower energy demand can lead to reduced water demand for thermal cooling at energy production facilities (medium confidence)+ Result in fuel poverty alleviation and improving the security of energy supply (high confidence)± Result in direct and indirect macroeco- nomic effects (GDP , employment, public budgets) associated with lower energy prices due to the reduced energy demand, energy efficiency invest- ments, and fostering innovation + Also result in improving labour productivity (high confidence)+ The development of ‘green buildings’ can foster innovation – Reduced energy demand can lead to early retirement of fossil energy infrastructure (medium confidence)± Can reduce or increase income inequalities (medium confidence)+ Eliminate major sources (both direct and indirect) of poor air quality (indoor and outdoor) (high confidence)+ Result in reduced consumption of natural resources (medium confidence)+ Reduce emissions and increase resilience high 2 train +8077 AR6_WGIII 2014 8 In this case, appropriate access policies should be designed to efficiently shield poor households from the burden of carbon taxation Economic impacts (synergies and trade-offs) are associated with reduced energy demand, resulting in lower energy prices, energy efficiency investments, fostering innovation, and improvements in labour productivity Efficient appliances+ Reduce poverty due to less energy expenditures and other financial savings – The distributional costs of some mitigation policies supporting energy efficiency may reduce the dispos- able income of the poor medium 1 train +8078 AR6_WGIII 2014 9 Also, improved cookstoves provide better food security (medium confidence)+ Improve health through better indoor air quality, fuel poverty alleviation, better ambient air quality, and reduction of the heat island effect (high confidence) + Reduce school absenteeism due to better indoor conditions, while fuel poverty alleviation increases the availble space at home for reading (medium confidence)+ Efficient cookstoves result in substantial time savings for women, thus increasing the time for rest, communication, education and productive activi- ties (medium confidence)+ Lower energy demand can lead to reduced water demand for thermal cooling at energy production facilities (medium confidence)+ Result in energy/ fuel poverty allevia- tion and improving the security of energy supply (high confidence)± Result in direct and indirect macroeco- nomic effects (GDP , employment, public budgets) associated with lower energy prices due to the reduced energy demand, energy efficiency invest- ments, and fostering innovation + Also result in improving labour productivity (high confidence)– Reduced energy demand can lead to early retirement of fossil energy infrastructure (medium confidence)± Can reduce or increase income inequalities + Efficient cookstoves result in substantial time savings for women and children, thus enhancing education and the development of productive activities (medium confidence)+ Eliminate major sources (both direct and indirect) of poor air quality (indoor and outdoor) (high confidence)+ Result in reduced consumption of natural resources – Possible risks due to the penetration of new, efficient appliances and early retirement of existing equipment (medium confidence)+ Reduce emissions and increase resil- ience (high confidence)+ Result in halting deforesta- tion through efficient cookstoves medium 1 train +8079 AR6_WGIII 2015 1 Sustainable Development Goals SectorSectoral mitigation optionsSDG 1 End poverty SDG 2 Zero hungerSDG 3 Good health and wellbeingSDG 4 Quality education SDG 5 Gender equalitySDG 6 Clean water and san- itationSDG 7 Affordable and clean energy SDG 8 Decent work and economic growthSDG 9 Industry, innovation and infrastructure SDG 10 Reduced inequalitiesSDG 11 Sustainable cities and communitiesSDG 12 Responsible consumption and productionSDG 13 Climate actionSDG 14 Life below waterSDG 15 Life on landSDG 16 Peace, justice and strong institutions SDG 17 PartnershipLine of sight (section num- bers, tables, figures, box)Remarks (context specificity/scale) BuildingsBuilding design and performance+ Reduce poverty due to less energy expenditures and other financial savings medium 1 train +8080 AR6_WGIII 2015 2 Also, green surfaces and urban farming (vertical, basement or unused buildings) contribute to local and resilient food production (medium confidence)+ Improve health through better indoor air quality, fuel poverty alleviation, better ambient air quality, and reduction of the heat island effect high 2 train +8081 AR6_WGIII 2015 3 Also, these measures result in water savings due to improved indoor conditions and lower space of dwellings (medium confidence)+ Result in fuel poverty alleviation and improving the security of energy supply (high confidence)± Result in direct and indirect macro- economic effects (GDP , employment, public budgets) associated with lower energy prices due to the reduced energy demand (high confidence)– Reduced energy demand can lead to early retirement of fossil energy infra- structure (medium confidence)+ Reduce income inequalities (low confidence)+ Eliminate major sources (both direct and indirect) of poor air quality (indoor and outdoor) (high confidence)+ Result in reduced consumption of natural resources (medium confidence)+ Reduce emissions and increase resilience (high confidence)+ Green roofs and walls strengthen urban biodiversity (medium confidence)+ Institutions that are effective, accountable and transparent are needed at all levels of government for boosting sufficiency measures (medium confidence)Section 9.8 and Table 9.5Economic impacts (synergies and trade-offs) are mainly associated with reduced energy demand and the resulting in lower energy prices On-site and nearby production and use of renewables+ Increase the productive time of women and children – The distribu- tional costs of some mitigation policies sup- porting RES (renewables) may reduce the disposable income of the poor (medium confidence)+ Improving energy access enhances agricultural produc- tivity and improves food security – Increased bioenergy production may restrict the available land for food production (medium confidence)+ Improve health through better indoor air quality, energy/fuel poverty alle- viation, better ambient air quality, and reduction of the heat island effect (high confidence) + Reduce school absenteeism due to better indoor conditions and enable people living in poor developing countries to read (medium confidence)+ Improved access to electricity and clean fuels in developing coun- tries will result in substantial time savings for women, thus increasing the time for rest, communication, education and productive activi- ties medium 1 train +8082 AR6_WGIII 2015 4 Also, improved access to electric- ity is necessary to treat water in homes – Switch to bioenergy may increase water use compared to existing conditions (medium confidence)+ Result in energy poverty alleviation and improving the security of energy supply – Risks of reduced energy access, in cases where the distributional costs of mitigation increase the energy costs (medium confidence)± Result in direct and indirect macro- economic effects (GDP , employment, public budgets) asso- ciated with lower energy prices due to the reduced energy demand, RES invest- ments, improved energy access and fostering innovation (high confidence)+ Adoption of RES and smart grids helps in infrastruc- ture improvement and expansion – Increased RES penetration can lead to early retirement of fossil energy infrastructure (high confidence)± Can reduce or increase income inequalities + Improved access to electricity and clean fuels in developing countries will result in substantial time savings for women and children, enhanc- ing education and the development of productive activities (medium confidence)+ Eliminate major sources (both direct and indirect) of poor air quality (indoor and outdoor) (high confidence)+ Result in reduced consump- tion of natural resources (medium confidence)+ Reduce emissions and increase resil- ience (high confidence)+ Result in halting deforesta- tion through improved access to electricity and clean fuels (medium confidence)+ Improved access to electric lighting can improve safety (particularly for women and children) + Institutions that are effective, accountable and transparent are needed at all levels of government for providing energy access and promoting modern renewables (medium confidence)+ The develop- ment of zero energy buildings requires, among other things, capacity building and citizen par - ticipation, as well as monitoring of the achieve- ments medium 1 train +8083 AR6_WGIII 2015 6 Some of the trade-offs are mainly related to the switch to bioenergy, which may restrict the available land for food production and increase water consumption Economic impacts (synergies and trade-offs) are associated with reduced demand for fossil fuels, resulting in lower energy prices, RES investments, fostering innovation, and improvements in energy access Change in construction methods and circular economy+ Improve health through better labour conditions (medium confidence)± The change in construction methods and the devlopment of circular business models can lead to reduced or increased water demand, as a trade-off (medium confidence)+ Result in energy/ fuel poverty allevia- tion and improving the security of energy supply (high confidence)± Result in direct and indirect macro- economic effects (GDP , employment, public budgets) associated with development of smarter construction methods and circular business models + Also result in improving labour productivity (high confidence)+ The develop- ment of smarter construction methods and circular business models can foster innovation (high confidence)+ Result in reduced consumption of natural and scarce resources, waste generation and enviromental impacts (high confidence) + Result in reduced consumption of natural and scarce resources (high confidence)+ Reduce emissions and increase resilience (high confidence)+ The change in construction methods and the development of circular business models requires a better integra- tion and partner - ship between stakeholders (high confidence)Sections 9.4 and 9.5Economic impacts (synergies and trade-offs) are associated with development of smarter construction methods, circular business models, and improvements in labour productivity Change in construction materials+ Bio-based and natural materials, e.g., raw earth, can improve indoor air quality and brings the concept of biophilia – Bio-based and natural materials can be more susceptible to the appearance of biological organisms that can cause health problems (medium confidence)± The change in construc- tion materials can lead to reduced or increased water demand, as a trade-off (medium confidence)+ Result in energy/ fuel poverty allevia- tion and improving the security of energy supply (high confidence)± Result in direct and indirect macro- economic effects (GDP , employment, public budgets) asso- ciated with the use and development of green construction materials (medium confidence) + The use and development of green construction materials can foster innovation (high confidence)+ Result in reduced consumption of natural and scarce resources, waste generation and enviromental impacts (high confidence)+ Result in reduced consumption of natural and scarce resources (high confidence)+ Reduce emissions, removal and storage of CO 2 (bio-based materials) and increase resilience (high confi- dence)– Bio-based materials can increase the pressure and competition for land use (medium confidence)+ The change in construction materials requires a better integration and partnership between stake- holders high 2 train +8084 AR6_WGIII 2016 1 Sustainable Development Goals SectorSectoral mitigation optionsSDG 1 End poverty SDG 2 Zero hungerSDG 3 Good health and wellbeingSDG 4 Quality education SDG 5 Gender equalitySDG 6 Clean water and sanitationSDG 7 Affordable and clean energy SDG 8 Decent work and economic growthSDG 9 Industry, innovation and infrastructure SDG 10 Reduced inequalitiesSDG 11 Sustainable cities and communitiesSDG 12 Responsible consumption and productionSDG 13 Climate action SDG 14 Life below waterSDG 15 Life on landSDG 16 Peace, justice and strong institutions SDG 17 PartnershipLine of sight (section numbers , tables, figures , box)Remarks (context specificity/ scale) TransportFuel efficiency- light duty vehicle+ Improved efficiency reduces costs and makes transport more affordable (high confidence)± Land use for wind energy needs to be coordinated based on local circumstances, otherwise can have negative implications on food security (medium confidence)+ Reduce air pollution/ improve air quality (high confidence)+ Can support the global rate of improvement in overall energy efficiency (high confidence)+ Creation of new jobs due to new investment in fuel efficiency (medium confidence)+ Can reduce air pollution in cities (high confidence)+ Reduction of GHG emissions (high confidence)+ Reduce demand for land needed to produce transportation fuels medium 1 train +8085 AR6_WGIII 2016 7 Strengthened institutional capacity that also supports the scale and coordination of the mitigation options can increase these synergies Electric light- duty vehicles (LDVs)± Land use for solar energy needs to be coordinated based on local circumstances, otherwise can have negative implica- tions on food security medium 1 train +8086 AR6_WGIII 2016 8 BEVs could create jobs associated with the EV supply chain, and could create jobs to build and operate the associated infrastructure low 0 test +8087 AR6_WGIII 2016 9 Large-scale electrification of LDVs requires expansion of low-carbon power systems, while charging or battery swapping infrastructure is needed for some segments (high confidence)± Significant equity issues with EVs in the transition period can be overcome with programmes, for example, by expanding public charging infra- structure (medium confidence)+ Can reduce air and noise pollution in cities (high confidence)± Could increase demand for critical minerals but increased recycling can mitigate this risk (medium confidence)+ Reduction of GHG emissions (high confidence)Sections 10.3, 10.4 and 10.8 Shift to public transport+ Affordable transport access for all + Improve access to health, educa- tion, and other social services lowering the cost of services needed by the low-income/ poor high 2 train +8088 AR6_WGIII 2016 10 The reason for the trade-off is competition between food and bioenergy crops (medium confidence)+ Access to healthcare; reduce air pollution/ inrease air quality (high confidence) + Affordable public transport can increase access to educational resources (high confidence)+ Affordable trans- port access for all (high confidence)+ Improves energy efficiency of transport and makes it more affordable (high confidence)+ Role of transport for economic and human development (high confidence)± Needs adequate infrastructure; in develop- ing countries weather conditions and unreliable connectivity affect the lack of incentives to improve existing public transportation (high confidence)+ Improved invest- ments in public transit increase equity in transport access (high confidence)+ Sustainable transport systems for cities; facili- tates universal access to public transport (high confidence) + Could support positive economic links between urban and peri-urban areas (high confidence) + Can reduce air pollution in cities (high confidence)+ Reduced material consumption during production of vehicles and their operations (high confidence)+ Reduction of GHG emissions (high confidence)Sections 10.2 and 10.8, Table 10.3 Shift to bikes, ebikes and non motorised transport+ Affordable transport access for all + Improve access to health, educa- tion, and other social services, lowering the cost of services needed by the low-income/ poor (high confidence)± Could lead to fisheries damage if not properly managed (medium confidence)± Reduce air pollution; increases physical activity leading to reduced health mor - tality – Traffic crashes discourage the use of bikes + Scaling up active modes (through careful local urban design and transport planning) can reduce gender inequi- ties in access to basic services, healthcare and education (medium confidence)+ Scaling up active modes (through careful local urban design and transport planning) can reduce gender inequities in access to basic services, health- care and educa- tion (medium confidence)+ Scaling up active modes (through careful local urban design and transport planning) can reduce gender inequities in access to basic services (medium confidence)+ Saves energy (high confidence)+ Increases employ- ment opportunities, demand for bike repair shops, bike parking (medium confidence)+ Needs adequate infrastructure + Opportunities including digitalisation, the Internet of Things and also ‘big data' (high confidence)+ Access to bicycle lanes or cycle tracks increases the odds of female commuters using bicycles (medium confidence)+ Compact, polycentric cities where active transport is most viable can enhance inclusive and sustainable urbanisation (high confidence) + Can reduce air pollution in cities (high confidence)+ Reduced material consumption during production of vehicles and their operations (high confidence)+ Reduction of GHG emissions (high confidence)+ Preserve land that would have been otherwise used to construct and maintain parking garages and surface parking lots (medium confidence)Sections 10.2 and 10.8, Table 10.3 Fuel efficiency- heavy duty vehicle+ Improved efficiency reduces costs and makes transport more affordable (high confidence)+ Reduce air pollution/ improve air quality (high confidence)+ Improved efficiency reduces costs and makes transport more affordable (high confidence)+ Can support the global rate of improvement in overall energy efficiency (high confidence)+ Creation of new jobs due to new investment in fuel efficiency (medium confidence)+ Reduction of GHG emissions (high confidence)+ Reduce demand for land needed to produce transportation fuels medium 1 train +8089 AR6_WGIII 2017 1 Sustainable Development Goals SectorSectoral mitigation optionsSDG 1 End poverty SDG 2 Zero hungerSDG 3 Good health and wellbeingSDG 4 Quality education SDG 5 Gender equalitySDG 6 Clean water and sanitationSDG 7 Affordable and clean energy SDG 8 Decent work and economic growthSDG 9 Industry, innovation and infrastructure SDG 10 Reduced inequalitiesSDG 11 Sustainable cities and communitiesSDG 12 Responsible consumption and productionSDG 13 Climate action SDG 14 Life below waterSDG 15 Life on landSDG 16 Peace, justice and strong institutions SDG 17 PartnershipLine of sight (section numbers , tables, figures , box)Remarks (context specificity/ scale) TransportFuel shift (including electricity)- heavy duty vehicles (HDVs)+ Reduce air pollution/improve air quality (medium confidence)+ Some alternative fuels can help increase the share of renewable energy in the global energy mix (medium confidence)+ Could create jobs associated with the supply chain of new fuels + Could create jobs to build and operate the associated infrastructure (low confidence)+ R&D is critical for new fuels and to test the full life cycle costs of various heavy vehicle options; need to invest in supporting infrastructure (high confidence)± Electric vehicles (EVs) and fuel cell vehicles (FCVs) for HDVs could increase demand for critical minerals but increased recycling can mitigate this risk (medium confidence)+ Reduction of GHG emissions high 2 train +8090 AR6_WGIII 2017 7 Strengthened institutional capacity that also supports the scale and coordination of the mitigation options can increase these synergiesShipping efficiency, logistics optimisation, new fuels+ Some alternative fuels can help increase the share of renewable energy in the global energy mix (medium confidence) + Can support the global rate of improvement in over - all energy efficiency (medium confidence)+ Could create jobs associated with the supply chain of new fuels + Could create jobs to build and operate the associated infrastructure (low confidence)+ R&D is critical for new fuels and to test the full life cycle costs of various heavy vehicle options; need to invest in supporting infrastructure (high confidence)+ Reduction of GHG emissions (high confidence)Sections 10.6 and 10.8 Aviation- energy efficiency, new fuels+ Some alternative fuels can help increase the share of renewable energy in the global energy mix (medium confidence) + Can support the global rate of improvement in over - all energy efficiency (medium confidence)+ Could create jobs associated with the supply chain of new fuels + Could create jobs to build and operate the associated infrastructure (low confidence)+ R&D is critical for new fuels and to test the full life cycle costs of various heavy vehicle options; need to invest in supporting infrastructure (high confidence)+ Reduction of GHG emissions (high confidence)Sections 10.5 and 10.8 Biofuels– Using land to produce biofuels could put stress on global food systems + Could increase incomes for farmers and support invest- ments in rural infra- structure medium 1 train +8091 AR6_WGIII 2017 8 The biofuels supply chain may nega- tively impact air quality (e.g., due to increased fertiliser use) (medium confidence)+ Can help increase the share of renew- able energy in the global energy mix (high confidence)+ Could create jobs associated with the supply chain of biofuels + Could create jobs to build and operate the associated infrastructure (low confidence)+ R&D is critical for new fuels and to test the full life cycle costs of various heavy vehicle options; need to invest in supporting infrastructure (high confidence)+ Could reduce air pollution in cities (medium confidence)+ Reduction of GHG emissions (high confidence)± Could increase eutrophication in water bodies (high confidence)± Additional land use for biofuels may increase pressue on biodiversity high 2 train +8092 AR6_WGIII 2018 1 Sustainable Development Goals SectorSectoral mitigation optionsSDG 1 End poverty SDG 2 Zero hungerSDG 3 Good health and wellbeingSDG 4 Quality education SDG 5 Gender equalitySDG 6 Clean water and sanitationSDG 7 Affordable and clean energy SDG 8 Decent work and economic growthSDG 9 Industry, innovation and infrastructure SDG 10 Reduced inequalitiesSDG 11 Sustainable cities and communitiesSDG 12 Responsible consumption and productionSDG 13 Climate action SDG 14 Life below waterSDG 15 Life on landSDG 16 Peace, justice and strong institutions SDG 17 PartnershipLine of sight (section numbers , tables, figures , box)Remarks (context specificity/scale) IndustryEnergy efficiency+ Reduce air pollution (medium confidence)+ Enhances security in clean energy (high confidence)+ Employment opportunities in a green economy (high confidence) + Industrial innovation through new technologies (high confidence)+ Contributes to climate action through efficient use of energy (high confidence)Section 11.5.3Heavily dependent on technology and so the scale of the continous co-benefits across regions would depend on the extent and ease of technological transfer Material efficiency and demand reduction+ Reduce the pressures on water bodies (low confidence)+ New Business Models generate employment oppor - tunities (medium confidence) – Reduction in national sales tax revenue in medium term (low confidence)+ Infrastructural development to support mitigation option (medium confidence)+ Environmental stewardship (medium confidence)+ Contributes to climate action through reduced consumption (high confidence)Section 11.5.3The scale of the co-benefits achieved through material efficiency would depend on the extent to which the transition from traditional to requisite new business models can be achieved Circular material flows+ Reduce air pollution (medium confidence)+ Increase use of waste as resource (high confidence)+ Improved energy efficiency as key CE practice (medium confidence)+ Job opportunities through new business models (medium confidence)+ Public environ- mental awareness (medium confidence)+Enhances environmental benefits + Increase use of waste as resource (high confidence)+ Studies reported direct relationship between CE and SDG 14 (high confidence)+ Enhances biodiversity protec- tion on land (low confidence)+ Improved social relations between industrial sectors and local societies (medium confidence)Section 11.5.3Successful implementation of transformational new business models is required to scale up and derive extended co-benefits through the CE strategy Electrification+ Supports poverty alleviation strategies (high confidence)+ Improved food security – Fuel switching to options such as biomass and bioenergy can have negative impact on food prices (medium confidence)+ Supports delivery of health services + Improves indoor air quality compared to biomass use (high confidence)+ Reduces energy-related hurdles domestically affecting women (high confidence)+ Decarbonisation of grid when fuel is switched to cleaner sources (high confidence)+ Increased economic activity and employment (high confidence)+ Contributes to climate action through switching to renewables (high confidence)– Negative impact on SDG 15 [fuel switching to options such as biomass and bioenergy] (high confidence)Sections 11.5.3 and 6.7.7The extent of the co-benefits experience on social system would be relative as it would be dependent on their current access to energy CCS and carbon capture and utilisation (CCU)+ Control of non-CO 2 pollutants (such as sulphur dioxide) – increase of non- CO2 pollutants (such as particulate matter, nitrogen oxide and ammonia) (high confidence)– Deployment of CCS and CCU would require increased water consumption (high confidence)+ Decarbonisation of energy production through utilisation of CO2 (high confidence) – Deployment of CCS and CCU would require high energy demand (high confidence) + Diversified employment prospects (low confidence)+ Direct foreign investment and know-how (medium confidence)+ Deployment of CCS and CCU would contribute to enhancing the sustainability of cities (high confidence)+ Contributes to climate action through carbon capture (high confidence)– Deployment of CCS and CCU would require additional land use high 2 train +8093 AR6_WGIII 2058 21 For pathways that limit warming to 2°C (>67%) or lower, using GWP100 to inform cost-effective abatement choices between gases would achieve these long-term temperature goals at close to least global cost within a few percent high 2 test \ No newline at end of file