Climate change

Page semi-protected
Listen to this article
Source: Wikipedia, the free encyclopedia.
(Redirected from Global warming)

The global map shows sea temperature rises of 0.5 to 1 degree Celsius; land temperature rises of 1 to 2 degree Celsius; and Arctic temperature rises of up to 4 degrees Celsius.
Changes in surface air temperature over the past 50 years.[1] The Arctic has warmed the most, and temperatures on land have generally increased more than sea surface temperatures.
Earth's average surface air temperature has increased almost 1.5 °C (about 2.5 °F) since the Industrial Revolution. Natural forces cause some variability, but the 20-year average shows the progressive influence of human activity.[2]

In common usage, climate change describes global warming—the ongoing increase in global average temperature—and its effects on Earth's climate system. Climate change in a broader sense also includes previous long-term changes to Earth's climate. The current rise in global average temperature is more rapid than previous changes, and is primarily caused by humans burning fossil fuels.[3][4] Fossil fuel use, deforestation, and some agricultural and industrial practices add to greenhouse gases, notably carbon dioxide and methane.[5] Greenhouse gases absorb some of the heat that the Earth radiates after it warms from sunlight. Larger amounts of these gases trap more heat in Earth's lower atmosphere, causing global warming.

Climate change has an increasingly large impact on the environment. Deserts are expanding, while heat waves and wildfires are becoming more common.[6] Amplified warming in the Arctic has contributed to thawing permafrost, retreat of glaciers and sea ice decline.[7] Higher temperatures are also causing more intense storms, droughts, and other weather extremes.[8] Rapid environmental change in mountains, coral reefs, and the Arctic is forcing many species to relocate or become extinct.[9] Even if efforts to minimise future warming are successful, some effects will continue for centuries. These include ocean heating, ocean acidification and sea level rise.[10]

Climate change threatens people with increased flooding, extreme heat, increased food and water scarcity, more disease, and economic loss. Human migration and conflict can also be a result.[11] The World Health Organization (WHO) calls climate change the greatest threat to global health in the 21st century.[12] Societies and ecosystems will experience more severe risks without action to limit warming.[13] Adapting to climate change through efforts like flood control measures or drought-resistant crops partially reduces climate change risks, although some limits to adaptation have already been reached.[14] Poorer communities are responsible for a small share of global emissions, yet have the least ability to adapt and are most vulnerable to climate change.[15][16]

Bobcat Fire in Monrovia, CA, September 10, 2020
Bleached colony of Acropora coral
A dry lakebed in California, which is experiencing its worst megadrought in 1,200 years.[17]
Examples of some effects of climate change: Wildfire intensified by heat and drought, bleaching of corals occurring more often due to marine heatwaves, and worsening droughts compromising water supplies.

Many climate change impacts have been felt in recent years, with 2023 the warmest on record at +1.48 °C (2.66 °F).[18] Additional warming will increase these impacts and can trigger tipping points, such as melting all of the Greenland ice sheet.[19] Under the 2015 Paris Agreement, nations collectively agreed to keep warming "well under 2 °C". However, with pledges made under the Agreement, global warming would still reach about 2.7 °C (4.9 °F) by the end of the century.[20] Limiting warming to 1.5 °C will require halving emissions by 2030 and achieving net-zero emissions by 2050.[21]

Strategies to phase out fossil fuels involve conserving energy, generating clean and sustainable electricity, and using electricity to power transportation, heat buildings, and operate industrial facilities. The electricity supply can be made cleaner and more plentiful by vastly increasing the deployment of wind, and solar power, alongside other forms of renewable energy and nuclear power.[22][23] Carbon can also be removed from the atmosphere, for instance by increasing forest cover and farming with methods that capture carbon in soil.[24]

Terminology

Before the 1980s it was unclear whether the warming effect of increased greenhouse gases was stronger than the cooling effect of airborne particulates in air pollution. Scientists used the term inadvertent climate modification to refer to human impacts on the climate at this time.[25]

In the 1980s, the terms global warming and climate change became more common, often being used interchangeably.[26] Scientifically, global warming refers only to increased surface warming, while climate change describes both global warming and its effects on Earth's climate system, such as precipitation changes.[25] Climate change can also be used more broadly to include changes to the climate that have happened throughout Earth's history.[27] Global warming—used as early as 1975[28]—became the more popular term after NASA climate scientist James Hansen used it in his 1988 testimony in the U.S. Senate.[29] Since the 2000s, climate change has increased usage.[30]

Various scientists, politicians and media now use the terms climate crisis or climate emergency to talk about climate change, and global heating instead of global warming.[31]

Global temperature rise

Global surface temperature reconstruction over the last 2000 years using proxy data from tree rings, corals, and ice cores in blue.[32] Directly observed data is in red.[33]

Temperature records prior to global warming

Prior to human evolution the record includes hotter temperatures and occasional abrupt changes, such as the Paleocene–Eocene Thermal Maximum 55.5 million years ago.[34]

Over the last few million years Human beings evolved in a climate that cycled through ice ages, with global average temperature ranging between current levels and 5–6 °C colder than today.[35][36] Historical patterns of warming and cooling, like the Medieval Warm Period and the Little Ice Age, did not occur at the same time across different regions. Temperatures may have reached as high as those of the late 20th century in a limited set of regions.[37] Climate information for that period comes from climate proxies, such as trees and ice cores.[38]

Warming since the Industrial Revolution

In recent decades, new high temperature records have substantially outpaced new low temperature records on a growing portion of Earth's surface.[39]
There has been an increase in ocean heat content during recent decades as the oceans absorb over 90% of the heat from global warming.[40]

Around 1850 thermometer records began to provide global coverage.[41] Between the 18th century and 1970 there was little net warming, as the warming impact of greenhouse gas emissions was offset by cooling from sulfur dioxide emissions. Sulfur dioxide causes acid rain, but it also produces sulfate aerosols in the atmosphere, which reflect sunlight and cause so-called global dimming. After 1970, the increasing accumulation of greenhouse gases and controls on sulfur pollution led to a marked increase in temperature.[42][43][44]

Multiple independent datasets all show worldwide increases in surface temperature,[45] at a rate of around 0.2 °C per decade.[46] The 2013-2022 decade warmed to an average 1.15 °C [1.00–1.25 °C] compared to the pre-industrial baseline (1850–1900).[47] Not every single year was warmer than the last: internal climate variability processes can make any year 0.2 °C warmer or colder than the average.[48] From 1998 to 2013, negative phases of two such processes, Pacific Decadal Oscillation (PDO)[49] and Atlantic Multidecadal Oscillation (AMO).[50] caused a so-called "global warming hiatus".[51] After the hiatus, the opposite occurred, with years like 2023 exhibiting temperatures well above even the recent average.[52] This is why the temperature change is defined in terms of a 20-year average, which minimises the noise of hot and cold years and decadal climate patterns, and detects the long-term signal.[53]: 5 [54]

A wide range of other observations reinforce the evidence of warming.[55][56] The upper atmosphere is cooling, because greenhouse gases are trapping heat near the Earth's surface, and so less heat is radiating into space.[57] Warming reduces average snow cover and forces the retreat of glaciers. At the same time, warming also causes greater evaporation from the oceans, leading to more atmospheric humidity, more and heavier precipitation.[58] Plants are flowering earlier in spring, and thousands of animal species have been permanently moving to cooler areas.[59]

Differences by region

Different regions of the world warm at different rates. The pattern is independent of where greenhouse gases are emitted, because the gases persist long enough to diffuse across the planet. Since the pre-industrial period, the average surface temperature over land regions has increased almost twice as fast as the global-average surface temperature.[60] This is because oceans lose more heat by evaporation and oceans can store a lot of heat.[61] The thermal energy in the global climate system has grown with only brief pauses since at least 1970, and over 90% of this extra energy has been stored in the ocean.[62][63] The rest has heated the atmosphere, melted ice, and warmed the continents.[64]

The Northern Hemisphere and the North Pole have warmed much faster than the South Pole and Southern Hemisphere. The Northern Hemisphere not only has much more land, but also more seasonal snow cover and sea ice. As these surfaces flip from reflecting a lot of light to being dark after the ice has melted, they start absorbing more heat.[65] Local black carbon deposits on snow and ice also contribute to Arctic warming.[66] Arctic surface temperatures are increasing between three and four times faster than in the rest of the world.[67][68][69] Melting of ice sheets near the poles weakens both the Atlantic and the Antarctic limb of thermohaline circulation, which further changes the distribution of heat and precipitation around the globe.[70][71][72][73]

Future global temperatures

CMIP6 multi-model projections of global surface temperature changes for the year 2090 relative to the 1850–1900 average. The current trajectory for warming by the end of the century is roughly half way between these two extremes.[20][74][75]

The World Meteorological Organization estimates a 66% chance of global temperatures exceeding 1.5 °C warming from the preindustrial baseline for at least one year between 2023 and 2027.[76][77] Because the IPCC uses a 20 year average to define global temperature changes, a single year exceeding 1.5 °C does not break the limit.

The IPCC expects the 20-year average global temperature to exceed +1.5 °C in the early 2030s.[78] The IPCC Sixth Assessment Report (2023) included projections that by 2100 global warming is very likely to reach 1.0-1.8 °C under a scenario with very low emissions of greenhouse gases, 2.1-3.5 °C under an intermediate emissions scenario, or 3.3-5.7 °C under a very high emissions scenario.[79] In the intermediate and high emission scenarios, the warming will continue past 2100.[80][81]

The remaining carbon budget for staying beneath certain temperature increases is determined by modelling the carbon cycle and climate sensitivity to greenhouse gases.[82] According to the IPCC, global warming can be kept below 1.5 °C with a two-thirds chance if emissions after 2018 do not exceed 420 or 570 gigatonnes of CO2. This corresponds to 10 to 13 years of current emissions. There are high uncertainties about the budget. For instance, it may be 100 gigatonnes of CO2 equivalent smaller due to CO2 and methane release from permafrost and wetlands.[83] However, it is clear that fossil fuel resources need to be proactively kept in the ground to prevent substantial warming. Otherwise, their shortages would not occur until the emissions have already locked in significant long-term impacts.[84]

Causes of recent global temperature rise

Drivers of climate change from 1850–1900 to 2010–2019. There was no significant contribution from internal variability or solar and volcanic drivers.

The climate system experiences various cycles on its own which can last for years, decades or even centuries. For example, El Niño events cause short-term spikes in surface temperature while La Niña events cause short term cooling.[85] Their relative frequency can affect global temperature trends on a decadal timescale.[86] Other changes are caused by an imbalance of energy from external forcings.[87] Examples of these include changes in the concentrations of greenhouse gases, solar luminosity, volcanic eruptions, and variations in the Earth's orbit around the Sun.[88]

To determine the human contribution to climate change, unique "fingerprints" for all potential causes are developed and compared with both observed patterns and known internal climate variability.[89] For example, solar forcing—whose fingerprint involves warming the entire atmosphere—is ruled out because only the lower atmosphere has warmed.[90] Atmospheric aerosols produce a smaller, cooling effect. Other drivers, such as changes in albedo, are less impactful.[91]

Greenhouse gases

CO2 concentrations over the last 800,000 years as measured from ice cores[92][93][94][95] (blue/green) and directly[96] (black)

Greenhouse gases are transparent to sunlight, and thus allow it to pass through the atmosphere to heat the Earth's surface. The Earth radiates it as heat, and greenhouse gases absorb a portion of it. This absorption slows the rate at which heat escapes into space, trapping heat near the Earth's surface and warming it over time.[97]

While water vapour (≈50%) and clouds (≈25%) are the biggest contributors to the greenhouse effect, they primarily change as a function of temperature and are therefore mostly considered to be feedbacks that change climate sensitivity. On the other hand, concentrations of gases such as CO2 (≈20%), tropospheric ozone,[98] CFCs and nitrous oxide are added or removed independently from temperature, and are therefore considered to be external forcings that change global temperatures.[99]

Before the Industrial Revolution, naturally-occurring amounts of greenhouse gases caused the air near the surface to be about 33 °C warmer than it would have been in their absence.[100][101] Human activity since the Industrial Revolution, mainly extracting and burning fossil fuels (coal, oil, and natural gas),[102] has increased the amount of greenhouse gases in the atmosphere, resulting in a radiative imbalance. In 2019, the concentrations of CO2 and methane had increased by about 48% and 160%, respectively, since 1750.[103] These CO2 levels are higher than they have been at any time during the last 2 million years. Concentrations of methane are far higher than they were over the last 800,000 years.[104]

The Global Carbon Project shows how additions to CO2 since 1880 have been caused by different sources ramping up one after another.

Global anthropogenic greenhouse gas emissions in 2019 were equivalent to 59 billion tonnes of CO2. Of these emissions, 75% was CO2, 18% was methane, 4% was nitrous oxide, and 2% was fluorinated gases.[105] CO2 emissions primarily come from burning fossil fuels to provide energy for transport, manufacturing, heating, and electricity.[5] Additional CO2 emissions come from deforestation and industrial processes, which include the CO2 released by the chemical reactions for making cement, steel, aluminum, and fertiliser.[106] Methane emissions come from livestock, manure, rice cultivation, landfills, wastewater, and coal mining, as well as oil and gas extraction.[107] Nitrous oxide emissions largely come from the microbial decomposition of fertiliser.[108]

While methane only lasts in the atmosphere for an average of 12 years,[109] CO2 lasts much longer. The Earth's surface absorbs CO2 as part of the carbon cycle. While plants on land and in the ocean absorb most excess emissions of CO2 every year, that CO2 is returned back to the atmosphere when biological matter is digested, burns, or decays.[110] Land-surface carbon sink processes, such as carbon fixation in the soil and photosynthesis, remove about 29% of annual global CO2 emissions.[111] The ocean has absorbed 20 to 30% of emitted CO2 over the last 2 decades.[112] CO2 is only removed from the atmosphere for the long term when it is stored in the Earth's crust, which is a process that can take millions of years to complete.[110]

Land surface changes

The rate of global tree cover loss has approximately doubled since 2001, to an annual loss approaching an area the size of Italy.[113]

According to Food and Agriculture Organization, around 30% of Earth's land area is largely unusable for humans (glaciers, deserts, etc.), 26% is forests, 10% is shrubland and 34% is agricultural land.[114] Deforestation is the main land use change contributor to global warming,[115] as the destroyed trees release CO2, and are not replaced by new trees, removing that carbon sink.[24] Between 2001 and 2018, 27% of deforestation was from permanent clearing to enable agricultural expansion for crops and livestock. Another 24% has been lost to temporary clearing under the shifting cultivation agricultural systems. 26% was due to logging for wood and derived products, and wildfires have accounted for the remaining 23%.[116] Some forests have not been fully cleared, but were already degraded by these impacts. Restoring these forests also recovers their potential as a carbon sink.[117]

Local vegetation cover impacts how much of the sunlight gets reflected back into space (albedo), and how much heat is lost by evaporation. For instance, the change from a dark forest to grassland makes the surface lighter, causing it to reflect more sunlight. Deforestation can also modify the release of chemical compounds that influence clouds, and by changing wind patterns.[118] In tropic and temperate areas the net effect is to produce significant warming, and forest restoration can make local temperatures cooler.[117] At latitudes closer to the poles, there is a cooling effect as forest is replaced by snow-covered (and more reflective) plains.[118] Globally, these increases in surface albedo have been the dominant direct influence on temperature from land use change. Thus, land use change to date is estimated to have a slight cooling effect.[119]

Other factors

Aerosols and clouds

Air pollution, in the form of aerosols, affects the climate on a large scale.[120] Aerosols scatter and absorb solar radiation. From 1961 to 1990, a gradual reduction in the amount of sunlight reaching the Earth's surface was observed. This phenomenon is popularly known as global dimming,[121] and is primarily attributed to sulfate aerosols produced by the combustion of fossil fuels with heavy sulfur concentrations like coal and bunker fuel.[44] Smaller contributions come from black carbon, organic carbon from combustion of fossil fuels and biofuels, and from anthropogenic dust.[122][43][123][124][125] Globally, aerosols have been declining since 1990 due to pollution controls, meaning that they no longer mask greenhouse gas warming as much.[126][44]

Aerosols also have indirect effects on the Earth's energy budget. Sulfate aerosols act as cloud condensation nuclei and lead to clouds that have more and smaller cloud droplets. These clouds reflect solar radiation more efficiently than clouds with fewer and larger droplets.[127] They also reduce the growth of raindrops, which makes clouds more reflective to incoming sunlight.[128] Indirect effects of aerosols are the largest uncertainty in radiative forcing.[129]

While aerosols typically limit global warming by reflecting sunlight, black carbon in soot that falls on snow or ice can contribute to global warming. Not only does this increase the absorption of sunlight, it also increases melting and sea-level rise.[130] Limiting new black carbon deposits in the Arctic could reduce global warming by 0.2 °C by 2050.[131] The effect of decreasing sulfur content of fuel oil for ships since 2020[132] is estimated to cause an additional 0.05 °C increase in global mean temperature by 2050.[133]

Solar and volcanic activity

The Fourth National Climate Assessment ("NCA4", USGCRP, 2017) includes charts illustrating that neither solar nor volcanic activity can explain the observed warming.[134][135]

As the Sun is the Earth's primary energy source, changes in incoming sunlight directly affect the climate system.[129] Solar irradiance has been measured directly by satellites,[136] and indirect measurements are available from the early 1600s onwards.[129] Since 1880, there has been no upward trend in the amount of the Sun's energy reaching the Earth, in contrast to the warming of the lower atmosphere (the troposphere).[137] The upper atmosphere (the stratosphere) would also be warming if the Sun was sending more energy to Earth, but instead, it has been cooling.[90] This is consistent with greenhouse gases preventing heat from leaving the Earth's atmosphere.[138]

Explosive volcanic eruptions can release gases, dust and ash that partially block sunlight and reduce temperatures, or they can send water vapor into the atmosphere, which adds to greenhouse gases and increases temperatures.[139] These impacts on temperature only last for several years, because both water vapor and volcanic material have low persistence in the atmosphere.[140] volcanic CO2 emissions are more persistent, but they are equivalent to less than 1% of current human-caused CO2 emissions.[141] Volcanic activity still represents the single largest natural impact (forcing) on temperature in the industrial era. Yet, like the other natural forcings, it has had negligible impacts on global temperature trends since the Industrial Revolution.[142]

Climate change feedbacks

Sea ice reflects 50% to 70% of incoming sunlight, while the ocean, being darker, reflects only 6%. As an area of sea ice melts and exposes more ocean, more heat is absorbed by the ocean, raising temperatures that melt still more ice. This is a positive feedback process.[143]

The response of the climate system to an initial forcing is modified by feedbacks: increased by "self-reinforcing" or "positive" feedbacks and reduced by "balancing" or "negative" feedbacks.[144] The main reinforcing feedbacks are the water-vapour feedback, the ice–albedo feedback, and the net effect of clouds.[145][146] The primary balancing mechanism is radiative cooling, as Earth's surface gives off more heat to space in response to rising temperature.[147] In addition to temperature feedbacks, there are feedbacks in the carbon cycle, such as the fertilising effect of CO2 on plant growth.[148]

Uncertainty over feedbacks, particularly cloud cover,[149] is the major reason why different climate models project different magnitudes of warming for a given amount of emissions.[150] As air warms, it can hold more moisture. Water vapour, as a potent greenhouse gas, holds heat in the atmosphere.[145] If cloud cover increases, more sunlight will be reflected back into space, cooling the planet. If clouds become higher and thinner, they act as an insulator, reflecting heat from below back downwards and warming the planet.[151]

Another major feedback is the reduction of snow cover and sea ice in the Arctic, which reduces the reflectivity of the Earth's surface.[152] More of the Sun's energy is now absorbed in these regions, contributing to amplification of Arctic temperature changes.[153] Arctic amplification is also thawing permafrost, which releases methane and CO2 into the atmosphere.[154] Climate change can also cause methane releases from wetlands, marine systems, and freshwater systems.[155] Overall, climate feedbacks are expected to become increasingly positive.[156]

Around half of human-caused CO2 emissions have been absorbed by land plants and by the oceans.[157] This fraction is not static and if future CO2 emissions decrease, the Earth will be able to absorb up to around 70%. If they increase substantially, it'll still absorb more carbon than now, but the overall fraction will decrease to below 40%.[158] This is because climate change increases droughts and heat waves that eventually inhibit plant growth on land, and soils will release more carbon from dead plants when they are warmer.[159][160] The rate at which oceans absorb atmospheric carbon will be lowered as they become more acidic and experience changes in thermohaline circulation and phytoplankton distribution.[161][162][71]

Modelling

Energy flows between space, the atmosphere, and Earth's surface. Most sunlight passes through the atmosphere to heat the Earth's surface, then greenhouse gases absorb most of the heat the Earth radiates in response. Adding to greenhouse gases increases this insulating effect, causing an energy imbalance that heats the planet up.

A climate model is a representation of the physical, chemical and biological processes that affect the climate system.[163] Models include natural processes like changes in the Earth's orbit, historical changes in the Sun's activity, and volcanic forcing.[164] Models are used to estimate the degree of warming future emissions will cause when accounting for the strength of climate feedbacks,[165][166]. Models also predict the circulation of the oceans, the annual cycle of the seasons, and the flows of carbon between the land surface and the atmosphere.[167]

The physical realism of models is tested by examining their ability to simulate contemporary or past climates.[168] Past models have underestimated the rate of Arctic shrinkage[169] and underestimated the rate of precipitation increase.[170] Sea level rise since 1990 was underestimated in older models, but more recent models agree well with observations.[171] The 2017 United States-published National Climate Assessment notes that "climate models may still be underestimating or missing relevant feedback processes".[172] Additionally, climate models may be unable to adequately predict short-term regional climatic shifts.[173]

A subset of climate models add societal factors to a physical climate model. These models simulate how population, economic growth, and energy use affect—and interact with—the physical climate. With this information, these models can produce scenarios of future greenhouse gas emissions. This is then used as input for physical climate models and carbon cycle models to predict how atmospheric concentrations of greenhouse gases might change.[174][175] Depending on the socioeconomic scenario and the mitigation scenario, models produce atmospheric CO2 concentrations that range widely between 380 and 1400 ppm.[176]

Impacts

The sixth IPCC Assessment Report projects changes in average soil moisture that can disrupt agriculture and ecosystems. A reduction in soil moisture by one standard deviation means that average soil moisture will approximately match the ninth driest year between 1850 and 1900 at that location.

Environmental effects

The environmental effects of climate change are broad and far-reaching, affecting oceans, ice, and weather. Changes may occur gradually or rapidly. Evidence for these effects comes from studying climate change in the past, from modelling, and from modern observations.[177] Since the 1950s, droughts and heat waves have appeared simultaneously with increasing frequency.[178] Extremely wet or dry events within the monsoon period have increased in India and East Asia.[179] Monsoonal precipitation over the Northern Hemisphere has increased since 1980.[180] The rainfall rate and intensity of hurricanes and typhoons is likely increasing,[181] and the geographic range likely expanding poleward in response to climate warming.[182] Frequency of tropical cyclones has not increased as a result of climate change.[183]

Historical sea level reconstruction and projections up to 2100 published in 2017 by the U.S. Global Change Research Program[184]

Global sea level is rising as a consequence of thermal expansion and the melting of glaciers and ice sheets. Between 1993 and 2020, the rise increased over time, averaging 3.3 ± 0.3 mm per year.[185] Over the 21st century, the IPCC projects 32–62 cm of sea level rise under a low emission scenario, 44–76 cm under an intermediate one and 65–101 cm under a very high emission scenario.[186] Marine ice sheet instability processes in Antarctica may add substantially to these values,[187] including the possibility of a 2-meter sea level rise by 2100 under high emissions.[188]

Climate change has led to decades of shrinking and thinning of the Arctic sea ice.[189] While ice-free summers are expected to be rare at 1.5 °C degrees of warming, they are set to occur once every three to ten years at a warming level of 2 °C.[190] Higher atmospheric CO2 concentrations cause more CO2 to dissolve in the oceans, which is making them more acidic.[191] Because oxygen is less soluble in warmer water,[192] its concentrations in the ocean are decreasing, and dead zones are expanding.[193]

Tipping points and long-term impacts

Different levels of global warming may cause different parts of Earth's climate system to reach tipping points that cause transitions to different states.[194][195]

Greater degrees of global warming increase the risk of passing through 'tipping points'—thresholds beyond which certain major impacts can no longer be avoided even if temperatures return to their previous state.[196][197] For instance, the Greenland ice sheet is already melting, but if global warming reaches levels between 1.7 °C and 2.3 °C, its melting will continue until it fully disappears. If the warming is later reduced to 1.5 °C or less, it will still lose a lot more ice than if the warming was never allowed to reach the threshold in the first place.[198] While the ice sheets would melt over millennia, other tipping points would occur faster and give societies less time to respond. The collapse of major ocean currents like the Atlantic meridional overturning circulation (AMOC), and irreversible damage to key ecosystems like the Amazon rainforest and coral reefs can unfold in a matter of decades.[195]

The long-term effects of climate change on oceans include further ice melt, ocean warming, sea level rise, ocean acidification and ocean deoxygenation.[199] The timescale of long term impacts are centuries to millennia due to CO2's long atmospheric lifetime.[200] When net emissions stabilise surface air temperatures will also stabilise, but oceans and ice caps will continue to absorb excess heat from the atmosphere. The result is an estimated total sea level rise of 2.3 metres per degree Celsius (4.2 ft/°F) after 2000 years.[201] Oceanic CO2 uptake is slow enough that ocean acidification will also continue for hundreds to thousands of years.[202] Deep oceans (below 2,000 metres (6,600 ft)) are also already committed to losing over 10% of their dissolved oxygen by the warming which occurred to date.[203] Further, West Antarctic ice sheet appears committed to practically irreversible melting, which would increase the sea levels by at least 3.3 m (10 ft 10 in) over approximately 2000 years.[195][204][205][206][207][208][209][210]

Nature and wildlife

Recent warming has driven many terrestrial and freshwater species poleward and towards higher altitudes.[211] Higher atmospheric CO2 levels and an extended growing season have resulted in global greening. However, heatwaves and drought have reduced ecosystem productivity in some regions. The future balance of these opposing effects is unclear.[212] Climate change has contributed to the expansion of drier climate zones, such as the expansion of deserts in the subtropics.[213] The size and speed of global warming is making abrupt changes in ecosystems more likely.[214] Overall, it is expected that climate change will result in the extinction of many species.[215]

The oceans have heated more slowly than the land, but plants and animals in the ocean have migrated towards the colder poles faster than species on land.[216] Just as on land, heat waves in the ocean occur more frequently due to climate change, harming a wide range of organisms such as corals, kelp, and seabirds.[217] Ocean acidification makes it harder for marine calcifying organisms such as mussels, barnacles and corals to produce shells and skeletons; and heatwaves have bleached coral reefs.[218] Harmful algal blooms enhanced by climate change and eutrophication lower oxygen levels, disrupt food webs and cause great loss of marine life.[219] Coastal ecosystems are under particular stress. Almost half of global wetlands have disappeared due to climate change and other human impacts.[220]

Climate change impacts on the environment

Humans

Extreme weather will be progressively more common as the Earth warms.[225]

The effects of climate change are impacting humans everywhere in the world.[226] Impacts can be observed on all continents and ocean regions,[227] with low-latitude, less developed areas facing the greatest risk.[228] Continued warming has potentially "severe, pervasive and irreversible impacts" for people and ecosystems.[229] The risks are unevenly distributed, but are generally greater for disadvantaged people in developing and developed countries.[230]

Food and health

The World Health Organization (WHO) calls climate change the greatest threat to global health in the 21st century.[231] Extreme weather leads to injury and loss of life.[232] Various infectious diseases are more easily transmitted in a warmer climate, such as dengue fever and malaria.[233] Crop failures can lead to food shortages and malnutrition, particularly effecting children.[234] Both children and older people are vulnerable to extreme heat.[235] The WHO has estimated that between 2030 and 2050, climate change would cause around 250,000 additional deaths per year. They assessed deaths from heat exposure in elderly people, increases in diarrhea, malaria, dengue, coastal flooding, and childhood malnutrition.[236] By 2100, 50% to 75% of the global population may face climate conditions that are life-threatening due to combined effects of extreme heat and humidity.[237]

Climate change is affecting food security. It has caused reduction in global yields of maize, wheat, and soybeans between 1981 and 2010.[238] Future warming could further reduce global yields of major crops.[239] Crop production will probably be negatively affected in low-latitude countries, while effects at northern latitudes may be positive or negative.[240] Up to an additional 183 million people worldwide, particularly those with lower incomes, are at risk of hunger as a consequence of these impacts.[241] Climate change also impacts fish populations. Globally, less will be available to be fished.[242] Regions dependent on glacier water, regions that are already dry, and small islands have a higher risk of water stress due to climate change.[243]

Livelihoods and inequality

Economic damages due to climate change may be severe and there is a chance of disastrous consequences.[244] Severe impacts are expected in South-East Asia and sub-Saharan Africa, where most of the local inhabitants are dependent upon natural and agricultural resources.[245][246] Heat stress can prevent outdoor labourers from working. If warming reaches 4 °C then labour capacity in those regions could be reduced by 30 to 50%.[247] The World Bank estimates that between 2016 and 2030, climate change could drive over 120 million people into extreme poverty without adaptation.[248]

Inequalities based on wealth and social status have worsened due to climate change.[249] Major difficulties in mitigating, adapting, and recovering to climate shocks are faced by marginalised people who have less control over resources.[250][245] Indigenous people, who are subsistent on their land and ecosystems, will face endangerment to their wellness and lifestyles due to climate change.[251] An expert elicitation concluded that the role of climate change in armed conflict has been small compared to factors such as socio-economic inequality and state capabilities.[252]

While women are not inherently more at risk from climate change and shocks, limits on women's resources and discriminatory gender norms constrain their adaptive capacity and resilience.[253] For example, women's work burdens, including hours worked in agriculture, tend to decline less than men's during climate shocks such as heat stress.[253]

Climate migration

Low-lying islands and coastal communities are threatened by sea level rise, which makes urban flooding more common. Sometimes, land is permanently lost to the sea.[254] This could lead to statelessness for people in island nations, such as the Maldives and Tuvalu.[255] In some regions, the rise in temperature and humidity may be too severe for humans to adapt to.[256] With worst-case climate change, models project that almost one-third of humanity might live in Sahara-like uninhabitable and extremely hot climates.[257]

These factors can drive climate or environmental migration, within and between countries.[11] More people are expected to be displaced because of sea level rise, extreme weather and conflict from increased competition over natural resources. Climate change may also increase vulnerability, leading to "trapped populations" who are not able to move due to a lack of resources.[258]

Climate change impacts on people

Reducing and recapturing emissions

Global greenhouse gas emission scenarios, based on policies and pledges as of November 2021

Climate change can be mitigated by reducing the rate at which greenhouse gases are emitted into the atmosphere, and by increasing the rate at which carbon dioxide is removed from the atmosphere.[264] In order to limit global warming to less than 1.5 °C global greenhouse gas emissions needs to be net-zero by 2050, or by 2070 with a 2 °C target.[83] This requires far-reaching, systemic changes on an unprecedented scale in energy, land, cities, transport, buildings, and industry.[265]

The United Nations Environment Programme estimates that countries need to triple their pledges under the Paris Agreement within the next decade to limit global warming to 2 °C. An even greater level of reduction is required to meet the 1.5 °C goal.[266] With pledges made under the Paris Agreement as of October 2021, global warming would still have a 66% chance of reaching about 2.7 °C (range: 2.2–3.2 °C) by the end of the century.[20] Globally, limiting warming to 2 °C may result in higher economic benefits than economic costs.[267]

Although there is no single pathway to limit global warming to 1.5 or 2 °C,[268] most scenarios and strategies see a major increase in the use of renewable energy in combination with increased energy efficiency measures to generate the needed greenhouse gas reductions.[269] To reduce pressures on ecosystems and enhance their carbon sequestration capabilities, changes would also be necessary in agriculture and forestry,[270] such as preventing deforestation and restoring natural ecosystems by reforestation.[271]

Other approaches to mitigating climate change have a higher level of risk. Scenarios that limit global warming to 1.5 °C typically project the large-scale use of carbon dioxide removal methods over the 21st century.[272] There are concerns, though, about over-reliance on these technologies, and environmental impacts.[273] Solar radiation modification (SRM) is also a possible supplement to deep reductions in emissions. However, SRM raises significant ethical and legal concerns, and the risks are imperfectly understood.[274]

Clean energy

Coal, oil, and natural gas remain the primary global energy sources even as renewables have begun rapidly increasing.[275]
Wind and solar power, Germany

Renewable energy is key to limiting climate change.[276] For decades, fossil fuels have accounted for roughly 80% of the world's energy use.[277] The remaining share has been split between nuclear power and renewables (including hydropower, bioenergy, wind and solar power and geothermal energy).[278] Fossil fuel use is expected to peak in absolute terms prior to 2030 and then to decline, with coal use experiencing the sharpest reductions.[279] Renewables represented 75% of all new electricity generation installed in 2019, nearly all solar and wind.[280] Other forms of clean energy, such as nuclear and hydropower, currently have a larger share of the energy supply. However, their future growth forecasts appear limited in comparison.[281]

While solar panels and onshore wind are now among the cheapest forms of adding new power generation capacity in many locations,[282] green energy policies are needed to achieve a rapid transition from fossil fuels to renewables.[283] To achieve carbon neutrality by 2050, renewable energy would become the dominant form of electricity generation, rising to 85% or more by 2050 in some scenarios. Investment in coal would be eliminated and coal use nearly phased out by 2050.[284][285]

Electricity generated from renewable sources would also need to become the main energy source for heating and transport.[286] Transport can switch away from internal combustion engine vehicles and towards electric vehicles, public transit, and active transport (cycling and walking).[287][288] For shipping and flying, low-carbon fuels would reduce emissions.[287] Heating could be increasingly decarbonised with technologies like heat pumps.[289]

There are obstacles to the continued rapid growth of clean energy, including renewables. For wind and solar, there are environmental and land use concerns for new projects.[290] Wind and solar also produce energy intermittently and with seasonal variability. Traditionally, hydro dams with reservoirs and conventional power plants have been used when variable energy production is low. Going forward, battery storage can be expanded, energy demand and supply can be matched, and long-distance transmission can smooth variability of renewable outputs.[276] Bioenergy is often not carbon-neutral and may have negative consequences for food security.[291] The growth of nuclear power is constrained by controversy around radioactive waste, nuclear weapon proliferation, and accidents.[292][293] Hydropower growth is limited by the fact that the best sites have been developed, and new projects are confronting increased social and environmental concerns.[294]

Low-carbon energy improves human health by minimising climate change as well as reducing air pollution deaths,[295] which were estimated at 7 million annually in 2016.[296] Meeting the Paris Agreement goals that limit warming to a 2 °C increase could save about a million of those lives per year by 2050, whereas limiting global warming to 1.5 °C could save millions and simultaneously increase energy security and reduce poverty.[297] Improving air quality also has economic benefits which may be larger than mitigation costs.[298]

Energy conservation

Reducing energy demand is another major aspect of reducing emissions.[299] If less energy is needed, there is more flexibility for clean energy development. It also makes it easier to manage the electricity grid, and minimises carbon-intensive infrastructure development.[300] Major increases in energy efficiency investment will be required to achieve climate goals, comparable to the level of investment in renewable energy.[301] Several COVID-19 related changes in energy use patterns, energy efficiency investments, and funding have made forecasts for this decade more difficult and uncertain.[302]

Strategies to reduce energy demand vary by sector. In the transport sector, passengers and freight can switch to more efficient travel modes, such as buses and trains, or use electric vehicles.[303] Industrial strategies to reduce energy demand include improving heating systems and motors, designing less energy-intensive products, and increasing product lifetimes.[304] In the building sector the focus is on better design of new buildings, and higher levels of energy efficiency in retrofitting.[305] The use of technologies like heat pumps can also increase building energy efficiency.[306]

Agriculture and industry

Taking into account direct and indirect emissions, industry is the sector with the highest share of global emissions. Data as of 2019 from the IPCC.

Agriculture and forestry face a triple challenge of limiting greenhouse gas emissions, preventing the further conversion of forests to agricultural land, and meeting increases in world food demand.[307] A set of actions could reduce agriculture and forestry-based emissions by two thirds from 2010 levels. These include reducing growth in demand for food and other agricultural products, increasing land productivity, protecting and restoring forests, and reducing greenhouse gas emissions from agricultural production.[308]

On the demand side, a key component of reducing emissions is shifting people towards plant-based diets.[309] Eliminating the production of livestock for meat and dairy would eliminate about 3/4ths of all emissions from agriculture and other land use.[310] Livestock also occupy 37% of ice-free land area on Earth and consume feed from the 12% of land area used for crops, driving deforestation and land degradation.[311]

Steel and cement production are responsible for about 13% of industrial CO2 emissions. In these industries, carbon-intensive materials such as coke and lime play an integral role in the production, so that reducing CO2 emissions requires research into alternative chemistries.[312]

Carbon sequestration

Most CO2 emissions have been absorbed by carbon sinks, including plant growth, soil uptake, and ocean uptake (2020 Global Carbon Budget).

Natural carbon sinks can be enhanced to sequester significantly larger amounts of CO2 beyond naturally occurring levels.[313] Reforestation and afforestation (planting forests where there were none before) are among the most mature sequestration techniques, although the latter raises food security concerns.[314] Farmers can promote sequestration of carbon in soils through practices such as use of winter cover crops, reducing the intensity and frequency of tillage, and using compost and manure as soil amendments.[315] Forest and landscape restoration yields many benefits for the climate, including greenhouse gas emissions sequestration and reduction.[117] Restoration/recreation of coastal wetlands, prairie plots and seagrass meadows increases the uptake of carbon into organic matter.[316][317] When carbon is sequestered in soils and in organic matter such as trees, there is a risk of the carbon being re-released into the atmosphere later through changes in land use, fire, or other changes in ecosystems.[318]

Where energy production or CO2-intensive heavy industries continue to produce waste CO2, the gas can be captured and stored instead of released to the atmosphere. Although its current use is limited in scale and expensive,[319] carbon capture and storage (CCS) may be able to play a significant role in limiting CO2 emissions by mid-century.[320] This technique, in combination with bioenergy (BECCS) can result in net negative emissions as CO2 is drawn from the atmosphere.[321] It remains highly uncertain whether carbon dioxide removal techniques will be able to play a large role in limiting warming to 1.5 °C. Policy decisions that rely on carbon dioxide removal increase the risk of global warming rising beyond international goals.[322]

Adaptation

Adaptation is "the process of adjustment to current or expected changes in climate and its effects".[323]: 5  Without additional mitigation, adaptation cannot avert the risk of "severe, widespread and irreversible" impacts.[324] More severe climate change requires more transformative adaptation, which can be prohibitively expensive.[325] The capacity and potential for humans to adapt is unevenly distributed across different regions and populations, and developing countries generally have less.[326] The first two decades of the 21st century saw an increase in adaptive capacity in most low- and middle-income countries with improved access to basic sanitation and electricity, but progress is slow. Many countries have implemented adaptation policies. However, there is a considerable gap between necessary and available finance.[327]

Adaptation to sea level rise consists of avoiding at-risk areas, learning to live with increased flooding, and building flood controls. If that fails, managed retreat may be needed.[328] There are economic barriers for tackling dangerous heat impact. Avoiding strenuous work or having air conditioning is not possible for everybody.[329] In agriculture, adaptation options include a switch to more sustainable diets, diversification, erosion control, and genetic improvements for increased tolerance to a changing climate.[330] Insurance allows for risk-sharing, but is often difficult to get for people on lower incomes.[331] Education, migration and early warning systems can reduce climate vulnerability.[332] Planting mangroves or encouraging other coastal vegetation can buffer storms.[333][334]

Ecosystems adapt to climate change, a process that can be supported by human intervention. By increasing connectivity between ecosystems, species can migrate to more favourable climate conditions. Species can also be introduced to areas acquiring a favorable climate. Protection and restoration of natural and semi-natural areas helps build resilience, making it easier for ecosystems to adapt. Many of the actions that promote adaptation in ecosystems, also help humans adapt via ecosystem-based adaptation. For instance, restoration of natural fire regimes makes catastrophic fires less likely, and reduces human exposure. Giving rivers more space allows for more water storage in the natural system, reducing flood risk. Restored forest acts as a carbon sink, but planting trees in unsuitable regions can exacerbate climate impacts.[335]

There are synergies but also trade-offs between adaptation and mitigation.[336] An example for synergy is increased food productivity, which has large benefits for both adaptation and mitigation.[337] An example of a trade-off is that increased use of air conditioning allows people to better cope with heat, but increases energy demand. Another trade-off example is that more compact urban development may reduce emissions from transport and construction, but may also increase the urban heat island effect, exposing people to heat-related health risks.[338]

Examples of adaptation methods

Policies and politics

The Climate Change Performance Index ranks countries by greenhouse gas emissions (40% of score), renewable energy (20%), energy use (20%), and climate policy (20%).
  High
  Medium
  Low
  Very low

Countries that are most vulnerable to climate change have typically been responsible for a small share of global emissions. This raises questions about justice and fairness.[339] Limiting global warming makes it much easier to achieve the UN's Sustainable Development Goals, such as eradicating poverty and reducing inequalities. The connection is recognised in Sustainable Development Goal 13 which is to "take urgent action to combat climate change and its impacts".[340] The goals on food, clean water and ecosystem protection have synergies with climate mitigation.[341]

The geopolitics of climate change is complex. It has often been framed as a free-rider problem, in which all countries benefit from mitigation done by other countries, but individual countries would lose from switching to a low-carbon economy themselves. Sometimes mitigation also has localised benefits though. For instance, the benefits of a coal phase-out to public health and local environments exceed the costs in almost all regions.[342] Furthermore, net importers of fossil fuels win economically from switching to clean energy, causing net exporters to face stranded assets: fossil fuels they cannot sell.[343]

Policy options

A wide range of policies, regulations, and laws are being used to reduce emissions. As of 2019, carbon pricing covers about 20% of global greenhouse gas emissions.[344] Carbon can be priced with carbon taxes and emissions trading systems.[345] Direct global fossil fuel subsidies reached $319 billion in 2017, and $5.2 trillion when indirect costs such as air pollution are priced in.[346] Ending these can cause a 28% reduction in global carbon emissions and a 46% reduction in air pollution deaths.[347] Money saved on fossil subsidies could be used to support the transition to clean energy instead.[348] More direct methods to reduce greenhouse gases include vehicle efficiency standards, renewable fuel standards, and air pollution regulations on heavy industry.[349] Several countries require utilities to increase the share of renewables in power production.[350]

Climate justice

Policy designed through the lens of climate justice tries to address human rights issues and social inequality. According to proponents of climate justice, the costs of climate adaptation should be paid by those most responsible for climate change, while the beneficiaries of payments should be those suffering impacts. One way this can be addressed in practice is to have wealthy nations pay poorer countries to adapt.[351]

Oxfam found that in 2023 the wealthiest 10% of people were responsible for 50% of global emissions, while the bottom 50% were responsible for just 8%.[352] Production of emissions is another way to look at responsibility: under that approach, the top 21 fossil fuel companies would owe cumulative climate reparations of $5.4 trillion over the period 2025–2050.[353] To achieve a just transition, people working in the fossil fuel sector would also need other jobs, and their communities would need investments.[354]

International climate agreements

Since 2000, rising CO2 emissions in China and the rest of world have surpassed the output of the United States and Europe.[355]
Per person, the United States generates CO2 at a far faster rate than other primary regions.[355]

Nearly all countries in the world are parties to the 1994 United Nations Framework Convention on Climate Change (UNFCCC).[356] The goal of the UNFCCC is to prevent dangerous human interference with the climate system.[357] As stated in the convention, this requires that greenhouse gas concentrations are stabilised in the atmosphere at a level where ecosystems can adapt naturally to climate change, food production is not threatened, and economic development can be sustained.[358] The UNFCCC does not itself restrict emissions but rather provides a framework for protocols that do. Global emissions have risen since the UNFCCC was signed.[359] Its yearly conferences are the stage of global negotiations.[360]

The 1997 Kyoto Protocol extended the UNFCCC and included legally binding commitments for most developed countries to limit their emissions.[361] During the negotiations, the G77 (representing developing countries) pushed for a mandate requiring developed countries to "[take] the lead" in reducing their emissions,[362] since developed countries contributed most to the accumulation of greenhouse gases in the atmosphere. Per-capita emissions were also still relatively low in developing countries and developing countries would need to emit more to meet their development needs.[363]

The 2009 Copenhagen Accord has been widely portrayed as disappointing because of its low goals, and was rejected by poorer nations including the G77.[364] Associated parties aimed to limit the global temperature rise to below 2 °C.[365] The Accord set the goal of sending $100 billion per year to developing countries for mitigation and adaptation by 2020, and proposed the founding of the Green Climate Fund.[366] As of 2020, only 83.3 billion were delivered. Only in 2023 the target is expected to be achieved.[367]

In 2015 all UN countries negotiated the Paris Agreement, which aims to keep global warming well below 2.0 °C and contains an aspirational goal of keeping warming under 1.5 °C.[368] The agreement replaced the Kyoto Protocol. Unlike Kyoto, no binding emission targets were set in the Paris Agreement. Instead, a set of procedures was made binding. Countries have to regularly set ever more ambitious goals and reevaluate these goals every five years.[369] The Paris Agreement restated that developing countries must be financially supported.[370] As of October 2021, 194 states and the European Union have signed the treaty and 191 states and the EU have ratified or acceded to the agreement.[371]

The 1987 Montreal Protocol, an international agreement to stop emitting ozone-depleting gases, may have been more effective at curbing greenhouse gas emissions than the Kyoto Protocol specifically designed to do so.[372] The 2016 Kigali Amendment to the Montreal Protocol aims to reduce the emissions of hydrofluorocarbons, a group of powerful greenhouse gases which served as a replacement for banned ozone-depleting gases. This made the Montreal Protocol a stronger agreement against climate change.[373]

National responses

In 2019, the United Kingdom parliament became the first national government to declare a climate emergency.[374] Other countries and jurisdictions followed suit.[375] That same year, the European Parliament declared a "climate and environmental emergency".[376] The European Commission presented its European Green Deal with the goal of making the EU carbon-neutral by 2050.[377] In 2021, the European Commission released its "Fit for 55" legislation package, which contains guidelines for the car industry; all new cars on the European market must be zero-emission vehicles from 2035.[378]

Major countries in Asia have made similar pledges: South Korea and Japan have committed to become carbon-neutral by 2050, and China by 2060.[379] While India has strong incentives for renewables, it also plans a significant expansion of coal in the country.[380] Vietnam is among very few coal-dependent fast developing countries that pledged to phase out unabated coal power by the 2040s or as soon as possible thereafter.[381]

As of 2021, based on information from 48 national climate plans, which represent 40% of the parties to the Paris Agreement, estimated total greenhouse gas emissions will be 0.5% lower compared to 2010 levels, below the 45% or 25% reduction goals to limit global warming to 1.5 °C or 2 °C, respectively.[382]

Society

Denial and misinformation

Data has been cherry picked from short periods to falsely assert that global temperatures are not rising. Blue trendlines show short periods that mask longer-term warming trends (red trendlines). Blue rectangle with blue dots shows the so-called global warming hiatus.[383]

Public debate about climate change has been strongly affected by climate change denial and misinformation, which originated in the United States and has since spread to other countries, particularly Canada and Australia. Climate change denial has originated from fossil fuel companies, industry groups, conservative think tanks, and contrarian scientists.[384] Like the tobacco industry, the main strategy of these groups has been to manufacture doubt about climate-change related scientific data and results.[385] People who hold unwarranted doubt about climate change are called climate change "skeptics", although "contrarians" or "deniers" are more appropriate terms.[386]

There are different variants of climate denial: some deny that warming takes place at all, some acknowledge warming but attribute it to natural influences, and some minimise the negative impacts of climate change.[387] Manufacturing uncertainty about the science later developed into a manufactured controversy: creating the belief that there is significant uncertainty about climate change within the scientific community in order to delay policy changes.[388] Strategies to promote these ideas include criticism of scientific institutions,[389] and questioning the motives of individual scientists.[387] An echo chamber of climate-denying blogs and media has further fomented misunderstanding of climate change.[390]

Public awareness and opinion

The public substantially underestimates the degree of scientific consensus that humans are causing climate change.[391] Studies from 2019 to 2021[392][4][393] found scientific consensus to range from 98.7 to 100%.

Climate change came to international public attention in the late 1980s.[394] Due to media coverage in the early 1990s, people often confused climate change with other environmental issues like ozone depletion.[395] In popular culture, the climate fiction movie The Day After Tomorrow (2004) and the Al Gore documentary An Inconvenient Truth (2006) focused on climate change.[394]

Significant regional, gender, age and political differences exist in both public concern for, and understanding of, climate change. More highly educated people, and in some countries, women and younger people, were more likely to see climate change as a serious threat.[396] Partisan gaps also exist in many countries,[397] and countries with high CO2 emissions tend to be less concerned.[398] Views on causes of climate change vary widely between countries.[399] Concern has increased over time,[397] to the point where in 2021 a majority of citizens in many countries express a high level of worry about climate change, or view it as a global emergency.[400] Higher levels of worry are associated with stronger public support for policies that address climate change.[401]

Climate movement

Climate protests demand that political leaders take action to prevent climate change. They can take the form of public demonstrations, fossil fuel divestment, lawsuits and other activities.[402] Prominent demonstrations include the School Strike for Climate. In this initiative, young people across the globe have been protesting since 2018 by skipping school on Fridays, inspired by Swedish teenager Greta Thunberg.[403] Mass civil disobedience actions by groups like Extinction Rebellion have protested by disrupting roads and public transport.[404]

Litigation is increasingly used as a tool to strengthen climate action from public institutions and companies. Activists also initiate lawsuits which target governments and demand that they take ambitious action or enforce existing laws on climate change.[405] Lawsuits against fossil-fuel companies generally seek compensation for loss and damage.[406]

History

Early discoveries

This 1912 article succinctly describes the greenhouse effect, how burning coal creates carbon dioxide to cause global warming and climate change.[407]

Scientists in the 19th century such as Alexander von Humboldt began to foresee the effects of climate change.[408][409][410][411] In the 1820s, Joseph Fourier proposed the greenhouse effect to explain why Earth's temperature was higher than the Sun's energy alone could explain. Earth's atmosphere is transparent to sunlight, so sunlight reaches the surface where it is converted to heat. However, the atmosphere is not transparent to heat radiating from the surface, and captures some of that heat, which in turn warms the planet.[412]

In 1856 Eunice Newton Foote demonstrated that the warming effect of the Sun is greater for air with water vapour than for dry air, and that the effect is even greater with carbon dioxide (CO2). She concluded that "An atmosphere of that gas would give to our earth a high temperature..."[413][414]

Studying what would become known as the greenhouse effect, Tyndall's pre-1861 ratio spectrophotometer measured how much various gases in a tube absorb and emit infrared radiation—which humans experience as heat.

Starting in 1859,[415] John Tyndall established that nitrogen and oxygen—together totaling 99% of dry air—are transparent to radiated heat. However, water vapour and gases such as methane and carbon dioxide absorb radiated heat and re-radiate that heat into the atmosphere. Tyndall proposed that changes in the concentrations of these gases may have caused climatic changes in the past, including ice ages.[416]

Svante Arrhenius noted that water vapour in air continuously varied, but the CO2 concentration in air was influenced by long-term geological processes. Warming from increased CO2 levels would increase the amount of water vapour, amplifying warming in a positive feedback loop. In 1896, he published the first climate model of its kind, projecting that halving CO2 levels could have produced a drop in temperature initiating an ice age. Arrhenius calculated the temperature increase expected from doubling CO2 to be around 5–6 °C.[417] Other scientists were initially skeptical and believed that the greenhouse effect was saturated so that adding more CO2 would make no difference, and that the climate would be self-regulating.[418] Beginning in 1938, Guy Stewart Callendar published evidence that climate was warming and CO2 levels were rising,[419] but his calculations met the same objections.[418]

Development of a scientific consensus

Scientific consensus on causation: Academic studies of scientific agreement on human-caused global warming among climate experts (2010–2015) reflect that the level of consensus correlates with expertise in climate science.[420] A 2019 study found scientific consensus to be at 100%,[421] and a 2021 study concluded that consensus exceeded 99%.[422] Another 2021 study found that 98.7% of climate experts indicated that the Earth is getting warmer mostly because of human activity.[423]

In the 1950s, Gilbert Plass created a detailed computer model that included different atmospheric layers and the infrared spectrum. This model predicted that increasing CO2 levels would cause warming. Around the same time, Hans Suess found evidence that CO2 levels had been rising, and Roger Revelle showed that the oceans would not absorb the increase. The two scientists subsequently helped Charles Keeling to begin a record of continued increase, which has been termed the "Keeling Curve".[418] Scientists alerted the public,[424] and the dangers were highlighted at James Hansen's 1988 Congressional testimony.[29] The Intergovernmental Panel on Climate Change (IPCC), set up in 1988 to provide formal advice to the world's governments, spurred interdisciplinary research.[425] As part of the IPCC reports, scientists assess the scientific discussion that takes place in peer-reviewed journal articles.[426]

There is a near-complete scientific consensus that the climate is warming and that this is caused by human activities. As of 2019, agreement in recent literature reached over 99%.[421][422] No scientific body of national or international standing disagrees with this view.[427] Consensus has further developed that some form of action should be taken to protect people against the impacts of climate change. National science academies have called on world leaders to cut global emissions.[428] The 2021 IPCC Assessment Report stated that it is "unequivocal" that climate change is caused by humans.[422]

See also

References

  1. ^ "GISS Surface Temperature Analysis (v4)". NASA. Retrieved 12 January 2024.
  2. ^ IPCC AR6 WG1 2021, SPM-7
  3. ^ IPCC SR15 Ch1 2018, p. 54: Since 1970 the global average temperature has been rising at a rate of 1.7 °C per century, compared to a long-term decline over the past 7,000 years at a baseline rate of 0.01 °C per century (NOAA, 2016; Marcott et al., 2013). These global-level rates of human-driven change far exceed the rates of change driven by geophysical or biosphere forces that have altered the Earth System trajectory in the past (e.g., Summerhayes, 2015; Foster et al., 2017); even abrupt geophysical events do not approach current rates of human-driven change.
  4. ^ a b Lynas, Mark; Houlton, Benjamin Z.; Perry, Simon (19 October 2021). "Greater than 99% consensus on human caused climate change in the peer-reviewed scientific literature". Environmental Research Letters. 16 (11): 114005. Bibcode:2021ERL....16k4005L. doi:10.1088/1748-9326/ac2966. S2CID 239032360.
  5. ^ a b Our World in Data, 18 September 2020
  6. ^ IPCC SRCCL 2019, p. 7: Since the pre-industrial period, the land surface air temperature has risen nearly twice as much as the global average temperature (high confidence). Climate change... contributed to desertification and land degradation in many regions (high confidence).; IPCC SRCCL 2019, p. 45: Climate change is playing an increasing role in determining wildfire regimes alongside human activity (medium confidence), with future climate variability expected to enhance the risk and severity of wildfires in many biomes such as tropical rainforests (high confidence).
  7. ^ IPCC SROCC 2019, p. 16: Over the last decades, global warming has led to widespread shrinking of the cryosphere, with mass loss from ice sheets and glaciers (very high confidence), reductions in snow cover (high confidence) and Arctic sea ice extent and thickness (very high confidence), and increased permafrost temperature (very high confidence).
  8. ^ IPCC AR6 WG1 Ch11 2021, p. 1517
  9. ^ EPA (19 January 2017). "Climate Impacts on Ecosystems". Archived from the original on 27 January 2018. Retrieved 5 February 2019. Mountain and arctic ecosystems and species are particularly sensitive to climate change... As ocean temperatures warm and the acidity of the ocean increases, bleaching and coral die-offs are likely to become more frequent.
  10. ^ IPCC SR15 Ch1 2018, p. 64: Sustained net zero anthropogenic emissions of CO2 and declining net anthropogenic non-CO2 radiative forcing over a multi-decade period would halt anthropogenic global warming over that period, although it would not halt sea level rise or many other aspects of climate system adjustment.
  11. ^ a b Cattaneo et al. 2019; IPCC AR6 WG2 2022, pp. 15, 53
  12. ^ IPCC AR5 SYR 2014, pp. 13–16; WHO, Nov 2015: "Climate change is the greatest threat to global health in the 21st century. Health professionals have a duty of care to current and future generations. You are on the front line in protecting people from climate impacts – from more heat-waves and other extreme weather events; from outbreaks of infectious diseases such as malaria, dengue and cholera; from the effects of malnutrition; as well as treating people who are affected by cancer, respiratory, cardiovascular and other non-communicable diseases caused by environmental pollution."
  13. ^ IPCC AR6 WG2 2022, p. 19
  14. ^ IPCC AR6 WG2 2022, pp. 21–26, 2504; IPCC AR6 SYR SPM 2023, pp. 8–9: "Effectiveness15 of adaptation in reducing climate risks16 is documented for specific contexts, sectors and regions (high confidence)...Soft limits to adaptation are currently being experienced by small-scale farmers and households along some low-lying coastal areas (medium confidence) resulting from financial, governance, institutional and policy constraints (high confidence). Some tropical, coastal, polar and mountain ecosystems have reached hard adaptation limits (high confidence). Adaptation does not prevent all losses and damages, even with effective adaptation and before reaching soft and hard limits (high confidence)."
  15. ^ Tietjen, Bethany (2 November 2022). "Loss and damage: Who is responsible when climate change harms the world's poorest countries?". The Conversation. Retrieved 30 August 2023.
  16. ^ "Climate Change 2022: Impacts, Adaptation and Vulnerability". IPCC. 27 February 2022. Retrieved 30 August 2023.
  17. ^ Ivanova, Irina (2 June 2022). "California is rationing water amid its worst drought in 1,200 years". CBS News.
  18. ^ Poyntin, Mark; Rivault, Erwan (10 January 2024). "2023 confirmed as world's hottest year on record". BBC. Retrieved 13 January 2024.
  19. ^ IPCC AR6 WG1 Technical Summary 2021, p. 71
  20. ^ a b c United Nations Environment Programme 2021, p. 36: "A continuation of the effort implied by the latest unconditional NDCs and announced pledges is at present estimated to result in warming of about 2.7 °C (range: 2.2–3.2 °C) with a 66 per cent chance."
  21. ^ IPCC SR15 Ch2 2018, pp. 95–96: In model pathways with no or limited overshoot of 1.5 °C, global net anthropogenic CO2 emissions decline by about 45% from 2010 levels by 2030 (40–60% interquartile range), reaching net zero around 2050 (2045–2055 interquartile range); IPCC SR15 2018, p. 17, SPM C.3:All pathways that limit global warming to 1.5 °C with limited or no overshoot project the use of carbon dioxide removal (CDR) on the order of 100–1000 GtCO2 over the 21st century. CDR would be used to compensate for residual emissions and, in most cases, achieve net negative emissions to return global warming to 1.5 °C following a peak (high confidence). CDR deployment of several hundreds of GtCO2 is subject to multiple feasibility and sustainability constraints (high confidence).; Rogelj et al. 2015; Hilaire et al. 2019
  22. ^ United Nations Environment Programme 2019, p. xxiii, Table ES.3; Teske, ed. 2019, p. xxvii, Fig.5.
  23. ^ United Nations Environment Programme 2019, Table ES.3 & p. 49; NREL 2017, pp. vi, 12
  24. ^ a b IPCC SRCCL Summary for Policymakers 2019, p. 18
  25. ^ a b NASA, 5 December 2008.
  26. ^ NASA, 7 July 2020; Shaftel 2016: "'Climate change' and 'global warming' are often used interchangeably but have distinct meanings. ... Global warming refers to the upward temperature trend across the entire Earth since the early 20th century ... Climate change refers to a broad range of global phenomena ...[which] include the increased temperature trends described by global warming."; Associated Press, 22 September 2015: "The terms global warming and climate change can be used interchangeably. Climate change is more accurate scientifically to describe the various effects of greenhouse gases on the world because it includes extreme weather, storms and changes in rainfall patterns, ocean acidification and sea level.".
  27. ^ IPCC AR5 SYR Glossary 2014, p. 120: "Climate change refers to a change in the state of the climate that can be identified (e.g., by using statistical tests) by changes in the mean and/or the variability of its properties and that persists for an extended period, typically decades or longer. Climate change may be due to natural internal processes or external forcings such as modulations of the solar cycles, volcanic eruptions and persistent anthropogenic changes in the composition of the atmosphere or in land use."
  28. ^ Broeker, Wallace S. (8 August 1975). "Climatic Change: Are We on the Brink of a Pronounced Global Warming?". Science. 189 (4201): 460–463. Bibcode:1975Sci...189..460B. doi:10.1126/science.189.4201.460. JSTOR 1740491. PMID 17781884. S2CID 16702835.
  29. ^ a b Weart "The Public and Climate Change: The Summer of 1988", "News reporters gave only a little attention ...".
  30. ^ Joo et al. 2015.
  31. ^ Hodder & Martin 2009; BBC Science Focus Magazine, 3 February 2020
  32. ^ Neukom et al. 2019b.
  33. ^ "Global Annual Mean Surface Air Temperature Change". NASA. Retrieved 23 February 2020.
  34. ^ IPCC AR5 WG1 Ch5 2013, pp. 389, 399–400: "The PETM [around 55.5–55.3 million years ago] was marked by ... global warming of 4 °C to 7 °C ... Deglacial global warming occurred in two main steps from 17.5 to 14.5 ka [thousand years ago] and 13.0 to 10.0 ka."
  35. ^ Michon, Scott. "What's the hottest the Earth's ever been?". SMITHSONIAN INSTITUTION. Retrieved 6 August 2023.
  36. ^ Michon, Scott. "What's the coldest the Earth's ever been?". SMITHSONIAN INSTITUTION. Retrieved 6 August 2023.
  37. ^ IPCC AR5 WG1 Ch5 2013, p. 386; Neukom et al. 2019a
  38. ^ IPCC SR15 Ch1 2018, p. 57: This report adopts the 51-year reference period, 1850–1900 inclusive, assessed as an approximation of pre-industrial levels in AR5 ... Temperatures rose by 0.0 °C–0.2 °C from 1720–1800 to 1850–1900; Hawkins et al. 2017, p. 1844
  39. ^ "Mean Monthly Temperature Records Across the Globe / Timeseries of Global Land and Ocean Areas at Record Levels for September from 1951-2023". NCEI.NOAA.gov. National Centers for Environmental Information (NCEI) of the National Oceanic and Atmospheric Administration (NOAA). September 2023. Archived from the original on 14 October 2023. (change "202309" in URL to see years other than 2023, and months other than 09=September)
  40. ^ Top 700 meters: Lindsey, Rebecca; Dahlman, Luann (6 September 2023). "Climate Change: Ocean Heat Content". climate.gov. National Oceanic and Atmospheric Administration (NOAA). Archived from the original on 29 October 2023.Top 2000 meters: "Ocean Warming / Latest Measurement: December 2022 / 345 (± 2) zettajoules since 1955". NASA.gov. National Aeronautics and Space Administration. Archived from the original on 20 October 2023.
  41. ^ IPCC AR5 WG1 Summary for Policymakers 2013, pp. 4–5: "Global-scale observations from the instrumental era began in the mid-19th century for temperature and other variables ... the period 1880 to 2012 ... multiple independently produced datasets exist."
  42. ^ Mooney, Chris; Osaka, Shannon (26 December 2023). "Is climate change speeding up? Here's what the science says". The Washington Post. Retrieved 18 January 2024.
  43. ^ a b "Global 'Sunscreen' Has Likely Thinned, Report NASA Scientists". NASA. 15 March 2007.
  44. ^ a b c Quaas, Johannes; Jia, Hailing; Smith, Chris; Albright, Anna Lea; Aas, Wenche; Bellouin, Nicolas; Boucher, Olivier; Doutriaux-Boucher, Marie; Forster, Piers M.; Grosvenor, Daniel; Jenkins, Stuart; Klimont, Zbigniew; Loeb, Norman G.; Ma, Xiaoyan; Naik, Vaishali; Paulot, Fabien; Stier, Philip; Wild, Martin; Myhre, Gunnar; Schulz, Michael (21 September 2022). "Robust evidence for reversal of the trend in aerosol effective climate forcing". Atmospheric Chemistry and Physics. 22 (18): 12221–12239. Bibcode:2022ACP....2212221Q. doi:10.5194/acp-22-12221-2022. hdl:20.500.11850/572791. S2CID 252446168.
  45. ^ EPA 2016: The U.S. Global Change Research Program, the National Academy of Sciences, and the Intergovernmental Panel on Climate Change (IPCC) have each independently concluded that warming of the climate system in recent decades is "unequivocal". This conclusion is not drawn from any one source of data but is based on multiple lines of evidence, including three worldwide temperature datasets showing nearly identical warming trends as well as numerous other independent indicators of global warming (e.g. rising sea levels, shrinking Arctic sea ice).
  46. ^ IPCC SR15 Ch1 2018, p. 81.
  47. ^ Earth System Science Data 2023, p. 2306
  48. ^ Samset, B. H.; Fuglestvedt, J. S.; Lund, M. T. (7 July 2020). "Delayed emergence of a global temperature response after emission mitigation". Nature Communications. 11 (1): 3261. Bibcode:2020NatCo..11.3261S. doi:10.1038/s41467-020-17001-1. hdl:11250/2771093. PMC 7341748. PMID 32636367. At the time of writing, that translated into 2035–2045, where the delay was mostly due to the impacts of the around 0.2 °C of natural, interannual variability of global mean surface air temperature
  49. ^ Seip, Knut L.; Grøn, ø.; Wang, H. (31 August 2023). "Global lead-lag changes between climate variability series coincide with major phase shifts in the Pacific decadal oscillation". Theoretical and Applied Climatology. 154 (3–4): 1137–1149. Bibcode:2023ThApC.154.1137S. doi:10.1007/s00704-023-04617-8. hdl:11250/3088837. ISSN 0177-798X. S2CID 261438532.
  50. ^ Yao, Shuai-Lei; Huang, Gang; Wu, Ren-Guang; Qu, Xia (January 2016). "The global warming hiatus—a natural product of interactions of a secular warming trend and a multi-decadal oscillation". Theoretical and Applied Climatology. 123 (1–2): 349–360. Bibcode:2016ThApC.123..349Y. doi:10.1007/s00704-014-1358-x. ISSN 0177-798X. S2CID 123602825. Retrieved 20 September 2023.
  51. ^ Xie, Shang-Ping; Kosaka, Yu (June 2017). "What Caused the Global Surface Warming Hiatus of 1998–2013?". Current Climate Change Reports. 3 (2): 128–140. Bibcode:2017CCCR....3..128X. doi:10.1007/s40641-017-0063-0. ISSN 2198-6061. S2CID 133522627. Retrieved 20 September 2023.
  52. ^ "Global temperature exceeds 2 °C above pre-industrial average on 17 November". Copernicus. 21 November 2023. Retrieved 31 January 2024. While exceeding the 2 °C threshold for a number of days does not mean that we have breached the Paris Agreement targets, the more often that we exceed this threshold, the more serious the cumulative effects of these breaches will become.
  53. ^ IPCC, 2021: Summary for Policymakers. In: Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change [Masson-Delmotte, V., P. Zhai, A. Pirani, S.L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M. I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J. B. R. Matthews, T. K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, and B. Zhou (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, New York, US, pp. 3−32, doi:10.1017/9781009157896.001.
  54. ^ McGrath, Matt (17 May 2023). "Global warming set to break key 1.5C limit for first time". BBC News. Retrieved 31 January 2024. The researchers stress that temperatures would have to stay at or above 1.5C for 20 years to be able to say the Paris agreement threshold had been passed.
  55. ^ Kennedy et al. 2010, p. S26. Figure 2.5.
  56. ^ Loeb et al. 2021.
  57. ^ "Global Warming". NASA JPL. 3 June 2010. Retrieved 11 September 2020. Satellite measurements show warming in the troposphere but cooling in the stratosphere. This vertical pattern is consistent with global warming due to increasing greenhouse gases but inconsistent with warming from natural causes.
  58. ^ Kennedy et al. 2010, pp. S26, S59–S60; USGCRP Chapter 1 2017, p. 35.
  59. ^ IPCC AR6 WG2 2022, pp. 257–260
  60. ^ IPCC SRCCL Summary for Policymakers 2019, p. 7
  61. ^ Sutton, Dong & Gregory 2007.
  62. ^ "Climate Change: Ocean Heat Content". Noaa Climate.gov. NOAA. 2018. Archived from the original on 12 February 2019. Retrieved 20 February 2019.
  63. ^ IPCC AR5 WG1 Ch3 2013, p. 257: "Ocean warming dominates the global energy change inventory. Warming of the ocean accounts for about 93% of the increase in the Earth's energy inventory between 1971 and 2010 (high confidence), with warming of the upper (0 to 700 m) ocean accounting for about 64% of the total.
  64. ^ von Schuckman, K.; Cheng, L.; Palmer, M. D.; Hansen, J.; et al. (7 September 2020). "Heat stored in the Earth system: where does the energy go?". Earth System Science Data. 12 (3): 2013–2041. Bibcode:2020ESSD...12.2013V. doi:10.5194/essd-12-2013-2020. hdl:20.500.11850/443809.
  65. ^ NOAA, 10 July 2011.
  66. ^ United States Environmental Protection Agency 2016, p. 5: "Black carbon that is deposited on snow and ice darkens those surfaces and decreases their reflectivity (albedo). This is known as the snow/ice albedo effect. This effect results in the increased absorption of radiation that accelerates melting."
  67. ^ "Arctic warming three times faster than the planet, report warns". Phys.org. 20 May 2021. Retrieved 6 October 2022.
  68. ^ Rantanen, Mika; Karpechko, Alexey Yu; Lipponen, Antti; Nordling, Kalle; Hyvärinen, Otto; Ruosteenoja, Kimmo; Vihma, Timo; Laaksonen, Ari (11 August 2022). "The Arctic has warmed nearly four times faster than the globe since 1979". Communications Earth & Environment. 3 (1): 168. Bibcode:2022ComEE...3..168R. doi:10.1038/s43247-022-00498-3. hdl:11250/3115996. ISSN 2662-4435. S2CID 251498876.
  69. ^ "The Arctic is warming four times faster than the rest of the world". 14 December 2021. Retrieved 6 October 2022.
  70. ^ Liu, Wei; Fedorov, Alexey V.; Xie, Shang-Ping; Hu, Shineng (26 June 2020). "Climate impacts of a weakened Atlantic Meridional Overturning Circulation in a warming climate". Science Advances. 6 (26): eaaz4876. Bibcode:2020SciA....6.4876L. doi:10.1126/sciadv.aaz4876. PMC 7319730. PMID 32637596.
  71. ^ a b Pearce, Fred (18 April 2023). "New Research Sparks Concerns That Ocean Circulation Will Collapse". Retrieved 3 February 2024.
  72. ^ Lee, Sang-Ki; Lumpkin, Rick; Gomez, Fabian; Yeager, Stephen; Lopez, Hosmay; Takglis, Filippos; Dong, Shenfu; Aguiar, Wilton; Kim, Dongmin; Baringer, Molly (13 March 2023). "Human-induced changes in the global meridional overturning circulation are emerging from the Southern Ocean". Communications Earth & Environment. 4 (1): 69. Bibcode:2023ComEE...4...69L. doi:10.1038/s43247-023-00727-3.
  73. ^ "NOAA Scientists Detect a Reshaping of the Meridional Overturning Circulation in the Southern Ocean". NOAA. 29 March 2023.
  74. ^ Schuur, Edward A.G.; Abbott, Benjamin W.; Commane, Roisin; Ernakovich, Jessica; Euskirchen, Eugenie; Hugelius, Gustaf; Grosse, Guido; Jones, Miriam; Koven, Charlie; Leshyk, Victor; Lawrence, David; Loranty, Michael M.; Mauritz, Marguerite; Olefeldt, David; Natali, Susan; Rodenhizer, Heidi; Salmon, Verity; Schädel, Christina; Strauss, Jens; Treat, Claire; Turetsky, Merritt (2022). "Permafrost and Climate Change: Carbon Cycle Feedbacks From the Warming Arctic". Annual Review of Environment and Resources. 47: 343–371. doi:10.1146/annurev-environ-012220-011847. Medium-range estimates of Arctic carbon emissions could result from moderate climate emission mitigation policies that keep global warming below 3 °C (e.g., RCP4.5). This global warming level most closely matches country emissions reduction pledges made for the Paris Climate Agreement...
  75. ^ Phiddian, Ellen (5 April 2022). "Explainer: IPCC Scenarios". Cosmos. Retrieved 30 September 2023. "The IPCC doesn't make projections about which of these scenarios is more likely, but other researchers and modellers can. The Australian Academy of Science, for instance, released a report last year stating that our current emissions trajectory had us headed for a 3 °C warmer world, roughly in line with the middle scenario. Climate Action Tracker predicts 2.5 to 2.9 °C of warming based on current policies and action, with pledges and government agreements taking this to 2.1 °C.
  76. ^ McGrath, Matt (17 May 2023). "Global warming set to break key 1.5C limit for first time". BBC. Retrieved 17 May 2023.
  77. ^ Harvey, Fiona (17 May 2023). "World likely to breach 1.5C climate threshold by 2027, scientists warn". The Guardian. Retrieved 17 May 2023.
  78. ^ https://www.ipcc.ch/report/ar6/wg1/downloads/report/IPCC_AR6_WGI_Full_Report.pdf#page=955
  79. ^ IPCC AR6 WG1 Summary for Policymakers 2021, p. SPM-17
  80. ^ Meinshausen, Malte; Smith, S. J.; Calvin, K.; Daniel, J. S.; Kainuma, M. L. T.; Lamarque, J-F.; Matsumoto, K.; Montzka, S. A.; Raper, S. C. B.; Riahi, K.; Thomson, A.; Velders, G. J. M.; van Vuuren, D.P. P. (2011). "The RCP greenhouse gas concentrations and their extensions from 1765 to 2300". Climatic Change. 109 (1–2): 213–241. Bibcode:2011ClCh..109..213M. doi:10.1007/s10584-011-0156-z. ISSN 0165-0009.
  81. ^ Lyon, Christopher; Saupe, Erin E.; Smith, Christopher J.; Hill, Daniel J.; Beckerman, Andrew P.; Stringer, Lindsay C.; Marchant, Robert; McKay, James; Burke, Ariane; O'Higgins, Paul; Dunhill, Alexander M.; Allen, Bethany J.; Riel-Salvatore, Julien; Aze, Tracy (2021). "Climate change research and action must look beyond 2100". Global Change Biology. 28 (2): 349–361. doi:10.1111/gcb.15871. hdl:20.500.11850/521222. ISSN 1365-2486. PMID 34558764. S2CID 237616583.
  82. ^ Rogelj et al. 2019
  83. ^ a b IPCC SR15 Summary for Policymakers 2018, p. 12
  84. ^ IPCC AR5 WG3 Ch5 2014, pp. 379–380.
  85. ^ Brown, Patrick T.; Li, Wenhong; Xie, Shang-Ping (27 January 2015). "Regions of significant influence on unforced global mean surface air temperature variability in climate models: Origin of global temperature variability". Journal of Geophysical Research: Atmospheres. 120 (2): 480–494. doi:10.1002/2014JD022576. hdl:10161/9564.
  86. ^ Trenberth, Kevin E.; Fasullo, John T. (December 2013). "An apparent hiatus in global warming?". Earth's Future. 1 (1): 19–32. Bibcode:2013EaFut...1...19T. doi:10.1002/2013EF000165.
  87. ^ National Research Council 2012, p. 9
  88. ^ IPCC AR5 WG1 Ch10 2013, p. 916.
  89. ^ Knutson 2017, p. 443; IPCC AR5 WG1 Ch10 2013, pp. 875–876
  90. ^ a b USGCRP 2009, p. 20.
  91. ^ IPCC AR6 WG1 Summary for Policymakers 2021, p. 7
  92. ^ Lüthi, Dieter; Le Floch, Martine; Bereiter, Bernhard; Blunier, Thomas; Barnola, Jean-Marc; Siegenthaler, Urs; Raynaud, Dominique; Jouzel, Jean; Fischer, Hubertus; Kawamura, Kenji; Stocker, Thomas F. (May 2005). "High-resolution carbon dioxide concentration record 650,000–800,000 years before present". Nature. 453 (7193): 379–382. Bibcode:2008Natur.453..379L. doi:10.1038/nature06949. ISSN 0028-0836. PMID 18480821. S2CID 1382081.
  93. ^ Fischer, Hubertus; Wahlen, Martin; Smith, Jesse; Mastroianni, Derek; Deck, Bruce (12 March 1999). "Ice Core Records of Atmospheric CO 2 Around the Last Three Glacial Terminations". Science. 283 (5408): 1712–1714. Bibcode:1999Sci...283.1712F. doi:10.1126/science.283.5408.1712. ISSN 0036-8075. PMID 10073931.
  94. ^ Indermühle, Andreas; Monnin, Eric; Stauffer, Bernhard; Stocker, Thomas F.; Wahlen, Martin (1 March 2000). "Atmospheric CO 2 concentration from 60 to 20 kyr BP from the Taylor Dome Ice Core, Antarctica". Geophysical Research Letters. 27 (5): 735–738. Bibcode:2000GeoRL..27..735I. doi:10.1029/1999GL010960. S2CID 18942742.
  95. ^ Etheridge, D.; Steele, L.; Langenfelds, R.; Francey, R.; Barnola, J.-M.; Morgan, V. (1998). "Historical CO2 Records from the Law Dome DE08, DE08-2, and DSS Ice Cores". Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory. U.S. Department of Energy. Retrieved 20 November 2022.
  96. ^ Keeling, C.; Whorf, T. (2004). "Atmospheric CO2 Records from Sites in the SIO Air Sampling Network". Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory. U.S. Department of Energy. Retrieved 20 November 2022.
  97. ^ NASA. "The Causes of Climate Change". Climate Change: Vital Signs of the Planet. Archived from the original on 8 May 2019. Retrieved 8 May 2019.
  98. ^ Ozone acts as a greenhouse gas in the lowest layer of the atmosphere, the troposphere (as opposed to the stratospheric ozone layer). Wang, Shugart & Lerdau 2017
  99. ^ Schmidt et al. 2010; USGCRP Climate Science Supplement 2014, p. 742
  100. ^ IPCC AR4 WG1 Ch1 2007, FAQ1.1: "To emit 240 W m−2, a surface would have to have a temperature of around −19 °C. This is much colder than the conditions that actually exist at the Earth's surface (the global mean surface temperature is about 14 °C).
  101. ^ ACS. "What Is the Greenhouse Effect?". Archived from the original on 26 May 2019. Retrieved 26 May 2019.
  102. ^ The Guardian, 19 February 2020.
  103. ^ WMO 2021, p. 8.
  104. ^ IPCC AR6 WG1 Technical Summary 2021, p. TS-35.
  105. ^ IPCC AR6 WG3 Summary for Policymakers 2022, Figure SPM.1.
  106. ^ Olivier & Peters 2019, p. 17; Our World in Data, 18 September 2020; EPA 2020: Greenhouse gas emissions from industry primarily come from burning fossil fuels for energy, as well as greenhouse gas emissions from certain chemical reactions necessary to produce goods from raw materials; "Redox, extraction of iron and transition metals". Hot air (oxygen) reacts with the coke (carbon) to produce carbon dioxide and heat energy to heat up the furnace. Removing impurities: The calcium carbonate in the limestone thermally decomposes to form calcium oxide. calcium carbonate → calcium oxide + carbon dioxide; Kvande 2014: Carbon dioxide gas is formed at the anode, as the carbon anode is consumed upon reaction of carbon with the oxygen ions from the alumina (Al2O3). Formation of carbon dioxide is unavoidable as long as carbon anodes are used, and it is of great concern because CO2 is a greenhouse gas
  107. ^ EPA 2020; Global Methane Initiative 2020: Estimated Global Anthropogenic Methane Emissions by Source, 2020: Enteric fermentation (27%), Manure Management (3%), Coal Mining (9%), Municipal Solid Waste (11%), Oil & Gas (24%), Wastewater (7%), Rice Cultivation (7%)
  108. ^ EPA 2019: Agricultural activities, such as fertilizer use, are the primary source of N2O emissions; Davidson 2009: 2.0% of manure nitrogen and 2.5% of fertilizer nitrogen was converted to nitrous oxide between 1860 and 2005; these percentage contributions explain the entire pattern of increasing nitrous oxide concentrations over this period
  109. ^ "Understanding methane emissions". International Energy Agency.
  110. ^ a b Riebeek, Holli (16 June 2011). "The Carbon Cycle". Earth Observatory. NASA. Archived from the original on 5 March 2016. Retrieved 5 April 2018.
  111. ^ IPCC SRCCL Summary for Policymakers 2019, p. 10
  112. ^ IPCC SROCC Ch5 2019, p. 450.
  113. ^ World Resources Institute, 31 March 2021
  114. ^ Ritchie & Roser 2018
  115. ^ The Sustainability Consortium, 13 September 2018; UN FAO 2016, p. 18.
  116. ^ Curtis et al. 2018
  117. ^ a b c Garrett, L.; Lévite, H.; Besacier, C.; Alekseeva, N.; Duchelle, M. (2022). The key role of forest and landscape restoration in climate action. Rome: FAO. doi:10.4060/cc2510en. ISBN 978-92-5-137044-5.
  118. ^ a b World Resources Institute, 8 December 2019
  119. ^ IPCC SRCCL Ch2 2019, p. 172: "The global biophysical cooling alone has been estimated by a larger range of climate models and is −0.10 ± 0.14 °C; it ranges from −0.57 °C to +0.06 °C ... This cooling is essentially dominated by increases in surface albedo: historical land cover changes have generally led to a dominant brightening of land"
  120. ^ Haywood 2016, p. 456; McNeill 2017; Samset et al. 2018.
  121. ^ IPCC AR5 WG1 Ch2 2013, p. 183.
  122. ^ He et al. 2018; Storelvmo et al. 2016
  123. ^ "Aerosol pollution has caused decades of global dimming". American Geophysical Union. 18 February 2021. Archived from the original on 27 March 2023. Retrieved 18 December 2023.
  124. ^ Xia, Wenwen; Wang, Yong; Chen, Siyu; Huang, Jianping; Wang, Bin; Zhang, Guang J.; Zhang, Yue; Liu, Xiaohong; Ma, Jianmin; Gong, Peng; Jiang, Yiquan; Wu, Mingxuan; Xue, Jinkai; Wei, Linyi; Zhang, Tinghan (2022). "Double Trouble of Air Pollution by Anthropogenic Dust". Environmental Science & Technology. 56 (2): 761–769. Bibcode:2022EnST...56..761X. doi:10.1021/acs.est.1c04779. hdl:10138/341962. PMID 34941248. S2CID 245445736.
  125. ^ "Global Dimming Dilemma". 4 June 2020.
  126. ^ Wild et al. 2005; Storelvmo et al. 2016; Samset et al. 2018.
  127. ^ Twomey 1977.
  128. ^ Albrecht 1989.
  129. ^ a b c USGCRP Chapter 2 2017, p. 78.
  130. ^ Ramanathan & Carmichael 2008; RIVM 2016.
  131. ^ Sand et al. 2015
  132. ^ IMO 2020 – cutting sulphur oxide emissions
  133. ^ CarbonBrief, 2023
  134. ^ "Climate Science Special Report: Fourth National Climate Assessment, Volume I - Chapter 3: Detection and Attribution of Climate Change". science2017.globalchange.gov. U.S. Global Change Research Program (USGCRP): 1–470. 2017. Archived from the original on 23 September 2019. Adapted directly from Fig. 3.3.
  135. ^ Wuebbles, D.J.; Fahey, D.W.; Hibbard, K.A.; Deangelo, B.; Doherty, S.; Hayhoe, K.; Horton, R.; Kossin, J.P.; Taylor, P.C.; Waple, A.M.; Yohe, C.P. (23 November 2018). "Climate Science Special Report / Fourth National Climate Assessment (NCA4), Volume I /Executive Summary / Highlights of the Findings of the U.S. Global Change Research Program Climate Science Special Report". globalchange.gov. U.S. Global Change Research Program: 1–470. doi:10.7930/J0DJ5CTG. Archived from the original on 14 June 2019.
  136. ^ National Academies 2008, p. 6
  137. ^ "Is the Sun causing global warming?". Climate Change: Vital Signs of the Planet. Archived from the original on 5 May 2019. Retrieved 10 May 2019.
  138. ^ IPCC AR4 WG1 Ch9 2007, pp. 702–703; Randel et al. 2009.
  139. ^ Greicius, Tony (2 August 2022). "Tonga eruption blasted unprecedented amount of water into stratosphere". NASA Global Climate Change. Retrieved 18 January 2024. Massive volcanic eruptions like Krakatoa and Mount Pinatubo typically cool Earth's surface by ejecting gases, dust, and ash that reflect sunlight back into space. In contrast, the Tonga volcano didn't inject large amounts of aerosols into the stratosphere, and the huge amounts of water vapor from the eruption may have a small, temporary warming effect, since water vapor traps heat. The effect would dissipate when the extra water vapor cycles out of the stratosphere and would not be enough to noticeably exacerbate climate change effects.
  140. ^ USGCRP Chapter 2 2017, p. 79
  141. ^ Fischer & Aiuppa 2020.
  142. ^ USGCRP Chapter 2 2017, p. 79
  143. ^ "Thermodynamics: Albedo". NSIDC. Archived from the original on 11 October 2017. Retrieved 10 October 2017.
  144. ^ "The study of Earth as an integrated system". Vitals Signs of the Planet. Earth Science Communications Team at NASA's Jet Propulsion Laboratory / California Institute of Technology. 2013. Archived from the original on 26 February 2019.
  145. ^ a b USGCRP Chapter 2 2017, pp. 89–91.
  146. ^ IPCC AR6 WG1 Technical Summary 2021, p. 58: The net effect of changes in clouds in response to global warming is to amplify human-induced warming, that is, the net cloud feedback is positive (high confidence)
  147. ^ USGCRP Chapter 2 2017, pp. 89–90.
  148. ^ IPCC AR5 WG1 2013, p. 14
  149. ^ IPCC AR6 WG1 Technical Summary 2021, pp. 58, 59: clouds remain the largest contribution to overall uncertainty in climate feedbacks
  150. ^ Wolff et al. 2015: "the nature and magnitude of these feedbacks are the principal cause of uncertainty in the response of Earth's climate (over multi-decadal and longer periods) to a particular emissions scenario or greenhouse gas concentration pathway."
  151. ^ Williams, Ceppi & Katavouta 2020.
  152. ^ NASA, 28 May 2013.
  153. ^ Cohen et al. 2014.
  154. ^ a b Turetsky et al. 2019
  155. ^ Dean et al. 2018.
  156. ^ IPCC AR6 WG1 Technical Summary 2021, p. 58: 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.
  157. ^ Climate.gov, 23 June 2022:"Carbon cycle experts estimate that natural "sinks"—processes that remove carbon from the atmosphere—on land and in the ocean absorbed the equivalent of about half of the carbon dioxide we emitted each year in the 2011–2020 decade."
  158. ^ IPCC AR6 WG1 Technical Summary 2021, p. TS-122, Box TS.5, Figure 1
  159. ^ Melillo et al. 2017: Our first-order estimate of a warming-induced loss of 190 Pg of soil carbon over the 21st century is equivalent to the past two decades of carbon emissions from fossil fuel burning.
  160. ^ IPCC SRCCL Ch2 2019, pp. 133, 144.
  161. ^ USGCRP Chapter 2 2017, pp. 93–95.
  162. ^ Liu, Y.; Moore, J. K.; Primeau, F.; Wang, W. L. (22 December 2022). "Reduced CO2 uptake and growing nutrient sequestration from slowing overturning circulation". Nature Climate Change. 13: 83–90. doi:10.1038/s41558-022-01555-7. OSTI 2242376. S2CID 255028552.
  163. ^ IPCC AR5 SYR Glossary 2014, p. 120.
  164. ^ Carbon Brief, 15 January 2018, "What are the different types of climate models?"
  165. ^ Wolff et al. 2015
  166. ^ Carbon Brief, 15 January 2018, "Who does climate modelling around the world?"
  167. ^ Carbon Brief, 15 January 2018, "What is a climate model?"
  168. ^ IPCC AR4 WG1 Ch8 2007, FAQ 8.1.
  169. ^ Stroeve et al. 2007; National Geographic, 13 August 2019
  170. ^ Liepert & Previdi 2009.
  171. ^ Rahmstorf et al. 2007; Mitchum et al. 2018
  172. ^ USGCRP Chapter 15 2017.
  173. ^ Hébert, R.; Herzschuh, U.; Laepple, T. (31 October 2022). "Millennial-scale climate variability over land overprinted by ocean temperature fluctuations". Nature Geoscience. 15 (1): 899–905. Bibcode:2022NatGe..15..899H. doi:10.1038/s41561-022-01056-4. PMC 7614181. PMID 36817575.
  174. ^ Carbon Brief, 15 January 2018, "What are the inputs and outputs for a climate model?"
  175. ^ Matthews et al. 2009
  176. ^ Carbon Brief, 19 April 2018; Meinshausen 2019, p. 462.
  177. ^ Hansen et al. 2016; Smithsonian, 26 June 2016.
  178. ^ USGCRP Chapter 15 2017, p. 415.
  179. ^ Scientific American, 29 April 2014; Burke & Stott 2017.
  180. ^ Liu, Fei; Wang, Bin; Ouyang, Yu; Wang, Hui; Qiao, Shaobo; Chen, Guosen; Dong, Wenjie (19 April 2022). "Intraseasonal variability of global land monsoon precipitation and its recent trend". npj Climate and Atmospheric Science. 5 (1): 30. Bibcode:2022npCAS...5...30L. doi:10.1038/s41612-022-00253-7. ISSN 2397-3722.
  181. ^ USGCRP Chapter 9 2017, p. 260.
  182. ^ Studholme, Joshua; Fedorov, Alexey V.; Gulev, Sergey K.; Emanuel, Kerry; Hodges, Kevin (29 December 2021). "Poleward expansion of tropical cyclone latitudes in warming climates". Nature Geoscience. 15: 14–28. doi:10.1038/s41561-021-00859-1. S2CID 245540084.
  183. ^ "Hurricanes and Climate Change". Center for Climate and Energy Solutions. 10 July 2020.
  184. ^ NOAA 2017.
  185. ^ WMO 2021, p. 12.
  186. ^ IPCC AR6 WG2 2022, p. 1302
  187. ^ DeConto & Pollard 2016
  188. ^ Bamber et al. 2019.
  189. ^ Zhang et al. 2008
  190. ^ IPCC SROCC Summary for Policymakers 2019, p. 18
  191. ^ Doney et al. 2009.
  192. ^ Deutsch et al. 2011
  193. ^ IPCC SROCC Ch5 2019, p. 510; "Climate Change and Harmful Algal Blooms". EPA. 5 September 2013. Retrieved 11 September 2020.
  194. ^ "Tipping Elements – big risks in the Earth System". Potsdam Institute for Climate Impact Research. Retrieved 31 January 2024.
  195. ^ a b c Armstrong McKay, David I.; Staal, Arie; Abrams, Jesse F.; Winkelmann, Ricarda; Sakschewski, Boris; Loriani, Sina; Fetzer, Ingo; Cornell, Sarah E.; Rockström, Johan; Lenton, Timothy M. (9 September 2022). "Exceeding 1.5 °C global warming could trigger multiple climate tipping points". Science. 377 (6611): eabn7950. doi:10.1126/science.abn7950. hdl:10871/131584. ISSN 0036-8075. PMID 36074831. S2CID 252161375.
  196. ^ IPCC SR15 Ch3 2018, p. 283.
  197. ^ Pearce, Rosamund; Prater, Tom (10 February 2020). "Nine Tipping Points That Could Be Triggered by Climate Change". CarbonBrief. Retrieved 27 May 2022.
  198. ^ Bochow, Nils; Poltronieri, Anna; Robinson, Alexander; Montoya, Marisa; Rypdal, Martin; Boers, Niklas (18 October 2023). "Overshooting the critical threshold for the Greenland ice sheet". Nature. 622 (7983): 528–536. Bibcode:2023Natur.622..528B. doi:10.1038/s41586-023-06503-9. PMC 10584691. PMID 37853149.
  199. ^ IPCC AR6 WG1 Summary for Policymakers 2021, p. 21
  200. ^ IPCC AR5 WG1 Ch12 2013, pp. 88–89, FAQ 12.3
  201. ^ Smith et al. 2009; Levermann et al. 2013
  202. ^ IPCC AR5 WG1 Ch12 2013, p. 1112.
  203. ^ Oschlies, Andreas (16 April 2021). "A committed fourfold increase in ocean oxygen loss". Nature Communications. 12 (1): 2307. Bibcode:2021NatCo..12.2307O. doi:10.1038/s41467-021-22584-4. PMC 8052459. PMID 33863893.
  204. ^ Voosen, Paul (18 December 2018). "Discovery of recent Antarctic ice sheet collapse raises fears of a new global flood". Science. Retrieved 28 December 2018.
  205. ^ Turney, Chris S. M.; Fogwill, Christopher J.; Golledge, Nicholas R.; McKay, Nicholas P.; Sebille, Erik van; Jones, Richard T.; Etheridge, David; Rubino, Mauro; Thornton, David P.; Davies, Siwan M.; Ramsey, Christopher Bronk (11 February 2020). "Early Last Interglacial ocean warming drove substantial ice mass loss from Antarctica". Proceedings of the National Academy of Sciences. 117 (8): 3996–4006. Bibcode:2020PNAS..117.3996T. doi:10.1073/pnas.1902469117. ISSN 0027-8424. PMC 7049167. PMID 32047039.
  206. ^ Carlson, Anders E; Walczak, Maureen H; Beard, Brian L; Laffin, Matthew K; Stoner, Joseph S; Hatfield, Robert G (10 December 2018). Absence of the West Antarctic ice sheet during the last interglaciation. American Geophysical Union Fall Meeting.
  207. ^ Lau, Sally C. Y.; Wilson, Nerida G.; Golledge, Nicholas R.; Naish, Tim R.; Watts, Phillip C.; Silva, Catarina N. S.; Cooke, Ira R.; Allcock, A. Louise; Mark, Felix C.; Linse, Katrin (21 December 2023). "Genomic evidence for West Antarctic Ice Sheet collapse during the Last Interglacial" (PDF). Science. 382 (6677): 1384–1389. Bibcode:2023Sci...382.1384L. doi:10.1126/science.ade0664. PMID 38127761. S2CID 266436146.
  208. ^ AHMED, Issam. "Antarctic octopus DNA reveals ice sheet collapse closer than thought". phys.org. Retrieved 23 December 2023.
  209. ^ A. Naughten, Kaitlin; R. Holland, Paul; De Rydt, Jan (23 October 2023). "Unavoidable future increase in West Antarctic ice-shelf melting over the twenty-first century". Nature Climate Change. 13 (11): 1222–1228. Bibcode:2023NatCC..13.1222N. doi:10.1038/s41558-023-01818-x. S2CID 264476246.
  210. ^ Poynting, Mark (24 October 2023). "Sea-level rise: West Antarctic ice shelf melt 'unavoidable'". BBC. Retrieved 26 October 2023.
  211. ^ IPCC SR15 Ch3 2018, p. 218.
  212. ^ IPCC SRCCL Ch2 2019, p. 133.
  213. ^ IPCC SRCCL Summary for Policymakers 2019, p. 7; Zeng & Yoon 2009.
  214. ^ Turner et al. 2020, p. 1.
  215. ^ Urban 2015.
  216. ^ Poloczanska et al. 2013; Lenoir et al. 2020
  217. ^ Smale et al. 2019
  218. ^ IPCC SROCC Summary for Policymakers 2019, p. 13.
  219. ^ IPCC SROCC Ch5 2019, p. 510
  220. ^ IPCC SROCC Ch5 2019, p. 451.
  221. ^ "Coral Reef Risk Outlook". National Oceanic and Atmospheric Administration. 2 January 2012. Retrieved 4 April 2020. At present, local human activities, coupled with past thermal stress, threaten an estimated 75 percent of the world's reefs. By 2030, estimates predict more than 90% of the world's reefs will be threatened by local human activities, warming, and acidification, with nearly 60% facing high, very high, or critical threat levels.
  222. ^ Carbon Brief, 7 January 2020.
  223. ^ IPCC AR5 WG2 Ch28 2014, p. 1596: "Within 50 to 70 years, loss of hunting habitats may lead to elimination of polar bears from seasonally ice-covered areas, where two-thirds of their world population currently live."
  224. ^ "What a changing climate means for Rocky Mountain National Park". National Park Service. Retrieved 9 April 2020.
  225. ^ IPCC AR6 WG1 Summary for Policymakers 2021, p. SPM-23, Fig. SPM.6
  226. ^ Lenton, Timothy M.; Xu, Chi; Abrams, Jesse F.; Ghadiali, Ashish; Loriani, Sina; Sakschewski, Boris; Zimm, Caroline; Ebi, Kristie L.; Dunn, Robert R.; Svenning, Jens-Christian; Scheffer, Marten (2023). "Quantifying the human cost of global warming". Nature Sustainability. 6 (10): 1237–1247. Bibcode:2023NatSu...6.1237L. doi:10.1038/s41893-023-01132-6. hdl:10871/132650.
  227. ^ IPCC AR5 WG2 Ch18 2014, pp. 983, 1008
  228. ^ IPCC AR5 WG2 Ch19 2014, p. 1077.
  229. ^ IPCC AR5 SYR Summary for Policymakers 2014, p. 8, SPM 2
  230. ^ IPCC AR5 SYR Summary for Policymakers 2014, p. 13, SPM 2.3
  231. ^ WHO, Nov 2015
  232. ^ IPCC AR5 WG2 Ch11 2014, pp. 720–723
  233. ^ Watts et al. 2019, pp. 1836, 1848.
  234. ^ Costello et al. 2009; Watts et al. 2015; IPCC AR5 WG2 Ch11 2014, p. 713
  235. ^ Watts et al. 2019, pp. 1841, 1847.
  236. ^ WHO 2014: "Under a base case socioeconomic scenario, we estimate approximately 250 000 additional deaths due to climate change per year between 2030 and 2050. These numbers do not represent a prediction of the overall impacts of climate change on health, since we could not quantify several important causal pathways."
  237. ^ IPCC AR6 WG2 2022, p. 988
  238. ^ IPCC SRCCL Ch5 2019, p. 451.
  239. ^ Zhao et al. 2017; IPCC SRCCL Ch5 2019, p. 439
  240. ^ IPCC AR5 WG2 Ch7 2014, p. 488
  241. ^ IPCC SRCCL Ch5 2019, p. 462
  242. ^ IPCC SROCC Ch5 2019, p. 503.
  243. ^ Holding et al. 2016; IPCC AR5 WG2 Ch3 2014, pp. 232–233.
  244. ^ DeFries et al. 2019, p. 3; Krogstrup & Oman 2019, p. 10.
  245. ^ a b Women's leadership and gender equality in climate action and disaster risk reduction in Africa − A call for action. Accra: FAO & The African Risk Capacity (ARC) Group. 2021. doi:10.4060/cb7431en. ISBN 978-92-5-135234-2. S2CID 243488592.
  246. ^ IPCC AR5 WG2 Ch13 2014, pp. 796–797
  247. ^ IPCC AR6 WG2 2022, p. 725
  248. ^ Hallegatte et al. 2016, p. 12.
  249. ^ IPCC AR5 WG2 Ch13 2014, p. 796.
  250. ^ Grabe, Grose and Dutt, 2014; FAO, 2011; FAO, 2021a; Fisher and Carr, 2015; IPCC, 2014; Resurrección et al., 2019; UNDRR, 2019; Yeboah et al., 2019.
  251. ^ "Climate Change | United Nations For Indigenous Peoples". United Nations Department of Economic and Social Affairs. Retrieved 29 April 2022.
  252. ^ Mach et al. 2019.
  253. ^ a b The status of women in agrifood systems - Overview. Rome: FAO. 2023. doi:10.4060/cc5060en. S2CID 258145984.
  254. ^ IPCC SROCC Ch4 2019, p. 328.
  255. ^ UNHCR 2011, p. 3.
  256. ^ Matthews 2018, p. 399.
  257. ^ Balsari, Dresser & Leaning 2020
  258. ^ Flavell 2014, p. 38; Kaczan & Orgill-Meyer 2020
  259. ^ Serdeczny et al. 2016.
  260. ^ IPCC SRCCL Ch5 2019, pp. 439, 464.
  261. ^ National Oceanic and Atmospheric Administration. "What is nuisance flooding?". Retrieved 8 April 2020.
  262. ^ Kabir et al. 2016.
  263. ^ Van Oldenborgh et al. 2019.
  264. ^ IPCC AR5 SYR Glossary 2014, p. 125.
  265. ^ IPCC SR15 Summary for Policymakers 2018, p. 15
  266. ^ United Nations Environment Programme 2019, p. XX
  267. ^ IPCC AR6 WG3 2022, p. 300: The global benefits of pathways limiting warming to 2 °C (>67%) outweigh global mitigation costs over the 21st century, if aggregated economic impacts of climate change are at the moderate to high end of the assessed range, and a weight consistent with economic theory is given to economic impacts over the long term. This holds true even without accounting for benefits in other sustainable development dimensions or nonmarket damages from climate change (medium confidence).
  268. ^ IPCC SR15 Ch2 2018, p. 109.
  269. ^ Teske, ed. 2019, p. xxiii.
  270. ^ World Resources Institute, 8 August 2019
  271. ^ IPCC SR15 Ch3 2018, p. 266: Where reforestation is the restoration of natural ecosystems, it benefits both carbon sequestration and conservation of biodiversity and ecosystem services.
  272. ^ Bui et al. 2018, p. 1068; IPCC SR15 Summary for Policymakers 2018, p. 17
  273. ^ IPCC SR15 2018, p. 34; IPCC SR15 Summary for Policymakers 2018, p. 17
  274. ^ IPCC SR15 Ch4 2018, pp. 347–352
  275. ^ Friedlingstein et al. 2019
  276. ^ a b United Nations Environment Programme 2019, p. 46; Vox, 20 September 2019; Sepulveda, Nestor A.; Jenkins, Jesse D.; De Sisternes, Fernando J.; Lester, Richard K. (2018). "The Role of Firm Low-Carbon Electricity Resources in Deep Decarbonization of Power Generation". Joule. 2 (11): 2403–2420. doi:10.1016/j.joule.2018.08.006.
  277. ^ IEA World Energy Outlook 2023, pp. 18
  278. ^ REN21 2020, p. 32, Fig.1.
  279. ^ IEA World Energy Outlook 2023, pp. 18, 26
  280. ^ The Guardian, 6 April 2020.
  281. ^ IEA 2021, p. 57, Fig 2.5; Teske et al. 2019, p. 180, Table 8.1
  282. ^ Our World in Data-Why did renewables become so cheap so fast?; IEA – Projected Costs of Generating Electricity 2020
  283. ^ "IPCC Working Group III report: Mitigation of Climate Change". Intergovernmental Panel on Climate Change. 4 April 2022. Retrieved 19 January 2024.
  284. ^ IPCC SR15 Ch2 2018, p. 131, Figure 2.15
  285. ^ Teske 2019, pp. 409–410.
  286. ^ United Nations Environment Programme 2019, p. XXIII, Table ES.3; Teske, ed. 2019, p. xxvii, Fig.5.
  287. ^ a b IPCC SR15 Ch2 2018, pp. 142–144; United Nations Environment Programme 2019, Table ES.3 & p. 49
  288. ^ "Transport emissions". Climate action. European Commission. 2016. Archived from the original on 10 October 2021. Retrieved 2 January 2022.
  289. ^ IPCC AR5 WG3 Ch9 2014, p. 697; NREL 2017, pp. vi, 12
  290. ^ Berrill et al. 2016.
  291. ^ IPCC SR15 Ch4 2018, pp. 324–325.
  292. ^ Gill, Matthew; Livens, Francis; Peakman, Aiden. "Nuclear Fission". In Letcher (2020), pp. 147–149.
  293. ^ Horvath, Akos; Rachlew, Elisabeth (January 2016). "Nuclear power in the 21st century: Challenges and possibilities". Ambio. 45 (Suppl 1): S38–49. Bibcode:2016Ambio..45S..38H. doi:10.1007/s13280-015-0732-y. ISSN 1654-7209. PMC 4678124. PMID 26667059.
  294. ^ "Hydropower". iea.org. International Energy Agency. Retrieved 12 October 2020. Hydropower generation is estimated to have increased by over 2% in 2019 owing to continued recovery from drought in Latin America as well as strong capacity expansion and good water availability in China (...) capacity expansion has been losing speed. This downward trend is expected to continue, due mainly to less large-project development in China and Brazil, where concerns over social and environmental impacts have restricted projects.
  295. ^ Watts et al. 2019, p. 1854; WHO 2018, p. 27
  296. ^ Watts et al. 2019, p. 1837; WHO 2016
  297. ^ WHO 2018, p. 27; Vandyck et al. 2018; IPCC SR15 2018, p. 97: "Limiting warming to 1.5 °C can be achieved synergistically with poverty alleviation and improved energy security and can provide large public health benefits through improved air quality, preventing millions of premature deaths. However, specific mitigation measures, such as bioenergy, may result in trade-offs that require consideration."
  298. ^ IPCC AR6 WG3 2022, p. 300
  299. ^ IPCC SR15 Ch2 2018, p. 97
  300. ^ IPCC AR5 SYR Summary for Policymakers 2014, p. 29; IEA 2020b
  301. ^ IPCC SR15 Ch2 2018, p. 155, Fig. 2.27
  302. ^ IEA 2020b
  303. ^ IPCC SR15 Ch2 2018, p. 142
  304. ^ IPCC SR15 Ch2 2018, pp. 138–140
  305. ^ IPCC SR15 Ch2 2018, pp. 141–142
  306. ^ IPCC AR5 WG3 Ch9 2014, pp. 686–694.
  307. ^ World Resources Institute, December 2019, p. 1
  308. ^ World Resources Institute, December 2019, pp. 1, 3
  309. ^ IPCC SRCCL 2019, p. 22, B.6.2
  310. ^ IPCC SRCCL Ch5 2019, pp. 487, 488, FIGURE 5.12 Humans on a vegan exclusive diet would save about 7.9 GtCO2 equivalent per year by 2050 IPCC AR6 WG1 Technical Summary 2021, p. 51 Agriculture, Forestry and Other Land Use used an average of 12 GtCO2 per year between 2007 and 2016 (23% of total anthropogenic emissions).
  311. ^ IPCC SRCCL Ch5 2019, pp. 82, 162, FIGURE 1.1
  312. ^ "Low and zero emissions in the steel and cement industries" (PDF). pp. 11, 19–22.
  313. ^ World Resources Institute, 8 August 2019: IPCC SRCCL Ch2 2019, pp. 189–193.
  314. ^ Kreidenweis et al. 2016
  315. ^ National Academies of Sciences, Engineering, and Medicine 2019, pp. 95–102
  316. ^ National Academies of Sciences, Engineering, and Medicine 2019, pp. 45–54
  317. ^ Nelson, J. D. J.; Schoenau, J. J.; Malhi, S. S. (1 October 2008). "Soil organic carbon changes and distribution in cultivated and restored grassland soils in Saskatchewan". Nutrient Cycling in Agroecosystems. 82 (2): 137–148. Bibcode:2008NCyAg..82..137N. doi:10.1007/s10705-008-9175-1. ISSN 1573-0867. S2CID 24021984.
  318. ^ Ruseva et al. 2020
  319. ^ IPCC SR15 Ch4 2018, pp. 326–327; Bednar, Obersteiner & Wagner 2019; European Commission, 28 November 2018, p. 188
  320. ^ Bui et al. 2018, p. 1068.
  321. ^ IPCC AR5 SYR 2014, p. 125; Bednar, Obersteiner & Wagner 2019.
  322. ^ IPCC SR15 2018, p. 34
  323. ^ IPCC, 2022: Summary for Policymakers [H.-O. Pörtner, D.C. Roberts, E.S. Poloczanska, K. Mintenbeck, M. Tignor, A. Alegría, M. Craig, S. Langsdorf, S. Löschke, V. Möller, A. Okem (eds.)]. In: Climate Change 2022: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change [H.-O. Pörtner, D.C. Roberts, M. Tignor, E.S. Poloczanska, K. Mintenbeck, A. Alegría, M. Craig, S. Langsdorf, S. Löschke, V. Möller, A. Okem, B. Rama (eds.)]. Cambridge University Press, Cambridge and New York, pp. 3–33, doi:10.1017/9781009325844.001.
  324. ^ IPCC AR5 SYR 2014, p. 17.
  325. ^ IPCC SR15 Ch4 2018, pp. 396–397.
  326. ^ IPCC AR4 WG2 Ch19 2007, p. 796.
  327. ^ UNEP 2018, pp. xii–xiii.
  328. ^ Stephens, Scott A.; Bell, Robert G.; Lawrence, Judy (2018). "Developing signals to trigger adaptation to sea-level rise". Environmental Research Letters. 13 (10). 104004. Bibcode:2018ERL....13j4004S. doi:10.1088/1748-9326/aadf96. ISSN 1748-9326.
  329. ^ Matthews 2018, p. 402.
  330. ^ IPCC SRCCL Ch5 2019, p. 439.
  331. ^ Surminski, Swenja; Bouwer, Laurens M.; Linnerooth-Bayer, Joanne (2016). "How insurance can support climate resilience". Nature Climate Change. 6 (4): 333–334. Bibcode:2016NatCC...6..333S. doi:10.1038/nclimate2979. ISSN 1758-6798.
  332. ^ IPCC SR15 Ch4 2018, pp. 336–337.
  333. ^ "Mangroves against the storm". Shorthand. Retrieved 20 January 2023.
  334. ^ "How marsh grass could help protect us from climate change". World Economic Forum. 24 October 2021. Retrieved 20 January 2023.
  335. ^ Morecroft, Michael D.; Duffield, Simon; Harley, Mike; Pearce-Higgins, James W.; et al. (2019). "Measuring the success of climate change adaptation and mitigation in terrestrial ecosystems". Science. 366 (6471): eaaw9256. doi:10.1126/science.aaw9256. ISSN 0036-8075. PMID 31831643. S2CID 209339286.
  336. ^ Berry, Pam M.; Brown, Sally; Chen, Minpeng; Kontogianni, Areti; et al. (2015). "Cross-sectoral interactions of adaptation and mitigation measures". Climate Change. 128 (3): 381–393. Bibcode:2015ClCh..128..381B. doi:10.1007/s10584-014-1214-0. hdl:10.1007/s10584-014-1214-0. ISSN 1573-1480. S2CID 153904466.
  337. ^ IPCC AR5 SYR 2014, p. 54.
  338. ^ Sharifi, Ayyoob (2020). "Trade-offs and conflicts between urban climate change mitigation and adaptation measures: A literature review". Journal of Cleaner Production. 276: 122813. doi:10.1016/j.jclepro.2020.122813. ISSN 0959-6526. S2CID 225638176.
  339. ^ IPCC AR5 SYR Summary for Policymakers 2014, p. 17, Section 3
  340. ^ IPCC SR15 Ch5 2018, p. 447; United Nations (2017) Resolution adopted by the General Assembly on 6 July 2017, Work of the Statistical Commission pertaining to the 2030 Agenda for Sustainable Development (A/RES/71/313)
  341. ^ IPCC SR15 Ch5 2018, p. 477.
  342. ^ Rauner et al. 2020
  343. ^ Mercure et al. 2018
  344. ^ World Bank, June 2019, p. 12, Box 1
  345. ^ Union of Concerned Scientists, 8 January 2017; Hagmann, Ho & Loewenstein 2019.
  346. ^ Watts et al. 2019, p. 1866
  347. ^ UN Human Development Report 2020, p. 10
  348. ^ International Institute for Sustainable Development 2019, p. iv
  349. ^ ICCT 2019, p. iv; Natural Resources Defense Council, 29 September 2017
  350. ^ National Conference of State Legislators, 17 April 2020; European Parliament, February 2020
  351. ^ Gabbatiss, Josh; Tandon, Ayesha (4 October 2021). "In-depth Q&A: What is 'climate justice'?". Carbon Brief. Retrieved 16 October 2021.
  352. ^ Khalfan, Ashfaq; Lewis, Astrid Nilsson; Aguilar, Carlos; Persson, Jacqueline; Lawson, Max; Dab, Nafkote; Jayoussi, Safa; Acharya, Sunil (November 2023). "Climate Equality: A planet for the 99%" (PDF). Oxfam Digital Repository. Oxfam GB. doi:10.21201/2023.000001. Retrieved 18 December 2023.
  353. ^ Grasso, Marco; Heede, Richard (19 May 2023). "Time to pay the piper: Fossil fuel companies' reparations for climate damages". One Earth. 6 (5): 459–463. Bibcode:2023OEart...6..459G. doi:10.1016/j.oneear.2023.04.012. hdl:10281/416137. S2CID 258809532.
  354. ^ Carbon Brief, 4 Jan 2017.
  355. ^ a b Friedlingstein et al. 2019, Table 7.
  356. ^ UNFCCC, "What is the United Nations Framework Convention on Climate Change?"
  357. ^ UNFCCC 1992, Article 2.
  358. ^ IPCC AR4 WG3 Ch1 2007, p. 97.
  359. ^ EPA 2019.
  360. ^ UNFCCC, "What are United Nations Climate Change Conferences?"
  361. ^ Kyoto Protocol 1997; Liverman 2009, p. 290.
  362. ^ Dessai 2001, p. 4; Grubb 2003.
  363. ^ Liverman 2009, p. 290.
  364. ^ Müller 2010; The New York Times, 25 May 2015; UNFCCC: Copenhagen 2009; EUobserver, 20 December 2009.
  365. ^ UNFCCC: Copenhagen 2009.
  366. ^ Conference of the Parties to the Framework Convention on Climate Change. Copenhagen. 7–18 December 2009. un document= FCCC/CP/2009/L.7. Archived from the original on 18 October 2010. Retrieved 24 October 2010.
  367. ^ Bennett, Paige (2 May 2023). "High-Income Nations Are on Track Now to Meet $100 Billion Climate Pledges, but They're Late". Ecowatch. Retrieved 10 May 2023.
  368. ^ Paris Agreement 2015.
  369. ^ Climate Focus 2015, p. 3; Carbon Brief, 8 October 2018.
  370. ^ Climate Focus 2015, p. 5.
  371. ^ "Status of Treaties, United Nations Framework Convention on Climate Change". United Nations Treaty Collection. Retrieved 13 October 2021.; Salon, 25 September 2019.
  372. ^ Goyal et al. 2019
  373. ^ Yeo, Sophie (10 October 2016). "Explainer: Why a UN climate deal on HFCs matters". Carbon Brief. Retrieved 10 January 2021.
  374. ^ BBC, 1 May 2019; Vice, 2 May 2019.
  375. ^ The Verge, 27 December 2019.
  376. ^ The Guardian, 28 November 2019
  377. ^ Politico, 11 December 2019.
  378. ^ "European Green Deal: Commission proposes transformation of EU economy and society to meet climate ambitions". European Commission. 14 July 2021.
  379. ^ The Guardian, 28 October 2020
  380. ^ "India". Climate Action Tracker. 15 September 2021. Retrieved 3 October 2021.
  381. ^ Do, Thang Nam; Burke, Paul J. (2023). "Phasing out coal power in a developing country context: Insights from Vietnam". Energy Policy. 176 (May 2023 113512): 113512. Bibcode:2023EnPol.17613512D. doi:10.1016/j.enpol.2023.113512. hdl:1885/286612. S2CID 257356936.
  382. ^ UN NDC Synthesis Report 2021, pp. 4–5; UNFCCC Press Office (26 February 2021). "Greater Climate Ambition Urged as Initial NDC Synthesis Report Is Published". Retrieved 21 April 2021.
  383. ^ Stover 2014.
  384. ^ Dunlap & McCright 2011, pp. 144, 155; Björnberg et al. 2017
  385. ^ Oreskes & Conway 2010; Björnberg et al. 2017
  386. ^ O'Neill & Boykoff 2010; Björnberg et al. 2017
  387. ^ a b Björnberg et al. 2017
  388. ^ Dunlap & McCright 2015, p. 308.
  389. ^ Dunlap & McCright 2011, p. 146.
  390. ^ Harvey et al. 2018
  391. ^ "Public perceptions on climate change" (PDF). PERITIA Trust EU – The Policy Institute of King's College London. June 2022. p. 4. Archived (PDF) from the original on 15 July 2022.
  392. ^ Powell, James (20 November 2019). "Scientists Reach 100% Consensus on Anthropogenic Global Warming". Bulletin of Science, Technology & Society. 37 (4): 183–184. doi:10.1177/0270467619886266. S2CID 213454806.
  393. ^ Myers, Krista F.; Doran, Peter T.; Cook, John; Kotcher, John E.; Myers, Teresa A. (20 October 2021). "Consensus revisited: quantifying scientific agreement on climate change and climate expertise among Earth scientists 10 years later". Environmental Research Letters. 16 (10): 104030. Bibcode:2021ERL....16j4030M. doi:10.1088/1748-9326/ac2774. S2CID 239047650.
  394. ^ a b Weart "The Public and Climate Change (since 1980)"
  395. ^ Newell 2006, p. 80; Yale Climate Connections, 2 November 2010
  396. ^ Pew 2015, p. 10.
  397. ^ a b Pew 2020.
  398. ^ Pew 2015, p. 15.
  399. ^ Yale 2021, p. 7.
  400. ^ Yale 2021, p. 9; UNDP 2021, p. 15.
  401. ^ Smith & Leiserowitz 2013, p. 943.
  402. ^ Gunningham 2018.
  403. ^ The Guardian, 19 March 2019; Boulianne, Lalancette & Ilkiw 2020.
  404. ^ Deutsche Welle, 22 June 2019.
  405. ^ Connolly, Kate (29 April 2021). "'Historic' German ruling says climate goals not tough enough". The Guardian. Retrieved 1 May 2021.
  406. ^ Setzer & Byrnes 2019.
  407. ^ "Coal Consumption Affecting Climate". Rodney and Otamatea Times, Waitemata and Kaipara Gazette. Warkworth, New Zealand. 14 August 1912. p. 7. Text was earlier published in Popular Mechanics, March 1912, p. 341.
  408. ^ Nord, D.C. (2020). Nordic Perspectives on the Responsible Development of the Arctic: Pathways to Action. Springer Polar Sciences. Springer International Publishing. p. 51. ISBN 978-3-030-52324-4. Retrieved 11 March 2023.
  409. ^ Mukherjee, A.; Scanlon, B.R.; Aureli, A.; Langan, S.; Guo, H.; McKenzie, A.A. (2020). Global Groundwater: Source, Scarcity, Sustainability, Security, and Solutions. Elsevier Science. p. 331. ISBN 978-0-12-818173-7. Retrieved 11 March 2023.
  410. ^ von Humboldt, A.; Wulf, A. (2018). Selected Writings of Alexander von Humboldt: Edited and Introduced by Andrea Wulf. Everyman's Library Classics Series. Knopf Doubleday Publishing Group. p. 10. ISBN 978-1-101-90807-5. Retrieved 11 March 2023.
  411. ^ Erdkamp, P.; Manning, J.G.; Verboven, K. (2021). Climate Change and Ancient Societies in Europe and the Near East: Diversity in Collapse and Resilience. Palgrave Studies in Ancient Economies. Springer International Publishing. p. 6. ISBN 978-3-030-81103-7. Retrieved 11 March 2023.
  412. ^ Archer & Pierrehumbert 2013, pp. 10–14
  413. ^ Foote, Eunice (November 1856). "Circumstances affecting the Heat of the Sun's Rays". The American Journal of Science and Arts. 22: 382–383. Retrieved 31 January 2016 – via Google Books.
  414. ^ Huddleston 2019
  415. ^ Tyndall 1861.
  416. ^ Archer & Pierrehumbert 2013, pp. 39–42; Fleming 2008, Tyndall
  417. ^ Lapenis 1998.
  418. ^ a b c Weart "The Carbon Dioxide Greenhouse Effect"; Fleming 2008, Arrhenius
  419. ^ Callendar 1938; Fleming 2007.
  420. ^ Cook, John; Oreskes, Naomi; Doran, Peter T.; Anderegg, William R. L.; et al. (2016). "Consensus on consensus: a synthesis of consensus estimates on human-caused global warming". Environmental Research Letters. 11 (4): 048002. Bibcode:2016ERL....11d8002C. doi:10.1088/1748-9326/11/4/048002. hdl:1983/34949783-dac1-4ce7-ad95-5dc0798930a6.
  421. ^ a b Powell, James (20 November 2019). "Scientists Reach 100% Consensus on Anthropogenic Global Warming". Bulletin of Science, Technology & Society. 37 (4): 183–184. doi:10.1177/0270467619886266. S2CID 213454806. Retrieved 15 November 2020.
  422. ^ a b c Lynas, Mark; Houlton, Benjamin Z; Perry, Simon (2021). "Greater than 99% consensus on human caused climate change in the peer-reviewed scientific literature". Environmental Research Letters. 16 (11): 114005. Bibcode:2021ERL....16k4005L. doi:10.1088/1748-9326/ac2966. ISSN 1748-9326. S2CID 239032360.
  423. ^ Myers, Krista F.; Doran, Peter T.; Cook, John; Kotcher, John E.; Myers, Teresa A. (20 October 2021). "Consensus revisited: quantifying scientific agreement on climate change and climate expertise among Earth scientists 10 years later". Environmental Research Letters. 16 (10): 104030. Bibcode:2021ERL....16j4030M. doi:10.1088/1748-9326/ac2774. S2CID 239047650.
  424. ^ Weart "Suspicions of a Human-Caused Greenhouse (1956–1969)"
  425. ^ Weart 2013, p. 3567.
  426. ^ Royal Society 2005.
  427. ^ National Academies 2008, p. 2; Oreskes 2007, p. 68; Gleick, 7 January 2017
  428. ^ Joint statement of the G8+5 Academies (2009); Gleick, 7 January 2017.

Sources

 This article incorporates text from a free content work. Licensed under CC BY-SA 3.0 (license statement/permission). Text taken from The status of women in agrifood systems – Overview​, FAO, FAO.

IPCC reports

Fourth Assessment Report

Fifth Assessment report

Special Report: Global Warming of 1.5 °C

Special Report: Climate change and Land

Special Report: The Ocean and Cryosphere in a Changing Climate

Sixth Assessment Report

Other peer-reviewed sources

Books, reports and legal documents

Non-technical sources

External links

Listen to this article (1 hour and 16 minutes)
Spoken Wikipedia icon
This audio file was created from a revision of this article dated 30 October 2021 (2021-10-30), and does not reflect subsequent edits.