Climate change in the Arctic

Due to climate change in the Arctic, this polar region is expected to become "profoundly different" by 2050.[1]: 2321  The speed of change is "among the highest in the world",[1]: 2321  with the rate of warming being 3-4 times faster than the global average.[2][3][4][5] This warming has already resulted in the profound Arctic sea ice decline, the accelerating melting of the Greenland ice sheet and the thawing of the permafrost landscape.[1]: 2321 [6] These ongoing transformations are expected to be irreversible for centuries or even millennia.[1]: 2321 

Average decadal extent and area of the Arctic Ocean sea ice since 1979.
July 2012 melting event in Greenland
2020 Siberia heatwave
Coastal erosion caused by permafrost thaw in Alaska
Arctic sea ice extent and area have declined every decade since the start of start of satellite observations in 1979: Greenland ice sheet had experienced a "massive melting event" in 2012, which reoccurred in 2019 and 2021; Satellite image of the extremely anomalous 2020 Siberian heatwave; Permafrost thaw is leading to severe erosion, like in this coastal location in Alaska

Natural life in the Arctic is affected greatly. As the tundra warms, its soil becomes more hospitable to earthworms and larger plants,[7] and the boreal forests spread to the north - yet this also makes the landscape more prone to wildfires, which take longer to recover from than in the other regions. Beavers also take advantage of this warming to colonize the Arctic rivers, and their dams contributing to methane emissions due to the increase in stagnant waters. The Arctic Ocean has experienced a large increase in the marine primary production as warmer waters and less shade from sea ice benefit phytoplankton.[8] At the same time, it is already less alkaline than the rest of the global ocean, so ocean acidification caused by the increasing CO2 concentrations is more severe, threatening some forms of zooplankton such as pteropods.[1]: 2328 

The Arctic Ocean is expected to see its first ice-free events in the near future - most likely before 2050, and potentially in the late 2020s or early 2030s.[9] This would have no precedent in the last 700,000 years.[10][11] Some sea ice regrows every Arctic winter, but such events are expected to occur more and more frequently as the warming increases. This has great implications for the fauna species which are dependent on sea ice, such as polar bears. For humans, trade routes across the ocean will become more convenient. Yet, multiple countries have infrastructure in the Arctic which is worth billions of dollars, and it is threatened with collapse as the underlying permafrost thaws. The Arctic's indigenous people have a long relationship with its icy conditions, and face the loss of their cultural heritage.

Further, there are numerous implications which go beyond the Arctic region. Sea ice loss not only enhances warming in the Arctic but also adds to global temperature increase through the ice-albedo feedback. Permafrost thaw results in emissions of CO2 and methane that are comparable to those of major countries. Greenland melting is a significant contributor to global sea level rise. If the warming exceeds - or thereabouts, there is a significant risk of the entire ice sheet being lost over an estimated 10,000 years, adding up to to global sea levels. Warming in the Arctic may affect the stability of the jet stream, and thus the extreme weather events in midlatitude regions, but there is only "low confidence" in that hypothesis.

Impacts on the physical environment edit

Warming edit

 
The image above shows where average air temperatures (October 2010 – September 2011) were up to 2 degrees Celsius above (red) or below (blue) the long-term average (1981–2010).

The period of 1995–2005 was the warmest decade in the Arctic since at least the 17th century, with temperatures 2 °C (3.6 °F) above the 1951–1990 average.[12] Alaska and western Canada's temperature rose by 3 to 4 °C (5.40 to 7.20 °F) during that period.[13] 2013 research has shown that temperatures in the region haven't been as high as they currently are since at least 44,000 years ago and perhaps as long as 120,000 years ago.[14][15] Since 2013, Arctic annual mean surface air temperature (SAT) has been at least 1 °C (1.8 °F) warmer than the 1981-2010 mean.

In 2016, there were extreme anomalies from January to February with the temperature in the Arctic being estimated to be between 4–5.8 °C (7.2–10.4 °F) more than it was between 1981 and 2010.[16] In 2020, mean SAT was 1.9 °C (3.4 °F) warmer than the 1981–2010 average.[17] On 20 June 2020, for the first time, a temperature measurement was made inside the Arctic Circle of 38 °C, more than 100 °F. This kind of weather was expected in the region only by 2100. In March, April and May the average temperature in the Arctic was 10 °C (18.0 °F) higher than normal.[18][19] This heat wave, without human – induced warming, could happen only one time in 80,000 years, according to an attribution study published in July 2020. It is the strongest link of a weather event to anthropogenic climate change that had been ever found, for now.[20]

Arctic amplification edit

 
Potential regional warming caused by the loss of all land ice outside of East Antarctica, and by the disappearance of Arctic sea ice every year starting from June. While plausible, consistent sea ice loss would likely require relatively high warming, and the loss of all ice in Greenland would require multiple millennia.

Snow– and ice–albedo feedback have a substantial effect on regional temperatures. In particular, the presence of ice cover and sea ice makes the North Pole and the South Pole colder than they would have been without it.[21] Consequently, recent Arctic sea ice decline is one of the primary factors behind the Arctic warming nearly four times faster than the global average since 1979 (the year when continuous satellite readings of the Arctic sea ice began), in a phenomenon known as Arctic amplification.[22]

Modelling studies show that strong Arctic amplification only occurs during the months when significant sea ice loss occurs, and that it largely disappears when the simulated ice cover is held fixed.[23] Conversely, the high stability of ice cover in Antarctica, where the thickness of the East Antarctic ice sheet allows it to rise nearly 4 km above the sea level, means that this continent has experienced very little net warming over the past seven decades, most of which was concentrated in West Antarctica.[24][25][26] Ice loss in the Antarctic and its contribution to sea level rise is instead driven overwhelmingly by the warming of the Southern Ocean, which had absorbed 35–43% of the total heat taken up by all oceans between 1970 and 2017.[27]

Ice–albedo feedback also has a smaller, but still notable effect on the global temperatures. Arctic sea ice decline between 1979 and 2011 is estimated to have been responsible for 0.21 watts per square meter (W/m2) of radiative forcing, which is equivalent to a quarter of radiative forcing from CO2[28] increases over the same period. When compared to cumulative increases in greenhouse gas radiative forcing since the start of the Industrial Revolution, it is equivalent to the estimated 2019 radiative forcing from nitrous oxide (0.21 W/m2), nearly half of 2019 radiative forcing from methane (0.54 W/m2) and 10% of the cumulative CO2 increase (2.16 W/m2).[29] Between 1992 and 2015, this effect was partly offset by the growth in sea ice cover around Antarctica, which produced cooling of about 0.06 W/m2 per decade. However, Antarctic sea ice had also begun to decline afterwards, and the combined role of changes in ice cover between 1992 and 2018 is equivalent to 10% of all the anthropogenic greenhouse gas emissions.[30]
 
The dark ocean surface reflects only 6 percent of incoming solar radiation, while sea ice reflects 50 to 70 percent.[31]

The Arctic was historically described as warming twice as fast as the global average,[32] but this estimate was based on older observations which missed the more recent acceleration. By 2021, enough data was available to show that the Arctic had warmed three times as fast as the globe - 3.1°C between 1971 and 2019, as opposed to the global warming of 1°C over the same period.[33] Moreover, this estimate defines the Arctic as everything above 60th parallel north, or a full third of the Northern Hemisphere: in 2021–2022, it was found that since 1979, the warming within the Arctic Circle itself (above the 66th parallel) has been nearly four times faster than the global average.[34][35] Within the Arctic Circle itself, even greater Arctic amplification occurs in the Barents Sea area, with hotspots around West Spitsbergen Current: weather stations located on its path record decadal warming up to seven times faster than the global average.[36][37] This has fuelled concerns that unlike the rest of the Arctic sea ice, ice cover in the Barents Sea may permanently disappear even around 1.5 degrees of global warming.[38][39]

The acceleration of Arctic amplification has not been linear: a 2022 analysis found that it occurred in two sharp steps, with the former around 1986, and the latter after 2000.[40] The first acceleration is attributed to the increase in anthropogenic radiative forcing in the region, which is in turn likely connected to the reductions in stratospheric sulfur aerosols pollution in Europe in the 1980s in order to combat acid rain. Since sulphate aerosols have a cooling effect, their absence is likely to have increased Arctic temperatures by up to 0.5 degrees Celsius.[41][42] The second acceleration has no known cause,[33] which is why it did not show up in any climate models. It is likely to be an example of multi-decadal natural variability, like the suggested link between Arctic temperatures and Atlantic Multi-decadal Oscillation (AMO),[43] in which case it can be expected to reverse in the future. However, even the first increase in Arctic amplification was only accurately simulated by a fraction of the current CMIP6 models.[40]

Precipitation edit

An observed impact of climate change is aa strong increase in the number of lightnings in the Arctic. Lightnings increase the risk for wildfires.[44] rise of 1.5 degrees in global temperature from the pre-industrial level could change the type of precipitation in the Arctic from snow to rain in summer and autumn.[45]

Cryosphere loss edit

 
On average, climate change has lowered the thickness of land ice with every year, and reduced the extent of sea ice cover.[46]

Sea ice edit

 
1870–2009 Northern Hemisphere sea ice extent in million square kilometers. Blue shading indicates the pre-satellite era; data then is less reliable.

Sea ice in the Arctic region has declined in recent decades in area and volume due to climate change. It has been melting more in summer than it refreezes in winter. Global warming, caused by greenhouse gas forcing is responsible for the decline in Arctic sea ice. The decline of sea ice in the Arctic has been accelerating during the early twenty‐first century, with a decline rate of 4.7% per decade (it has declined over 50% since the first satellite records).[47][48][49] It is also thought that summertime sea ice will cease to exist sometime during the 21st century.[50]

The region is at its warmest in at least 4,000 years[51] and the Arctic-wide melt season has lengthened at a rate of five days per decade (from 1979 to 2013), dominated by a later autumn freeze-up.[52] The IPCC Sixth Assessment Report (2021) stated that Arctic sea ice area will likely drop below 1 million km2 in at least some Septembers before 2050.[53] In September 2020, the US National Snow and Ice Data Center reported that the Arctic sea ice in 2020 had melted to an extent of 3.74 million km2, its second-smallest extent since records began in 1979.[54] Earth lost 28 trillion tonnes of ice between 1994 and 2017, with Arctic sea ice accounting for 7.6 trillion tonnes of this loss. The rate of ice loss has risen by 57% since the 1990s.[55]

Greenland ice sheet edit

 
2023 projections of how much the Greenland ice sheet may shrink from its present extent by the year 2300 under the worst possible climate change scenario (upper half) and of how much faster its remaining ice will be flowing in that case (lower half)

Greenland has had major glaciers and ice caps for at least 18 million years,[56] but a single ice sheet first covered most of the island some 2.6 million years ago.[57] Since then, it has both grown[58][59] and contracted significantly.[60][61][62] The oldest known ice on Greenland is about 1 million years old.[63] Due to anthropogenic greenhouse gas emissions, the ice sheet is now the warmest it has been in the past 1000 years,[64] and is losing ice at the fastest rate in at least the past 12,000 years.[65]

Every summer, parts of the surface melt and ice cliffs calve into the sea. Normally the ice sheet would be replenished by winter snowfall,[66] but due to global warming the ice sheet is melting two to five times faster than before 1850,[67] and snowfall has not kept up since 1996.[68] If the Paris Agreement goal of staying below 2 °C (3.6 °F) is achieved, melting of Greenland ice alone would still add around 6 cm (2+12 in) to global sea level rise by the end of the century. If there are no reductions in emissions, melting would add around 13 cm (5 in) by 2100,[69]: 1302  with a worst-case of about 33 cm (13 in).[70] For comparison, melting has so far contributed 1.4 cm (12 in) since 1972,[71] while sea level rise from all sources was 15–25 cm (6–10 in)) between 1901 and 2018.[72]: 5 

A narrated tour about Greenland's ice sheet.
If all 2,900,000 cubic kilometres (696,000 cu mi) of the ice sheet were to melt, it would increase global sea levels by ~7.4 m (24 ft).[73] Global warming between 1.7 °C (3.1 °F) and 2.3 °C (4.1 °F) would likely make this melting inevitable.[74] However, 1.5 °C (2.7 °F) would still cause ice loss equivalent to 1.4 m (4+12 ft) of sea level rise,[75] and more ice will be lost if the temperatures exceed that level before declining.[74] If global temperatures continue to rise, the ice sheet will likely disappear within 10,000 years.[76][77] At very high warming, its future lifetime goes down to around 1,000 years.[70]

Biological environment edit

Impacts on Arctic flora edit

 
Western Hemisphere Arctic Vegetation Index Trend
 
Eastern Hemisphere Vegetation Index Trend

Climate change is expected to have a strong effect on the Arctic's flora, some of which is already being seen.[78] NASA and NOAA have continuously monitored Arctic vegetation with satellite instruments such as Moderate Resolution Imaging Spectroradiometer (MODIS) and Advanced very-high-resolution radiometer (AVHRR).[79] Their data allows scientists to calculate so-called "Arctic greening" and "Arctic browning".[80] From 1985 to 2016, greening has occurred in 37.3% of all sites sampled in the tundra, whereas browning was observed only in 4.7% of the sites - typically the ones that were still experiencing cooling and drying, as opposed to warming and wettening for the rest.[81]

This expansion of vegetation in the Arctic is not equivalent across types of vegetation. A major trend has been from shrub-type plants taking over areas previously dominated by moss and lichens. This change contributes to the consideration that the tundra biome is currently experiencing the most rapid change of any terrestrial biomes on the planet.[82][83] The direct impact on mosses and lichens is unclear as there exist very few studies at species level, but climate change is more likely to cause increased fluctuation and more frequent extreme events.[84] While shrubs may increase in range and biomass, warming may also cause a decline in cushion plants such as moss campion, and since cushion plants act as facilitator species across trophic levels and fill important ecological niches in several environments, this could cause cascading effects in these ecosystems that could severely affect the way in which they function and are structured.[85]

The expansion of these shrubs can also have strong effects on other important ecological dynamics, such as the albedo effect.[86] These shrubs change the winter surface of the tundra from undisturbed, uniform snow to mixed surface with protruding branches disrupting the snow cover,[87] this type of snow cover has a lower albedo effect, with reductions of up to 55%, which contributes to a positive feedback loop on regional and global climate warming.[87] This reduction of the albedo effect means that more radiation is absorbed by plants, and thus, surface temperatures increase, which could disrupt current surface-atmosphere energy exchanges and affect thermal regimes of permafrost.[87] Carbon cycling is also being affected by these changes in vegetation, as parts of the tundra increase their shrub cover, they behave more like boreal forests in terms of carbon cycling.[88] This is speeding up the carbon cycle, as warmer temperatures lead to increased permafrost thawing and carbon release, but also carbon capturing from plants that have increased growth.[88] It is not certain whether this balance will go in one direction or the other, but studies have found that it is more likely that this will eventually lead to increased CO2 in the atmosphere.[88]

However, boreal forests, particularly those in North America, showed a different response to warming. Many boreal forests greened, but the trend was not as strong as it was for tundra of the circumpolar Arctic, mostly characterized by shrub expansion and increased growth.[89] In North America, some boreal forests actually experienced browning over the study period. Droughts, increased forest fire activity, animal behavior, industrial pollution, and a number of other factors may have contributed to browning.[80]


Impacts on terrestrial fauna edit

 
Projected change in polar bear habitat from 2001–2010 to 2041–2050

The northward shift of the subarctic climate zone is allowing animals that are adapted to that climate to move into the far north, where they are replacing species that are more adapted to a pure Arctic climate. Where the Arctic species are not being replaced outright, they are often interbreeding with their southern relations. Among slow-breeding vertebrate species, this usually has the effect of reducing the genetic diversity of the genus. Another concern is the spread of infectious diseases, such as brucellosis or phocine distemper virus, to previously untouched populations. This is a particular danger among marine mammals who were previously segregated by sea ice.[90]

On 3 April 2007, the National Wildlife Federation urged the United States Congress to place polar bears under the Endangered Species Act.[91] Four months later, the United States Geological Survey completed a year-long study[92] which concluded in part that the floating Arctic sea ice will continue its rapid shrinkage over the next 50 years, consequently wiping out much of the polar bear habitat. The bears would disappear from Alaska, but would continue to exist in the Canadian Arctic Archipelago and areas off the northern Greenland coast.[93] Secondary ecological effects are also resultant from the shrinkage of sea ice; for example, polar bears are denied their historic length of seal hunting season due to late formation and early thaw of pack ice.

Similarly, Arctic warming negatively affects the foraging and breeding ecology many other species of arctic marine mammals, such as walruses,[94] seals, foxes or reindeers.[95]  In July 2019, 200 Svalbard reindeer were found starved to death apparently due to low precipitation related to climate change.[96]

In the short-term, climate warming may have neutral or positive effects on the nesting cycle of many Arctic-breeding shorebirds.[97]

Marine ecosystems edit

In terms of aquatic vegetation, reduction of sea ice has boosted the productivity of phytoplankton by about twenty percent over the past thirty years. However, the effect on marine ecosystems is unclear, since the larger types of phytoplankton, which are the preferred food source of most zooplankton, do not appear to have increased as much as the smaller types. So far, Arctic phytoplankton have not had a significant impact on the global carbon cycle.[98] In summer, the melt ponds on young and thin ice have allowed sunlight to penetrate the ice, in turn allowing ice algae to bloom in unexpected concentrations, although it is unknown just how long this phenomenon has been occurring, or what its effect on broader ecological cycles is.[99]

Greenhouse gas emissions from the Arctic edit

Permafrost thaw edit

 
Permafrost thaw ponds on Baffin Island

Permafrost is an important component of hydrological systems and ecosystems within the Arctic landscape.[100] In the Northern Hemisphere the terrestrial permafrost domain comprises around 18 million km2.[101] Within this permafrost region, the total soil organic carbon (SOC) stock is estimated to be 1,460-1,600 Pg (where 1 Pg = 1 billion tons), which constitutes double the amount of carbon currently contained in the atmosphere.[102][103]

As recent warming deepens the active layer subject to permafrost thaw, this exposes formerly stored carbon to biogenic processes which facilitate its entrance into the atmosphere as carbon dioxide and methane.[104] Because carbon emissions from permafrost thaw contribute to the same warming which facilitates the thaw, it is a well-known example of a positive climate change feedback,[105] and because widespread permafrost thaw is effectively irreversible, it is also considered one of tipping points in the climate system.[106]

In the northern circumpolar region, permafrost contains organic matter equivalent to 1400–1650 billion tons of pure carbon, which was built up over thousands of years. This amount equals almost half of all organic material in all soils,[107][104] and it is about twice the carbon content of the atmosphere, or around four times larger than the human emissions of carbon between the start of the Industrial Revolution and 2011.[108] Further, most of this carbon (~1,035 billion tons) is stored in what is defined as the near-surface permafrost, no deeper than 3 metres (9.8 ft) below the surface.[107][104] However, only a fraction of this stored carbon is expected to enter the atmosphere.[109] In general, the volume of permafrost in the upper 3 m of ground is expected to decrease by about 25% per 1 °C (1.8 °F) of global warming,[110]: 1283  yet even under the RCP8.5 scenario associated with over 4 °C (7.2 °F) of global warming by the end of the 21st century,[111] about 5% to 15% of permafrost carbon is expected to be lost "over decades and centuries".[104]
 
Nine probable scenarios of greenhouse gas emissions from permafrost thaw during the 21st century, which show a limited, moderate and intense CO2 and CH4 emission response to low, medium and high-emission Representative Concentration Pathways. The vertical bar uses emissions of selected large countries as a comparison: the right-hand side of the scale shows their cumulative emissions since the start of the Industrial Revolution, while the left-hand side shows each country's cumulative emissions for the rest of the 21st century if they remained unchanged from their 2019 levels.[104]

Altogether, it is expected that cumulative greenhouse gas emissions from permafrost thaw will be smaller than the cumulative anthropogenic emissions, yet still substantial on a global scale, with some experts comparing them to emissions caused by deforestation.[104] The IPCC Sixth Assessment Report estimates that carbon dioxide and methane released from permafrost could amount to the equivalent of 14–175 billion tonnes of carbon dioxide per 1 °C (1.8 °F) of warming.[110]: 1237  For comparison, by 2019, annual anthropogenic emissions of carbon dioxide alone stood around 40 billion tonnes.[110]: 1237  A major review published in the year 2022 concluded that if the goal of preventing 2 °C (3.6 °F) of warming was realized, then the average annual permafrost emissions throughout the 21st century would be equivalent to the year 2019 annual emissions of Russia. Under RCP4.5, a scenario considered close to the current trajectory and where the warming stays slightly below 3 °C (5.4 °F), annual permafrost emissions would be comparable to year 2019 emissions of Western Europe or the United States, while under the scenario of high global warming and worst-case permafrost feedback response, they would approach year 2019 emissions of China.[104]

Fewer studies have attempted to describe the impact directly in terms of warming. A 2018 paper estimated that if global warming was limited to 2 °C (3.6 °F), gradual permafrost thaw would add around 0.09 °C (0.16 °F) to global temperatures by 2100,[112] while a 2022 review concluded that every 1 °C (1.8 °F) of global warming would cause 0.04 °C (0.072 °F) and 0.11 °C (0.20 °F) from abrupt thaw by the year 2100 and 2300. Around 4 °C (7.2 °F) of global warming, abrupt (around 50 years) and widespread collapse of permafrost areas could occur, resulting in an additional warming of 0.2–0.4 °C (0.36–0.72 °F).[106][113]

Black carbon edit

Black carbon deposits (from the combustion of heavy fuel oil (HFO) of Arctic shipping) absorb solar radiation in the atmosphere and strongly reduce the albedo when deposited on snow and ice, thus accelerating the effect of the melting of snow and sea ice.[114] A 2013 study quantified that gas flaring at petroleum extraction sites contributed over 40% of the black carbon deposited in the Arctic.[115][116] 2019 research attributed the majority (56%) of Arctic surface black carbon to emissions from Russia, followed by European emissions, and Asia also being a large source.[117][114] In 2015, research suggested that reducing black carbon emissions and short-lived greenhouse gases by roughly 60 percent by 2050 could cool the Arctic up to 0.2 °C.[118] However, a 2019 study indicates that "Black carbon emissions will continuously rise due to increased shipping activities", specifically fishing vessels.[119]

The number of wildfires in the Arctic Circle has increased. In 2020, Arctic wildfire CO2 emissions broke a new record, peaking at 244 megatonnes of carbon dioxide emitted.[120]  This is due to the burning of peatlands, carbon-rich soils that originate from the accumulation of waterlogged plants which are mostly found at Arctic latitudes.[120] These peatlands are becoming more likely to burn as temperatures increase, but their own burning and releasing of CO2 contributes to their own likelihood of burning in a positive feedback loop.[120]The smoke from wildfires defined as "brown carbon" also increases arctic warming, with its warming effect is around 30% that of black carbon. As wildfires increases with warming this create a positive feedback loop.[121]

Methane clathrate deposits edit

 
Methane clathrate is released as gas into the surrounding water column or soils when ambient temperature increases
The clathrate gun hypothesis is a proposed explanation for the periods of rapid warming during the Quaternary. The hypothesis is that changes in fluxes in upper intermediate waters in the ocean caused temperature fluctuations that alternately accumulated and occasionally released methane clathrate on upper continental slopes. This would have had an immediate impact on the global temperature, as methane is a much more powerful greenhouse gas than carbon dioxide. Despite its atmospheric lifetime of around 12 years, methane's global warming potential is 72 times greater than that of carbon dioxide over 20 years, and 25 times over 100 years (33 when accounting for aerosol interactions).[122] It is further proposed that these warming events caused the Bond Cycles and individual interstadial events, such as the Dansgaard–Oeschger interstadials.[123]
In 2018, a perspective piece devoted to tipping points in the climate system suggested that the climate change contribution from methane hydrates would be "negligible" by the end of the century, but could amount to 0.4–0.5 °C (0.72–0.90 °F) on the millennial timescales.[124] In 2021, the IPCC Sixth Assessment Report no longer included methane hydrates in the list of potential tipping points, and says that "it is very unlikely that CH4 emissions from clathrates will substantially warm the climate system over the next few centuries."[125] The report had also linked terrestrial hydrate deposites to gas emission craters discovered in the Yamal Peninsula in Siberia, Russia beginning in July 2014,[126] but noted that since terrestrial gas hydrates predominantly form at a depth below 200 metres, a substantial response within the next few centuries can be ruled out.[125] Likewise, a 2022 assessment of tipping points described methane hydrates as a "threshold-free feedback" rather than a tipping point.[127][128]

Effects on other parts of the world edit

On ocean circulation edit

 
Modelled 21st century warming under the "intermediate" global warming scenario (top). The potential collapse of the subpolar gyre in this scenario (middle). The collapse of the entire Atlantic Meriditional Overturning Circulation (bottom).
The Atlantic meridional overturning circulation (AMOC) is the "main current system in the South and North Atlantic Oceans".[129]: 2238  It is a component of Earth's oceanic circulation system and plays an important role in the climate system. The AMOC includes currents at the surface as well as at great depths in the Atlantic Ocean. These currents are driven by changes in the atmospheric weather as well as by changes in temperature and salinity. They collectively make up one half of the global thermohaline circulation that encompasses the flow of major ocean currents. The other half is the Southern Ocean overturning circulation.[130]
AMOC has not always existed. For much of the Earth's history, overturning circulation in the Northern Hemisphere used to occur in the North Pacific. Paleoclimate evidence shows that the shift from the Pacific to Atlantic overturning circulation had occurred 34 million years ago, at the Eocene-Oligocene transition, when the Arctic-Atlantic gateway had closed.[131] This closure fundamentally changed the thermohaline circulation structure - yet some researchers have suggested that climate change may end up reversing this shift and re-establish the Pacific circulation after the AMOC shuts down.[132][133] This is because it affects the AMOC in two major ways - by making surface waters warmer as an inevitable consequence of Earth's energy imbalance, and by making them less saline due to the addition of large quantities of fresh water from melting ice (mainly from Greenland), and through increasing precipitation over the North Atlantic. Both would increase the difference between the surface and lower layers, and thus make the upwelling and downwelling which drives the circulation more difficult.[134]
Severe weakening of the AMOC may lead to an outright collapse of the circulation, which would not be easily reversible and thus constitute one of the tipping points in the climate system.[135] A collapse would substantially lower the average temperature and amount of rain and snowfall in Europe.[136][137] It would also potentially raise the frequency of extreme weather events and have other severe effects.[138][139] Gold-standard Earth system models indicate that a collapse is unlikely, and would only become plausible if high levels of warming (≥4 °C (7.2 °F))[136] are sustained well after the year 2100.[140][141][142] Some paleoceanographic research seems to support this idea.[143][144] However, certain researchers fear that the complex models are too stable,[145] and lower-complexity projections pointing to an earlier collapse are more accurate.[146][147] One of those projections suggests that AMOC collapse could happen around 2057,[148] but many scientists are skeptical of the claim.[149] Some research also suggests that the Southern Ocean overturning circulation may be more prone to collapse than the AMOC.[150][138]

In 2021, the IPCC Sixth Assessment Report again assessed that the AMOC is very likely to decline within the 21st century, and expressed high confidence that changes to it would be reversible within centuries if the warming was reversed.[151]: 19  Unlike the Fifth Assessment Report, it had only expressed medium confidence rather than high confidence in AMOC avoiding a collapse before the end of the century. This reduction in confidence was likely influenced by several review studies drawing attention to the circulation stability bias within general circulation models,[152][153] as well as simplified ocean modelling studies suggesting that the AMOC may be more vulnerable to abrupt change than what the larger-scale models suggest.[146]

In 2022, an extensive assessment of all potential climate tipping points identified 16 plausible climate tipping points, including a collapse of the AMOC. It suggested that a collapse would most likely be triggered by 4 °C (7.2 °F) of global warming, but that there's enough uncertainty to suggest it could be triggered at warming levels as low as 1.4 °C (2.5 °F), or as high as 8 °C (14 °F). Likewise, it estimates that once AMOC collapse is triggered, it would most likely take place over 50 years, but the entire range is between 15 and 300 years.[136][154] That assessment also treated the collapse of the Northern Subpolar Gyre as a potential separate tipping point, which could occur at between 1.1 °C (2.0 °F) degrees and 3.8 °C (6.8 °F) (although this is only simulated by a fraction of climate models). The most likely figure is 1.8 °C (3.2 °F), and once triggered, the collapse of the gyre would most likely take 10 years from start to end, with a range between 5 and 50 years. The loss of this convection is estimated to lower the global temperature by 0.5 °C (0.90 °F), while the average temperature in Europe decreases by around 3 °C (5.4 °F). There are also substantial impacts on regional precipitation.[136][154]

On mid-latitude weather edit

Since the early 2000s, climate models have consistently identified that global warming will gradually push jet streams poleward. In 2008, this was confirmed by observational evidence, which proved that from 1979 to 2001, the northern jet stream moved northward at an average rate of 2.01 kilometres (1.25 mi) per year, with a similar trend in the Southern Hemisphere jet stream.[155][156] Climate scientists have hypothesized that the jet stream will also gradually weaken as a result of global warming. Trends such as Arctic sea ice decline, reduced snow cover, evapotranspiration patterns, and other weather anomalies have caused the Arctic to heat up faster than other parts of the globe, in what is known as the Arctic amplification. In 2021-2022, it was found that since 1979, the warming within the Arctic Circle has been nearly four times faster than the global average,[157][158] and some hotspots in the Barents Sea area warmed up to seven times faster than the global average.[159][160] While the Arctic remains one of the coldest places on Earth today, the temperature gradient between it and the warmer parts of the globe will continue to diminish with every decade of global warming as the result of this amplification. If this gradient has a strong influence on the jet stream, then it will eventually become weaker and more variable in its course, which would allow more cold air from the polar vortex to leak mid-latitudes and slow the progression of Rossby Waves, leading to more persistent and more extreme weather.

The hypothesis above is closely associated with Jennifer Francis, who had first proposed it in a 2012 paper co-authored by Stephen J. Vavrus.[161] While some paleoclimate reconstructions have suggested that the polar vortex becomes more variable and causes more unstable weather during periods of warming back in 1997,[162] this was contradicted by climate modelling, with PMIP2 simulations finding in 2010 that the Arctic oscillation was much weaker and more negative during the Last Glacial Maximum, and suggesting that warmer periods have stronger positive phase AO, and thus less frequent leaks of the polar vortex air.[163] However, a 2012 review in the Journal of the Atmospheric Sciences noted that "there [has been] a significant change in the vortex mean state over the twenty-first century, resulting in a weaker, more disturbed vortex.",[164] which contradicted the modelling results but fit the Francis-Vavrus hypothesis. Additionally, a 2013 study noted that the then-current CMIP5 tended to strongly underestimate winter blocking trends,[165] and other 2012 research had suggested a connection between declining Arctic sea ice and heavy snowfall during midlatitude winters.[166]

However, because the specific observations are considered short-term observations, there is considerable uncertainty in the conclusions. Climatology observations require several decades to definitively distinguish various forms of natural variability from climate trends.[167] This point was stressed by reviews in 2013[168] and in 2017.[169] A study in 2014 concluded that Arctic amplification significantly decreased cold-season temperature variability over the Northern Hemisphere in recent decades. Cold Arctic air intrudes into the warmer lower latitudes more rapidly today during autumn and winter, a trend projected to continue in the future except during summer, thus calling into question whether winters will bring more cold extremes.[170] A 2019 analysis of a data set collected from 35 182 weather stations worldwide, including 9116 whose records go beyond 50 years, found a sharp decrease in northern midlatitude cold waves since the 1980s.[171]

Moreover, a range of long-term observational data collected during 2010s and published in 2020s now suggests that the intensification of Arctic amplification since the early 2010s was not linked to significant changes on midlatitude atmospheric patterns.[172][173] State-of-the-art modelling research of PAMIP (Polar Amplification Model Intercomparison Project) improved upon the 2010 findings of PMIP2 - it did find that sea ice decline would weaken the jet stream and increase the probability of atmospheric blocking, but the connection was very minor, and typically insignificant next to interannual variability.[174][175] In 2022, a follow-up study found that while the PAMIP average had likely underestimated the weakening caused by sea ice decline by 1.2 to 3 times, even the corrected connection still amounts to only 10% of the jet stream's natural variability.[176]

Impacts on people edit

Territorial claims edit

Growing evidence that global warming is shrinking polar ice has added to the urgency of several nations' Arctic territorial claims in hopes of establishing resource development and new shipping lanes, in addition to protecting sovereign rights.[177]

As ice sea coverage decreases more and more, year on year, Arctic countries (Russia, Canada, Finland, Iceland, Norway, Sweden, the United States and Denmark representing Greenland) are making moves on the geopolitical stage to ensure access to potential new shipping lanes, oil and gas reserves, leading to overlapping claims across the region.[178] However, there is only one single land border dispute in the Arctic, with all others relating to the sea, that is Hans Island.[179]  This small uninhabited island lies in the Nares strait, between Canada's Ellesmere Island and the northern coast of Greenland. Its status comes from its geographical position, right between the equidistant boundaries determined in a 1973 treaty between Canada and Denmark.[179]  Even though both countries have acknowledged the possibility of splitting the island, no agreement on the island has been reached, with both nations still claiming it for themselves.[179]

There is more activity in terms of maritime boundaries between countries, where overlapping claims for internal waters, territorial seas and particularly Exclusive Economic Zones (EEZs) can cause frictions between nations. Currently, official maritime borders have an unclaimed triangle of international waters lying between them, that is at the centerpoint of international disputes.[178]

This unclaimed land can be obtainable by submitting a claim to the United Nations Convention on the Law of the Sea, these claims can be based on geological evidence that continental shelves extend beyond their current maritime borders and into international waters.[178]

Some overlapping claims are still pending resolution by international bodies, such as a large portion containing the north pole that is both claimed by Denmark and Russia, with some parts of it also contested by Canada.[178] Another example is that of the Northwest Passage, globally recognized as international waters, but technically in Canadian waters.[178] This has led to Canada wanting to limit the number of ships that can go through for environmental reasons but the United States disputes that they have the authority to do so, favouring unlimited passage of vessels.[178]

Navigation edit

The Transpolar Sea Route is a future Arctic shipping lane running from the Atlantic Ocean to the Pacific Ocean across the center of the Arctic Ocean. The route is also sometimes called Trans-Arctic Route. In contrast to the Northeast Passage (including the Northern Sea Route) and the North-West Passage it largely avoids the territorial waters of Arctic states and lies in international high seas.[180]

Governments and private industry have shown a growing interest in the Arctic.[181] Major new shipping lanes are opening up: the northern sea route had 34 passages in 2011 while the Northwest Passage had 22 traverses, more than any time in history.[182] Shipping companies may benefit from the shortened distance of these northern routes. Access to natural resources will increase, including valuable minerals and offshore oil and gas.[183] Finding and controlling these resources will be difficult with the continually moving ice.[183] Tourism may also increase as less sea ice will improve safety and accessibility to the Arctic.[183]

The melting of Arctic ice caps is likely to increase traffic in and the commercial viability of the Northern Sea Route. One study, for instance, projects, "remarkable shifts in trade flows between Asia and Europe, diversion of trade within Europe, heavy shipping traffic in the Arctic and a substantial drop in Suez traffic. Projected shifts in trade also imply substantial pressure on an already threatened Arctic ecosystem."[184]

Infrastructure edit

 
Map of likely risk to infrastructure from permafrost thaw expected to occur by 2050.

As of 2021, there are 1162 settlements located directly atop the Arctic permafrost, which host an estimated 5 million people. By 2050, permafrost layer below 42% of these settlements is expected to thaw, affecting all their inhabitants (currently 3.3 million people).[185] Consequently, a wide range of infrastructure in permafrost areas is threatened by the thaw.[186][187]: 236  By 2050, it's estimated that nearly 70% of global infrastructure located in the permafrost areas would be at high risk of permafrost thaw, including 30–50% of "critical" infrastructure. The associated costs could reach tens of billions of dollars by the second half of the century.[188] Reducing greenhouse gas emissions in line with the Paris Agreement is projected to stabilize the risk after mid-century; otherwise, it'll continue to worsen.[189]

In Alaska alone, damages to infrastructure by the end of the century would amount to $4.6 billion (at 2015 dollar value) if RCP8.5, the high-emission climate change scenario, were realized. Over half stems from the damage to buildings ($2.8 billion), but there's also damage to roads ($700 million), railroads ($620 million), airports ($360 million) and pipelines ($170 million).[190] Similar estimates were done for RCP4.5, a less intense scenario which leads to around 2.5 °C (4.5 °F) by 2100, a level of warming similar to the current projections.[191] In that case, total damages from permafrost thaw are reduced to $3 billion, while damages to roads and railroads are lessened by approximately two-thirds (from $700 and $620 million to $190 and $220 million) and damages to pipelines are reduced more than ten-fold, from $170 million to $16 million. Unlike the other costs stemming from climate change in Alaska, such as damages from increased precipitation and flooding, climate change adaptation is not a viable way to reduce damages from permafrost thaw, as it would cost more than the damage incurred under either scenario.[190]

In Canada, Northwest Territories have a population of only 45,000 people in 33 communities, yet permafrost thaw is expected to cost them $1.3 billion over 75 years, or around $51 million a year. In 2006, the cost of adapting Inuvialuit homes to permafrost thaw was estimated at $208/m2 if they were built at pile foundations, and $1,000/m2 if they didn't. At the time, the average area of a residential building in the territory was around 100 m2. Thaw-induced damage is also unlikely to be covered by home insurance, and to address this reality, territorial government currently funds Contributing Assistance for Repairs and Enhancements (CARE) and Securing Assistance for Emergencies (SAFE) programs, which provide long- and short-term forgivable loans to help homeowners adapt. It is possible that in the future, mandatory relocation would instead take place as the cheaper option. However, it would effectively tear the local Inuit away from their ancestral homelands. Right now, their average personal income is only half that of the median NWT resident, meaning that adaptation costs are already disproportionate for them.[192]

By 2022, up to 80% of buildings in some Northern Russia cities had already experienced damage.[188] By 2050, the damage to residential infrastructure may reach $15 billion, while total public infrastructure damages could amount to 132 billion.[193] This includes oil and gas extraction facilities, of which 45% are believed to be at risk.[189]

Toxic pollution edit

 
Graphical representation of leaks from various toxic hazards caused by the thaw of formerly stable permafrost.[194]

For much of the 20th century, it was believed that permafrost would "indefinitely" preserve anything buried there, and this made deep permafrost areas popular locations for hazardous waste disposal. In places like Canada's Prudhoe Bay oil field, procedures were developed documenting the "appropriate" way to inject waste beneath the permafrost. This means that as of 2023, there are ~4500 industrial facilities in the Arctic permafrost areas which either actively process or store hazardous chemicals. Additionally, there are between 13,000 and 20,000 sites which have been heavily contaminated, 70% of them in Russia, and their pollution is currently trapped in the permafrost. About a fifth of both the industrial and the polluted sites (1000 and 2200–4800) are expected to start thawing in the future even if the warming does not increase from its 2020 levels. Only about 3% more sites would start thawing between now and 2050 under the climate change scenario consistent with the Paris Agreement goals, RCP2.6, but by 2100, about 1100 more industrial facilities and 3500 to 5200 contaminated sites are expected to start thawing even then. Under the very high emission scenario RCP8.5, 46% of industrial and contaminated sites would start thawing by 2050, and virtually all of them would be affected by the thaw by 2100.[194] Organochlorines and other persistent organic pollutants are of a particular concern, due to their potential to repeatedly reach local communities after their re-release through biomagnification in fish. At worst, future generations born in the Arctic would enter life with weakened immune systems due to pollutants accumulating across generations.[195]

 
Distribution of toxic substances currently located at various permafrost sites in Alaska, by sector. The number of fish skeletons represents the toxicity of each substance.[194]

A notable example of pollution risks associated with permafrost was the 2020 Norilsk oil spill, caused by the collapse of diesel fuel storage tank at Norilsk-Taimyr Energy's thermal power plant No. 3. It spilled 6,000 tonnes of fuel into the land and 15,000 into the water, polluting Ambarnaya, Daldykan and many smaller rivers on Taimyr Peninsula, even reaching lake Pyasino, which is a crucial water source in the area. State of emergency at the federal level was declared.[196][197] The event has been described as the second-largest oil spill in modern Russian history.[198][199]

Another issue associated with permafrost thaw is the release of natural mercury deposits. An estimated 800,000 tons of mercury are frozen in the permafrost soil. According to observations, around 70% of it is simply taken up by vegetation after the thaw.[195] However, if the warming continues under RCP8.5, then permafrost emissions of mercury into the atmosphere would match the current global emissions from all human activities by 2200. Mercury-rich soils also pose a much greater threat to humans and the environment if they thaw near rivers. Under RCP8.5, enough mercury will enter the Yukon River basin by 2050 to make its fish unsafe to eat under the EPA guidelines. By 2100, mercury concentrations in the river will double. Contrastingly, even if mitigation is limited to RCP4.5 scenario, mercury levels will increase by about 14% by 2100, and will not breach the EPA guidelines even by 2300.[200]
In 2021, research claimed that there must be mineral deposits of mercury (a highly toxic heavy metal) beneath the southwestern ice sheet, because of the exceptional concentrations in meltwater entering the local fjords. If confirmed, these concentrations would have equalled up to 10% of mercury in all of the world's rivers.[201][202] In 2024, a follow-up study found only "very low" concentrations in meltwater from 21 locations. It concluded that the 2021 findings were best explained by accidental sample contamination with mercury(II) chloride, used by the first team of researchers as a reagent.[203] However, there is still a risk of toxic waste being released from Camp Century, formerly a United States military site built to carry nuclear weapons for the Project Iceworm. The project was cancelled, but the site was never cleaned up, and it now threatens to pollute the meltwater with nuclear waste, 20,000 liters of chemical waste and 24 million liters of untreated sewage as the melt progresses.[204][205]

Impacts on indigenous peoples edit

As climate change speeds up, it is having more and more of a direct impact on societies around the world. This is particularly true of people that live in the Arctic, where increases in temperature are occurring at faster rates than at other latitudes in the world, and where traditional ways of living, deeply connected with the natural arctic environment are at particular risk of environmental disruption caused by these changes.[183]

The warming of the atmosphere and ecological changes that come alongside it presents challenges to local communities such as the Inuit. Hunting, which is a major way of survival for some small communities, will be changed with increasing temperatures.[206] The reduction of sea ice will cause certain species populations to decline or even become extinct.[183] Inuit communities are deeply reliant on seal hunting, which is dependent on sea ice flats, where seals are hunted.[207]

Unsuspected changes in river and snow conditions will cause herds of animals, including reindeer, to change migration patterns, calving grounds, and forage availability.[183] In good years, some communities are fully employed by the commercial harvest of certain animals.[206] The harvest of different animals fluctuates each year and with the rise of temperatures it is likely to continue changing and creating issues for Inuit hunters, as unpredictability and disruption of ecological cycles further complicate life in these communities, which already face significant problems, such as Inuit communities being the poorest and most unemployed of North America.[207]

Other forms of transportation in the Arctic have seen negative impacts from the current warming, with some transportation routes and pipelines on land being disrupted by the melting of ice.[183] Many Arctic communities rely on frozen roadways to transport supplies and travel from area to area.[183] The changing landscape and unpredictability of weather is creating new challenges in the Arctic.[208] Researchers have documented historical and current trails created by the Inuit in the Pan Inuit Trails Atlas, finding that the change in sea ice formation and breakup has resulted in changes to the routes of trails created by the Inuit.[209]

Adaptation edit

Research edit

Individual countries within the Arctic zone, Canada, Denmark (Greenland), Finland, Iceland, Norway, Russia, Sweden, and the United States (Alaska) conduct independent research through a variety of organizations and agencies, public and private, such as Russia's Arctic and Antarctic Research Institute. Countries who do not have Arctic claims, but are close neighbors, conduct Arctic research as well, such as the Chinese Arctic and Antarctic Administration (CAA). The United States's National Oceanic and Atmospheric Administration (NOAA) produces an Arctic Report Card annually, containing peer-reviewed information on recent observations of environmental conditions in the Arctic relative to historical records.[210][211] International cooperative research between nations has also become increasingly important:

The 2021 Arctic Monitoring and Assessment Programme (AMAP) report by an international team of more than 60 experts, scientists, and indigenous knowledge keepers from Arctic communities, was prepared from 2019 to 2021.[215]: vii  It is a follow-up report of the 2017 assessment, "Snow, Water, Ice and Permafrost in the Arctic" (SWIPA).[215]: vii  The 2021 IPCC AR6 WG1 Technical Report confirmed that "[o]bserved and projected warming" were ""strongest in the Arctic".[216]: 29  According to an 11 August 2022 article published in Nature, there have been numerous reports that the Arctic is warming from twice to three times as fast as the global average since 1979, but the co-authors cautioned that the recent report of the "four-fold Arctic warming ratio" was potentially an "extremely unlikely event".[217] The annual mean Arctic Amplification (AA) index had "reached values exceeding four" from c. 2002 through 2022, according to a July 2022 article in Geophysical Research Letters.[218]: 1 [219]

The 14 December 2021 16th Arctic Report Card produced by the United States's National Oceanic and Atmospheric Administration (NOAA) and released annually, examined the "interconnected physical, ecological and human components" of the circumpolar Arctic.[220][45] The report said that the 12 months between October 2020 and September 2021 were the "seventh warmest over Arctic land since the record began in 1900".[220] The 2017 report said that the melting ice in the warming Arctic was unprecedented in the past 1500 years.[210][211] NOAA's State of the Arctic Reports, starting in 2006, updates some of the records of the original 2004 and 2005 Arctic Climate Impact Assessment (ACIA) reports by the intergovernmental Arctic Council and the non-governmental International Arctic Science Committee.[221]

A 2022 United Nations Environment Programme (UNEP) report "Spreading Like Wildfire: The Rising Threat Of Extraordinary Landscape Fires" said that smoke from wildfires around the world created a positive feedback loop that is a contributing factor to Arctic melting.[222][121] The 2020 Siberian heatwave was "associated with extensive burning in the Arctic Circle".[222]: 36  Report authors said that this extreme heat event was the first to demonstrate that it would have been "almost impossible" without anthropogenic emissions and climate change.[223][222]: 36 

See also edit

References edit

  1. ^ a b c d e Constable, A.J.; Harper, S.; Dawson, J.; Holsman, K.; Mustonen, T.; Piepenburg, D.; Rost, B. (2022). "Cross-Chapter Paper 6: Polar Regions". Climate Change 2022: Impacts, Adaptation and Vulnerability. 2021: 2319–2367. Bibcode:2021AGUFM.U13B..05K. doi:10.1017/9781009325844.023.
  2. ^ "Arctic warming three times faster than the planet, report warns". Phys.org. 20 May 2021. Retrieved 6 October 2022.
  3. ^ "The Arctic is warming four times faster than the rest of the world". 14 December 2021. Retrieved 6 October 2022.
  4. ^ 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): 1–10. doi:10.1038/s43247-022-00498-3. hdl:11250/3115996. ISSN 2662-4435. S2CID 251498876.
  5. ^ Chylek, Petr; Folland, Chris; Klett, James D.; Wang, Muyin; Hengartner, Nick; Lesins, Glen; Dubey, Manvendra K. (25 June 2022). "Annual Mean Arctic Amplification 1970–2020: Observed and Simulated by CMIP6 Climate Models". Geophysical Research Letters. 49 (13). doi:10.1029/2022GL099371. S2CID 250097858.
  6. ^ Shepherd, Andrew; Ivins, Erik; Rignot, Eric; Smith, Ben; van den Broeke, Michiel; Velicogna, Isabella; Whitehouse, Pippa; Briggs, Kate; Joughin, Ian; Krinner, Gerhard; Nowicki, Sophie (12 March 2020). "Mass balance of the Greenland Ice Sheet from 1992 to 2018". Nature. 579 (7798): 233–239. doi:10.1038/s41586-019-1855-2. hdl:2268/242139. ISSN 1476-4687. PMID 31822019. S2CID 219146922. Archived from the original on 23 October 2022. Retrieved 23 October 2022.
  7. ^ Lindsey, Rebecca (18 January 2012). "Shrub Takeover One Sign of Arctic Change". ClimateWatch Magazine. NOAA. Retrieved 19 January 2012.
  8. ^ Kennedy, Caitlyn (1 December 2011). "Sea Ice Declines Boost Arctic Phytoplankton Productivity". ClimateWatch Magazine. NOAA. Retrieved 20 January 2012.
  9. ^ Jahn, Alexandra; Holland, Marika M.; Kay, Jennifer E. (5 March 2024). "Projections of an ice-free Arctic Ocean". Nature Reviews Earth & Environment. 5 (3): 164–176. doi:10.1038/s43017-023-00515-9.
  10. ^ Overpeck, Jonathan T.; Sturm, Matthew; Francis, Jennifer A.; et al. (23 August 2005). "Arctic System on Trajectory to New, Seasonally Ice-Free State". Eos, Transactions, American Geophysical Union. 86 (34): 309–316. Bibcode:2005EOSTr..86..309O. doi:10.1029/2005EO340001.
  11. ^ Butt, F. A.; H. Drange; A. Elverhoi; O. H. Ottera; A. Solheim (2002). "The Sensitivity of the North Atlantic Arctic Climate System to Isostatic Elevation Changes, Freshwater and Solar Forcings" (PDF). Quaternary Science Reviews. 21 (14–15): 1643–1660. doi:10.1016/S0277-3791(02)00018-5. OCLC 108566094. Archived from the original (PDF) on 10 September 2008.
  12. ^ Przybylak, Rajmund (2007). "Recent air-temperature changes in the Arctic". Annals of Glaciology. 46 (1): 316–324. Bibcode:2007AnGla..46..316P. doi:10.3189/172756407782871666. S2CID 129155170.
  13. ^ Arctic Climate Impact Assessment (2004): Arctic Climate Impact Assessment. Cambridge University Press, ISBN 0-521-61778-2, siehe online Archived 28 June 2013 at the Wayback Machine
  14. ^ Arctic Temperatures Highest in at Least 44,000 Years, Livescience, 24 October 2013
  15. ^ Miller, G. H.; Lehman, S. J.; Refsnider, K. A.; Southon, J. R.; Zhong, Y. (2013). "Unprecedented recent summer warmth in Arctic Canada". Geophysical Research Letters. 40 (21): 5745–5751. Bibcode:2013GeoRL..40.5745M. doi:10.1002/2013GL057188. S2CID 128849141.
  16. ^ Yu, Yining; Xiao, Wanxin; Zhang, Zhilun; Cheng, Xiao; Hui, Fengming; Zhao, Jiechen (17 July 2021). "Evaluation of 2-m Air Temperature and Surface Temperature from ERA5 and ERA-I Using Buoy Observations in the Arctic during 2010–2020". Remote Sensing. 13 (Polar Sea Ice: Detection, Monitoring and Modeling): 2813. Bibcode:2021RemS...13.2813Y. doi:10.3390/rs13142813.
  17. ^ "Surface Air Temperature". Arctic Program. October 2020. Retrieved 18 May 2021.
  18. ^ Rosane, Olivia (22 June 2020). "A Siberian Town Just Hit 100 F Degrees". Ecowatch. Retrieved 23 June 2020.
  19. ^ King, Simon; Rowlatt, Justin (22 June 2020). "Arctic Circle sees 'highest-ever' recorded temperatures". BBC. Retrieved 23 June 2020.
  20. ^ Rowlatt, Justin (15 July 2020). "Climate change: Siberian heatwave 'clear evidence' of warming". BBC. Retrieved 17 July 2020.
  21. ^ Deser, Clara; Walsh, John E.; Timlin, Michael S. (1 February 2000). "Arctic Sea Ice Variability in the Context of Recent Atmospheric Circulation Trends". J. Climate. 13 (3): 617–633. Bibcode:2000JCli...13..617D. CiteSeerX 10.1.1.384.2863. doi:10.1175/1520-0442(2000)013<0617:ASIVIT>2.0.CO;2.
  22. ^ 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.
  23. ^ Dai, Aiguo; Luo, Dehai; Song, Mirong; Liu, Jiping (10 January 2019). "Arctic amplification is caused by sea-ice loss under increasing CO2". Nature Communications. 10 (1): 121. Bibcode:2019NatCo..10..121D. doi:10.1038/s41467-018-07954-9. PMC 6328634. PMID 30631051.
  24. ^ Singh, Hansi A.; Polvani, Lorenzo M. (10 January 2020). "Low Antarctic continental climate sensitivity due to high ice sheet orography". npj Climate and Atmospheric Science. 3. doi:10.1038/s41612-020-00143-w. S2CID 222179485.
  25. ^ Steig, Eric; Schneider, David; Rutherford, Scott; Mann, Michael E.; Comiso, Josefino; Shindell, Drew (1 January 2009). "Warming of the Antarctic ice-sheet surface since the 1957 International Geophysical Year". Arts & Sciences Faculty Publications.
  26. ^ Xin, Meijiao; Li, Xichen; Stammerjohn, Sharon E; Cai, Wenju; Zhu, Jiang; Turner, John; Clem, Kyle R; Song, Chentao; Wang, Wenzhu; Hou, Yurong (17 May 2023). "A broadscale shift in antarctic temperature trends". Climate Dynamics. 61 (9–10): 4623–4641. Bibcode:2023ClDy...61.4623X. doi:10.1007/s00382-023-06825-4. S2CID 258777741.
  27. ^ Auger, Matthis; Morrow, Rosemary; Kestenare, Elodie; Nordling, Kalle; Sallée, Jean-Baptiste; Cowley, Rebecca (21 January 2021). "Southern Ocean in-situ temperature trends over 25 years emerge from interannual variability". Nature Communications. 10 (1): 514. Bibcode:2021NatCo..12..514A. doi:10.1038/s41467-020-20781-1. PMC 7819991. PMID 33479205.
  28. ^ Pistone, Kristina; Eisenman, Ian; Ramanathan, Veerabhadran (2019). "Radiative Heating of an Ice-Free Arctic Ocean". Geophysical Research Letters. 46 (13): 7474–7480. Bibcode:2019GeoRL..46.7474P. doi:10.1029/2019GL082914. ISSN 1944-8007. S2CID 197572148.
  29. ^ Arias, Paola A.; Bellouin, Nicolas; Coppola, Erika; Jones, Richard G.; et al. (2021). "Technical Summary" (PDF). IPCC AR6 WG1. p. 76.
  30. ^ Riihelä, Aku; Bright, Ryan M.; Anttila, Kati (28 October 2021). "Recent strengthening of snow and ice albedo feedback driven by Antarctic sea-ice loss". Nature Geoscience. 14: 832–836. doi:10.1038/s41561-021-00841-x. hdl:11250/2830682.
  31. ^ "Thermodynamics: Albedo". NSIDC.
  32. ^ "Polar Vortex: How the Jet Stream and Climate Change Bring on Cold Snaps". InsideClimate News. 2 February 2018. Retrieved 24 November 2018.
  33. ^ a b "Arctic warming three times faster than the planet, report warns". Phys.org. 20 May 2021. Retrieved 6 October 2022.
  34. ^ 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): 1–10. doi:10.1038/s43247-022-00498-3. hdl:11250/3115996. ISSN 2662-4435. S2CID 251498876.
  35. ^ "The Arctic is warming four times faster than the rest of the world". 14 December 2021. Retrieved 6 October 2022.
  36. ^ Isaksen, Ketil; Nordli, Øyvind; et al. (15 June 2022). "Exceptional warming over the Barents area". Scientific Reports. 12 (1): 9371. doi:10.1038/s41598-022-13568-5. PMC 9200822. PMID 35705593. S2CID 249710630.
  37. ^ Damian Carrington (15 June 2022). "New data reveals extraordinary global heating in the Arctic". The Guardian. Retrieved 7 October 2022.
  38. ^ Armstrong McKay, David; Abrams, Jesse; Winkelmann, Ricarda; Sakschewski, Boris; Loriani, Sina; Fetzer, Ingo; Cornell, Sarah; Rockström, Johan; Staal, Arie; Lenton, Timothy (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.
  39. ^ Armstrong McKay, David (9 September 2022). "Exceeding 1.5°C global warming could trigger multiple climate tipping points – paper explainer". climatetippingpoints.info. Retrieved 2 October 2022.
  40. ^ a b Chylek, Petr; Folland, Chris; Klett, James D.; Wang, Muyin; Hengartner, Nick; Lesins, Glen; Dubey, Manvendra K. (25 June 2022). "Annual Mean Arctic Amplification 1970–2020: Observed and Simulated by CMIP6 Climate Models". Geophysical Research Letters. 49 (13). doi:10.1029/2022GL099371. S2CID 250097858.
  41. ^ Acosta Navarro, J.C.; Varma, V.; Riipinen, I.; Seland, Ø.; Kirkevåg, A.; Struthers, H.; Iversen, T.; Hansson, H.-C.; Ekman, A. M. L. (14 March 2016). "Amplification of Arctic warming by past air pollution reductions in Europe". Nature Geoscience. 9 (4): 277–281. Bibcode:2016NatGe...9..277A. doi:10.1038/ngeo2673.
  42. ^ Harvey, C. (14 March 2016). "How cleaner air could actually make global warming worse". Washington Post.
  43. ^ Chylek, Petr; Folland, Chris K.; Lesins, Glen; Dubey, Manvendra K.; Wang, Muyin (16 July 2009). "Arctic air temperature change amplification and the Atlantic Multidecadal Oscillation". Geophysical Research Letters. 36 (14): L14801. Bibcode:2009GeoRL..3614801C. CiteSeerX 10.1.1.178.6926. doi:10.1029/2009GL038777. S2CID 14013240.
  44. ^ Chao-Fong, Léonie (7 January 2021). "'Drastic' rise in high Arctic lightning has scientists worried". The Guardian. Retrieved 30 January 2022.
  45. ^ a b Druckenmiller, Matthew; Thoman, Rick; Moon, Twila (14 December 2021). "2021 Arctic Report Card reveals a (human) story of cascading disruptions, extreme events and global connections". The Conversation. Retrieved 30 January 2022.
  46. ^ Slater, Thomas; Lawrence, Isobel R.; Otosaka, Inès N.; Shepherd, Andrew; Gourmelen, Noel; Jakob, Livia; Tepes, Paul; Gilbert, Lin; Nienow, Peter (25 January 2021). "Review article: Earth's ice imbalance". The Cryosphere. 15 (1): 233–246   Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License. Bibcode:2021TCry...15..233S. doi:10.5194/tc-15-233-2021. hdl:20.500.11820/df343a4d-6b66-4eae-ac3f-f5a35bdeef04.
  47. ^ Huang, Yiyi; Dong, Xiquan; Bailey, David A.; Holland, Marika M.; Xi, Baike; DuVivier, Alice K.; Kay, Jennifer E.; Landrum, Laura L.; Deng, Yi (19 June 2019). "Thicker Clouds and Accelerated Arctic Sea Ice Decline: The Atmosphere-Sea Ice Interactions in Spring". Geophysical Research Letters. 46 (12): 6980–6989. Bibcode:2019GeoRL..46.6980H. doi:10.1029/2019gl082791. hdl:10150/634665. ISSN 0094-8276. S2CID 189968828.
  48. ^ Senftleben, Daniel; Lauer, Axel; Karpechko, Alexey (15 February 2020). "Constraining Uncertainties in CMIP5 Projections of September Arctic Sea Ice Extent with Observations". Journal of Climate. 33 (4): 1487–1503. Bibcode:2020JCli...33.1487S. doi:10.1175/jcli-d-19-0075.1. ISSN 0894-8755. S2CID 210273007.
  49. ^ Yadav, Juhi; Kumar, Avinash; Mohan, Rahul (21 May 2020). "Dramatic decline of Arctic sea ice linked to global warming". Natural Hazards. 103 (2): 2617–2621. Bibcode:2020NatHa.103.2617Y. doi:10.1007/s11069-020-04064-y. ISSN 0921-030X. S2CID 218762126.
  50. ^ "Ice in the Arctic is melting even faster than scientists expected, study finds". NPR.org. Retrieved 10 July 2022.
  51. ^ Fisher, David; Zheng, James; Burgess, David; Zdanowicz, Christian; Kinnard, Christophe; Sharp, Martin; Bourgeois, Jocelyne (March 2012). "Recent melt rates of Canadian arctic ice caps are the highest in four millennia". Global and Planetary Change. 84: 3–7. Bibcode:2012GPC....84....3F. doi:10.1016/j.gloplacha.2011.06.005.
  52. ^ J. C. Stroeve; T. Markus; L. Boisvert; J. Miller; A. Barrett (2014). "Changes in Arctic melt season and implications for sea ice loss". Geophysical Research Letters. 41 (4): 1216–1225. Bibcode:2014GeoRL..41.1216S. doi:10.1002/2013GL058951. S2CID 131673760.
  53. ^ IPCC AR6 WG1 Ch9 2021, p. 9-6, line 19
  54. ^ "Arctic summer sea ice second lowest on record: US researchers". phys.org. 21 September 2020.
  55. ^ Slater, T. S.; Lawrence, I. S.; Otosaka, I. N.; Shepherd, A.; Gourmelen, N.; Jakob, L.; Tepes, P.; Gilbert, L.; Nienow, P. (25 January 2021). "Review article: Earth's ice imbalance". The Cryosphere. 15 (1): 233–246. Bibcode:2021TCry...15..233S. doi:10.5194/tc-15-233-2021. hdl:20.500.11820/df343a4d-6b66-4eae-ac3f-f5a35bdeef04.
  56. ^ Thiede, Jörn; Jessen, Catherine; Knutz, Paul; Kuijpers, Antoon; Mikkelsen, Naja; Nørgaard-Pedersen, Niels; Spielhagen, Robert F (2011). "Millions of Years of Greenland Ice Sheet History Recorded in Ocean Sediments". Polarforschung. 80 (3): 141–159. hdl:10013/epic.38391.
  57. ^ Contoux, C.; Dumas, C.; Ramstein, G.; Jost, A.; Dolan, A.M. (15 August 2015). "Modelling Greenland ice sheet inception and sustainability during the Late Pliocene" (PDF). Earth and Planetary Science Letters. 424: 295–305. Bibcode:2015E&PSL.424..295C. doi:10.1016/j.epsl.2015.05.018. Archived (PDF) from the original on 8 November 2020. Retrieved 7 December 2023.
  58. ^ Knutz, Paul C.; Newton, Andrew M. W.; Hopper, John R.; Huuse, Mads; Gregersen, Ulrik; Sheldon, Emma; Dybkjær, Karen (15 April 2019). "Eleven phases of Greenland Ice Sheet shelf-edge advance over the past 2.7 million years" (PDF). Nature Geoscience. 12 (5): 361–368. Bibcode:2019NatGe..12..361K. doi:10.1038/s41561-019-0340-8. S2CID 146504179. Archived (PDF) from the original on 20 December 2023. Retrieved 7 December 2023.
  59. ^ Robinson, Ben (15 April 2019). "Scientists chart history of Greenland Ice Sheet for first time". The University of Manchester. Archived from the original on 7 December 2023. Retrieved 7 December 2023.
  60. ^ Reyes, Alberto V.; Carlson, Anders E.; Beard, Brian L.; Hatfield, Robert G.; Stoner, Joseph S.; Winsor, Kelsey; Welke, Bethany; Ullman, David J. (25 June 2014). "South Greenland ice-sheet collapse during Marine Isotope Stage 11". Nature. 510 (7506): 525–528. Bibcode:2014Natur.510..525R. doi:10.1038/nature13456. PMID 24965655. S2CID 4468457.
  61. ^ Christ, Andrew J.; Bierman, Paul R.; Schaefer, Joerg M.; Dahl-Jensen, Dorthe; Steffensen, Jørgen P.; Corbett, Lee B.; Peteet, Dorothy M.; Thomas, Elizabeth K.; Steig, Eric J.; Rittenour, Tammy M.; Tison, Jean-Louis; Blard, Pierre-Henri; Perdrial, Nicolas; Dethier, David P.; Lini, Andrea; Hidy, Alan J.; Caffee, Marc W.; Southon, John (30 March 2021). "A multimillion-year-old record of Greenland vegetation and glacial history preserved in sediment beneath 1.4 km of ice at Camp Century". Proceedings of the National Academy of Sciences. 118 (13): e2021442118. Bibcode:2021PNAS..11821442C. doi:10.1073/pnas.2021442118. ISSN 0027-8424. PMC 8020747. PMID 33723012.
  62. ^ Gautier, Agnieszka (29 March 2023). "How and when did the Greenland Ice Sheet form?". National Snow and Ice Data Center. Archived from the original on 28 May 2023. Retrieved 5 December 2023.
  63. ^ Yau, Audrey M.; Bender, Michael L.; Blunier, Thomas; Jouzel, Jean (15 July 2016). "Setting a chronology for the basal ice at Dye-3 and GRIP: Implications for the long-term stability of the Greenland Ice Sheet". Earth and Planetary Science Letters. 451: 1–9. Bibcode:2016E&PSL.451....1Y. doi:10.1016/j.epsl.2016.06.053.
  64. ^ Hörhold, M.; Münch, T.; Weißbach, S.; Kipfstuhl, S.; Freitag, J.; Sasgen, I.; Lohmann, G.; Vinther, B.; Laepple, T. (18 January 2023). "Modern temperatures in central–north Greenland warmest in past millennium". Nature. 613 (7506): 525–528. Bibcode:2014Natur.510..525R. doi:10.1038/nature13456. PMID 24965655. S2CID 4468457.
  65. ^ Briner, Jason P.; Cuzzone, Joshua K.; Badgeley, Jessica A.; Young, Nicolás E.; Steig, Eric J.; Morlighem, Mathieu; Schlegel, Nicole-Jeanne; Hakim, Gregory J.; Schaefer, Joerg M.; Johnson, Jesse V.; Lesnek, Alia J.; Thomas, Elizabeth K.; Allan, Estelle; Bennike, Ole; Cluett, Allison A.; Csatho, Beata; de Vernal, Anne; Downs, Jacob; Larour, Eric; Nowicki, Sophie (30 September 2020). "Rate of mass loss from the Greenland Ice Sheet will exceed Holocene values this century". Nature. 586 (7827): 70–74. Bibcode:2020Natur.586...70B. doi:10.1038/s41586-020-2742-6. PMID 32999481. S2CID 222147426.
  66. ^ Noël, B.; van Kampenhout, L.; Lenaerts, J. T. M.; van de Berg, W. J.; van den Broeke, M. R. (19 January 2021). "A 21st Century Warming Threshold for Sustained Greenland Ice Sheet Mass Loss". Geophysical Research Letters. 48 (5): e2020GL090471. Bibcode:2021GeoRL..4890471N. doi:10.1029/2020GL090471. hdl:2268/301943. S2CID 233632072.
  67. ^ "Special Report on the Ocean and Cryosphere in a Changing Climate: Executive Summary". IPCC. Archived from the original on 8 November 2023. Retrieved 5 December 2023.
  68. ^ Stendel, Martin; Mottram, Ruth (22 September 2022). "Guest post: How the Greenland ice sheet fared in 2022". Carbon Brief. Archived from the original on 22 October 2022. Retrieved 22 October 2022.
  69. ^ Fox-Kemper, B.; Hewitt, H.T.; Xiao, C.; Aðalgeirsdóttir, G.; Drijfhout, S.S.; Edwards, T.L.; Golledge, N.R.; Hemer, M.; Kopp, R.E.; Krinner, G.; Mix, A. (2021). Masson-Delmotte, V.; Zhai, P.; Pirani, A.; Connors, S.L.; Péan, C.; Berger, S.; Caud, N.; Chen, Y.; Goldfarb, L. (eds.). "Chapter 9: Ocean, Cryosphere and Sea Level Change" (PDF). Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK and New York, NY, US. Archived (PDF) from the original on 24 October 2022. Retrieved 22 October 2022.
  70. ^ a b Aschwanden, Andy; Fahnestock, Mark A.; Truffer, Martin; Brinkerhoff, Douglas J.; Hock, Regine; Khroulev, Constantine; Mottram, Ruth; Khan, S. Abbas (19 June 2019). "Contribution of the Greenland Ice Sheet to sea level over the next millennium". Science Advances. 5 (6): 218–222. Bibcode:2019SciA....5.9396A. doi:10.1126/sciadv.aav9396. PMC 6584365. PMID 31223652.
  71. ^ Mouginot, Jérémie; Rignot, Eric; Bjørk, Anders A.; van den Broeke, Michiel; Millan, Romain; Morlighem, Mathieu; Noël, Brice; Scheuchl, Bernd; Wood, Michael (20 March 2019). "Forty-six years of Greenland Ice Sheet mass balance from 1972 to 2018". Proceedings of the National Academy of Sciences. 116 (19): 9239–9244. Bibcode:2019PNAS..116.9239M. doi:10.1073/pnas.1904242116. PMC 6511040. PMID 31010924.
  72. ^ IPCC, 2021: Summary for Policymakers Archived 11 August 2021 at the Wayback Machine. 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 Archived 26 May 2023 at the Wayback Machine [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, NY, US, pp. 3–32, doi:10.1017/9781009157896.001.
  73. ^ Cite error: The named reference Greenland ice sheet BBC2017 was invoked but never defined (see the help page).
  74. ^ a b 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.
  75. ^ Christ, Andrew J.; Rittenour, Tammy M.; Bierman, Paul R.; Keisling, Benjamin A.; Knutz, Paul C.; Thomsen, Tonny B.; Keulen, Nynke; Fosdick, Julie C.; Hemming, Sidney R.; Tison, Jean-Louis; Blard, Pierre-Henri; Steffensen, Jørgen P.; Caffee, Marc W.; Corbett, Lee B.; Dahl-Jensen, Dorthe; Dethier, David P.; Hidy, Alan J.; Perdrial, Nicolas; Peteet, Dorothy M.; Steig, Eric J.; Thomas, Elizabeth K. (20 July 2023). "Deglaciation of northwestern Greenland during Marine Isotope Stage 11". Science. 381 (6655): 330–335. Bibcode:2023Sci...381..330C. doi:10.1126/science.ade4248. PMID 37471537. S2CID 259985096.
  76. ^ Armstrong McKay, David; Abrams, Jesse; Winkelmann, Ricarda; Sakschewski, Boris; Loriani, Sina; Fetzer, Ingo; Cornell, Sarah; Rockström, Johan; Staal, Arie; Lenton, Timothy (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. Archived from the original on 14 November 2022. Retrieved 22 October 2022.
  77. ^ Armstrong McKay, David (9 September 2022). "Exceeding 1.5°C global warming could trigger multiple climate tipping points – paper explainer". climatetippingpoints.info. Archived from the original on 18 July 2023. Retrieved 2 October 2022.
  78. ^ Bjorkman, Anne D.; García Criado, Mariana; Myers-Smith, Isla H.; Ravolainen, Virve; Jónsdóttir, Ingibjörg Svala; Westergaard, Kristine Bakke; Lawler, James P.; Aronsson, Mora; Bennett, Bruce; Gardfjell, Hans; Heiðmarsson, Starri (30 March 2019). "Status and trends in Arctic vegetation: Evidence from experimental warming and long-term monitoring". Ambio. 49 (3): 678–692. doi:10.1007/s13280-019-01161-6. ISSN 0044-7447. PMC 6989703. PMID 30929249.
  79. ^ Gutman, G.Garik (February 1991). "Vegetation indices from AVHRR: An update and future prospects". Remote Sensing of Environment. 35 (2–3): 121–136. Bibcode:1991RSEnv..35..121G. doi:10.1016/0034-4257(91)90005-q. ISSN 0034-4257.
  80. ^ a b Sonja, Myers-Smith, Isla H. Kerby, Jeffrey T. Phoenix, Gareth K. Bjerke, Jarle W. Epstein, Howard E. Assmann, Jakob J. John, Christian Andreu-Hayles, Laia Angers-Blondin, Sandra Beck, Pieter S. A. Berner, Logan T. Bhatt, Uma S. Bjorkman, Anne D. Blok, Daan Bryn, Anders Christiansen, Casper T. Cornelissen, J. Hans C. Cunliffe, Andrew M. Elmendorf, Sarah C. Forbes, Bruce C. Goetz, Scott J. Hollister, Robert D. de Jong, Rogier Loranty, Michael M. Macias-Fauria, Marc Maseyk, Kadmiel Normand, Signe Olofsson, Johan Parker, Thomas C. Parmentier, Frans-Jan W. Post, Eric Schaepman-Strub, Gabriela Stordal, Frode Sullivan, Patrick F. Thomas, Haydn J. D. Tommervik, Hans Treharne, Rachael Tweedie, Craig E. Walker, Donald A. Wilmking, Martin Wipf (2020). Complexity revealed in the greening of the Arctic. Umeå universitet, Institutionen för ekologi, miljö och geovetenskap. OCLC 1234747430.{{cite book}}: CS1 maint: multiple names: authors list (link)
  81. ^ Berner, Logan T.; Massey, Richard; Jantz, Patrick; Forbes, Bruce C.; Macias-Fauria, Marc; Myers-Smith, Isla; Kumpula, Timo; Gauthier, Gilles; Andreu-Hayles, Laia; Gaglioti, Benjamin V.; Burns, Patrick (December 2020). "Summer warming explains widespread but not uniform greening in the Arctic tundra biome". Nature Communications. 11 (1): 4621. Bibcode:2020NatCo..11.4621B. doi:10.1038/s41467-020-18479-5. ISSN 2041-1723. PMC 7509805. PMID 32963240.
  82. ^ Martin, Andrew; Petrokofsky, Gillian (24 May 2018). "Shrub growth and expansion in the Arctic tundra: an assessment of controlling factors using an evidence-based approach". Proceedings of the 5th European Congress of Conservation Biology. Jyväskylä: Jyvaskyla University Open Science Centre. doi:10.17011/conference/eccb2018/108642. S2CID 134164370.
  83. ^ Myers-Smith, Isla H.; Hik, David S. (25 September 2017). "Climate warming as a driver of tundra shrubline advance". Journal of Ecology. 106 (2): 547–560. doi:10.1111/1365-2745.12817. hdl:20.500.11820/f12e7d9d-1c24-4b5f-ad86-96715e071c7b. ISSN 0022-0477. S2CID 90390767.
  84. ^ Alatalo, Juha M.; Jägerbrand, Annika K.; Molau, Ulf (14 August 2014). "Climate change and climatic events: community-, functional- and species-level responses of bryophytes and lichens to constant, stepwise, and pulse experimental warming in an alpine tundra". Alpine Botany. 124 (2): 81–91. doi:10.1007/s00035-014-0133-z. ISSN 1664-2201. S2CID 6665119.
  85. ^ Alatalo, Juha M; Little, Chelsea J (22 March 2014). "Simulated global change: contrasting short and medium term growth and reproductive responses of a common alpine/Arctic cushion plant to experimental warming and nutrient enhancement". SpringerPlus. 3 (1): 157. doi:10.1186/2193-1801-3-157. ISSN 2193-1801. PMC 4000594. PMID 24790813.
  86. ^ Loranty, Michael M; Goetz, Scott J; Beck, Pieter S A (1 April 2011). "Tundra vegetation effects on pan-Arctic albedo". Environmental Research Letters. 6 (2): 024014. Bibcode:2011ERL.....6b4014L. doi:10.1088/1748-9326/6/2/024014. ISSN 1748-9326. S2CID 250681995.
  87. ^ a b c Belke-Brea, M.; Domine, F.; Barrere, M.; Picard, G.; Arnaud, L. (15 January 2020). "Impact of Shrubs on Winter Surface Albedo and Snow Specific Surface Area at a Low Arctic Site: In Situ Measurements and Simulations". Journal of Climate. 33 (2): 597–609. Bibcode:2020JCli...33..597B. doi:10.1175/jcli-d-19-0318.1. ISSN 0894-8755. S2CID 210295151.
  88. ^ a b c Jeong, Su-Jong; Bloom, A. Anthony; Schimel, David; Sweeney, Colm; Parazoo, Nicholas C.; Medvigy, David; Schaepman-Strub, Gabriela; Zheng, Chunmiao; Schwalm, Christopher R.; Huntzinger, Deborah N.; Michalak, Anna M. (July 2018). "Accelerating rates of Arctic carbon cycling revealed by long-term atmospheric CO 2 measurements". Science Advances. 4 (7): eaao1167. Bibcode:2018SciA....4.1167J. doi:10.1126/sciadv.aao1167. ISSN 2375-2548. PMC 6040845. PMID 30009255.
  89. ^ Martin, Andrew C.; Jeffers, Elizabeth S.; Petrokofsky, Gillian; Myers-Smith, Isla; Macias-Fauria, Marc (August 2017). "Shrub growth and expansion in the Arctic tundra: An assessment of controlling factors using an evidence-based approach". Environmental Research Letters. 12 (8): 085007. Bibcode:2017ERL....12h5007M. doi:10.1088/1748-9326/aa7989. S2CID 134164370.
  90. ^ Struzik, Ed (14 February 2011). "Arctic Roamers: The Move of Southern Species into Far North". Environment360. Yale University. Retrieved 19 July 2016. Grizzly bears mating with polar bears. Red foxes out-competing Arctic foxes. Exotic diseases making their way into once-isolated polar realms. These are just some of the worrisome phenomena now occurring as Arctic temperatures soar and the Arctic Ocean, a once-impermeable barrier, melts.
  91. ^ "Protection For Polar Bears Urged By National Wildlife Federation". Science Daily. 3 April 2008. Retrieved 3 April 2008.
  92. ^ DeWeaver, Eric; U.S. Geological Survey (2007). "Uncertainty in Climate Model Projections of Arctic Sea Ice Decline: An Evaluation Relevant to Polar Bears" (PDF). United States Department of the Interior. OCLC 183412441. Archived from the original (PDF) on 9 May 2009.
  93. ^ Broder, John; Revkin, Andrew C. (8 July 2007). "Warming Is Seen as Wiping Out Most Polar Bears". The New York Times. Retrieved 23 September 2007.
  94. ^ "Walruses in a Time of Climate Change". Arctic Program. 14 July 2016. Retrieved 19 May 2021.
  95. ^ Descamps, Sébastien; Aars, Jon; Fuglei, Eva; Kovacs, Kit M.; Lydersen, Christian; Pavlova, Olga; Pedersen, Åshild Ø.; Ravolainen, Virve; Strøm, Hallvard (28 June 2016). "Climate change impacts on wildlife in a High Arctic archipelago – Svalbard, Norway". Global Change Biology. 23 (2): 490–502. doi:10.1111/gcb.13381. ISSN 1354-1013. PMID 27250039. S2CID 34897286.
  96. ^ More Than 200 Reindeer Found Dead in Norway, Starved by Climate Change By Mindy Weisberger. Live Science, 29 July 2019
  97. ^ Weiser, E.L.; Brown, S.C.; Lanctot, R.B.; River Gates, H.; Abraham, K.F.; et al. (2018). "Effects of environmental conditions on reproductive effort and nest success of Arctic-breeding shorebirds". Ibis. 160 (3): 608–623. doi:10.1111/ibi.12571. hdl:10919/99313. S2CID 53514207.
  98. ^ Lee, Sang H.; Whitledge, Terry E.; Kang, Sung-Ho (25 August 2009). "Carbon Uptake Rates of Sea Ice Algae and Phytoplankton under Different Light Intensities in a Landfast Sea Ice Zone, Barrow, Alaska". Arctic. 61 (3). doi:10.14430/arctic25. ISSN 1923-1245.
  99. ^ Wu, Qiang (24 December 2019). "Satellite observations of unprecedented phytoplankton blooms in the Southern Ocean". The Cryosphere Discuss. doi:10.5194/tc-2019-282-sc1. S2CID 243147775.
  100. ^ "Terrestrial Permafrost". Arctic Program. 24 October 2017. Retrieved 18 May 2021.
  101. ^ Sayedi, Sayedeh Sara; Abbott, Benjamin W; Thornton, Brett F; Frederick, Jennifer M; Vonk, Jorien E; Overduin, Paul; Schädel, Christina; Schuur, Edward A G; Bourbonnais, Annie; Demidov, Nikita; Gavrilov, Anatoly (1 December 2020). "Subsea permafrost carbon stocks and climate change sensitivity estimated by expert assessment". Environmental Research Letters. 15 (12): B027-08. Bibcode:2020AGUFMB027...08S. doi:10.1088/1748-9326/abcc29. ISSN 1748-9326. S2CID 234515282.
  102. ^ Hugelius, G.; Strauss, J.; Zubrzycki, S.; Harden, J. W.; Schuur, E. A. G.; Ping, C.-L.; Schirrmeister, L.; Grosse, G.; Michaelson, G. J.; Koven, C. D.; O'Donnell, J. A. (1 December 2014). "Estimated stocks of circumpolar permafrost carbon with quantified uncertainty ranges and identified data gaps". Biogeosciences. 11 (23): 6573–6593. Bibcode:2014BGeo...11.6573H. doi:10.5194/bg-11-6573-2014. ISSN 1726-4189. S2CID 14158339.
  103. ^ "Permafrost and the Global Carbon Cycle". Arctic Program. 31 October 2019. Retrieved 18 May 2021.
  104. ^ a b c d e f g 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. S2CID 252986002.
  105. ^ Natali, Susan M.; Holdren, John P.; Rogers, Brendan M.; Treharne, Rachael; Duffy, Philip B.; Pomerance, Rafe; MacDonald, Erin (10 December 2020). "Permafrost carbon feedbacks threaten global climate goals". Biological Sciences. 118 (21). doi:10.1073/pnas.2100163118. PMC 8166174. PMID 34001617.
  106. ^ a b Armstrong McKay, David; Abrams, Jesse; Winkelmann, Ricarda; Sakschewski, Boris; Loriani, Sina; Fetzer, Ingo; Cornell, Sarah; Rockström, Johan; Staal, Arie; Lenton, Timothy (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.
  107. ^ a b Tarnocai, C.; Canadell, J.G.; Schuur, E.A.G.; Kuhry, P.; Mazhitova, G.; Zimov, S. (June 2009). "Soil organic carbon pools in the northern circumpolar permafrost region". Global Biogeochemical Cycles. 23 (2): GB2023. Bibcode:2009GBioC..23.2023T. doi:10.1029/2008gb003327.
  108. ^ Schuur; et al. (2011). "High risk of permafrost thaw". Nature. 480 (7375): 32–33. Bibcode:2011Natur.480...32S. doi:10.1038/480032a. PMID 22129707. S2CID 4412175.
  109. ^ Bockheim, J.G. & Hinkel, K.M. (2007). "The importance of "Deep" organic carbon in permafrost-affected soils of Arctic Alaska". Soil Science Society of America Journal. 71 (6): 1889–92. Bibcode:2007SSASJ..71.1889B. doi:10.2136/sssaj2007.0070N. Archived from the original on 17 July 2009. Retrieved 5 June 2010.
  110. ^ a b c Fox-Kemper, B., H.T. Hewitt, C. Xiao, G. Aðalgeirsdóttir, S.S. Drijfhout, T.L. Edwards, N.R. Golledge, M. Hemer, R.E. Kopp, G.  Krinner, A. Mix, D. Notz, S. Nowicki, I.S. Nurhati, L. Ruiz, J.-B. Sallée, A.B.A. Slangen, and Y. Yu, 2021: Chapter 9: Ocean, Cryosphere and Sea Level Change. 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, NY, USA, pp. 1211–1362, doi:10.1017/9781009157896.011.
  111. ^ IPCC: Table SPM-2, in: Summary for Policymakers (archived 16 July 2014), in: IPCC AR5 WG1 2013, p. 21
  112. ^ Schellnhuber, Hans Joachim; Winkelmann, Ricarda; Scheffer, Marten; Lade, Steven J.; Fetzer, Ingo; Donges, Jonathan F.; Crucifix, Michel; Cornell, Sarah E.; Barnosky, Anthony D. (2018). "Trajectories of the Earth System in the Anthropocene". Proceedings of the National Academy of Sciences. 115 (33): 8252–8259. Bibcode:2018PNAS..115.8252S. doi:10.1073/pnas.1810141115. ISSN 0027-8424. PMC 6099852. PMID 30082409.
  113. ^ Armstrong McKay, David (9 September 2022). "Exceeding 1.5°C global warming could trigger multiple climate tipping points – paper explainer". climatetippingpoints.info. Retrieved 2 October 2022.
  114. ^ a b Qi, Ling; Wang, Shuxiao (November 2019). "Sources of black carbon in the atmosphere and in snow in the Arctic". Science of the Total Environment. 691: 442–454. Bibcode:2019ScTEn.691..442Q. doi:10.1016/j.scitotenv.2019.07.073. ISSN 0048-9697. PMID 31323589. S2CID 198135020.
  115. ^ Stohl, A.; Klimont, Z.; Eckhardt, S.; Kupiainen, K.; Chevchenko, V.P.; Kopeikin, V.M.; Novigatsky, A.N. (2013), "Black carbon in the Arctic: the underestimated role of gas flaring and residential combustion emissions", Atmos. Chem. Phys., 13 (17): 8833–8855, Bibcode:2013ACP....13.8833S, doi:10.5194/acp-13-8833-2013
  116. ^ Stanley, Michael (10 December 2018). "Gas flaring: An industry practice faces increasing global attention" (PDF). World Bank. Archived from the original (PDF) on 15 February 2019. Retrieved 20 January 2020.
  117. ^ Zhu, Chunmao; Kanaya, Yugo; Takigawa, Masayuki; Ikeda, Kohei; Tanimoto, Hiroshi; Taketani, Fumikazu; Miyakawa, Takuma; Kobayashi, Hideki; Pisso, Ignacio (24 September 2019). "Flexpart v10.1 simulation of source contributions to Arctic black carbon". Atmospheric Chemistry and Physics. doi:10.5194/acp-2019-590. S2CID 204117555.
  118. ^ "The Race to Understand Black Carbon's Climate Impact". ClimateCentral. 2017. Archived from the original on 22 November 2017. Retrieved 21 May 2017.
  119. ^ Zhang, Qiang; Wan, Zheng; Hemmings, Bill; Abbasov, Faig (December 2019). "Reducing black carbon emissions from Arctic shipping: Solutions and policy implications". Journal of Cleaner Production. 241: 118261. doi:10.1016/j.jclepro.2019.118261. ISSN 0959-6526. S2CID 203303955.
  120. ^ a b c Witze, Alexandra (10 September 2020). "The Arctic is burning like never before — and that's bad news for climate change". Nature. 585 (7825): 336–337. Bibcode:2020Natur.585..336W. doi:10.1038/d41586-020-02568-y. ISSN 0028-0836. PMID 32913318. S2CID 221625701.
  121. ^ a b McGrath, Matt (19 March 2022). "Climate change: Wildfire smoke linked to Arctic melting". BBC. Retrieved 20 March 2022.
  122. ^ Shindell, Drew T.; Faluvegi, Greg; Koch, Dorothy M.; Schmidt, Gavin A.; Unger, Nadine; Bauer, Susanne E. (2009). "Improved attribution of climate forcing to emissions". Science. 326 (5953): 716–718. Bibcode:2009Sci...326..716S. doi:10.1126/science.1174760. PMID 19900930. S2CID 30881469.
  123. ^ Kennett, James P.; Cannariato, Kevin G.; Hendy, Ingrid L.; Behl, Richard J. (2003). Methane Hydrates in Quaternary Climate Change: The Clathrate Gun Hypothesis. Washington DC: American Geophysical Union. doi:10.1029/054SP. ISBN 978-0-87590-296-8.
  124. ^ Schellnhuber, Hans Joachim; Winkelmann, Ricarda; Scheffer, Marten; Lade, Steven J.; Fetzer, Ingo; Donges, Jonathan F.; Crucifix, Michel; Cornell, Sarah E.; Barnosky, Anthony D. (2018). "Trajectories of the Earth System in the Anthropocene". Proceedings of the National Academy of Sciences. 115 (33): 8252–8259. Bibcode:2018PNAS..115.8252S. doi:10.1073/pnas.1810141115. ISSN 0027-8424. PMC 6099852. PMID 30082409.
  125. ^ a b Fox-Kemper, B.; Hewitt, H.T.; Xiao, C.; Aðalgeirsdóttir, G.; Drijfhout, S.S.; Edwards, T.L.; Golledge, N.R.; Hemer, M.; Kopp, R.E.; Krinner, G.; Mix, A. (2021). Masson-Delmotte, V.; Zhai, P.; Pirani, A.; Connors, S.L.; Péan, C.; Berger, S.; Caud, N.; Chen, Y.; Goldfarb, L. (eds.). "Chapter 5: Global Carbon and other Biogeochemical Cycles and Feedbacks" (PDF). Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK and New York, NY, USA: 5. doi:10.1017/9781009157896.011.
  126. ^ Moskvitch, Katia (2014). "Mysterious Siberian crater attributed to methane". Nature. doi:10.1038/nature.2014.15649. S2CID 131534214. Archived from the original on 19 November 2014. Retrieved 4 August 2014.
  127. ^ Armstrong McKay, David; Abrams, Jesse; Winkelmann, Ricarda; Sakschewski, Boris; Loriani, Sina; Fetzer, Ingo; Cornell, Sarah; Rockström, Johan; Staal, Arie; Lenton, Timothy (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.
  128. ^ Armstrong McKay, David (9 September 2022). "Exceeding 1.5°C global warming could trigger multiple climate tipping points – paper explainer". climatetippingpoints.info. Retrieved 2 October 2022.
  129. ^ IPCC, 2021: Annex VII: Glossary [Matthews, J.B.R., V. Möller, R. van Diemen, J.S. Fuglestvedt, V. Masson-Delmotte, C.  Méndez, S. Semenov, A. Reisinger (eds.)]. 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, NY, USA, pp. 2215–2256, doi:10.1017/9781009157896.022.
  130. ^ "NOAA Scientists Detect a Reshaping of the Meridional Overturning Circulation in the Southern Ocean". NOAA. 29 March 2023.
  131. ^ Hutchinson, David; Coxall, Helen; O'Regan, Matt; Nilsson, Johan; Caballero, Rodrigo; de Boer, Agatha (23 March 2020). "Arctic closure as a trigger for Atlantic overturning at the Eocene-Oligocene Transition". EGU General Assembly Conference Abstracts: 7493. Bibcode:2020EGUGA..22.7493H. doi:10.5194/egusphere-egu2020-7493. S2CID 225974919.
  132. ^ Molina, Maria J.; Hu, Aixue; Meehl, Gerald A. (22 November 2021). "Response of Global SSTs and ENSO to the Atlantic and Pacific Meridional Overturning Circulations". Journal of Climate. 35 (1): 49–72. doi:10.1175/JCLI-D-21-0172.1. OSTI 1845078. S2CID 244228477.
  133. ^ Rahmstorf, Stefan (9 February 2024). "New study suggests the Atlantic overturning circulation AMOC "is on tipping course"". RealClimate.
  134. ^ Gierz, Paul (31 August 2015). "Response of Atlantic Overturning to future warming in a coupled atmosphere-ocean-ice sheet model". Geophysical Research Letters. 42 (16): 6811–6818. Bibcode:2015GeoRL..42.6811G. doi:10.1002/2015GL065276.
  135. ^ "Explainer: Nine 'tipping points' that could be triggered by climate change". Carbon Brief. 10 February 2020. Retrieved 4 September 2021.
  136. ^ a b c d Armstrong McKay, David; Abrams, Jesse; Winkelmann, Ricarda; Sakschewski, Boris; Loriani, Sina; Fetzer, Ingo; Cornell, Sarah; Rockström, Johan; Staal, Arie; Lenton, Timothy (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.
  137. ^ "Atlantic circulation collapse could cut British crop farming". Phys.org. 13 January 2020. Retrieved 3 October 2022.
  138. ^ a b Lenton, T. M.; Armstrong McKay, D.I.; Loriani, S.; Abrams, J.F.; Lade, S.J.; Donges, J.F.; Milkoreit, M.; Powell, T.; Smith, S.R.; Zimm, C.; Buxton, J.E.; Daube, Bruce C.; Krummel, Paul B.; Loh, Zoë; Luijkx, Ingrid T. (2023). The Global Tipping Points Report 2023 (Report). University of Exeter.
  139. ^ Hansen, J.; Sato, M.; Hearty, P.; Ruedy, R.; Kelley, M.; et al. (23 July 2015). "Ice melt, sea level rise and superstorms: evidence from paleoclimate data, climate modeling, and modern observations that 2 °C global warming is highly dangerous" (PDF). Atmospheric Chemistry and Physics Discussions. 15 (14): 20059–20179. Bibcode:2015ACPD...1520059H. doi:10.5194/acpd-15-20059-2015.
  140. ^ Liu, Wei; Xie, Shang-Ping; Liu, Zhengyu; Zhu, Jiang (4 January 2017). "Overlooked possibility of a collapsed Atlantic Meridional Overturning Circulation in warming climate". Science Advances. 3 (1): e1601666. Bibcode:2017SciA....3E1666L. doi:10.1126/sciadv.1601666. PMC 5217057. PMID 28070560.
  141. ^ Bakker, P; Schmittner, A; Lenaerts, JT; Abe-Ouchi, A; Bi, D; van den Broeke, MR; Chan, WL; Hu, A; Beadling, RL; Marsland, SJ; Mernild, SH; Saenko, OA; Swingedouw, D; Sullivan, A; Yin, J (11 November 2016). "Fate of the Atlantic Meridional Overturning Circulation: Strong decline under continued warming and Greenland melting". Geophysical Research Letters. 43 (23): 12, 252–12, 260. Bibcode:2016GeoRL..4312252B. doi:10.1002/2016GL070457. hdl:10150/622754. S2CID 133069692.
  142. ^ Sigmond, Michael; Fyfe, John C.; Saenko, Oleg A.; Swart, Neil C. (1 June 2020). "Ongoing AMOC and related sea-level and temperature changes after achieving the Paris targets". Nature Climate Change. 10 (7): 672–677. Bibcode:2020NatCC..10..672S. doi:10.1038/s41558-020-0786-0. S2CID 219175812.
  143. ^ He, Feng; Clark, Peter U. (7 April 2022). "Freshwater forcing of the Atlantic Meridional Overturning Circulation revisited". Nature Climate Change. 12 (5): 449–454. Bibcode:2022NatCC..12..449H. doi:10.1038/s41558-022-01328-2. S2CID 248004571.
  144. ^ Kim, Soong-Ki; Kim, Hyo-Jeong; Dijkstra, Henk A.; An, Soon-Il (11 February 2022). "Slow and soft passage through tipping point of the Atlantic Meridional Overturning Circulation in a changing climate". npj Climate and Atmospheric Science. 5 (13). Bibcode:2022npCAS...5...13K. doi:10.1038/s41612-022-00236-8. S2CID 246705201.
  145. ^ Valdes, Paul (2011). "Built for stability". Nature Geoscience. 4 (7): 414–416. Bibcode:2011NatGe...4..414V. doi:10.1038/ngeo1200. ISSN 1752-0908.
  146. ^ a b Lohmann, Johannes; Ditlevsen, Peter D. (2 March 2021). "Risk of tipping the overturning circulation due to increasing rates of ice melt". Proceedings of the National Academy of Sciences. 118 (9): e2017989118. Bibcode:2021PNAS..11817989L. doi:10.1073/pnas.2017989118. ISSN 0027-8424. PMC 7936283. PMID 33619095.
  147. ^ Boers, Niklas (August 2021). "Observation-based early-warning signals for a collapse of the Atlantic Meridional Overturning Circulation" (PDF). Nature Climate Change. 11 (8): 680–688. Bibcode:2021NatCC..11..680B. doi:10.1038/s41558-021-01097-4. S2CID 236930519.
  148. ^ Ditlevsen, Peter; Ditlevsen, Susanne (25 July 2023). "Warning of a forthcoming collapse of the Atlantic meridional overturning circulation". Nature Communications. 14 (1): 4254. arXiv:2304.09160. Bibcode:2023NatCo..14.4254D. doi:10.1038/s41467-023-39810-w. ISSN 2041-1723. PMC 10368695. PMID 37491344.
  149. ^ "expert reaction to paper warning of a collapse of the Atlantic meridional overturning circulation". Science Media Centre. 25 July 2023. Retrieved 11 August 2023.
  150. ^ 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.
  151. ^ IPCC, 2019: Summary for Policymakers. In: IPCC Special Report on the Ocean and Cryosphere in a Changing Climate [H.-O. Pörtner, D.C. Roberts, V. Masson-Delmotte, P. Zhai, M. Tignor, E. Poloczanska, K. Mintenbeck, A. Alegría, M.  Nicolai, A. Okem, J. Petzold, B. Rama, N.M. Weyer (eds.)]. Cambridge University Press, Cambridge, UK and New York, NY, USA. doi:10.1017/9781009157964.001.
  152. ^ Mecking, J.V.; Drijfhout, S.S.; Jackson, L.C.; Andrews, M.B. (1 January 2017). "The effect of model bias on Atlantic freshwater transport and implications for AMOC bi-stability". Tellus A: Dynamic Meteorology and Oceanography. 69 (1): 1299910. Bibcode:2017TellA..6999910M. doi:10.1080/16000870.2017.1299910. S2CID 133294706.
  153. ^ Weijer, W.; Cheng, W.; Drijfhout, S. S.; Fedorov, A. V.; Hu, A.; Jackson, L. C.; Liu, W.; McDonagh, E. L.; Mecking, J. V.; Zhang, J. (2019). "Stability of the Atlantic Meridional Overturning Circulation: A Review and Synthesis". Journal of Geophysical Research: Oceans. 124 (8): 5336–5375. Bibcode:2019JGRC..124.5336W. doi:10.1029/2019JC015083. ISSN 2169-9275. S2CID 199807871.
  154. ^ a b Armstrong McKay, David (9 September 2022). "Exceeding 1.5°C global warming could trigger multiple climate tipping points – paper explainer". climatetippingpoints.info. Retrieved 2 October 2022.
  155. ^ Archer, Cristina L.; Caldeira, Ken (18 April 2008). "Historical trends in the jet streams". Geophysical Research Letters. 35 (8). Bibcode:2008GeoRL..35.8803A. doi:10.1029/2008GL033614. S2CID 59377392.
  156. ^ "Jet stream found to be permanently drifting north". Associated Press. 18 April 2008. Archived from the original on 17 August 2016. Retrieved 7 October 2022.
  157. ^ 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.
  158. ^ "The Arctic is warming four times faster than the rest of the world". Science Magazine. 14 December 2021. Retrieved 6 October 2022.
  159. ^ Isaksen, Ketil; Nordli, Øyvind; et al. (15 June 2022). "Exceptional warming over the Barents area". Scientific Reports. 12 (1): 9371. Bibcode:2022NatSR..12.9371I. doi:10.1038/s41598-022-13568-5. PMC 9200822. PMID 35705593.
  160. ^ Damian Carrington (15 June 2022). "New data reveals extraordinary global heating in the Arctic". The Guardian. Retrieved 7 October 2022.
  161. ^ Francis, Jennifer A.; Vavrus, Stephen J. (2012). "Evidence linking Arctic amplification to extreme weather in mid-latitudes". Geophysical Research Letters. 39 (6): L06801. Bibcode:2012GeoRL..39.6801F. CiteSeerX 10.1.1.419.8599. doi:10.1029/2012GL051000. S2CID 15383119.
  162. ^ Zielinski, G.; Mershon, G. (1997). "Paleoenvironmental implications of the insoluble microparticle record in the GISP2 (Greenland) ice core during the rapidly changing climate of the Pleistocene-Holocene transition". Bulletin of the Geological Society of America. 109 (5): 547–559. Bibcode:1997GSAB..109..547Z. doi:10.1130/0016-7606(1997)109<0547:piotim>2.3.co;2.
  163. ^ Lue, J.-M.; Kim, S.-J.; Abe-Ouchi, A.; Yu, Y.; Ohgaito, R. (2010). "Arctic Oscillation during the Mid-Holocene and Last Glacial Maximum from PMIP2 Coupled Model Simulations". Journal of Climate. 23 (14): 3792–3813. Bibcode:2010JCli...23.3792L. doi:10.1175/2010JCLI3331.1. S2CID 129156297.
  164. ^ Mitchell, Daniel M.; Osprey, Scott M.; Gray, Lesley J.; Butchart, Neal; Hardiman, Steven C.; Charlton-Perez, Andrew J.; Watson, Peter (August 2012). "The Effect of Climate Change on the Variability of the Northern Hemisphere Stratospheric Polar Vortex". Journal of the Atmospheric Sciences. 69 (8): 2608–2618. Bibcode:2012JAtS...69.2608M. doi:10.1175/jas-d-12-021.1. ISSN 0022-4928. S2CID 122783377.
  165. ^ Masato, Giacomo; Hoskins, Brian J.; Woollings, Tim (2013). "Winter and Summer Northern Hemisphere Blocking in CMIP5 Models". Journal of Climate. 26 (18): 7044–7059. Bibcode:2013JCli...26.7044M. doi:10.1175/JCLI-D-12-00466.1.
  166. ^ Liu, Jiping; Curry, Judith A.; Wang, Huijun; Song, Mirong; Horton, Radley M. (27 February 2012). "Impact of declining Arctic sea ice on winter snowfall". PNAS. 109 (11): 4074–4079. Bibcode:2012PNAS..109.4074L. doi:10.1073/pnas.1114910109. PMC 3306672. PMID 22371563.
  167. ^ Weng, H. (2012). "Impacts of multi-scale solar activity on climate. Part I: Atmospheric circulation patterns and climate extremes". Advances in Atmospheric Sciences. 29 (4): 867–886. Bibcode:2012AdAtS..29..867W. doi:10.1007/s00376-012-1238-1. S2CID 123066849.
  168. ^ James E. Overland (8 December 2013). "Atmospheric science: Long-range linkage". Nature Climate Change. 4 (1): 11–12. Bibcode:2014NatCC...4...11O. doi:10.1038/nclimate2079.
  169. ^ Seviour, William J.M. (14 April 2017). "Weakening and shift of the Arctic stratospheric polar vortex: Internal variability or forced response?". Geophysical Research Letters. 44 (7): 3365–3373. Bibcode:2017GeoRL..44.3365S. doi:10.1002/2017GL073071. hdl:1983/caf74781-222b-4735-b171-8842cead4086. S2CID 131938684.
  170. ^ Screen, James A. (15 June 2014). "Arctic amplification decreases temperature variance in northern mid- to high-latitudes". Nature Climate Change. 4 (7): 577–582. Bibcode:2014NatCC...4..577S. doi:10.1038/nclimate2268. hdl:10871/15095.
  171. ^ van Oldenborgh, Geert Jan; Mitchell-Larson, Eli; Vecchi, Gabriel A.; de Vries, Hylke; Vautar, Robert; Otto, Friederike (22 November 2019). "Cold waves are getting milder in the northern midlatitudes". Environmental Research Letters. 14 (11): 114004. Bibcode:2019ERL....14k4004V. doi:10.1088/1748-9326/ab4867. S2CID 204420462.
  172. ^ Blackport, Russell; Screen, James A.; van der Wiel, Karin; Bintanja, Richard (September 2019). "Minimal influence of reduced Arctic sea ice on coincident cold winters in mid-latitudes". Nature Climate Change. 9 (9): 697–704. Bibcode:2019NatCC...9..697B. doi:10.1038/s41558-019-0551-4. hdl:10871/39784. S2CID 199542188.
  173. ^ Blackport, Russell; Screen, James A. (February 2020). "Insignificant effect of Arctic amplification on the amplitude of midlatitude atmospheric waves". Science Advances. 6 (8): eaay2880. Bibcode:2020SciA....6.2880B. doi:10.1126/sciadv.aay2880. PMC 7030927. PMID 32128402.
  174. ^ Streffing, Jan; Semmler, Tido; Zampieri, Lorenzo; Jung, Thomas (24 September 2021). "Response of Northern Hemisphere Weather and Climate to Arctic Sea Ice Decline: Resolution Independence in Polar Amplification Model Intercomparison Project (PAMIP) Simulations". Journal of Climate. 34 (20): 8445–8457. Bibcode:2021JCli...34.8445S. doi:10.1175/JCLI-D-19-1005.1. S2CID 239631549.
  175. ^ Paul Voosen (12 May 2021). "Landmark study casts doubt on controversial theory linking melting Arctic to severe winter weather". Science Magazine. Retrieved 7 October 2022.
  176. ^ Smith, D.M.; Eade, R.; Andrews, M.B.; et al. (7 February 2022). "Robust but weak winter atmospheric circulation response to future Arctic sea ice loss". Nature Communications. 13 (1): 727. Bibcode:2022NatCo..13..727S. doi:10.1038/s41467-022-28283-y. PMC 8821642. PMID 35132058. S2CID 246637132.
  177. ^ Eckel, Mike (20 September 2007). "Russia: Tests Show Arctic Ridge Is Ours". The Washington Post. Associated Press. Retrieved 21 September 2007.[dead link]
  178. ^ a b c d e f "Territorial Claims in the Arctic Circle: An Explainer". The Observer. Retrieved 19 May 2021.
  179. ^ a b c "Evolution of Arctic Territorial Claims and Agreements: A Timeline (1903–Present) • Stimson Center". Stimson Center. 15 September 2013. Retrieved 19 May 2021.
  180. ^ Humpert, Malte; Raspotnik, Andreas (2012). "The Future of Shipping Along the Transpolar Sea Route" (PDF). The Arctic Yearbook. 1 (1): 281–307. Archived from the original (PDF) on 21 January 2016. Retrieved 18 November 2015.
  181. ^ "As The Earth Warms, The Lure Of The Arctic's Natural Resources Grows". 18 March 2019.
  182. ^ Byers, Michael. "Melting Arctic brings new opportunities". aljazeera.com.
  183. ^ a b c d e f g h Hassol, Susan Joy (2004). Impacts of a warming Arctic (Reprinted ed.). Cambridge, UK: Cambridge University Press. ISBN 978-0-521-61778-9.
  184. ^ Bekkers, Eddy; Francois, Joseph F.; Rojas-Romagosa, Hugo (1 December 2016). "Melting Ice Caps and the Economic Impact of Opening the Northern Sea Route" (PDF). The Economic Journal. 128 (610): 1095–1127. doi:10.1111/ecoj.12460. ISSN 1468-0297. S2CID 55162828.
  185. ^ Ramage, Justine; Jungsberg, Leneisja; Wang, Shinan; Westermann, Sebastian; Lantuit, Hugues; Heleniak, Timothy (6 January 2021). "Population living on permafrost in the Arctic". Population and Environment. 43: 22–38. doi:10.1007/s11111-020-00370-6. S2CID 254938760.
  186. ^ Nelson, F. E.; Anisimov, O. A.; Shiklomanov, N. I. (1 July 2002). "Climate Change and Hazard Zonation in the Circum-Arctic Permafrost Regions". Natural Hazards. 26 (3): 203–225. doi:10.1023/A:1015612918401. S2CID 35672358.
  187. ^ Barry, Roger Graham; Gan, Thian-Yew (2021). The global cryosphere past, present and future (Second revised ed.). Cambridge, United Kingdom: Cambridge University Press. ISBN 978-1-108-48755-9. OCLC 1256406954.
  188. ^ a b Hjort, Jan; Streletskiy, Dmitry; Doré, Guy; Wu, Qingbai; Bjella, Kevin; Luoto, Miska (11 January 2022). "Impacts of permafrost degradation on infrastructure". Nature Reviews Earth & Environment. 3 (1): 24–38. Bibcode:2022NRvEE...3...24H. doi:10.1038/s43017-021-00247-8. hdl:10138/344541. S2CID 245917456.
  189. ^ a b Hjort, Jan; Karjalainen, Olli; Aalto, Juha; Westermann, Sebastian; Romanovsky, Vladimir E.; Nelson, Frederick E.; Etzelmüller, Bernd; Luoto, Miska (11 December 2018). "Degrading permafrost puts Arctic infrastructure at risk by mid-century". Nature Communications. 9 (1): 5147. Bibcode:2018NatCo...9.5147H. doi:10.1038/s41467-018-07557-4. PMC 6289964. PMID 30538247.
  190. ^ a b Melvin, April M.; Larsen, Peter; Boehlert, Brent; Neumann, James E.; Chinowsky, Paul; Espinet, Xavier; Martinich, Jeremy; Baumann, Matthew S.; Rennels, Lisa; Bothner, Alexandra; Nicolsky, Dmitry J.; Marchenko, Sergey S. (26 December 2016). "Climate change damages to Alaska public infrastructure and the economics of proactive adaptation". Proceedings of the National Academy of Sciences. 114 (2): E122–E131. doi:10.1073/pnas.1611056113. PMC 5240706. PMID 28028223.
  191. ^ "The CAT Thermometer". Retrieved 25 April 2023.
  192. ^ Tsui, Emily (4 March 2021). "Reducing Individual Costs of Permafrost Thaw Damage in Canada's Arctic". The Arctic Institute.
  193. ^ Melnikov, Vladimir; Osipov, Victor; Brouchkov, Anatoly V.; Falaleeva, Arina A.; Badina, Svetlana V.; Zheleznyak, Mikhail N.; Sadurtdinov, Marat R.; Ostrakov, Nikolay A.; Drozdov, Dmitry S.; Osokin, Alexei B.; Sergeev, Dmitry O.; Dubrovin, Vladimir A.; Fedorov, Roman Yu. (24 January 2022). "Climate warming and permafrost thaw in the Russian Arctic: potential economic impacts on public infrastructure by 2050". Natural Hazards. 112: 231–251. doi:10.1007/s11069-021-05179-6. S2CID 246211747.
  194. ^ a b c Langer, Morit; Schneider von Deimling, Thomas; Westermann, Sebastian; Rolph, Rebecca; Rutte, Ralph; Antonova, Sofia; Rachold, Volker; Schultz, Michael; Oehme, Alexander; Grosse, Guido (28 March 2023). "Thawing permafrost poses environmental threat to thousands of sites with legacy industrial contamination". Nature Communications. 14 (1): 1721. Bibcode:2023NatCo..14.1721L. doi:10.1038/s41467-023-37276-4. PMC 10050325. PMID 36977724.
  195. ^ a b Miner, Kimberley R.; D'Andrilli, Juliana; Mackelprang, Rachel; Edwards, Arwyn; Malaska, Michael J.; Waldrop, Mark P.; Miller, Charles E. (30 September 2021). "Emergent biogeochemical risks from Arctic permafrost degradation". Nature Climate Change. 11 (1): 809–819. Bibcode:2021NatCC..11..809M. doi:10.1038/s41558-021-01162-y. S2CID 238234156.
  196. ^ "Diesel fuel spill in Norilsk in Russia's Arctic contained". TASS. Moscow, Russia. 5 June 2020. Retrieved 7 June 2020.
  197. ^ Max Seddon (4 June 2020). "Siberia fuel spill threatens Moscow's Arctic ambitions". Financial Times. Archived from the original on 10 December 2022.
  198. ^ Nechepurenko, Ivan (5 June 2020), "Russia Declares Emergency After Arctic Oil Spill", New York Times
  199. ^ Antonova, Maria (5 June 2020). "Russia Says Melting Permafrost Is Behind The Massive Arctic Fuel Spill". Science Daily. Retrieved 19 July 2020.
  200. ^ Schaefer, Kevin; Elshorbany, Yasin; Jafarov, Elchin; Schuster, Paul F.; Striegl, Robert G.; Wickland, Kimberly P.; Sunderland, Elsie M. (16 September 2020). "Potential impacts of mercury released from thawing permafrost". Nature Communications. 11 (1): 4650. Bibcode:2020NatCo..11.4650S. doi:10.1038/s41467-020-18398-5. PMC 7494925. PMID 32938932.
  201. ^ Hawkings, Jon R.; Linhoff, Benjamin S.; Wadham, Jemma L.; Stibal, Marek; Lamborg, Carl H.; Carling, Gregory T.; Lamarche-Gagnon, Guillaume; Kohler, Tyler J.; Ward, Rachael; Hendry, Katharine R.; Falteisek, Lukáš; Kellerman, Anne M.; Cameron, Karen A.; Hatton, Jade E.; Tingey, Sarah; Holt, Amy D.; Vinšová, Petra; Hofer, Stefan; Bulínová, Marie; Větrovský, Tomáš; Meire, Lorenz; Spencer, Robert G. M. (24 May 2021). "Large subglacial source of mercury from the southwestern margin of the Greenland Ice Sheet". Nature Geoscience. 14 (5): 496–502. Bibcode:2021NatGe..14..496H. doi:10.1038/s41561-021-00753-w.
  202. ^ Walther, Kelcie (15 July 2021). "As the Greenland Ice Sheet Retreats, Mercury is Being Released From the Bedrock Below". Columbia Climate School. Archived from the original on 23 December 2023. Retrieved 23 December 2023.
  203. ^ Jørgensen, Christian Juncher; Søndergaard, Jens; Larsen, Martin Mørk; Kjeldsen, Kristian Kjellerup; Rosa, Diogo; Sapper, Sarah Elise; Heimbürger-Boavida, Lars-Eric; Kohler, Stephen G.; Wang, Feiyue; Gao, Zhiyuan; Armstrong, Debbie; Albers, Christian Nyrop (26 January 2024). "Large mercury release from the Greenland Ice Sheet invalidated". Science Advances. 10 (4). doi:10.1126/sciadv.adi7760.
  204. ^ Colgan, William; Machguth, Horst; MacFerrin, Mike; Colgan, Jeff D.; van As, Dirk; MacGregor, Joseph A. (4 August 2016). "The abandoned ice sheet base at Camp Century, Greenland, in a warming climate". Geophysical Research Letters. 43 (15): 8091–8096. Bibcode:2016GeoRL..43.8091C. doi:10.1002/2016GL069688.
  205. ^ Rosen, Julia (4 August 2016). "Mysterious, ice-buried Cold War military base may be unearthed by climate change". Science Magazine. Archived from the original on 15 January 2024. Retrieved 23 December 2023.
  206. ^ a b Berkes, Fikret; Jolly, Dyanna (2001). "Adapting to climate change: social-ecological resilience in a Canadian western Arctic community" (PDF). Conservation Ecology. 5 (2).
  207. ^ a b Farquhar, Samantha D. (18 March 2020). "Inuit Seal Hunting in Canada: Emerging Narratives in an Old Controversy". Arctic. 73 (1): 13–19. doi:10.14430/arctic69833. ISSN 1923-1245. S2CID 216308832.
  208. ^ Timonin, Andrey (2021). "Climate Change in the Arctic and Future Directions for Adaptation: Views From Non-Arctic States". SSRN Electronic Journal. doi:10.2139/ssrn.3802303. ISSN 1556-5068. S2CID 233756936.
  209. ^ Rogers, Sarah (13 June 2014). "New online atlas tracks Nunavut's centuries-old Inuit trails". Nunatsiaq News. Retrieved 19 May 2021.
  210. ^ a b Freedman, Andrew (12 December 2017). "Arctic warming, ice melt 'unprecedented' in at least the past 1,500 years". Mashable. Retrieved 13 December 2017.
  211. ^ a b "Arctic Report Card: Update for 2017; Arctic shows no sign of returning to reliably frozen region of recent past decades". NOAA. Retrieved 13 December 2017.
  212. ^ "ESA's ice mission CryoSat-2". esa.int. 11 September 2008. Retrieved 15 June 2009.
  213. ^ Wininger, Corinne (26 October 2007). "E SF, VR, FORMAS sign MOU to promote Global Environmental Change Research". innovations-report.de. Retrieved 26 November 2007.
  214. ^ "Arctic Change". International Study of Arctic Change.
  215. ^ a b AMAP Arctic Climate Change Update 2021: Key Trends and Impacts. Arctic Monitoring and Assessment Programme (AMAP) (Report). Tromsø, Norway. 2021. pp. viii + 148. ISBN 978-82-7971-201-5.
  216. ^ Arias, Paola A.; Bellouin, Nicolas; Coppola, Erika; Jones, Richard G.; et al. (2021). "Technical Summary" (PDF). IPCC AR6 WG1. p. 76.
  217. ^ 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. ISSN 2662-4435. S2CID 251498876.
  218. ^ Chylek, Petr; Folland, Chris; Klett, James D.; Wang, Muyin; Hengartner, Nick; Lesins, Glen; Dubey, Manvendra K. (16 July 2022). "Annual Mean Arctic Amplification 1970–2020: Observed and Simulated by CMIP6 Climate Models". Geophysical Research Letters. 49 (13). Bibcode:2022GeoRL..4999371C. doi:10.1029/2022GL099371. ISSN 0094-8276. S2CID 250097858. via Wikipedia Library and EBSCOhost
  219. ^ "Arctic temperatures are increasing four times faster than global warming". Los Alamos National Laboratory. Retrieved 18 July 2022.
  220. ^ a b Rapid and pronounced warming continues to drive the evolution of the Arctic environment (Report). Arctic Report Card: Update for 2021. NOAA.
  221. ^ Impacts of a Warming Arctic: Arctic Climate Impact Assessment. Arctic Climate Impact Assessment (ACIA) (Report). Overview report. Cambridge University Press. 15 October 2004. p. 140. ISBN 0-521-61778-2.
  222. ^ a b c Spreading like Wildfire – The Rising Threat of Extraordinary Landscape Fires. United Nations Environment Programme (UNEP) (Report). A UNEP Rapid Response Assessment. Nairobi, Kenya. 2022. p. 122.
  223. ^ Ciavarella, A.; Cotterill, D.; Stott, P. (2021). "Prolonged Siberian heat of 2020 almost impossible without human influence". Climatic Change. 166 (9): 9. Bibcode:2021ClCh..166....9C. doi:10.1007/s10584-021-03052-w. PMC 8550097. PMID 34720262. S2CID 233875870.


Works cited edit

Further reading edit

External links edit