1. Introduction
Climate change is causing significant environmental changes in the Arctic, with far-reaching consequences beyond its borders [
1]. The Arctic has become a major area of global climate research in recent decades, owing to its increased sensitivity to warming and crucial role in controlling planetary climate systems [
2,
3,
4,
5,
6,
7,
8]. The Arctic is warming at a rate nearly four times faster than the global average, a phenomenon known as Arctic amplification [
9]. This phenomenon is attributed to complex feedback mechanisms, such as the ice–albedo effect and shifts in atmospheric and oceanic dynamics [
10,
11,
12]. These processes accelerate regional warming and contribute to systemic changes in Arctic ecosystems, atmospheric circulation, and global climate regulation [
6,
13].
The rapid loss of Arctic Sea ice is one of the most prominent manifestations of environmental changes. Recent studies have documented unprecedented reductions in sea ice extent, with projections indicating a seasonally ice-free Arctic Ocean as early as the mid-21st century [
5,
6,
14]. This phenomenon has a cascading effect on global climate systems, affecting oceanic circulation, weather patterns, and the severity of extreme weather events [
15,
16,
17]. Permafrost thawing presents a significant concern by releasing substantial amounts of stored carbon, thereby amplifying the greenhouse effect and altering regional and global carbon dynamics [
18,
19,
20,
21]. These environmental changes are accompanied by significant ecological transformations. Rising temperatures and increased vegetation drive Arctic greening, which coexists with the localised browning produced by extreme weather events and permafrost degradation [
22,
23]. As sub-Arctic species migrate northward, both marine and terrestrial species face significant phenological changes and increased competition, posing substantial challenges to the Arctic endemics [
24,
25,
26]. Changes in hydrological cycles, such as increased precipitation and decreased snow cover, are changing freshwater ecosystems and affecting regional energy balances [
27,
28].
Understanding the dynamics of these changes is important not only for forecasting the course of Arctic change, but also for assessing their implications for global climate systems, biodiversity, and communities. Indigenous communities in the Arctic are particularly sensitive to these changes [
29,
30,
31], facing disruptions to traditional subsistence practices, infrastructure damage caused by permafrost thaw, and challenges from shifting ecological and socioeconomic conditions [
32,
33,
34,
35]. The opening of Arctic shipping routes and increased potential for resource extraction demonstrate its growing geopolitical importance [
15,
36].
The Arctic has a population of approximately 4 million people, of whom approximately 10% are Indigenous, living in Alaska, Canada, Finland, Greenland, Norway, Russia, and Sweden [
37,
38]. Different nations refer to Indigenous Peoples in a diversity of ways, including “native” in Alaska, First Nations, Inuit, and Metis in Canada (and also “aboriginal”), Sámi in Finland, Norway, Sweden, and Russia, Inuit in Greenland, and Aleuts and Nenets, among others, in Russia [
39]. Indigenous Peoples of the Arctic encompass a heterogeneity of cultures, worldviews, traditions, ethnicities, and biophysical environments [
40]. This spectrum of identity is often underpinned by a strong relationship and cultural commitment to the significance of ancestral lands. In many regions, Indigenous Peoples continue to live in small, remote settlements with livelihoods and cultural identities closely linked to traditional hunting, fishing, herding, and trapping activities.
Despite significant advances in understanding Arctic climate mechanisms (
Table 1), significant information gaps persist. Key uncertainties remain in feedback mechanisms, such as permafrost–carbon interactions, the function of short-lived climatic forcers, and the impact of Arctic changes on mid-latitude weather patterns [
18,
41]. In addition, the inadequate inclusion of Indigenous knowledge in research and governance frameworks impedes the development of effective and equitable adaptation strategies [
32,
42]. Addressing these key gaps is important, considering the rapid pace of environmental change in the Arctic and its global implications. This review synthesises insights from recent studies on Arctic climate change and its impacts, providing a comprehensive overview of emerging trends and challenges. By integrating findings from diverse disciplines, this paper aims to advance our understanding of Arctic processes, identify important research priorities, and facilitate the development of strategies to mitigate and adapt to the rapid transformations that are occurring. In doing so, this review not only contributes to advancing Arctic science but also highlights the Arctic’s significant role in the broader context of global climate change research and policymaking. This study examines the transboundary impacts of climate change on Arctic, emphasising the critical need for joint mitigation and adaptation efforts.
2. Study Area Description
The Arctic encompasses the Earth’s northernmost part (
Figure 1), also known as the circumpolar north, and consists of the following eight countries: Canada, the United States (Alaska), Denmark (via Greenland), Russia, Norway, Sweden, Finland, and Iceland [
62]. While the Arctic lacks a universally accepted southern boundary, for the purposes of this study, we adopt the Arctic Monitoring and Assessment Programme (AMAP) definition. This definition includes all areas north of the Arctic Circle (66° 33′N) and extends into sub-Arctic zones based on bioclimatic and ecological characteristics, particularly the presence of permafrost, tundra, and boreal forest systems [
39,
63]. This delineation is consistent with other pan-Arctic ecological studies investigating vegetation and environmental change (e.g., [
64,
65]). It is characterised by cold winters and cool summers, with lowest temperatures between −54 and −46 °C, and monthly average temperature is below +10 °C [
66,
67]. The Arctic, which was historically dominated by vast sea ice, is experiencing rapid environmental change, including significant ice loss due to climate warming [
68]. The Arctic is ecologically significant, serving as a habitat for species, such as polar bears and Arctic foxes, and functioning as a global climate regulator [
69]. Rising temperatures are caused by mechanisms including ice–albedo feedback and shifting atmospheric and oceanic patterns [
70]. Since 1979, these changes have resulted in a 13% per decade loss in Arctic Sea ice, permafrost thaw, and biodiversity disruptions, all of which have cascading consequences on ecosystems and global climate patterns [
71,
72]. The Arctic is home to four million people, including about 40 Indigenous groups, such as Alaska Natives, Inuit, Sámi, Nenets, Métis, Aleut, Yup’ik, Khanty, Evenk, and Chukchi, among others, constituting about 10% of the Arctic’s population [
37].
The Arctic possesses abundant natural resources, including oil, natural gas, wildlife, and fisheries, but the exploitation of these resources poses significant environmental and sociocultural challenges [
73]. The Arctic Council plays an important role in Arctic governance, integrating both national policies and international cooperation and promoting cooperation, coordination, and interaction among the Arctic States and their inhabitants [
74,
75].
3. Thematic Analysis and Methods
This review adopts an integrated approach to synthesise the recent literature and multidisciplinary perspectives on Arctic climate change and its implications. The methodology consists of the following three steps: (1) a thorough literature review; (2) thematic content analysis; and (3) a critical synthesis of emerging trends and knowledge gaps. The literature evaluation started with a search of peer-reviewed journals, reports from international organisations, and Indigenous knowledge repositories. Databases including Scopus, Web of Science, and PubMed (
Figure 2) were used to provide a diversified source base. We prioritised studies over the last two decades (2000–2025) focused on climate dynamics, environmental changes, socioeconomic implications, and Arctic-specific governance frameworks.
The main inclusion criteria were relevance to the Arctic and interdisciplinary perspectives. Indigenous knowledge systems, as documented in grey literature and community-based studies, were used to examine sociocultural components that are sometimes disregarded in mainstream research.
The collected literature was methodically organised into major themes, such as Arctic amplification, sea ice dynamics, permafrost thaw, ecosystem alterations, socioeconomic implications, and policy responses. A thematic content analysis was carried out to analyse recurring trends, causal relationships, and emerging challenges across these themes. This method focused on feedback mechanisms, such as ice–albedo interactions and permafrost–carbon dynamics, as well as their consequences for global climate systems. Second, the study critically synthesised emerging patterns, linking findings to regional and global contexts. Visualisation tools, such as geographic information systems (GISs) and data-driven infographics, were used to show spatial and temporal trends in Arctic warming, sea ice loss, and ecosystem impact. Figures were created to represent various phenomena, such as Arctic amplification and cascading ecosystem impacts, assuring clarity and accessibility. Finally, knowledge gaps and future research goals were identified using a gap analysis approach. This includes comparing current findings to existing policy and governance frameworks, emphasising the importance of interdisciplinary and participatory approaches. The incorporation of Indigenous ecological knowledge and scientific approaches was identified as an important pathway for developing long-term adaption strategies.
4. Arctic Temperature Trends and Climate Drivers
In recent decades, the Arctic has experienced significant increases in temperature as a result of both rising greenhouse gas emissions and the Arctic’s unique feedback mechanisms [
4]. Historically, the vast ice and snow cover in the Arctic has worked as a mirror, reflecting most of the sun’s energy into space [
76,
77,
78]. However, when temperatures rise, this reflecting layer is gradually replaced by darker ocean water and exposed land, both of which absorb significantly more solar energy [
79]. This absorption accelerates warming, resulting in additional ice and snow loss—a cycle that continues to increase. Beyond surface changes, air dynamics have an important impact on Arctic temperature trends [
80]. Global warming-induced shifts in the polar vortex and jet stream are sending warmer air masses northward, resulting in peculiar patterns of seasonal variability and extreme occurrences in the region [
81]. Ocean currents also contribute to this warming: the input of warmer Atlantic waters interrupts the normal cooling processes in Arctic, causing faster sea ice melt and increasing the thermal instability [
82]. Surface feedback mechanisms, atmospheric changes, and ocean currents all work together to form an Arctic climate system that is extremely sensitive to change [
80]. The effects are not isolated; warming in the Arctic influences weather patterns across the Northern Hemisphere, drives sea-level rises, and increases the risks to biodiversity and coastal ecosystems worldwide [
1]. Understanding these trends and drivers is important, as they reveal not only the vulnerabilities of the Arctic but also the broader impacts of a rapidly changing climate.
Recent studies show that the Arctic is warming at considerably higher rates than the world average—a phenomenon known as Arctic amplification, in which it warms at nearly four times the global average [
9]. This warming trend is significantly caused by specific regional feedback processes and atmospheric–oceanic interactions [
83]. Historical data suggest that Arctic temperatures have risen by about 2.3 °C since the early twentieth century, greatly exceeding world norms [
84]. Observations from 1979 to 2021 show that a significant portion of the Arctic Ocean warmed at a rate exceeding 0.75 °C per decade [
9]. Ground-based records, particularly from coastal and island stations, indicate that this trend accelerated significantly during the 1970s, coinciding with substantial increases in industrial emissions [
85]. Arctic amplification is caused by a complex combination of local feedback, including sea ice–albedo effects, increased heat and moisture transport from lower latitudes, and increased cloud cover [
86,
87].
The “tropically excited Arctic warming mechanism” (TEAM) is an important driver of Arctic amplification, connecting increased tropical convection with Arctic warming. TEAM suggests that enhanced convection in the Pacific warm pool boosts poleward Rossby wave propagation, thereby transferring heat and moisture to the Arctic [
88,
89]. This transport not only enhances downwelling longwave radiation (DLW), but it also contributes to rising Arctic temperatures [
88,
90,
91]. Atmospheric latent heat (LH) transfer has been demonstrated to be more efficient in warming the Arctic compared to dry static energy (DSE) transit [
92]. This effect is amplified by the “water vapour triple effect,” in which moisture boosts the greenhouse effect before condensing, resulting in additional warming via increased cloudiness [
93,
94,
95]. The role of maritime heat movement is another important consideration. Observations indicate that oceanic heat inflow through the Fram Strait and Barents Sea has increased, transporting warm Atlantic waters into the Arctic [
96]. This heat contributes directly to sea ice melt and local warming, allowing for longer periods of open water and diminishing sea ice extent [
97,
98]. Climate models also support this trend, indicating that under continued warming scenarios, ocean heat transport into the Arctic will continue to be a major driver of Arctic amplification, contributing significantly to long-term sea ice decline and temperature increases [
99,
100,
101].
Episodic phenomena, such as air mass changes and moisture invasions from the middle latitudes, also contribute significantly to Arctic warming. These events bring warm, moist air masses, resulting in transient increases in temperature and humidity, thereby enhancing Arctic amplification [
102]. These episodic changes in cloud cover and specific humidity, as well as shifts in seasonal energy budgets, highlight the importance of moisture transport in Arctic warming [
103,
104]. As these factors interact with specific Arctic feedback mechanisms, they collectively accelerate the rate of regional warming, thereby exacerbating the impacts of global climate change on the Arctic. Since 1979, satellite data have provided important insights into the seasonal variations of Arctic Sea ice. These data reveal a significant decline in Arctic Sea ice extent, averaging approximately 13% per decade, which is closely associated with rising temperatures in both surface air and sea surface environments [
71,
105]. These patterns are investigated using a variety of data sources, including passive microwave satellites, thermal infrared sensors, and reanalysis models, which blend historical records with current atmospheric data to account for fluctuations [
106,
107,
108]. The thorough integration of these approaches indicates not only a significant temperature rise but also changes in ocean heat distribution and cloud cover—factors that influence warming rates throughout the region.
4.1. Climate Drivers in the ARCTIC
The Arctic climate is shaped by three major drivers that accelerate warming more rapidly than in other regions. First, greenhouse gases, such as carbon dioxide (CO
2) and methane (CH
4), trap heat in the atmosphere, raising surface temperatures [
109,
110,
111]. Second, albedo changes due to melting snow and sea ice expose darker land and ocean surfaces, which absorb more solar radiation. This reduction in reflectivity intensifies warming in a self-reinforcing cycle [
44,
112]. Third, polar vortex dynamics are increasingly disrupted by Arctic warming. As the temperature gradient between the Arctic and mid-latitudes narrows, the polar vortex weakens and becomes more erratic, allowing cold Arctic air to move southward and allowing warmer air to enter the Arctic [
113,
114,
115]. Together, these climate drivers contribute to the unique and amplified warming pattern observed across the Arctic region.
4.2. Anthropogenic Drivers
Anthropogenic activities have significantly contributed to climate change in the Arctic. The primary driver of this change is the increased emission of greenhouse gases (GHGs), particularly carbon dioxide (CO
2) and methane (CH
4), resulting from fossil fuel combustion, industrial activities, and deforestation [
4,
116]. These emissions enhance the greenhouse effect, trapping heat in the atmosphere and leading to rising temperatures [
117]. Global warming caused by anthropogenic greenhouse gas emissions is leading to a significant reduction in the Arctic’s sea ice [
7]. Black carbon, a short-lived climate pollutant produced from incomplete combustion of fossil fuels and biomass, settles on Arctic ice, reducing its reflectivity (albedo) and accelerating melting [
118].
In conjunction with the effects of climate change, various other human-induced stressors, such as air and water pollution, overfishing, depletion of the ozone layer, habitat modification, and pollution resulting from resource extraction, as well as the escalating pressure on land and resources, are affecting life in the Arctic. The increase in wildfire seasons in the Arctic has made forest fires more intense and destructive [
119], resulting in deforestation and loss of forest cover [
120] due to a combination of climatic and non-climatic factors, predominantly human-caused climate change [
121]. Industrial activities, such as oil and gas extraction in the Arctic, contribute to environmental degradation and climate instability [
122]. The expansion of shipping routes, enabled by sea ice loss, increases emissions of pollutants, further warming the atmosphere [
123]. A notable example is the increasing presence of liquefied natural gas tankers, which produce methane leaks during transport, further amplifying Arctic warming [
124]. Deforestation and land-use changes, particularly in boreal forests bordering the Arctic, reduce carbon sequestration capacity and alter regional climate patterns [
125]. The combination of these human-induced factors is accelerating changes in Arctic ecosystems, leading to biodiversity loss, disruptions in Indigenous livelihoods, and shifts in global climate patterns due to altered oceanic and atmospheric circulation [
72].
Natural resource extraction is increasing in the Arctic and is a significant contributing factor to the rise in Arctic shipping. This can be observed in the case of increasing iron ore extraction in the Arctic [
126]. The transportation of bulk carriers carrying large quantities of various products, such as food grains, ores, coal, and cement, has increased substantially. Shipping activity increased by 37% from 2013 to 2023 and the distance sailed by all the vessels increased by 75% [
127]. Shipping activities in the Arctic have significant effects on biogeophysical systems, including the atmosphere and the ocean. In the atmosphere, shipping contributes to greenhouse gas emissions and the release of black carbon, which accelerates ice melting [
128]. In the ocean, shipping causes pollution, overharvesting of marine resources, disturbances to marine species, underwater noise pollution, and the introduction of invasive species [
129]. These impacts pose risks to both the environment and society, particularly for the Indigenous Peoples who rely on these systems [
130]. The Arctic is estimated to contain approximately 13% (90 billion barrels) of the world’s undiscovered conventional oil resources and 30% of its undiscovered conventional natural gas resources [
131]. It supplies roughly 10% of its oil and 25% of its natural gas to the world [
132]. With the melting of Arctic ice, there is increased exploration and production of fossil fuels in the area, leading to climate change.
4.3. Feedback Mechanisms in the Arctic
The Arctic is vulnerable to a variety of feedback mechanisms that accelerate warming and alter climate dynamics [
4]. Ice–albedo feedback, which occurs when the ice cover shrinks, exposes darker ocean or land surfaces [
10]. This exposed surface absorbs more solar energy than ice, which reflects sunlight, resulting in increased heat and accelerated ice melting. This cycle of melting ice and increased heat absorption is an important driver of Arctic warming [
133]. The release of greenhouse gases from permafrost thawing is an important feedback mechanism [
134]. As permafrost thaws, the organic material trapped beneath it decomposes, emitting carbon dioxide and methane, both of which are important greenhouse gases [
18]. These emissions add to atmospheric warming, which increases permafrost thaw, resulting in a feedback loop that affects global greenhouse gas levels. Ocean circulation patterns also play an important part. As Arctic ice melts and freshwater flows into the ocean, it alters the salinity and density of seawater, potentially slowing the Atlantic Meridional Overturning Circulation (AMOC). A weaker AMOC could slow the transfer of warm water northward, affecting climatic patterns far beyond the Arctic [
80]. Together, these feedback mechanisms show the interdependence of Arctic changes, with each loop exacerbating warming and changing not only the Arctic but also the worldwide climate system.
4.4. Arctic Amplification (AA)
Arctic amplification is a process in which the Arctic region warms at a much higher rate than the global average, as supported by instrumental evidence, climate models, and paleoclimate proxy records [
135,
136]. Several connected mechanisms contribute to this rapid warming, including marine heating and the ice–albedo feedback mechanism (
Figure 3), which exposes darker ocean surfaces when sea ice melts, resulting in higher solar energy absorption [
4]. Atmospheric feedback mechanisms, such as Planck feedback, lapse-rate feedback, and cloud feedback, as well as ocean heat transport and meridional atmospheric moisture transport, also play important roles [
137,
138]. Recent decreases in air pollution in Europe [
139] have been related to increasing Arctic warming, and future reductions in Asian aerosols owing to mitigation policies may exacerbate this trend [
140]. Despite the well-known trends, there is little agreement on the exact magnitude of AA, with studies reporting Arctic warming at nearly twice, thrice, and four times the global average. The Arctic Monitoring and Assessment Programme reports a rate of warming three times faster than global warming from 1971 to 2019 [
141]. The differences stem from different definitions of AA, such as the time periods and geographical limits examined. While climate models, notably those in the Coupled Model Intercomparison Project phase 6 (CMIP6), have made progress in modelling Arctic climate evolution and sea ice dynamics, many continue to struggle to appropriately depict the sensitivity of sea ice loss to global temperature increases. Understanding whether climate models can effectively reproduce the reported scale of AA is important, given that sea ice loss is an important cause of the phenomena. Although previous research suggests that AA is weaker in climate models than in observations [
142], a complete comparison with the most recent observational datasets has yet to be undertaken. Arctic amplification is an important climate phenomenon with far-reaching implications for global climate systems, emphasising the need for further research and better climate models to support effective policy decisions targeted at minimising climate change impacts.
AA is a well-known phenomenon characterised by increased warming in the Arctic relative to the global average, primarily caused by such processes as sea ice–albedo feedback, which reduces the surface albedo as ice melts, exposing more ocean that absorbs heat [
143,
144]. Recent research has highlighted the subtle functions of cloud and water vapour feedback mechanisms, with increased cloud cover blocking incoming solar radiation while increasing downward longwave radiation, resulting in net warming [
145]. Temperature feedback, influenced by stable stratification of the Arctic atmosphere, leads to positive feedback via a lower lapse rate, resulting in higher warmth aloft [
146]. The movement of heat and water vapour from lower latitudes to the Arctic exacerbates regional warming [
10,
147]. Despite a thorough understanding of these mechanisms, quantifying their relative contributions is difficult, implying that Arctic amplification persists even when specific feedback mechanisms are suppressed, resulting in dynamic competition among feedback processes in climate models [
142].
5. Climate Change Impacts in the Arctic
5.1. Arctic Sea Ice Decline: Impacts on Ecosystems and Biogeochemical Cycles
The Arctic sea ice (ASI) is changing rapidly, with significant implications for both the regional and global ecology. The significant drop in ASI extent, area, and volume observed over the last few decades is evidence of human-induced climate change, underlining the Arctic’s vulnerability to rising temperatures [
148]. This trend is more than just a climatic anomaly; it is also closely related to biological and biogeochemical processes that are currently poorly understood, notably how ASI interacts with life-history events in Arctic biota. While the impacts of ASI loss, such as altered migration and breeding patterns in marine and terrestrial species, have been extensively documented [
149], the more complex indirect impacts remain a key area of investigation. These indirect consequences involve a series of interactions between many ecosystem components, such as altered predator–prey relationships and disruptions to food webs [
150].
The effects of ASI on biogeochemical cycles, notably mercury cycling, are a major concern [
151]. Rapid loss of sea ice is impacting mercury transport and cycling in Arctic ecosystems, with consequences for human and wildlife health [
152]. ASI depletion has been documented to alter primary productivity, food web structure, and mercury methylation and demethylation rates [
153]. The combination of these changes’ influences mercury bioaccumulation in Arctic species, potentially amplifying its harmful effects. These bottom-up processes, driven by changes in sea ice cover and primary production, are exacerbated by top-down influences, where animal behavioural changes, such as changing feeding patterns, can lead to increased mercury exposure [
153,
154]. These biogeochemical alterations raise concerns about future of the Arctic food webs and the long-term sustainability of ecosystems (
Figure 4). While studies focus on the immediate implications of sea ice loss on mercury cycling, the long-term ecological and biogeochemical impacts are still unknown.
The significant drop in the Arctic Sea ice volume, a vital indicator for assessing the severity of Arctic warming, is a major focus of recent climate research [
17,
155]. Unlike sea ice extent or thickness, sea ice volume is a more sensitive indication of climate change because it incorporates both the horizontal dispersion and vertical thickness of the ice [
156,
157]. Over the past 40 years, the volume of sea ice has declined by about 75% [
158]. Summer minimums have decreased from over 16,000 km
3 to under 4000 km
3, indicating a breakdown of the multi-year ice cover [
148]. This loss of multi-year ice (MYI), which is more resistant to melting than first-year ice (FYI), is a major contributor to the positive feedback loop of the Arctic, in which melting ice exposes darker ocean water, increasing warming and additional ice loss [
156]. As MYI disappears, the Arctic is increasingly dominated by FYI, which is less stable and more vulnerable to seasonal fluctuations. The transition from MYI to FYI has profound implications, affecting not only the climate system but also the region’s ecological and biogeochemical stability [
148,
159]. Given this tendency, there is growing fear that the Arctic may experience an ice-free summer within the next 15 years, with far-reaching implications for both regional and global climates [
160]. Paleoenvironmental records provide important information about the long-term variability of Arctic Sea ice and how it interacts with climate and ecosystems. Evidence from the East Greenland shelf reveals prior sea ice events, revealing transitions towards colder, lower-salinity conditions approximately 4.7 ka, indicating the start of neoglacial cooling [
161]. These data indicate that sea ice variability has historically been a key driver of Arctic climate and ecosystem changes. However, the current rate of sea ice loss is unprecedented, and the future course of Arctic ice loss is unpredictable. The increase in polar water input and Arctic Sea ice export could worsen cooling trends in the north Atlantic Ocean, thereby affecting global ocean circulation and weather patterns [
162]. Understanding these paleosignals in the context of present climate change is important for improving forecasts of future Arctic conditions and their implications.
The integrated effects of ASI loss on Arctic ecosystems, combined with the intricate interplay of biological and biogeochemical processes, highlight the need for a multidisciplinary approach. While the direct biological effects of sea ice decrease are rather well established [
150,
155], the indirect effects caused by changes in nutrient dynamics, food web interactions, and biogeochemical cycles necessitate further exploration. The significance of mercury cycling in Arctic ecosystems, particularly in response to climate-induced changes in sea ice cover, is a growing concern that necessitates collaborative study efforts. Comprehensive monitoring systems and models that include both biological and physical components are critical for improving our predictive understanding of these cascading impacts. The rapid reduction in Arctic Sea ice is not merely an indication of climate change, but also a precursor to significant ecological and biogeochemical changes. ASI loss has far-reaching and complex impacts, ranging from changes in food webs to an altered mercury cycle. The interdisciplinary research needed to understand these processes is critical for creating successful conservation and mitigation solutions.
5.2. Glacier Retreat and Ice Sheet Dynamics
The melting of glaciers and ice sheets in the Arctic, particularly the Greenland ice sheet, has increased in recent past, creating concerns about global sea level rise [
163]. The Greenland ice sheet is losing mass at an alarming rate, with yearly ice loss growing from around 50 billion tonnes in the 1990s to nearly 280 billion tonnes by 2018 [
164]. The glacial melting trend in the Arctic is driven by the combination of surface melting caused by rising temperatures and greater iceberg calving from glaciers into the ocean [
165,
166]. Satellite data reveal that surface melting accounts for around 60% of ice loss, which is exacerbated by rising air temperatures and changes in atmospheric circulation [
167,
168]. Short-term climate variability and extreme glaciological events influence the long-term evolution of the Antarctic and Greenland ice sheets, impacting projections of sea level rise [
168]. Projections indicate that if current trends continue, the Greenland ice sheet alone might be one of the major contributors to world sea levels by 2100 [
169]. The melting of other glaciers and ice caps throughout the world adds to rising sea levels [
170,
171]. The worldwide consequences of such sea-level rise are severe, affecting low-lying coastal regions and island nations, increasing the frequency of flooding, and displacing people [
172]. The influx of freshwater from melting glaciers might disturb ocean circulation patterns, potentially impacting climatic systems well beyond the Arctic [
173]. The continued loss of glaciers and ice sheets highlights the need for global climate action, as these changes not only endanger local ecosystems and livelihoods but also have far-reaching consequences for global climate stability and sea levels [
173].
5.3. Implications for Ocean Circulation
The loss of Arctic ice has implications for ocean circulation, particularly key systems, like thermohaline circulation and the Atlantic Meridional Overturning Circulation (AMOC) [
80]. As sea ice melts, the influx of freshwater affects the salinity and density of surface waters in the Arctic and North Atlantic [
174,
175]. This influx of freshwater has the potential to disturb the equilibrium required for thermohaline circulation, which drives ocean currents by varying temperature and salt levels [
176]. The decrease in salinity, combined with the warming of surface waters, can affect these currents, which are critical for transferring heat and nutrients around the world. The AMOC, a critical component of the global climate system, is particularly susceptible to changes in the Arctic [
177]. This circulation system is critical in transporting warm water from the tropics to higher latitudes, influencing weather patterns and climate throughout Europe and North America [
110]. Studies have reported that the continued melting of ice sheets and sea ice is contributing to a slowing in the AMOC, which is projected to diminish by up to 34% by 2100 if current trends continue [
178]. Such a deceleration could lead to harsher winters in Europe, altered precipitation patterns, and greater storminess in the North Atlantic region [
179]. The weakening of AMOC could also cause a sea-level rise along the United States’ east coast due to changes in ocean dynamics [
180]. These changes in ocean circulation have far-reaching implications, with the potential to disturb marine ecosystems, affect fish migration patterns, and have an impact on global weather systems [
181]. As ocean currents play an important role in controlling the Earth’s climate, the continuous loss of Arctic ice cover raises concerns about their stability and the wider implications for global climate patterns.
5.4. Permafrost Thawing and Carbon Release
The Arctic permafrost zone covers approximately 24% of the Northern Hemisphere’s geographical area, including Alaska, Canada, Siberia, and northern Scandinavia [
182]. Permafrost is defined as permanently frozen ground that can range in depth from a few meters to over 1500 m in some areas [
183]. Permafrost depth is determined by a variety of factors, including local climate, vegetation cover, and soil composition [
184]. Continuous permafrost locations have a thicker and more stable layer, whereas sporadic permafrost locations have greater depth variability and sensitivity to warming [
185]. Recent data show that permafrost is thawing at an accelerated rate as global temperatures rise, with some research suggesting that the Arctic permafrost area is warming by about 0.5 °C each decade [
186]. This enhanced thawing causes ground sinking, which poses issues for infrastructure and alters hydrological processes [
187]. Thawing permafrost can also produce significant changes in Arctic geomorphology, as melting ice creates thermokarsts, which are depressions or uneven terrain caused by the melting of ice-rich permafrost [
188]. These landscape alterations have the potential to destabilise ecosystems and habitat stability, exacerbating the regional impacts of climate change. The thawing of permafrost poses threats to the carbon cycle by releasing significant amounts of greenhouse gases, principally carbon dioxide (CO
2) and methane (CH
4). According to Strauss et al. [
189], permafrost holds roughly 1500 gigatons of organic carbon, which is comparable to twice the current atmospheric carbon stock. As permafrost thaws, the microbial breakdown of organic materials increases, releasing CO
2 [
190]. Anaerobic conditions in saturated areas promote the creation of methane, a more potent greenhouse gas than CO
2 in the short term [
191]. This release of greenhouse gases triggers a feedback loop in which higher air concentrations contribute to greater warming and permafrost thaw.
5.5. Impact on Arctic Ecosystems and Biodiversity
- (a)
Terrestrial ecosystem changes
Climate change is causing major changes to the Arctic’s terrestrial ecosystems. Vegetation patterns are changing drastically, with tundra habitats receding north and being replaced by shrubs and forests in many areas [
192,
193]. This phenomenon, known as “greening”, is caused by longer growing seasons and higher temperatures, allowing more productive flora to take over previously sparse Arctic environments [
194]. However, this vegetation transition has ecological implications. Herbivores, such as caribou, which rely on tundra-specific lichens, are experiencing reduced food supply and altered migration patterns, potentially contributing to declines in population [
195,
196,
197]. Arctic foxes and wolves, which rely on tundra-adapted food, are facing increased competition from boreal species moving into their habitats [
53,
198]. The cascading effects of these changes disrupt the balance of Arctic terrestrial food webs, endangering biodiversity and ecosystem resilience.
- (b)
Marine Ecosystem Responses
Arctic marine habitats are undergoing dramatic changes as a result of sea ice loss and rising waters. The timing and intensity of phytoplankton blooms, a key component of Arctic food webs, are being influenced by early ice melt and extended periods of open sea [
199]. These changes have far-reaching consequences for the ecosystem, affecting primary consumers, such as zooplankton, and the larger marine species that rely on them [
200]. Polar bears, walruses, and seals are highly susceptible to climate change. Polar bears, which rely on sea ice for hunting, are facing habitat fragmentation as ice melts earlier and freezes later in the year [
201]. Walruses and seals, which rely on stable ice platforms for breeding and resting, are also impacted, with reduced ice cover threatening their survival and reproductive success [
202,
203]. These disturbances highlight the interconnectivity of Arctic marine systems and the importance of preventing future climate implications [
204].
- (c)
Invasive species and biodiversity threats
Warming Arctic temperatures are allowing non-native species to establish themselves in formerly hostile regions, jeopardising the health of native ecosystems (
Figure 5). Invasive species, which are frequently introduced through increasing shipping and ballast water discharge, can outcompete local flora and fauna, disturbing existing ecological relationships [
196]. For example, invasive plant species can change soil composition and nutrient cycles, making it harder for native tundra plants to recover. Similarly, non-native marine species introduced by humans can dominate food webs, displacing native species and causing unexpected ecological repercussions [
205]. Some of the invasive species found in the Arctic are the American mink, Nootka lupine, European green crab, Japanese knotweed, and Japanese ghost shrimp [
69]. The biodiversity in the Arctic, which is already under threat from climate-induced habitat changes, faces additional threats from these invasive species, emphasising the importance of biosecurity controls and adaptive management solutions to protect these sensitive ecosystems.
5.6. Socioeconomic Implications and Indigenous Communities
The warming of the Arctic has severely affected the traditional livelihoods and cultural practices of Indigenous communities (
Figure 6), which are inextricably linked to the region’s natural ecosystems [
32]. Climate-induced changes in wildlife behaviour, habitat availability, and seasonal patterns are affecting subsistence activities, such as hunting, fishing, and gathering [
152,
206]. Earlier sea ice breakup restricts access to marine animals, such as seals and walruses, which are important for both food and cultural practices among Indigenous Peoples [
203]. Changes in caribou migration routes and population dynamics disrupt traditional hunting practices, creating food insecurity [
207,
208]. Cultural traditions based on the regularity of Arctic surroundings, such as storytelling, seasonal festivities, sewing, and craft-making, are also under threat, undermining intergenerational knowledge transfer and cultural identity [
209,
210,
211]. As the temperature increases, the health of Arctic people and infrastructure is expected to face more challenges [
152]. Thawing permafrost destabilises buildings, roadways, and essential infrastructure, resulting in costly repairs and increased safety risks [
212]. Changes in precipitation and temperature enhance health risks, such as respiratory problems from increasing wildfire smoke and waterborne infections caused by poor sanitation systems [
152,
154]. Food security is a significant issue, since disruptions in subsistence hunting and fishing force communities to rely on expensive and less nutritious imported foods [
213]. These increased vulnerabilities highlight the critical need for infrastructure adaptation and public health measures that are targeted to the challenges of Arctic communities. The melting of Arctic Sea ice has created access to previously unexplored oil, gas, and mineral riches, as well as the possibility of expanded shipping routes and tourism.
While these changes have potential economic advantages, they also represent major risks to the environment. Increased industrial activity can cause pollution, habitat degradation, and social impacts, which disproportionately affect Indigenous communities that rely on Arctic ecosystem stability and resources [
214,
215]. The possibility of transboundary conflicts over resource claims emphasises the need for strong governance frameworks that balance economic development, environmental conservation, and community rights [
205]. Despite these constraints, Arctic communities have demonstrated extraordinary resilience through adaptive techniques and policy initiatives. Indigenous knowledge systems, which emphasise sustainable resource use and community-based monitoring, are crucial in responding to climate change [
203,
216]. Co-management of wildlife resources, for example, and the development of renewable energy initiatives minimise reliance on foreign supply while also mitigating environmental harm [
217]. At the policy level, efforts, like the Arctic Council and national adaptation programs, aim to boost resilience by encouraging collaboration among local, national, and international stakeholders [
212]. These efforts underscore the need of combining scientific research with Indigenous knowledge to provide equitable and effective solutions [
218] for the Arctic’s rapidly changing climate.
5.7. Global Implications of Arctic Warming
The increased warming in the Arctic causes significant disturbances to global climate systems by disrupting air circulation and heat transfer processes [
219]. One major impact is the destabilisation of the polar jet stream, which controls weather patterns in the Northern Hemisphere [
220]. Arctic temperatures rise faster than at lower latitudes, weakening the jet stream, causing it to meander and stall [
221]. This has resulted in persistent extreme weather conditions, such as extended droughts, flooding, and unseasonal cold snaps in areas far from the Arctic [
222]. In addition to the jet stream, the Arctic’s role as a heat sink is dwindling, reducing the Earth’s ability to manage heat distribution and fuelling increasingly erratic global climate behaviour [
223]. Melting Arctic ice is becoming a major contributor to increasing global sea-level rise [
224]. Unlike thermal expansion of saltwater, glacier melt introduces massive amounts of freshwater into the ocean, affecting salinity and density gradients [
225]. These changes affect ocean circulation patterns, such as the Atlantic Meridional Overturning Circulation, which could cause localised cooling in Europe while exacerbating warming elsewhere [
226]. Coastal communities around the world face increased risk of flooding, infrastructure damage, and forced migration [
227]. The Arctic’s contribution to rising sea levels highlights the dual complication of managing local impacts while addressing broader global vulnerability. The Arctic is an important location of climatic feedback loops that might impede progress towards global climate targets. Permafrost thawing, which releases methane and carbon dioxide stored for millennia, results in a self-reinforcing warming cycle [
228]. The loss of sea ice exacerbates the albedo effect, as darker ocean surfaces absorb more solar radiation, accelerating regional and global temperature increases [
44]. A less commonly discussed feedback mechanism is the destabilisation of Arctic peatlands, which may release massive amounts of ancient carbon into the atmosphere [
229]. These feedback mechanisms not only exacerbate current warming, but also pose a risk of exceeding climate tipping thresholds, making global temperature stabilisation more difficult. Arctic warming has far-reaching economic and policy repercussions. Melting glaciers are revealing previously inaccessible resources, including fossil fuels, offering a challenge to climate mitigation [
72]. Increased Arctic resource extraction, motivated by economic interests, might undermine attempts to reduce global emissions if not properly controlled [
230]. Arctic warming emphasises the limitations of current climate models, which underestimate warming rates, complicating global climate projections and policy planning.
6. Recent Advances in Arctic Climate Change Research
The Arctic has emerged as an important target of climate change studies because of its increased sensitivity and rapid environmental changes. Over the last two decades, research has improved our understanding of Arctic climate dynamics, highlighting both natural and anthropogenic influences on its ecosystem. Recent studies have documented that the Arctic is warming faster than the global average due to feedback mechanisms, like the ice–albedo effect [
6,
231]. Winter warming trends are pronounced, leading to a significant decline in sea ice extent, with predictions of a seasonally ice-free Arctic Ocean by mid-century [
6]. This impacts global climate systems and accelerates feedback loops [
5,
15]. Permafrost thawing releases stored carbon, intensifying the greenhouse effect and increasing infrastructure vulnerability for Arctic communities [
19,
28].
Some studies highlight that warmer temperatures are causing widespread Arctic greening, which is an increase in vegetation productivity and biomass [
194,
232,
233,
234]. However, these changes are not uniform, with some areas undergoing browning or decreased productivity as a result of extreme weather events or permafrost thaw [
22]. Marine and terrestrial species are suffering range changes, changing phenologies, and greater competition, with sub-Arctic species pushing into Arctic environments, endangering endemic species [
24,
25]. The release of carbon from thawing permafrost and the growth of anoxic conditions contribute to the permafrost carbon feedback, which has the potential to exacerbate global warming [
18,
20]. The destabilisation of methane hydrates and higher emissions from wildfires exacerbate these impacts [
18,
28].
Research suggests that the Arctic hydrological cycle is becoming more intense, with increasing precipitation, transitions from snow to rain, and more frequent extreme weather events [
27,
28]. These changes disrupt freshwater ecosystems, alter river dynamics, and reduce snow cover, which have an impact on regional energy balances [
235,
236]. Atmospheric dynamics, such as the Atlantification of Arctic waters and increasing poleward heat movement, are changing regional climate patterns [
16,
32,
34,
52]. Economic losses are anticipated for Arctic infrastructure as permafrost thaws, threatening the infrastructure’s stability [
19,
21,
237,
238]. The opening of Arctic shipping routes and increased resource extraction prospects have heightened geopolitical concerns [
15,
36]. Advances in climate modelling have enhanced our understanding of Arctic dynamics, with high-resolution models better capturing the interactions between sea ice, atmospheric circulation, and ocean dynamics. However, uncertainties persist due to the complexity of feedback mechanisms [
45,
239]. Probabilistic forecasting systems, like IceNet, are improving the accuracy of sea ice projections, particularly for extreme events [
240]. Recent research highlights the necessity of incorporating Indigenous knowledge into climate governance to improve resilience and adaptation [
32,
42]. The Arctic’s geopolitical relevance requires international cooperation, as governance frameworks are crucial in addressing climate impacts, managing risks, and achieving climate policy targets [
14,
50,
75,
241].
7. Policy Responses and Mitigation Efforts
7.1. International Initiatives for Arctic Climate Protection
The susceptibility of the Arctic to climate change has led to several international efforts aimed at encouraging international cooperation, although the efficacy of these efforts varies [
242]. The Arctic Council, formed in 1996, is the premier intergovernmental platform for Arctic governments and Indigenous communities [
243,
244]. It has played an important role in producing critical reports, such as the Arctic Climate Impact Assessment (ACIA), which highlights the multifaceted effects of warming on the region [
245]. The Council supports research and sustainable resource use through initiatives, such as the Agreement on Enhancing International Arctic Scientific Cooperation (2017) and the Agreement to Prevent Unregulated Fishing in the Central Arctic Ocean (2018) [
75]. However, its inability to impose legally binding controls on carbon emissions or industrial activity limits its overall effectiveness. The Paris Agreement (2015), while global in scope, indirectly addresses Arctic warming by aiming to keep the global temperature rise to well below 2 °C [
246]. Arctic-specific solutions, such as enhanced financial support for adaptation in polar regions or tighter timelines for emissions reductions in high-latitude countries, are noticeably lacking [
247]. Similarly, the Polar Code, produced by the International Maritime Organisation, establishes rules for Arctic shipping but calls for broader coverage to offset the hazards of increased commercial activity, such as stronger limits on black carbon emissions and ballast water discharge [
248]. The lack of comprehensive agreements explicitly designed to address Arctic concerns, such as regulating oil and gas development or overseeing geoengineering experiments, emphasises the need for more international action. Proposals for a binding Arctic Climate Treaty, similar to the Antarctic Treaty System, remain contentious due to competing interests among Arctic states and industry [
249].
7.2. Regional Policies and Indigenous Involvement and Leadership
National policies in Arctic nations vary significantly, reflecting different priorities in balancing economic development and environmental conservation [
250]. Canada’s Arctic and Northern Policy Framework prioritises sustainable development, renewable energy transitions, and climate resilience, while also incorporating Indigenous knowledge into government policies to some extent [
251]. In contrast, Russia’s Arctic strategy focusses on resource extraction, particularly oil, gas, and minerals, and prioritises economic gains over environmental preservation [
252]. The Nordic countries have adopted ambitious green policies, such as transitioning to renewable energy and promoting marine protected areas, but there are still challenges in reducing emissions from the expanding maritime and tourism industries [
253]. Indigenous communities, important to the Arctic’s social and biological systems, are becoming more active in policymaking [
50]. Co-management initiatives demonstrate the importance of combining traditional ecological knowledge and scientific research to develop culturally sensitive adaptation strategies [
254]. Indigenous knowledge has proven especially useful in monitoring climate impacts, such as changes in species migration and permafrost degradation [
255]. However, the United Nations Declaration on the Rights of Indigenous Peoples (UNDRIP) emphasises the importance of increased political and financial empowerment in order to promote fair participation in decision making [
256]. Indigenous organisations advocate for greater influence in shaping regional and national policies, particularly those governing land use, resource extraction, and conservation [
40,
257,
258].
7.3. Mitigation Strategies to Reduce Arctic Warming
Mitigating Arctic warming demands an integrated approach combining traditional emission reduction strategies with novel technologies and regional conservation activities [
259]. Emission reduction remains crucial, with Arctic nations working to decarbonise their economy through renewable energy transitions, carbon pricing mechanisms, and commitments to net-zero targets [
260]. Policies to reduce black carbon emissions, a major contributor to Arctic warming, are critical [
261]. Regional initiatives, such as the Arctic Council Framework for Action on Enhanced Black Carbon and Methane Emissions Reductions, seek to restrict these pollutants, but enforcement is uneven [
262]. Geoengineering solutions, such as solar radiation management and Arctic ice restoration, are under investigation as prospective interventions, for example, using reflective aerosols to increase the Earth’s albedo or artificially recovering sea ice could help to decrease regional warming [
263,
264]. However, these methods pose ethical, environmental, and geopolitical risks, demanding stringent international control. Protected areas, such as the United States’ Arctic National Wildlife Refuge, are important for biodiversity conservation and carbon sequestration [
265]. Expanding these zones, particularly in maritime habitats, may lessen anthropogenic disruptions while providing refuge for vulnerable species. Cleaner shipping technologies and tougher rules on resource extraction activities help to achieve localised mitigation.
7.4. Arctic Governance and Adaptation Strategies
Climate change is causing significant environmental and socioeconomic changes in the Arctic, necessitating the implementation of strong governance frameworks and adaptation strategies. The combination of global climate dynamics, Indigenous knowledge systems, and geopolitical interests has made the Arctic a focus for scientific research and policy intervention. The Arctic Council serves as a major forum for collaboration among Arctic nations and Indigenous organisations [
266]. It has played an important role in addressing concerns, such as environmental preservation and sustainable development, but its non-binding agreements impede enforcement [
32]. Despite these efforts, gaps exist in the management of challenges, such as increased shipping, resource extraction, and environmental deterioration. For example, the rapid melting of sea ice has increased geopolitical competition for shipping routes and mineral resources, emphasising the need for stronger legal frameworks [
36,
50,
267]. International agreements, such as the United Nations Convention on the Law of the Sea, provide legal processes for settling disputes, particularly those involving marine boundaries [
268]. However, there are concerns that these frameworks are insufficient to handle the Arctic’s rapid environmental changes [
212]. The Arctic Fisheries Agreement of 2018, which outlaws unregulated fishing in the central Arctic Ocean, is a recent example of effective international cooperation [
269].
Indigenous Peoples, who have lived sustainably in the Arctic for centuries, play crucial roles in governance. However, their sovereignty and rights are frequently overlooked in decision-making processes [
270]. Studies emphasise the necessity of incorporating Indigenous knowledge into governance frameworks to achieve culturally suitable and sustainable adaptation policies [
31,
32,
42]. Collaborative governance approaches that emphasise resource co-management have showed potential, notably in mitigating the effects of climate change on traditional livelihoods [
51,
271,
272]. Fragmented policies and disparities in capacities among Arctic governments present considerable governance issues. Regional disparities in economic development, environmental interests, and legal frameworks impede coordinated responses to climate change [
42]. Resource extraction activities, which are often motivated by national interests, clash with conservation objectives, complicating governance efforts [
14].
Indigenous People are among the most susceptible to climate change, with traditional subsistence activities disrupted by changes in wildlife migration, sea ice loss, and permafrost thaw [
22,
32]. Adaptation measures that incorporate Indigenous knowledge, like community-based monitoring systems, have been beneficial in increasing resilience [
51,
273]. For example, collaborating on research with Indigenous stakeholders has resulted in a better knowledge and management of ecological changes [
30,
31,
32,
274]. Thawing permafrost and coastal erosion pose significant dangers to Arctic infrastructure, such as buildings, pipelines, and transportation networks [
19,
275]. Engineering solutions, such as permafrost-resistant materials and coastal defence systems, are being deployed, but their high cost and the distant locations of Arctic communities prevent widespread adoption [
276,
277,
278]. Policies that prioritise climate-resilient infrastructure are important for mitigating these risks [
152,
279,
280].
Ecosystem-based adaptation strategies prioritise biodiversity conservation and habitat restoration in order to preserve ecosystem services that benefit communities [
281]. Arctic wetlands, for example, play an important role in carbon sequestration and water control; hence, their preservation is a priority [
22]. Protecting vital habitats for migratory species has become an important adaptation strategy [
23]. Integrating climate adaptation into national and regional strategies is critical for good governance. Participatory scenario planning, which combines multiple stakeholder viewpoints, including those of Indigenous communities, has proven effective in developing adaptive management strategies [
282]. However, research shows that it frequently fails to fully integrate with Indigenous practices, emphasising the need for more inclusive approaches [
282]. Cooperative agreements, such as the Arctic Fisheries Agreement, illustrate the possibility for multilateral governance, but gaps exist in addressing the security implications of rising human activity in the region [
36,
50]. Strengthening international agreements on Arctic shipping, fishing, and resource management is important for balancing economic development and environmental sustainability.
Continuous monitoring with advanced technologies, such as satellite-based remote sensing, drones, and IoT-enabled sensors, is required for real-time data collection and processing [
283]. International collaboration, such as the Arctic Council, must be strengthened to enable coordinated research and policy implementation [
250]. Enhanced funding channels and collaborations are required to support Arctic-focused research activities [
284]. Raising worldwide awareness of Arctic changes and their repercussions can also help to mobilise support for these efforts.
8. Key Research Gaps
The Arctic has several research gaps that limit effective climate action and equitable policy development, particularly for Indigenous Peoples and the ecosystem. An important gap is the poor inclusion of Indigenous knowledge systems into larger scientific frameworks. Indigenous communities have built up extensive, place-based knowledge about Arctic ecosystems over generations, which is underutilised in scientific and policy discussions [
285]. The lack of this knowledge reduces the cultural significance and efficacy of adaptation strategies, particularly those aimed at traditional livelihoods, like hunting, fishing, and reindeer herding. The detailed overview of the key research gaps identified in this study is given in
Supplementary Table S1. Studies that combine Indigenous knowledge with science are critical for developing inclusive policies that empower communities and strengthen their resilience. Beyond Indigenous-specific gaps, larger ecological and socioeconomic uncertainties persist. There is a lack of understanding about Arctic marine ecosystems, specifically the combined effects of ocean acidification, changing sea ice dynamics, and altered food webs. Polar bears, walruses, and Arctic cod are among the species at risk due to these conditions, which threaten both ecological equilibrium and Indigenous sustenance. Economic gaps include a lack of comprehensive appraisals of growing industries, such as Arctic shipping and resource extraction. The effects of infrastructure expansion on vulnerable permafrost and ecosystems are poorly understood, leaving important concerns unanswered. These research gaps and future goals highlight the need for a comprehensive and collaborative approach to Arctic climate research, which ensures that scientific understanding keeps up with the rapid changes taking place in this crucial area. Addressing these gaps necessitates interdisciplinary research, more collaboration between Arctic and non-Arctic nations, and a focus on participatory techniques that incorporate Indigenous perspectives at all stages of development. By addressing these research gaps, the Arctic can serve as a paradigm for integrated, sustainable, and inclusive climate response.
Despite growing interest in Arctic climate change, several key research gaps remain. These include limited high-resolution data across space and seasons, insufficient integration of interacting climate drivers (e.g., greenhouse gases, albedo, and atmospheric dynamics), and unclear linkages between Arctic warming and mid-latitude weather extremes. Moreover, local-scale impacts on ecosystems and Indigenous communities are underexplored, and observational data—especially in remote and winter-dominated regions—remain sparse. This study contributes by synthesising multiple drivers and highlighting the Arctic’s role as a critical zone of interconnected global change.
9. Recommendations for Addressing Climate Change in the Arctic
The Arctic is warming at an unprecedented rate, affecting global climate systems, ecosystems, and local communities. These recommendations provide a comprehensive strategy for addressing climate change in the Arctic, incorporating risk assessment, monitoring, mitigation, community participation, and policy formulation to promote resilience and sustainability. Effective risk assessment is essential to tackle Arctic climate challenges. Geospatial analysis can reveal hotspots for permafrost degradation, glacier retreat, and carbon release, as well as vulnerabilities in Arctic biodiversity and infrastructure. Creating multi-hazard vulnerability maps that integrate risks, such as ice melt, coastal erosion, and storm surges, allows initiatives to be prioritised more effectively. By including socioeconomic scenarios in climate models, policymakers can assess risks to livelihoods, food security, and cultural traditions, allowing them to address both local and global implications. Robust monitoring and early warning systems are required to track climate dynamics in real time. Advanced remote sensing technology, such as AI-enhanced satellites and autonomous underwater drones, can provide useful information about ice dynamics, permafrost thaw, and ecosystem changes. Localised field investigations that combine scientific methods with indigenous knowledge are important for understanding changes in animal migration patterns, ecosystem health, and social aspects. Early warning systems, powered by AI and integrated into global networks, can predict extreme weather and ice collapse events, improving preparedness for downstream impacts, like sea-level rise.
Large-scale renewable energy projects, such as offshore wind farms and hydrogen energy hubs, can replace fossil fuels, while localised solar power systems lessen the need for diesel generators in remote locations. Restoring Arctic wetlands and peatlands is a natural carbon sequestration approach, and novel technologies, such as carbon capture and biochar synthesis, can further reduce emissions. Geoengineering options, such as maritime cloud brightening and artificial ice floe creation, must be investigated carefully, with international frameworks insuring ethical and effective implementation. Marine ecosystem preservation, including protected areas and sustainable fisheries management, is important for biodiversity and carbon balance. Community engagement is critical to Arctic climate action. Capacity-sharing is important to promote sustainable practices and prepare future leaders for climate science and resilience approaches. Arctic climate councils, which comprise Indigenous leaders, scientists, and policymakers, facilitate collaborative decision making and support inclusive governance. Integrating Indigenous ecological knowledge into climate policies is important to strengthen research and adaptation while encouraging shared ownership of climate solutions. The recommendations for addressing climate change impacts in the Arctic are illustrated in
Figure 7, while a detailed overview of each component of recommendations is given in
Supplementary Table S1.
Policy frameworks must be strengthened in order for Arctic climate governance to function effectively. A legally binding Arctic Climate Treaty could impose limitations on emissions and resource extraction, while strengthening the Polar Code will ensure more stringent management of shipping emissions and pollution. Arctic-specific climate funds can help fund renewable energy, conservation, and community resilience initiatives, but methods for compensating Indigenous communities for climate-related losses must be established. Regional and global collaborations, such as those organised by the Arctic Council, can help to coordinate action and knowledge sharing. Technological advancements provide chances to strengthen Arctic resilience. Climate-resilient infrastructure, such as floating structures and adaptable pipelines, can resist severe weather and permafrost melt. Decentralised renewable energy systems and innovative battery storage technologies can provide stable electricity in remote Arctic locations, while IoT-based monitoring systems track vulnerabilities in real time, allowing for quick responses to emerging problems. Restoring Arctic habitats is important for minimising climate impacts. Wildlife corridors can help species migrate, and rewilding degraded landscapes with native vegetation increases biodiversity and carbon sequestration. Developing resilient agricultural practices, such as cold-resistant crops and agroforestry, can promote food security while decreasing reliance on imported foods.
Education and awareness about Arctic climate challenges is critical. Global campaigns emphasising the Arctic’s role in the climate system, combined with contemporary techniques, such as multimedia and virtual reality, can reach a larger audience. Citizen science programs can enable Arctic people to participate in environmental monitoring while also encouraging collaboration between scientists and local volunteers to increase data accuracy and coverage. This paradigm emphasises the importance of a collaborative and integrated strategy to combating climate change in the Arctic. Combining science, technology, policy, and community engagement is important to protect the Arctic’s ecological, cultural, and climatic integrity while addressing its global relevance. Collaborative actions are important to ensuring a sustainable and resilient future for the Arctic and beyond.
10. Conclusions
This study highlights the critical function of the Arctic as a planetary health indicator and climate regulator, demonstrating the cascading effects of Arctic warming on global systems. The review shows that climate change in the Arctic is not an isolated phenomenon but rather the result of intricate interactions between regional feedback mechanisms that have a significant impact on ecosystems, socioeconomic systems, and the stability of the global climate. Permafrost thawing, ecosystem changes, and the rapid melting of sea ice necessitate immediate and decisive action on multiple fronts. Critical knowledge gaps still exist despite tremendous advancements in our understanding of the causes and effects of Arctic warming, especially when it comes to addressing the complex socioeconomic issues that Arctic communities face and incorporating Indigenous ecological knowledge with scientific frameworks. Interdisciplinary approaches must be given top priority in future studies in order to fully understand the intricate relationships between permafrost carbon cycles, ocean circulation, and extreme weather patterns. To improve the projections of Arctic climate trajectories and their worldwide impacts, sophisticated climate models that integrate probabilistic forecasting and high-resolution data are crucial. Longitudinal studies that evaluate the long-term sociocultural effects of climate change on Indigenous communities—including their traditional livelihoods, cultural heritage, and resilience strategies—are also urgently required.
The policy recommendations emerging from this study emphasize the need for an Arctic-specific climate governance framework, which should include legally binding agreements to control region-specific emissions, resource extraction, and shipping operations. To promote fair and successful adaptation plans, collaborative governance models that integrate Indigenous leadership and co-management principles are crucial. To guarantee a balanced approach to conservation and economic potential, international cooperation may be strengthened through organisations, like the Arctic Council, and sustainable development principles can be included into Arctic policies. Finally, to reduce the effects of Arctic climate change and promote resilience, transformative actions are needed. Among these are investments in renewable energy initiatives, ecosystem-based adaptation strategies, and climate-resilient infrastructure designed to meet the particular difficulties faced by the Arctic. Public partnerships and increased global understanding of the Arctic’s crucial role in climate systems can help to mobilise resources and support for these initiatives.
Author Contributions
Conceptualization, I.H.M., R.A. and J.D.F.; methodology, I.H.M., R.A., J.D.F. and A.R.H.; software, R.A.; validation, I.H.M., R.A., J.D.F. and A.R.H.; formal analysis, I.H.M., R.A., J.D.F. and A.R.H.; investigation, I.H.M. and R.A.; resources, I.H.M., R.A. and J.D.F.; data curation, I.H.M. and R.A.; writing—original draft preparation, I.H.M., R.A. and J.D.F.; writing—review and editing, I.H.M., R.A., J.D.F. and A.R.H.; visualization, R.A.; supervision, J.D.F.; project administration, I.H.M. and J.D.F.; funding acquisition, I.H.M. and J.D.F. All authors have read and agreed to the published version of the manuscript.
Funding
This research was funded by UK Research and Innovation (Canada-Inuit Nunangat-United Kingdom (CINUK) Arctic Research Programme’s IMAGINE Project: NE/X003868/1), the UK’s Department for Science, Innovation and Technology, Foreign, Commonwealth and Development Office (FCDO) and NERC Arctic Office (ARCWISE Project), and the ERC Advanced Grant (via the UKRI Horizon Europe guarantee scheme, EPSRC grant# EP/Z533385/1). The APC was funded by the University of Leeds, UK and MDPI.
Data Availability Statement
Data are contained within the article.
Conflicts of Interest
The authors declare no conflict of interest.
References
- Mosoni, C.; Hildén, M.; Fronzek, S.; Reyer, C.P.; Carter, T.R. Cross-border dimensions of Arctic climate change impacts and implications for Europe. Wiley Interdiscip. Rev. Clim. Chang. 2024, 15, e905. [Google Scholar] [CrossRef]
- Bonan, G.B.; Doney, S.C. Climate, ecosystems, and planetary futures: The challenge to predict life in Earth system models. Science 2018, 359, eaam8328. [Google Scholar] [CrossRef] [PubMed]
- Koenigk, T.; Key, J.; Vihma, T. Climate change in the Arctic. In Physics and Chemistry of the Arctic Atmosphere; Springer: Berlin/Heidelberg, Germany, 2020; pp. 673–705. [Google Scholar]
- Previdi, M.; Smith, K.L.; Polvani, L.M. Arctic amplification of climate change: A review of underlying mechanisms. Environ. Res. Lett. 2021, 16, 093003. [Google Scholar] [CrossRef]
- Shu, Q.; Wang, Q.; Årthun, M.; Wang, S.; Song, Z.; Zhang, M.; Qiao, F. Arctic Ocean Amplification in a warming climate in CMIP6 models. Sci. Adv. 2022, 8, eabn9755. [Google Scholar] [CrossRef] [PubMed]
- Ivanov, V. Arctic sea ice loss enhances the oceanic contribution to climate change. Atmosphere 2023, 14, 409. [Google Scholar] [CrossRef]
- Barkhordarian, A.; Nielsen, D.M.; Olonscheck, D.; Baehr, J. Arctic marine heatwaves forced by greenhouse gases and triggered by abrupt sea-ice melt. Commun. Earth Environ. 2024, 5, 57. [Google Scholar] [CrossRef]
- Nimma, D.; Devi, O.R.; Laishram, B.; Ramesh, J.V.N.; Boddupalli, S.; Ayyasamy, R.; Tirth, V.; Arabil, A. Implications of climate change on freshwater ecosystems and their biodiversity. Desalin. Water Treat. 2025, 321, 100889. [Google Scholar] [CrossRef]
- Rantanen, M.; Karpechko, A.Y.; Lipponen, A.; Nordling, K.; Hyvärinen, O.; Ruosteenoja, K.; Vihma, T.; Laaksonen, A. The Arctic has warmed nearly four times faster than the globe since 1979. Commun. Earth Environ. 2022, 3, 168. [Google Scholar] [CrossRef]
- You, Q.; Cai, Z.; Pepin, N.; Chen, D.; Ahrens, B.; Jiang, Z.; Wu, F.; Kang, S.; Zhang, R.; Wu, T.; et al. Warming amplification over the Arctic Pole and Third Pole: Trends, mechanisms and consequences. Earth-Sci. Rev. 2021, 217, 103625. [Google Scholar] [CrossRef]
- Cai, Q.; Wang, J.; Beletsky, D.; Overland, J.; Ikeda, M.; Wan, L. Accelerated decline of summer Arctic sea ice during 1850–2017 and the amplified Arctic warming during the recent decades. Environ. Res. Lett. 2021, 16, 034015. [Google Scholar] [CrossRef]
- England, M.R.; Eisenman, I.; Lutsko, N.J.; Wagner, T.J. The recent emergence of Arctic amplification. Geophys. Res. Lett. 2021, 48, e2021GL094086. [Google Scholar] [CrossRef]
- Choudhary, S.; Saalim, S.M.; Khare, N. Climate change over the Arctic: Impacts and assessment. In Understanding Present and Past Arctic Environments; Elsevier: Amsterdam, The Netherlands, 2021; pp. 1–14. [Google Scholar]
- Hwang, B.; Landy, J. Impacts of climate change on Arctic sea ice. MCCIP Sci. Rev. 2020, 2020, 208–227. [Google Scholar]
- Yaisien, Y.; Fayek, Y.; Sharawi, H. Climate Change and its Profound Effects on Marine Climate. Fusion Multidiscip. Res. Int. J. 2023, 4, 432–444. [Google Scholar]
- Jenkins, M.; Dai, A. The impact of Sea-Ice loss on Arctic climate feedbacks and their role for Arctic Amplification. Geophys. Res. Lett. 2021, 48, e2021GL094599. [Google Scholar] [CrossRef]
- Moore, G.W.K.; Steele, M.; Schweiger, A.J.; Zhang, J.; Laidre, K.L. Thick and old sea ice in the Beaufort Sea during summer 2020/21 was associated with enhanced transport. Commun. Earth Environ. 2022, 3, 198. [Google Scholar] [CrossRef]
- Miner, K.R.; Turetsky, M.R.; Malina, E.; Bartsch, A.; Tamminen, J.; McGuire, A.D.; Fix, A.; Sweeney, C.; Elder, C.D.; Miller, C.E. Permafrost carbon emissions in a changing Arctic. Nat. Rev. Earth Environ. 2022, 3, 55–67. [Google Scholar] [CrossRef]
- Suter, L.; Streletskiy, D.; Shiklomanov, N. Assessment of the cost of climate change impacts on critical infrastructure in the circumpolar Arctic. Polar Geogr. 2019, 42, 267–286. [Google Scholar] [CrossRef]
- Macdonald, R.W.; Harner, T.; Fyfe, J. Recent climate change in the Arctic and its impact on contaminant pathways and interpretation of temporal trend data. Sci. Total Environ. 2005, 342, 5–86. [Google Scholar] [CrossRef]
- Badina, S.; Pankratov, A. Assessment of the Impacts of Climate Change on the Russian Arctic Economy (including the Energy Industry). Energies 2022, 15, 2849. [Google Scholar] [CrossRef]
- Phoenix, G.K.; Treharne, R. Arctic greening and browning: Challenges and a cascade of complexities. Glob. Chang. Biol. 2022, 28, 3481–3483. [Google Scholar] [CrossRef]
- Lebrun, A.; Comeau, S.; Gazeau, F.; Gattuso, J.P. Impact of climate change on Arctic macroalgal communities. Glob. Planet. Chang. 2022, 219, 103980. [Google Scholar] [CrossRef]
- Brandt, S.; Wassmann, P.; Piepenburg, D. Revisiting the footprints of climate change in Arctic marine food webs: An assessment of knowledge gained since 2010. Front. Mar. 2023, 10, 1096222. [Google Scholar] [CrossRef]
- Wassmann, P.; Duarte, C.M.; Agusti, S.; Sejr, M.K. Footprints of climate change in the Arctic marine ecosystem. Glob. Chang. Biol. 2011, 17, 1235–1249. [Google Scholar] [CrossRef]
- Lebrun, M.; Vancoppenolle, M.; Madec, G.; Massonnet, F. Arctic sea-ice-free season projected to extend into autumn. Cryosphere 2019, 13, 79–96. [Google Scholar] [CrossRef]
- Gelfan, A.; Gustafsson, D.; Motovilov, Y.; Arheimer, B.; Kalugin, A.; Krylenko, I.; Lavrenov, A. Climate change impact on the water regime of two great Arctic rivers: Modeling and uncertainty issues. Clim. Chang. 2017, 141, 499–515. [Google Scholar] [CrossRef]
- McCrystall, M.R.; Stroeve, J.; Serreze, M.; Forbes, B.C.; Screen, J.A. New climate models reveal faster and larger increases in Arctic precipitation than previously projected. Nat. Commun. 2021, 12, 6765. [Google Scholar] [CrossRef] [PubMed]
- Vogel, B.; Bullock, R.C. Institutions, indigenous peoples, and climate change adaptation in the Canadian Arctic. GeoJournal 2021, 86, 2555–2572. [Google Scholar] [CrossRef]
- Bogdanova, E.; Filant, K.; Ivanova, M.; Romanenko, T.; Voronina, L.; Hossain, K.; Filant, P.; Andronov, S.; Lobanov, A. Strengthening collaboration of the Indigenous peoples in the Russian Arctic: Adaptation in the COVID-19 Pandemic Times. Sustainability 2022, 14, 3225. [Google Scholar] [CrossRef]
- Malik, I.H. Can political ecology be decolonised? A dialogue with Paul Robbins. GEO Geogr. Environ. 2024, 11, e00140. [Google Scholar] [CrossRef]
- Ford, J.D.; Pearce, T.; Canosa, I.V.; Harper, S. The rapidly changing Arctic and its societal implications. Wiley Interdiscip. Rev. Clim. Chang. 2021, 12, e735. [Google Scholar] [CrossRef]
- Malik, I.H.; Ford, J.D.; Winters, I.; Hunter, B.; Flowers, N.; Quincey, D.; Flowers, K.; Flowers, M.; Coombs, D.; Foltz-Vincent, C.; et al. Monitoring climate change impacts, Indigenous livelihoods, and adaptation: Perspectives from Inuit community of Hopedale, Nunatsiavut, Canada. Camb. Prism. Coast. Futures 2025, 1–31. [Google Scholar] [CrossRef]
- Eerkes-Medrano, L.; Huntington, H.P. Untold stories: Indigenous knowledge beyond the changing Arctic cryosphere. Front. Clim. 2021, 3, 675805. [Google Scholar] [CrossRef]
- Huntington, H.P.; Boyle, M.; Flowers, G.E.; Weatherly, J.W.; Hamilton, L.C.; Hinzman, L.; Gerlach, C.; Zulueta, R.; Nicolson, C.; Overpeck, J. The influence of human activity in the Arctic on climate and climate impacts. Clim. Chang. 2007, 82, 77–92. [Google Scholar] [CrossRef]
- Ng, A.K.; Andrews, J.; Babb, D.; Lin, Y.; Becker, A. Implications of climate change for shipping: Opening the Arctic seas. Wiley Interdiscip. Rev. Clim. Chang. 2018, 9, e507. [Google Scholar] [CrossRef]
- Malik, I.H.; Ford, J.D. Barriers and limits to adaptation in the Arctic. Curr. Opin. Environ. Sustain. 2025, 73, 101519. [Google Scholar] [CrossRef]
- Malik, I.H.; Ford, J.D. Understanding the Impacts of Arctic Climate Change Through the Lens of Political Ecology. Wiley Interdiscip. Rev. Clim. Chang. 2025, 16, e927. [Google Scholar] [CrossRef]
- AMAP. Arctic Climate Change Update 2021: Key Trends and Impacts. Summary for Policy-Makers; Arctic Monitoring and Assessment Programme (AMAP): Oslo, Norway, 2021. [Google Scholar]
- Inuit Circumpolar Council. Ethical and Equitable Engagement Synthesis Report: A Collection of Inuit Rules, Guidelines, Protocols, and Values for the Engagement of Inuit Communities and Indigenous Knowledge from Across Inuit Nunaat; Synthesis Report; International: Anadyr, Russia, 2021. [Google Scholar]
- Schmale, J.; Zieger, P.; Ekman, A.M. Aerosols in current and future Arctic climate. Nat. Clim. Chang. 2021, 11, 95–105. [Google Scholar] [CrossRef]
- Stephen, K. Societal impacts of a rapidly changing Arctic. Curr. Clim. Chang. Rep. 2018, 4, 223–237. [Google Scholar] [CrossRef]
- Giesse, C.; Notz, D.; Baehr, J. The shifting distribution of Arctic daily temperatures under global warming. Earths Future 2024, 12, e2024EF004961. [Google Scholar] [CrossRef]
- Zhang, X.; Jiao, Z.; Zhao, C.; Qu, Y.; Liu, Q.; Zhang, H.; Tong, Y.; Wang, C.; Li, S.; Guo, J.; et al. Review of land surface albedo: Variance characteristics, climate effect and management strategy. Remote Sens. 2022, 14, 1382. [Google Scholar] [CrossRef]
- Vavrus, S.J.; Holland, M.M.; Jahn, A.; Bailey, D.A.; Blazey, B.A. Twenty-first-century Arctic climate change in CCSM4. J. Clim. 2012, 25, 2696–2710. [Google Scholar] [CrossRef]
- Rixen, C.; Høye, T.T.; Macek, P.; Aerts, R.; Alatalo, J.M.; Anderson, J.T.; Arnold, P.A.; Barrio, I.C.; Bjerke, J.W.; Björkman, M.P.; et al. Winters are changing: Snow effects on Arctic and alpine tundra ecosystems. Arct. Sci. 2022, 8, 572–608. [Google Scholar] [CrossRef]
- Virkkala, A.-M.; Rogers, B.M.; Watts, J.D.; Arndt, K.A.; Potter, S.; Wargowsky, I.; Schuur, E.A.G.; See, C.R.; Mauritz, M.; Boike, J.; et al. Wildfires offset the increasing but spatially heterogeneous Arctic–boreal CO2 uptake. Nat. Clim. Chang. 2025, 15, 188–195. [Google Scholar] [CrossRef]
- Rodenhizer, H.; Belshe, F.; Celis, G.; Ledman, J.; Mauritz, M.; Goetz, S.; Sankey, T.; Schuur, E.A. Abrupt permafrost thaw accelerates carbon dioxide and methane release at a tussock tundra site. Arct. Antarct. Alp. Res. 2022, 54, 443–464. [Google Scholar] [CrossRef]
- Heijmans, M.M.P.D.; Magnússon, R.Í.; Lara, M.J.; Frost, G.V.; Myers-Smith, I.H.; van Huissteden, J.; Jorgenson, M.T.; Fedorov, A.N.; Epstein, H.E.; Lawrence, D.M.; et al. Tundra vegetation change and impacts on permafrost. Nat. Rev. Earth Environ. 2022, 3, 68–84. [Google Scholar] [CrossRef]
- Arruda, G.M.; Krutkowski, S. Social impacts of climate change and resource development in the Arctic: Implications for Arctic governance. J. Enterp. Communities People Places Glob. Econ. 2017, 11, 277–288. [Google Scholar] [CrossRef]
- Johnson, N.; Behe, C.; Danielsen, F.; Krümmel, E.M.; Nickels, S.; Pulsifer, P.L. Community-Based Monitoring and Indigenous Knowledge in a Changing Arctic: A Review for the Sustaining Arctic Observing Networks. 2016; Volume 9, 74p, Available online: https://iccalaska.org/wp-icc/wp-content/uploads/2016/05/Community-Based-Monitoring-and-Indigenous-Knowledge-in-a-Changing-Arctic_web.pdf (accessed on 15 January 2025).
- Ingvaldsen, R.B.; Assmann, K.M.; Primicerio, R.; Fossheim, M.; Polyakov, I.V.; Dolgov, A.V. Physical manifestations and ecological implications of Arctic Atlantification. Nat. Rev. Earth Environ. 2021, 2, 874–889. [Google Scholar] [CrossRef]
- Callaghan, T.V.; Björn, L.O.; Chapin, F.S., III; Chernov, Y.; Christensen, T.R.; Huntley, B.; Imms, R.; Johansson, M.; Riedlinger, D.J.; Jonasson, S.; et al. Arctic tundra and polar desert ecosystems. In Arctic Climate Impact Assessment; Symon, C., Arris, L., Heal, B., Eds.; Cambridge University Press: Cambridge, UK, 2005; pp. 243–352. [Google Scholar]
- Stern, G.A.; Macdonald, R.W.; Outridge, P.M.; Wilson, S.; Chételat, J.; Cole, A.; Hintelmann, H.; Bustnes, J.O.; Gabrielsen, G.W.; Verreault, J. Climate variability and temporal trends of persistent organic pollutants in the Arctic: A study of glaucous gulls. Environ. Sci. Technol. 2010, 44, 3155–3161. [Google Scholar]
- McKinney, M.A.; Chételat, J.; Burke, S.M.; Elliott, K.H.; Fernie, K.J.; Houde, M.; Kahilainen, K.K.; Letcher, R.J.; Morris, A.D.; Muir, D.C.G.; et al. Climate change and mercury in the Arctic: Biotic interactions. Sci. Total Environ. 2022, 834, 155221. [Google Scholar] [CrossRef]
- Dudley, J.P.; Hoberg, E.P.; Jenkins, E.J.; Parkinson, A.J. Climate Change in the North American Arctic: A One Health Perspective. Ecohealth 2015, 12, 713–725. [Google Scholar] [CrossRef]
- de Wit, C.A.; Vorkamp, K.; Muir, D. Influence of climate change on persistent organic pollutants and chemicals of emerging concern in the Arctic: State of knowledge and recommendations for future research. Environ. Sci. Process. Impacts 2022, 24, 1530–1543. [Google Scholar] [CrossRef]
- Ford, J.D.; McDowell, G.; Jones, J. The state of climate change adaptation in the Arctic. Environ. Res. Lett. 2014, 9, 104005. [Google Scholar] [CrossRef]
- Vincent, W.F. Arctic climate change: Local impacts, global consequences, and policy implications. In The Palgrave Handbook of Arctic Policy and Politics; Palgrave Macmillan: Cham, Switzerland, 2020; pp. 507–526. [Google Scholar]
- Sesana, E.; Gagnon, A.S.; Ciantelli, C.; Cassar, J.A.; Hughes, J.J. Climate change impacts on cultural heritage: A literature review. WIREs Clim. Chang. 2021, 12, e710. [Google Scholar] [CrossRef]
- Nicu, I.C.; Fatorić, S. Climate change impacts on immovable cultural heritage in polar regions: A systematic bibliometric review. WIREs Clim. Chang. 2023, 14, e822. [Google Scholar] [CrossRef]
- Heininen, L. State of the Arctic strategies and policies—A summary. Arct. Yearb. 2012, 2012, 2–47. [Google Scholar]
- Olsen, A. The Arctic Environment. In Ship Operations in Extreme Low Temperature Environments; Springer Nature: Cham, Switzerland, 2024; pp. 13–49. [Google Scholar]
- Epstein, H.E.; Myers-Smith, I.; Walker, D.A. Recent dynamics of arctic and sub-arctic vegetation. Environ. Res. Lett. 2013, 8, 015040. [Google Scholar] [CrossRef]
- Walker, D.A.; Daniëls, F.J.A.; Alsos, I.; Bhatt, U.S.; Breen, A.L.; Buchhorn, M.; Bültmann, H.; Druckenmiller, L.A.; Edwards, M.E.; Ehrich, D.; et al. Circumpolar Arctic vegetation: A hierarchic review and roadmap toward an internationally consistent approach to survey, archive and classify tundra plot data. Environ. Res. Lett. 2016, 11, 055005. [Google Scholar] [CrossRef]
- Arctic Centre. The Arctic Region. Arctic Centre University of Lapland. 2025. Available online: https://www.arcticcentre.org/EN/arcticregion (accessed on 15 February 2025).
- Poseidon Expeditions. Temperature in the Arctic Circle. 2025. Available online: https://poseidonexpeditions.com/about/articles/temperature-in-arctic-circle/ (accessed on 12 March 2025).
- Poltronieri, A.; Bochow, N.; Boers, N.; Rypdal, M. Increasing fluctuations in the Arctic summer sea ice cover are expected with future global warming. Environ. Res. Clim. 2024, 3, 035007. [Google Scholar] [CrossRef]
- Adeniran-Obey, S.O.; Kayode-Edwards, I.I.; Onwaeze, O.O. Arctic Marine Ecosystems. In Arctic Marine Ecotoxicology; Springer: Cham, Switzerland, 2024; pp. 45–69. [Google Scholar]
- Baxter, I.; Ding, Q. An optimal atmospheric circulation mode in the Arctic favoring strong Summertime Sea ice melting and ice–albedo feedback. J. Clim. 2022, 35, 6627–6645. [Google Scholar] [CrossRef]
- NSIDC. Sea Ice Today: Analyses and Daily Images of Sea Ice Conditions; National Snow and Ice Data Center: Boulder, CO, USA, 2025. [Google Scholar]
- Kayode-Edwards, J.O.; Kayode-Edwards, I.I.; Kayode-Edwards, D.O. Climate Change in the Arctic. In Arctic Marine Ecotoxicology; Springer: Cham, Switzerland, 2024. [Google Scholar]
- Hanaček, K.; Kröger, M.; Scheidel, A.; Rojas, F.; Martinez-Alier, J. On thin ice–The Arctic commodity extraction frontier and environmental conflicts. Ecol. Econ. 2022, 191, 107247. [Google Scholar] [CrossRef]
- Arctic Council Secretariat. The Arctic Council: A Quick Guide, 3rd ed.; Arctic Council Secretariat: Tromsø, Norway, 2023; 36p. [Google Scholar]
- Forbis, R., Jr.; Hayhoe, K. Does Arctic governance hold the key to achieving climate policy targets? Environ. Res. Lett. 2018, 13, 020201. [Google Scholar] [CrossRef]
- Comiso, J. Polar Oceans from Space; Springer Science & Business Media: Berlin, Germany, 2010; Volume 41. [Google Scholar]
- Hall, R.J.; Hanna, E.; Chen, L. Winter Arctic Amplification at the synoptic timescale, 1979–2018, its regional variation and response to tropical and extratropical variability. Clim. Dyn. 2021, 56, 457–473. [Google Scholar] [CrossRef]
- Nab, C.; Mallett, R.; Gregory, W.; Landy, J.; Lawrence, I.; Willatt, R.; Stroeve, J.; Tsamados, M. Synoptic variability in satellite altimeter-derived radar freeboard of Arctic sea ice. Geophys. Res. Lett. 2023, 50, e2022GL100696. [Google Scholar] [CrossRef]
- Huang, Y.; Zhang, W.; Zhang, S.; Jin, F.; Fang, C.; Ma, X.; Wang, J.; Mu, J. Systematical insights into distribution and characteristics of microplastics in near-surface waters from the East Asian Seas to the Arctic Central Basin. Sci. Total Environ. 2022, 814, 151923. [Google Scholar] [CrossRef] [PubMed]
- Timmermans, M.L.; Marshall, J. Understanding Arctic Ocean circulation: A review of ocean dynamics in a changing climate. J. Geophys. Res. Ocean. 2020, 125, e2018JC014378. [Google Scholar] [CrossRef]
- Marshall, J. The Natural Resources of the Arctic and International Law: How the International System Manages Arctic Resources. Columbia Undergrad. Law Rev. 2012, 7, 1. [Google Scholar]
- Wassmann, P.; Krause-Jensen, D.; Bluhm, B.A.; Janout, M. Towards a Unifying Pan-Arctic Perspective of the Contemporary and Future Arctic Ocean. Front. Mar. Sci. 2021, 8, 678420. [Google Scholar] [CrossRef]
- Li, Z.; Ding, Q.; Steele, M.; Schweiger, A. Recent upper Arctic Ocean warming expedited by summertime atmospheric processes. Nat. Commun. 2022, 13, 362. [Google Scholar] [CrossRef]
- IPCC. Climate Change 2021: The Physical Science Basis. Contribution of Working Group 1 to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Masson-Delmotte, V., Zhai, P., Pirani, A., Connors, S.L., Péan, C., Berger, S., Caud, N., Chen, Y., Huang, M., Zhou, B., et al., Eds.; Cambridge University Press: Cambridge, UK, 2021. [Google Scholar]
- Screen, J.A.; Simmonds, I. The central role of diminishing sea ice in recent Arctic temperature amplification. Nature 2010, 464, 1334–1337. [Google Scholar] [CrossRef]
- Serreze, M.C.; Barry, R.G. Processes and impacts of Arctic amplification: A research synthesis. Glob. Planet. Chang. 2011, 77, 85–96. [Google Scholar] [CrossRef]
- Graversen, R.G.; Burtu, M. Arctic amplification enhanced by latent energy transport of atmospheric planetary waves. Q. J. R. Meteorol. Soc. 2016, 142, 2046–2054. [Google Scholar] [CrossRef]
- Lee, S. Testing of the tropically excited Arctic warming mechanism (TEAM) with traditional El Niño and La Niña. J. Clim. 2012, 25, 4015–4022. [Google Scholar] [CrossRef]
- Flournoy, M.D.; Feldstein, S.B.; Lee, S.; Clothiaux, E.E. Exploring the tropically excited Arctic warming mechanism with station data: Links between tropical convection and Arctic downward infrared radiation. J. Atmos. Sci. 2016, 73, 1143–1158. [Google Scholar] [CrossRef]
- Ding, Q.; Wallace, J.M.; Battisti, D.S.; Steig, E.J.; Gallant, A.J.; Kim, H.J.; Geng, L. Tropical forcing of the recent rapid Arctic warming in northeastern Canada and Greenland. Nature 2014, 509, 209–212. [Google Scholar] [CrossRef]
- Clark, J.P.; Lee, S. The role of the tropically excited Arctic warming mechanism on the warm Arctic cold continent surface air temperature trend pattern. Geophys. Res. Lett. 2019, 46, 8490–8499. [Google Scholar] [CrossRef]
- Liang, Y.; Bi, H.; Lei, R.; Vihma, T.; Huang, H. Atmospheric latent energy transport pathways into the Arctic and their connections to sea ice loss during winter over the observational period. J. Clim. 2023, 36, 6695–6712. [Google Scholar] [CrossRef]
- Baggett, C.; Lee, S. An identification of the mechanisms that lead to Arctic warming during planetary-scale and synoptic-scale wave life cycles. J. Atmos. Sci. 2017, 74, 1859–1877. [Google Scholar] [CrossRef]
- Yoshimori, M.; Abe-Ouchi, A.; Laîné, A. The role of atmospheric heat transport and regional feedbacks in the Arctic warming at equilibrium. Clim. Dyn. 2017, 49, 3457–3472. [Google Scholar] [CrossRef]
- Graversen, R.G.; Langen, P.L. On the role of the atmospheric energy transport in 2 × CO2—Induced polar amplification in CESM1. J. Clim. 2019, 32, 3941–3956. [Google Scholar] [CrossRef]
- Oldenburg, D.; Kwon, Y.O.; Frankignoul, C.; Danabasoglu, G.; Yeager, S.; Kim, W.M. The Respective Roles of Ocean Heat Transport and Surface Heat Fluxes in Driving Arctic Ocean Warming and Sea Ice Decline. J. Clim. 2024, 37, 1431–1448. [Google Scholar] [CrossRef]
- Karcher, M.J.; Gerdes, R.; Kauker, F.; Köberle, C. Arctic warming: Evolution and spreading of the 1990s warm event in the Nordic seas and the Arctic Ocean. J. Geophys. Res. Ocean. 2003, 108. [Google Scholar] [CrossRef]
- Spielhagen, R.F.; Werner, K.; Sørensen, S.A.; Zamelczyk, K.; Kandiano, E.; Budeus, G.; Husum, K.; Marchitto, T.M.; Hald, M. Enhanced modern heat transfer to the Arctic by warm Atlantic water. Science 2011, 331, 450–453. [Google Scholar] [CrossRef] [PubMed]
- Holland, M.M.; Bitz, C.M.; Tremblay, B. Future abrupt reductions in the summer Arctic sea ice. Geophys. Res. Lett. 2006, 33. [Google Scholar] [CrossRef]
- Hwang, Y.T.; Frierson, D.M.; Kay, J.E. Coupling between Arctic feedbacks and changes in poleward energy transport. Geophys. Res. Lett. 2011, 38. [Google Scholar] [CrossRef]
- Mahlstein, I.; Knutti, R. Ocean heat transport as a cause for model uncertainty in projected Arctic warming. J. Clim. 2011, 24, 1451–1460. [Google Scholar] [CrossRef]
- Kirbus, B.; Tiedeck, S.; Camplani, A.; Chylik, J.; Crewell, S.; Dahlke, S.; Ebell, K.; Gorodetskaya, I.; Griesche, H.; Handorf, D.; et al. Surface impacts and associated mechanisms of a moisture intrusion into the Arctic observed in mid-April 2020 during MOSAiC. Front. Earth Sci. 2023, 11, 1147848. [Google Scholar] [CrossRef]
- Graham, R.M.; Cohen, L.; Petty, A.A.; Boisvert, L.N.; Rinke, A.; Hudson, S.R.; Nicolaus, M.; Granskog, M.A. Increasing frequency and duration of Arctic winter warming events. Geophys. Res. Lett. 2017, 44, 6974–6983. [Google Scholar] [CrossRef]
- Hegyi, B.M.; Taylor, P.C. The unprecedented 2016–2017 Arctic sea ice growth season: The crucial role of atmospheric rivers and longwave fluxes. Geophys. Res. Lett. 2018, 45, 5204–5212. [Google Scholar] [CrossRef]
- Serreze, M.C.; Stroeve, J.; Barrett, A.P.; Boisvert, L.N. Summer atmospheric circulation anomalies over the Arctic Ocean and their influences on September sea ice extent: A cautionary tale. J. Geophys. Res. Atmos. 2016, 121, 11–463. [Google Scholar] [CrossRef]
- Overland, J.; Dunlea, E.; Box, J.E.; Corell, R.; Forsius, M.; Kattsov, V.; Olsen, M.S.; Pawlak, J.; Reiersen, L.-O.; Wang, M. The urgency of Arctic change. Polar Sci. 2019, 21, 6–13. [Google Scholar] [CrossRef]
- Li, W.; Hsu, C.Y.; Tedesco, M. Advancing arctic sea ice remote sensing with ai and deep learning: Opportunities and challenges. Remote Sens. 2024, 16, 3764. [Google Scholar] [CrossRef]
- Shu, Y.; Cui, H.; Song, L.; Gan, L.; Xu, S.; Wu, J.; Zheng, C. Influence of sea ice on ship routes and speed along the Arctic Northeast Passage. Ocean Coast. Manag. 2024, 256, 107320. [Google Scholar] [CrossRef]
- Tuckett, P. The Distribution and Evolution of Antarctic Surface Meltwater. Ph.D. Dissertation, University of Sheffield, Sheffield, UK, 2023. [Google Scholar]
- Walsh, J.E. Intensified warming of the Arctic: Causes and impacts on middle latitudes. Glob. Planet. Chang. 2014, 117, 52–63. [Google Scholar] [CrossRef]
- Box, J.E.; Colgan, W.T.; Christensen, T.R.; Schmidt, N.M.; Lund, M.; Parmentier, F.-J.W.; Brown, R.; Bhatt, U.S.; Euskirchen, E.S.; Romanovsky, V.E.; et al. Key indicators of Arctic climate change: 1971–2017. Environ. Res. Lett. 2019, 14, 045010. [Google Scholar] [CrossRef]
- Williamson, S.N.; Menounos, B. The influence of forest fire aerosol and air temperature on glacier albedo, western North America. Remote Sens. Environ. 2021, 267, 112732. [Google Scholar] [CrossRef]
- Repina, I.A. Dynamic Meteorology. In Russian National Report: Meteorology and Atmospheric; 2023; p. 151. Available online: https://iugg.org/wp-content/uploads/2023/08/2019-2022_IAMAS_RU.pdf#page=151 (accessed on 15 January 2025).
- Moen-Larsen, N.; Gjerde, K.L. Changing or frozen narratives? In The Arctic in Russian Media and Expert Commentary, 2021–2022; 2023; Available online: https://nupi.brage.unit.no/nupi-xmlui/bitstream/handle/11250/3106592/NUPI_research_paper_2_2023_MoenLarsenGjerde.pdf?sequence=1 (accessed on 15 January 2025).
- Elias, S. Threats to the Arctic; Elsevier: Amsterdam, The Netherlands, 2021. [Google Scholar]
- Irrgang, A.M.; Bendixen, M.; Farquharson, L.M.; Baranskaya, A.V.; Erikson, L.H.; Gibbs, A.E.; Ogorodov, S.A.; Overduin, P.P.; Lantuit, H.; Grigoriev, M.N.; et al. Drivers, dynamics and impacts of changing Arctic coasts. Nat. Rev. Earth Environ. 2022, 3, 39–54. [Google Scholar] [CrossRef]
- Bruhwiler, L.; Parmentier, F.J.W.; Crill, P.; Leonard, M.; Palmer, P.I. The Arctic carbon cycle and its response to changing climate. Curr. Clim. Chang. Rep. 2021, 7, 14–34. [Google Scholar] [CrossRef]
- Taketani, F.; Miyakawa, T.; Takigawa, M.; Yamaguchi, M.; Komazaki, Y.; Mordovskoi, P.; Takashima, H.; Zhu, C.; Nishino, S.; Tohjima, Y.; et al. Characteristics of atmospheric black carbon and other aerosol particles over the Arctic Ocean in early autumn 2016: Influence from biomass burning as assessed with observed microphysical properties and model simulations. Sci. Total Environ. 2022, 848, 157671. [Google Scholar] [CrossRef]
- Senande-Rivera, M.; Insua-Costa, D.; Miguez-Macho, G. Spatial and temporal expansion of global wildland fire activity in response to climate change. Nat. Commun. 2022, 13, 1208. [Google Scholar] [CrossRef]
- van Wees, D.; van der Werf, G.R.; Randerson, J.T.; Andela, N.; Chen, Y.; Morton, D.C. The role of fire in global forest loss dynamics. Glob. Chang. Biol. 2021, 27, 2377–2391. [Google Scholar] [CrossRef]
- Greaves, W. Cities and human security in a warming Arctic. In Climate Change and Arctic Security: Searching for a Paradigm Shift; Springer International Publishing: Cham, Switzerland, 2019; pp. 61–89. [Google Scholar]
- Romasheva, N.; Dmitrieva, D. Energy resources exploitation in the russian arctic: Challenges and prospects for the sustainable development of the ecosystem. Energies 2021, 14, 8300. [Google Scholar] [CrossRef]
- Chen, J.; Kang, S.; Wu, A.; Chen, L. Projected emissions and climate impacts of Arctic shipping along the Northern Sea Route. Environ. Pollut. 2024, 341, 122848. [Google Scholar] [CrossRef]
- Ye, X.; Zhang, B.; Dawson, J.; Amon, C.D.; Ezechukwu, C.; Igwegbe, E.; Kang, Q.; Song, X.; Chen, B. Arctic Oceanic Carbon Cycle: A Comprehensive Review of Mechanisms, Regulations, and Models. Water 2024, 16, 1667. [Google Scholar] [CrossRef]
- Artaxo, P.; Hansson, H.-C.; Andreae, M.O.; Bäck, J.; Alves, E.G.; Barbosa, H.M.J.; Bender, F.; Bourtsoukidis, E.; Carbone, S.; Chi, J.; et al. Tropical and boreal forest atmosphere interactions: A review. Tellus. Ser. B Chem. Phys. Meteorol. 2022, 74, 24–163. [Google Scholar]
- Bidgood, A.K.; Hall, J. We Need to Talk About Mining in the Arctic. Earth Sci. Syst. Soc. 2024, 4, 10117. [Google Scholar] [CrossRef]
- PAME. Protection of the Arctic Marine Environment; Arctic Council: Ottawa, ON, Canada, 2024. [Google Scholar]
- IMO (International Maritime Organization). Further Shipping GHG Emission Reduction Measures Adopted; International Maritime Organization: London, UK, 2021; Available online: https://www.imo.org/en/MediaCentre/PressBriefings/pages/MEPC76.aspx (accessed on 12 January 2025).
- Stafford, K.M. The Changing Arctic Marine Soundscape. Arctic Report Card: Update for 2021; Global Ocean Monitoring and Observing (GOMO) Program, Office of Oceanic and Atmospheric Research: Silver Spring, MD, USA, 2021.
- Berkman, P.A.; Fiske, G.J.; Lorenzini, D.; Young, O.R.; Pletnikoff, K.; Grebmeier, J.M.; Fernandez, L.M.; Divine, L.M.; Causey, D.; Kapsar, K.E.; et al. Satellite Record of Pan-Arctic Maritime Ship Traffic. In Arctic Report Card 2022; Druckenmiller, M.L., Thoman, R.L., Moon, T.A., Eds.; NOAA: Washington, DC, USA, 2022. [Google Scholar]
- USGS. Circum–Arctic Resource Appraisal: Estimates of Undiscovered Oil and Gas North of the Arctic Circle. U.S. Geological Survey (USGS). 2008. Available online: https://pubs.usgs.gov/fs/2008/3049/fs2008-3049.pdf (accessed on 12 February 2025).
- WWF. Oil and Gas. Arctic World Wide Fund for Nature (WWF). 2025. Available online: https://www.arcticwwf.org/threats/oil-and-gas/ (accessed on 13 February 2025).
- Kreplin, H.N.; Santos Ferreira, C.S.; Destouni, G.; Keesstra, S.D.; Salvati, L.; Kalantari, Z. Arctic wetland system dynamics under climate warming. Wiley Interdiscip. Rev. Water 2021, 8, e1526. [Google Scholar] [CrossRef]
- Hayes, D.J.; Kicklighter, D.W.; McGuire, A.D.; Chen, M.; Zhuang, Q.; Yuan, F.; Melillo, J.M.; Wullschleger, S.D. The impacts of recent permafrost thaw on land–atmosphere greenhouse gas exchange. Environ. Res. Lett. 2014, 9, 045005. [Google Scholar] [CrossRef]
- Miller, G.H.; Alley, R.B.; Brigham-Grette, J.; Fitzpatrick, J.J.; Polyak, L.; Serreze, M.C.; White, J.W. Arctic amplification: Can the past constrain the future? Quat. Sci. Rev. 2010, 29, 1779–1790. [Google Scholar] [CrossRef]
- Somerville, P.; Marin, M. The Arctic is Warming Four Times Faster than the Global Average; Risk Frontiers: St Leonards, Australia, 2023. [Google Scholar]
- Ceppi, P.; Brient, F.; Zelinka, M.D.; Hartmann, D.L. Cloud feedback mechanisms and their representation in global climate models. Wiley Interdiscip. Rev. Clim. Chang. 2017, 8, e465. [Google Scholar] [CrossRef]
- Beer, E.; Eisenman, I. Revisiting the role of the water vapor and lapse rate feedbacks in the Arctic amplification of climate change. J. Clim. 2022, 35, 2975–2988. [Google Scholar] [CrossRef]
- Navarro, J.C.A.; Varma, V.; Riipinen, I.; Seland, Ø.; Kirkevåg, A.; Struthers, H.; Iversen, T.; Hansson, H.-C.; Ekman, A.M.L. Amplification of Arctic warming by past air pollution reductions in Europe. Nat. Geosci. 2016, 9, 277–281. [Google Scholar] [CrossRef]
- Merikanto, J.; Nordling, K.; Räisänen, P.; Räisänen, J.; O′Donnell, D.; Partanen, A.I.; Korhonen, H. How Asian aerosols impact regional surface temperatures across the globe. Atmos. Chem. Phys. 2020, 21, 5865–5881. [Google Scholar] [CrossRef]
- AMAP. Arctic Monitoring and Assessment Programme: Work Plan 2015–2017. Water Treat. 2015, 321, 100889. [Google Scholar]
- Pithan, F.; Mauritsen, T. Arctic amplification dominated by temperature feedbacks in contemporary climate models. Nat. Geosci. 2014, 7, 181–184. [Google Scholar] [CrossRef]
- Manabe, S.; Wetherald, R.T. The effects of doubling the CO2 concentration on the climate of a general circulation model. J. Atmos. Sci. 1975, 32, 3–15. [Google Scholar] [CrossRef]
- Perovich, D.K.; Light, B.; Eicken, H.; Jones, K.F.; Runciman, K.; Nghiem, S.V. Increasing solar heating of the Arctic Ocean and adjacent seas, 1979–2005: Attribution and role in the ice-albedo feedback. Geophys. Res. Lett. 2007, 34. [Google Scholar] [CrossRef]
- Kay, J.E.; L’Ecuyer, T. Observational constraints on Arctic Ocean clouds and radiative fluxes during the early 21st century. J. Geophys. Res. Atmos. 2013, 118, 7219–7236. [Google Scholar] [CrossRef]
- Dai, A.; Jenkins, M.T. Relationships among Arctic warming, sea-ice loss, stability, lapse rate feedback, and Arctic amplification. Clim. Dyn. 2023, 61, 5217–5232. [Google Scholar] [CrossRef]
- Polyakov, I.V.; Timokhov, L.A.; Alexeev, V.A.; Bacon, S.; Dmitrenko, I.A.; Fortier, L.; Frolov, I.E.; Gascard, J.-C.; Hansen, E.; Ivanov, V.V.; et al. Arctic Ocean warming contributes to reduced polar ice cap. J. Phys. Oceanogr. 2010, 40, 2743–2756. [Google Scholar] [CrossRef]
- Gascard, J.C.; Zhang, J.; Rafizadeh, M. Rapid decline of Arctic sea ice volume: Causes and consequences. Cryosphere Discuss. 2019, 2019, 1–29. [Google Scholar]
- Kuletz, K.J.; Ferguson, S.H.; Frederiksen, M.; Gallagher, C.P.; Hauser, D.D.W.; Hop, H.; Kovacs, K.M.; Lydersen, C.; Mosbech, A.; Seitz, A.C. A review of climate change impacts on migration patterns of marine vertebrates in Arctic and Subarctic ecosystems. Front. Environ. Sci. 2024, 12, 1434549. [Google Scholar] [CrossRef]
- Macias-Fauria, M.; Post, E. Effects of sea ice on Arctic biota: An emerging crisis discipline. Biol. Lett. 2018, 14, 20170702. [Google Scholar] [CrossRef] [PubMed]
- Dastoor, A.; Angot, H.; Bieser, J.; Christensen, J.H.; Douglas, T.A.; Heimbürger-Boavida, L.E.; Jiskra, M.; Zdanowicz, C. Arctic mercury cycling. Nat. Rev. Earth Environ. 2022, 3, 270–286. [Google Scholar] [CrossRef]
- Grigorieva, E.A. Climate Change and Human Health in the Arctic: A Review. Climate 2024, 12, 89. [Google Scholar] [CrossRef]
- Stern, H.L.; Schweiger, A.J.; Stark, M.; Zhang, J.; Steele, M.; Hwang, B. Seasonal evolution of the sea-ice floe size distribution in the Beaufort and Chukchi seas. Elem. Sci. Anthr. 2018, 6, 48. [Google Scholar] [CrossRef]
- Patz, J.A.; Grabow, M.L.; Limaye, V.S. When it rains, it pours: Future climate extremes and health. Ann. Glob. Health 2014, 80, 332–344. [Google Scholar] [CrossRef]
- Meier, W.N.; Hovelsrud, G.K.; van Oort, B.E.; Key, J.R.; Kovacs, K.M.; Michel, C.; Haas, C.; Granskog, M.A.; Gerland, S.; Perovich, D.K.; et al. Arctic sea ice in transformation: A review of recent observed changes and impacts on biology and human activity. Rev. Geophys. 2014, 52, 185–217. [Google Scholar] [CrossRef]
- Otterå, O.H.; Drange, H. A possible feedback mechanism involving the Arctic freshwater, the Arctic sea ice, and the north Atlantic drift. Adv. Atmos. Sci. 2004, 21, 784–801. [Google Scholar] [CrossRef]
- Vihma, T. Effects of Arctic sea ice decline on weather and climate: A review. Surv. Geophys. 2014, 35, 1175–1214. [Google Scholar] [CrossRef]
- Diebold, F.X.; Rudebusch, G.D.; Göbel, M.; Coulombe, P.G.; Zhang, B. When will Arctic sea ice disappear? Projections of area, extent, thickness, and volume. J. Econom. 2023, 236, 105479. [Google Scholar] [CrossRef]
- Babb, D.G.; Galley, R.J.; Kirillov, S.; Landy, J.C.; Howell, S.E.L.; Stroeve, J.C.; Meier, W.; Ehn, J.K.; Barber, D.G. The stepwise reduction of multiyear sea ice area in the Arctic Ocean since 1980. J. Geophys. Res. Ocean. 2023, 128, e2023JC020157. [Google Scholar] [CrossRef]
- Khare, N.; Khare, R. The Arctic: A Barometer of Global Climate Variability; Elsevier: Amsterdam, The Netherlands, 2021. [Google Scholar]
- Harning, D.J.; Andrews, J.T.; Belt, S.T.; Cabedo-Sanz, P.; Geirsdóttir, Á.; Dildar, N.; Miller, G.H.; Sepúlveda, J. Sea ice control on winter subsurface temperatures of the North Iceland Shelf during the Little Ice Age: A TEX 86 calibration case study. Paleoceanogr. Paleoclimatology 2019, 34, 1006–1021. [Google Scholar] [CrossRef]
- Jennings, A.E.; Knudsen, K.L.; Hald, M.; Hansen, C.V.; Andrews, J.T. A mid-Holocene shift in Arctic sea-ice variability on the East Greenland Shelf. Holocene 2002, 12, 49–58. [Google Scholar] [CrossRef]
- Golledge, N.R. Long-term projections of sea-level rise from ice sheets. Wiley Interdiscip. Rev. Clim. Chang. 2020, 11, e634. [Google Scholar] [CrossRef]
- The IMBIE Team. Mass balance of the Greenland Ice Sheet from 1992 to 2018. Nature 2020, 579, 233–239. [Google Scholar] [CrossRef]
- Benn, D.I.; Warren, C.R.; Mottram, R.H. Calving processes and the dynamics of calving glaciers. Earth-Sci. Rev. 2007, 82, 143–179. [Google Scholar] [CrossRef]
- Shahgedanova, M. Climate change and melting glaciers. In The Impacts of Climate Change; Elsevier: Amsterdam, The Netherlands, 2021; pp. 53–84. [Google Scholar]
- Nghiem, S.V.; Hall, D.K.; Mote, T.L.; Tedesco, M.; Albert, M.R.; Keegan, K.; Shuman, C.A.; DiGirolamo, N.E.; Neumann, G. The extreme melt across the Greenland ice sheet in 2012. Geophys. Res. Lett. 2012, 39. [Google Scholar] [CrossRef]
- Hanna, E.; Topál, D.; Box, J.E.; Buzzard, S.; Christie, F.D.W.; Hvidberg, C.; Morlighem, M.; De Santis, L.; Silvano, A.; Colleoni, F.; et al. Short- and long-term variability of the Antarctic and Greenland ice sheets. Nat. Rev. Earth Environ. 2024, 5, 193–210. [Google Scholar]
- Nicholls, R.J. Planning for the impacts of sea level rise. Oceanography 2011, 24, 144–157. [Google Scholar] [CrossRef]
- Jacob, T.; Wahr, J.; Pfeffer, W.T.; Swenson, S. Recent contributions of glaciers and ice caps to sea level rise. Nature 2012, 482, 514–518. [Google Scholar] [CrossRef]
- Hansen, J.; Sato, M.; Hearty, P.; Ruedy, R.; Kelley, M.; Masson-Delmotte, V.; Russell, G.; Tselioudis, G.; Cao, J.; Rignot, E.; et al. Ice melt, sea level rise and superstorms: Evidence from paleoclimate data, climate modeling, and modern observations that 2 C global warming could be dangerous. Atmos. Chem. Phys. 2016, 16, 3761–3812. [Google Scholar] [CrossRef]
- McLeman, R. Migration and displacement risks due to mean sea-level rise. Bull. At. Sci. 2018, 74, 148–154. [Google Scholar] [CrossRef]
- UiT The Arctic University of Norway. Melting Arctic Sea-Ice Could Affect Global Ocean Circulation. ScienceDaily, 27 October 2024. Available online: http://www.sciencedaily.com/releases/2024/10/241027205850.htm (accessed on 14 March 2024).
- Rasul, G.; Molden, D. The global social and economic consequences of mountain cryospheric change. Front. Environ. Sci. 2019, 7, 91. [Google Scholar] [CrossRef]
- Li, H.; Fedorov, A.V. Persistent freshening of the Arctic Ocean and changes in the North Atlantic salinity caused by Arctic sea ice decline. Clim. Dyn. 2021, 57, 2995–3013. [Google Scholar] [CrossRef]
- Rudels, B.; Hainbucher, D. On the formation and spreading of thermohaline intrusions in the Arctic Ocean. Geophysica 2020, 55, 23–59. [Google Scholar]
- Sévellec, F.; Fedorov, A.V.; Liu, W. Arctic sea-ice decline weakens the Atlantic meridional overturning circulation. Nat. Clim. Chang. 2017, 7, 604–610. [Google Scholar] [CrossRef]
- Caesar, L.; Rahmstorf, S.; Robinson, A.; Feulner, G.; Saba, V. Observed fingerprint of a weakening Atlantic Ocean overturning circulation. Nature 2018, 556, 191–196. [Google Scholar] [CrossRef]
- Kautz, L.A.; Martius, O.; Pfahl, S.; Pinto, J.G.; Ramos, A.M.; Sousa, P.M.; Woollings, T. Atmospheric blocking and weather extremes over the Euro-Atlantic sector—A review. Weather Clim. Dyn. Discuss. 2021, 2021, 1–43. [Google Scholar] [CrossRef]
- Little, C.M.; Hu, A.; Hughes, C.W.; McCarthy, G.D.; Piecuch, C.G.; Ponte, R.M.; Thomas, M.D. The relationship between US East Coast sea level and the Atlantic meridional overturning circulation: A review. J. Geophys. Res. Ocean. 2019, 124, 6435–6458. [Google Scholar] [CrossRef]
- Gissi, E.; Manea, E.; Mazaris, A.D.; Fraschetti, S.; Almpanidou, V.; Bevilacqua, S.; Coll, M.; Guarnieri, G.; Lloret-Lloret, E.; Pascual, M.; et al. A review of the combined effects of climate change and other local human stressors on the marine environment. Sci. Total Environ. 2021, 755, 142564. [Google Scholar] [CrossRef]
- Luo, Y.; Boudreau, B.P.; Mucci, A. Disparate acidification and calcium carbonate desaturation of deep and shallow waters of the Arctic Ocean. Nat. Commun. 2016, 7, 12821. [Google Scholar] [CrossRef]
- Murton, J.B. Permafrost and climate change. In Climate Change; Elsevier: Amsterdam, The Netherlands, 2021; pp. 281–326. [Google Scholar]
- Knoblauch, C.; Beer, C.; Liebner, S.; Grigoriev, M.N.; Pfeiffer, E.M. Methane production as key to the greenhouse gas budget of thawing permafrost. Nat. Clim. Chang. 2018, 8, 309–312. [Google Scholar] [CrossRef]
- Van Huissteden, J. Thawing Permafrost; Springer International Publishing: Cham, Switzerland, 2020. [Google Scholar]
- Brown, T.M.; Macdonald, R.W.; Muir, D.C.; Letcher, R.J. The distribution and trends of persistent organic pollutants and mercury in marine mammals from Canada’s Eastern Arctic. Sci. Total Environ. 2018, 618, 500–517. [Google Scholar] [CrossRef]
- Walvoord, M.A.; Kurylyk, B.L. Hydrologic impacts of thawing permafrost—A review. Vadose Zone J. 2016, 15, vzj2016-01. [Google Scholar] [CrossRef]
- Kokelj, S.V.; Jorgenson, M.T. Advances in thermokarst research. Permafr. Periglac. Process. 2013, 24, 108–119. [Google Scholar] [CrossRef]
- Strauss, J.; Biasi, C.; Sanders, T.; Abbott, B.W.; von Deimling, T.S.; Voigt, C.; Winkel, M.; Marushchak, M.E.; Kou, D.; Fuchs, M.; et al. A globally relevant stock of soil nitrogen in the Yedoma permafrost domain. Nat. Commun. 2022, 13, 6074. [Google Scholar] [CrossRef]
- Turetsky, M.R.; Abbott, B.W.; Jones, M.C.; Anthony, K.W.; Olefeldt, D.; Schuur, E.A.G.; Koven, C.; McGuire, A.D.; Grosse, G.; Kuhry, P.; et al. Permafrost collapse is accelerating carbon release. Nature 2019, 569, 32–34. [Google Scholar] [CrossRef]
- Smith, K.A.; Ball, T.; Conen, F.; Dobbie, K.E.; Massheder, J.; Rey, A. Exchange of greenhouse gases between soil and atmosphere: Interactions of soil physical factors and biological processes. Eur. J. Soil science 2018, 69, 10–20. [Google Scholar] [CrossRef]
- Callaghan, T.V.; Björn, L.O.; Chernov, Y.; Chapin, T.; Christensen, T.R.; Huntley, B.; Ims, R.A.; Johansson, M.; Jolly, D.; Jonasson, S.; et al. Biodiversity, distributions and adaptations of Arctic species in the context of environmental change. AMBIO A J. Hum. Environ. 2004, 33, 404–417. [Google Scholar] [CrossRef]
- Hope, A.G.; Waltari, E.; Malaney, J.L.; Payer, D.C.; Cook, J.A.; Talbot, S.L. Arctic biodiversity: Increasing richness accompanies shrinking refugia for a cold-associated tundra fauna. Ecosphere 2015, 6, 1–67. [Google Scholar] [CrossRef]
- Myers-Smith, I.H.; Kerby, J.T.; Phoenix, G.K.; Bjerke, J.W.; Epstein, H.E.; Assmann, J.J.; John, C.; Andreu-Hayles, L.; Angers-Blondin, S.; Beck, P.S.A.; et al. Complexity revealed in the greening of the Arctic. Nat. Clim. Chang. 2020, 10, 106–117. [Google Scholar] [CrossRef]
- Johnson, A.; Trant, A.; Hermanutz, L.; Davis, E.; Saunders, M.; Siegwart Collier, L.; Way, R.; Knight, T. Climate warming impacts tuttuk (caribou) forage availability in Tongait (Torngat) Mountains, Labrador. Arct. Sci. 2025, 11, 1–14. [Google Scholar] [CrossRef]
- Pechsiri, J.S.; Sattari, A.; Martinez, P.G.; Xuan, L. A review of the climate-change-impacts’ rates of change in the Arctic. J. Environ. Prot. 2010, 1, 59. [Google Scholar] [CrossRef]
- Knopp, J.A.; Levenstein, B.; Watson, A.; Ivanova, I.; Lento, J. Systematic review of documented Indigenous Knowledge of freshwater biodiversity in the circumpolar Arctic. Freshw. Biol. 2022, 67, 194–209. [Google Scholar] [CrossRef]
- Lai, S.; Warret Rodrigues, C.; Gallant, D.; Roth, J.D.; Berteaux, D. Red foxes at their northern edge: Competition with the Arctic fox and winter movements. J. Mammal. 2022, 103, 586–597. [Google Scholar] [CrossRef]
- Giesenhagen, H.C.; Detmer, A.E.; de Wall, J.; Weber, A.; Gradinger, R.R.; Jochem, F.J. How are Antarctic planktonic microbial food webs and algal blooms affected by melting of sea ice? Microcosm simulations. Aquat. Microb. Ecol. 1999, 20, 183–201. [Google Scholar] [CrossRef]
- Falardeau, M.; Bennett, E.M. Towards integrated knowledge of climate change in Arctic marine systems: A systematic literature review of multidisciplinary research. Arct. Sci. 2019, 6, 1–23. [Google Scholar] [CrossRef]
- Chen, J. Impacts of Sea Ice Loss on Polar Bear Diet, Prey Availability, Foraging Behaviors, and Human-Bear Interactions in the Arctic. Master’s Thesis, The University of San Francisco, San Francisco, CA, USA, 2022. [Google Scholar]
- Laidre, K.L.; Stern, H.; Kovacs, K.M.; Lowry, L.; Moore, S.E.; Regehr, E.V.; Ferguson, S.H.; Wiig, Ø.; Boveng, P.; Angliss, R.P.; et al. Arctic marine mammal population status, sea ice habitat loss, and conservation recommendations for the 21st century. Conserv. Biol. 2015, 29, 724–737. [Google Scholar] [CrossRef]
- Inuit Circumpolar Council Alaska. Coastal Monitoring Indigenous Knowledge Holders Meeting Report; Inuit Circumpolar Council Alaska: Ottawa, ON, Canada, 2016. [Google Scholar]
- Huntington, H.P.; Zagorsky, A.; Kaltenborn, B.P.; Shin, H.C.; Dawson, J.; Lukin, M.; Dahl, P.E.; Guo, P.; Thomas, D.N. Societal implications of a changing Arctic Ocean. Ambio 2022, 51, 298–306. [Google Scholar] [CrossRef]
- Larsen, J.N.; Fondahl, G. (Eds.) Arctic Human Development Report: Regional Processes and Global Linkages; Nordic Council of Ministers: København, Denmark, 2015. [Google Scholar]
- Gryba, R.; Huntington, H.P.; Von Duyke, A.L.; Adams, B.; Frantz, B.; Gatten, J.; Harcharek, Q.; Olemaun, H.; Sarren, R.; Skin, J.; et al. Indigenous Knowledge of bearded seal (Erignathus barbatus), ringed seal (Pusa hispida), and spotted seal (Phoca largha) behaviour and habitat use near Utqiaġvik, Alaska, USA. Arct. Sci. 2021, 7, 832–858. [Google Scholar] [CrossRef]
- Green, K.M.; Beaudreau, A.H.; Lukin, M.H.; Crowder, L.B. Climate change stressors and social-ecological factors mediating access to subsistence resources in Arctic Alaska. Ecol. Soc. 2021, 26. [Google Scholar] [CrossRef]
- Borish, D.; Cunsolo, A.; Snook, J.; Shiwak, I.; Wood, M.; Mauro, I.; HERD Caribou Project Steering Committee. “Caribou was the reason, and everything else happened after”: Effects of caribou declines on Inuit in Labrador, Canada. Glob. Environ. Chang. 2021, 68, 102268. [Google Scholar] [CrossRef]
- Joy, F.; Herrmann, T.M. Sacred Sites in the Arctic North and Beyond: The Challenges of Protecting Cultural Heritage and Living Traditions in a Multitude of Contexts and Cultures. Nord.-Mediterr. Icel. E J. Nord. Mediterr. Stud. 2022, 17, 1–8. [Google Scholar] [CrossRef]
- Hossain, K.; Raheem, D.; Cormier, S. Food Security Governance in the Arctic-Barents Region; Springer International Publishing: Cham, Switzerland, 2018. [Google Scholar]
- Emanuelsen, K.; Pearce, T.; Oakes, J.; Harper, S.L.; Ford, J.D. Sewing and Inuit women’s health in the Canadian Arctic. Soc. Sci. Med. 2020, 265, 113523. [Google Scholar] [CrossRef]
- Landauer, M.; Juhola, S. Loss and damage in the rapidly changing arctic. In Loss and Damage from Climate Change: Concepts, Methods and Policy Options; Springer: Cham, Switzerland, 2019; pp. 425–447. [Google Scholar]
- Hicks Pries, C.E.; Schuur, E.A.; Natali, S.M.; Crummer, K.G. Old soil carbon losses increase with ecosystem respiration in experimentally thawed tundra. Nat. Clim. Chang. 2016, 6, 214–218. [Google Scholar] [CrossRef]
- Vladimirova, V. Indigenous people living with waste and pollution in the Arctic. In Ecological Concerns in Transition: A comparative Study on Responses to Waste and Environmental Destruction in the Region; Södertörns Högskola: Huddinge, Sweden, 2023; pp. 45–58. [Google Scholar]
- Adeniran-Obey, S.O.; Imoobe, T.O. Impact of Climate Change on Arctic Marine. In Arctic Marine Ecotoxicology; Springer: Cham, Switzerland, 2024; pp. 283–316. [Google Scholar]
- Mercer, L.; Whalen, D.; Pokiak, D.L.; Lim, M.; Mann, P.J. Ensuring continuity and impact in Arctic monitoring: A solution-orientated model for community-based environmental research. Environ. Res. Ecol. 2023, 2, 045001. [Google Scholar] [CrossRef]
- Galappaththi, E.K.; Falardeau, M.; Harris, L.N.; Rocha, J.C.; Moore, J.S.; Berkes, F. Resilience-based steps for adaptive co-management of Arctic small-scale fisheries. Environ. Res. Lett. 2022, 17, 083004. [Google Scholar] [CrossRef]
- Coggins, S.; Ford, J.; Berrang-Ford, L.; Harper, S.L.; Hyams, K.; Paavola, J.; Satyal, P.; Arotoma-Rojas, I. Climate justice and Indigenous Peoples in the Arctic. Georget. J. Int. Aff. 2021. Available online: https://gjia.georgetown.edu/2021/02/23/indigenous-peoples-and-climate-justice-in-the-arctic/ (accessed on 15 February 2025).
- Sharapov, D. Arctic Ice Changes and Global Warming. In E3S Web of Conferences; EDP Sciences: Les Ulis, France, 2023; Volume 460, p. 08014. [Google Scholar]
- Collingwood, E.; Scaife, A.A.; Lu, H.; Sinha, B.; King, J.; Marsh, R.; Marshall, G. Meridional wind in the upper stratosphere: A source of winter NAO predictability. Geophys. Res. Lett. 2024, 51. [Google Scholar] [CrossRef]
- Francis, J.A.; Vavrus, S.J.; Cohen, J. Amplified Arctic warming and mid-latitude weather: New perspectives on emerging connections. Wiley Interdiscip. Rev. Clim. Chang. 2017, 8, e474. [Google Scholar] [CrossRef]
- Fagan, B. The Little Ice Age: How Climate Made History 1300–1850; Hachette: London, UK, 2019. [Google Scholar]
- Pilkey, O.H.; Pilkey, C.O.; Pilkey-Jarvis, L.P.; Longo, N.J.; Pilkey, K.C.; Dodson, F.B.; Hayes, H.L. Escaping Nature: How to Survive Global Climate Change; Duke University Press: Durham, NC, USA, 2024. [Google Scholar]
- Beckmann, J.; Winkelmann, R. Effects of extreme melt events on ice flow and sea level rise of the Greenland Ice Sheet. Cryosphere 2023, 17, 3083–3099. [Google Scholar] [CrossRef]
- Adkins, J.F.; McIntyre, K.; Schrag, D.P. The salinity, temperature, and δ18O of the glacial deep ocean. Science 2002, 298, 1769–1773. [Google Scholar] [CrossRef] [PubMed]
- Rahmstorf, S.; Box, J.E.; Feulner, G.; Mann, M.E.; Robinson, A.; Rutherford, S.; Schaffernicht, E.J. Exceptional twentieth-century slowdown in Atlantic Ocean overturning circulation. Nat. Clim. Chang. 2015, 5, 475–480. [Google Scholar] [CrossRef]
- Nguyen, M.A.; Ahrens, L.; Josefsson, S.; Gustavsson, J.; Laudon, H.; Wiberg, K. Seasonal trends and retention of polycyclic aromatic compounds (PACs) in a remote sub-Arctic catchment. Environ. Pollut. 2023, 333, 121992. [Google Scholar] [CrossRef]
- Schuur, E.A.G.; McGuire, A.D.; Schädel, C.; Grosse, G.; Harden, J.W.; Hayes, D.J.; Hugelius, G.; Koven, C.D.; Kuhry, P.; Lawrence, D.M.; et al. Climate change and the permafrost carbon feedback. Nature 2015, 520, 171–179. [Google Scholar] [CrossRef]
- Worrall, F.; Chapman, P.; Holden, J.; Evans, C.; Artz, R.; Smith, P.; Grayson, R. A Review of Current Evidence on Carbon Fluxes and Greenhouse Gas Emissions from UK Peatlands. 2011. Available online: https://nora.nerc.ac.uk/id/eprint/15889/1/jncc442_webfinal.pdf (accessed on 12 January 2025).
- Overland, J.E. Less climatic resilience in the Arctic. Weather Clim. Extrem. 2020, 30, 100275. [Google Scholar] [CrossRef]
- Bokuchava, D.D.; Semenov, V.A. Mechanisms of the early 20th century warming in the Arctic. Earth-Sci. Rev. 2021, 222, 103820. [Google Scholar] [CrossRef]
- Bhatt, U.S.; Walker, D.A.; Raynolds, M.K.; Bieniek, P.A.; Epstein, H.E.; Comiso, J.C.; Pinzon, J.E.; Tucker, C.J.; Polyakov, I.V. Recent declines in warming and vegetation greening trends over pan-Arctic tundra. Remote Sens. 2013, 5, 4229–4254. [Google Scholar] [CrossRef]
- Campioli, M.; Schmidt, N.M.; Albert, K.R.; Leblans, N.; Ro-Poulsen, H.; Michelsen, A. Does warming affect growth rate and biomass production of shrubs in the High Arctic? Plant Ecol. 2013, 214, 1049–1058. [Google Scholar] [CrossRef]
- Magnússon, R.Í.; Groten, F.; Bartholomeus, H.; van Huissteden, K.; Heijmans, M.M. Tundra browning in the Indigirka Lowlands (North-Eastern Siberia) explained by drought, floods and small-scale vegetation shifts. J. Geophys. Res. Biogeosci. 2023, 128, e2022JG007330. [Google Scholar] [CrossRef]
- Saros, J.E.; Arp, C.D.; Bouchard, F.; Comte, J.; Couture, R.-M.; Dean, J.F.; Lafrenière, M.J.; MacIntyre, S.; McGowan, S.; Rautio, M.; et al. Sentinel responses of Arctic freshwater systems to climate: Linkages, evidence, and a roadmap for future research. Arct. Sci. 2022, 9, 356–392. [Google Scholar] [CrossRef]
- Wrona, F.J.; Prowse, T.D.; Reist, J.D.; Hobbie, J.E.; Lévesque, L.M.; Vincent, W.F. Climate impacts on Arctic freshwater ecosystems and fisheries: Background, rationale and approach of the Arctic Climate Impact Assessment (ACIA). AMBIO A J. Hum. Environ. 2006, 35, 326–329. [Google Scholar] [CrossRef]
- Debortoli, N.S.; Clark, D.G.; Ford, J.D.; Sayles, J.S.; Diaconescu, E.P. An integrative climate change vulnerability index for Arctic aviation and marine transportation. Nat. Commun. 2019, 10, 2596. [Google Scholar] [CrossRef]
- Debortoli, N.S.; Pearce, T.; Ford, J.D. Estimating Future Costs for Infrastructure in the Proposed Canadian Northern Corridor at Risk from Climate Change; The School of Public Policy Publications: Calgary, AB, Canada, 2023; Volume 16. [Google Scholar] [CrossRef]
- Rackow, T.; Danilov, S.; Goessling, H.F.; Hellmer, H.H.; Sein, D.V.; Semmler, T.; Sidorenko, D.; Jung, T. Delayed Antarctic sea-ice decline in high-resolution climate change simulations. Nat. Commun. 2022, 13, 637. [Google Scholar] [CrossRef]
- Andersson, T.R.; Hosking, J.S.; Pérez-Ortiz, M.; Paige, B.; Elliott, A.; Russell, C.; Law, S.; Jones, D.C.; Wilkinson, J.; Phillips, T.; et al. Seasonal Arctic Sea ice forecasting with probabilistic deep learning. Nat. Commun. 2021, 12, 5124. [Google Scholar] [CrossRef] [PubMed]
- Young, O.R. Arctic tipping points: Governance in turbulent times. Ambio 2012, 41, 75–84. [Google Scholar] [CrossRef]
- Stokke, O.S.; Hønneland, G. International Cooperation and Arctic Governance; Regime Effectiveness and Northern Region Building: London, UK, 2007. [Google Scholar]
- Sakharov, A. Arctic Council as a regional governance institution. Int. Organ. Res. J. 2015, 10, 40–53. [Google Scholar]
- Wiseman, M.S. The future of the Arctic Council. In The Palgrave Handbook of Arctic Policy and Politics; Springer: Berlin/Heidelberg, Germany, 2020; pp. 439–452. [Google Scholar]
- ACIA. Arctic Climate Impact Assessment; ACIA Overview Report; Cambridge University Press: Cambridge, UK, 2005; p. 1020. [Google Scholar]
- Ganesan, A.L.; Schwietzke, S.; Poulter, B.; Arnold, T.; Lan, X.; Rigby, M.; Vogel, F.R.; van der Werf, G.R.; Janssens-Maenhout, G.; Boesch, H.; et al. Advancing scientific understanding of the global methane budget in support of the Paris Agreement. Glob. Biogeochem. Cycles 2019, 33, 1475–1512. [Google Scholar] [CrossRef]
- Gosnell, R.; Saunes, L. Report No. 2: Integrated Naval Deterrence in the Arctic Region—Strategic Options for Enhancing Regional Naval Cooperation. 2024. Available online: https://digital-commons.usnwc.edu/nasi/2/ (accessed on 10 January 2025).
- Chircop, A. The polar code and the arctic marine environment: Assessing the regulation of the environmental risks of shipping. Int. J. Mar. Coast. Law 2020, 35, 533–569. [Google Scholar] [CrossRef]
- Malloy, B.A. On thin ice: How a binding treaty regime can save the Arctic. Hastings West-Northwest J. Environ. Law Policy 2010, 16, 471. [Google Scholar]
- Heininen, L.; Everett, K.; Padrtova, B.; Reissell, A. Arctic Policies and Strategies—Analysis, Synthesis, and Trends; International Institute for Applied Systems Analysis: Laxenburg, Austria, 2020. [Google Scholar]
- Government of Canada. Canada’s Arctic and Northern Policy Framework; Government of Canada: Ottawa, ON, Canada, 2019.
- Canosa, I.V.; Biesbroek, R.; Ford, J.; McCarty, J.L.; Orttung, R.W.; Paavola, J.; Burnasheva, D. Wildfire adaptation in the Russian Arctic: A systematic policy review. Clim. Risk Manag. 2023, 39, 100481. [Google Scholar] [CrossRef]
- Meijer, M.W.; Wolk, T. Policy and Practice in Norwegian Green Transition; Nordregio, Norway, 2021. Available online: https://norceresearch.brage.unit.no/norceresearch-xmlui/bitstream/handle/11250/2770504/Rapport%2019-2021%2c%20Norce%20Samfunn.pdf?sequence=1&isAllowed=y (accessed on 10 December 2024).
- Peacock Stephanie, S.J.; Mavrot, F.; Tomaselli, M.; Hanke, A.; Fenton, H.; Nathoo, R.; Aleuy, O.A.; Di Francesco, J.; Aguilar, X.F.; Jutha, N.; et al. Linking co-monitoring to co-management: Bringing together local, traditional, and scientific knowledge in a wildlife status assessment framework. Arct. Sci. 2020, 6, 247–266. [Google Scholar] [CrossRef]
- Ksenofontov, S.S.; Petrov, A.N. Global Change Impacts on Indigenous Sustainability in Sakha Republic: A Synthesis of Knowledge. Sustainability 2024, 16, 1157. [Google Scholar] [CrossRef]
- Cowan, A. UNDRIP and the intervention: Indigenous self-determination, participation, and racial discrimination in the northern territory of Australia. Pac. Rim Law Policy J. 2013, 22, 247. [Google Scholar]
- Inuit Tapiriit Kanatami, Inuit Nunangat Food Security Strategy. 2021. Available online: https://www.itk.ca/wp-content/uploads/2021/07/ITK_Food-Security-Strategy-Report_English_PDF-Version.pdf (accessed on 10 January 2025).
- Perrin, A.D.; Ljubicic, G.; Ogden, A. Northern research policy contributions to Canadian Arctic sustainability. Sustainability 2021, 13, 12035. [Google Scholar] [CrossRef]
- Zhang, Q.; Wan, Z.; Hemmings, B.; Abbasov, F. Reducing black carbon emissions from Arctic shipping: Solutions and policy implications. J. Clean. Prod. 2019, 241, 118261. [Google Scholar] [CrossRef]
- Balgehshiri, S.K.M.; Zohuri, B. The impact of energy transition to net-zero emissions on the world economy and global strategies. J. Econ. Manag. Res. 2023, 4, 2–7. [Google Scholar] [CrossRef]
- Yang, X.; Zhang, Z.; Cui, Z.; Cai, S. International Regulatory Framework for Black Carbon Emissions from Arctic Shipping: Current Situation, Problems, and Development. Sustainability 2024, 16, 10656. [Google Scholar] [CrossRef]
- Wang, J.; Zhang, Y. The effectiveness of legal framework of Arctic vessel-source black carbon governance. Environ. Sci. Pollut. Res. 2024, 31, 40472–40494. [Google Scholar] [CrossRef]
- Moore, J.C.; Mettiäinen, I.; Wolovick, M.; Zhao, L.; Gladstone, R.; Chen, Y.; Kirchner, S.; Koivurova, T. Targeted geoengineering: Local interventions with global implications. Glob. Policy 2021, 12, 108–118. [Google Scholar] [CrossRef]
- Cvijanovic, I.; Caldeira, K.; MacMartin, D.G. Impacts of ocean albedo alteration on Arctic sea ice restoration and Northern Hemisphere climate. Environ. Res. Lett. 2015, 10, 044020. [Google Scholar] [CrossRef]
- Turner, A.C.; Young, M.A.; Moran, M.D.; McClung, M.R. Comprehensive valuation of the ecosystem services of the Arctic National Wildlife Refuge. Nat. Areas J. 2021, 41, 125–137. [Google Scholar] [CrossRef]
- Graczyk, P.; Koivurova, T. The Arctic Council. In Handbook of the Politics of the Arctic; Edward Elgar Publishing: Cheltenham, UK, 2015; pp. 298–327. [Google Scholar]
- Dawson, J.; Cook, A.; Holloway, J.; Copland, L. Analysis of changing levels of ice strengthening (ice class) among vessels operating in the Canadian Arctic over the past 30 years. Arctic 2022, 75, 413–430. [Google Scholar] [CrossRef]
- Boyle, A.E. Dispute settlement and the Law of the Sea Convention: Problems of fragmentation and jurisdiction. Int. Comp. Law Q. 1997, 46, 37–54. [Google Scholar] [CrossRef]
- Vogel, J. Analyzing the Economic and Political Implications of the Port State Measures Agreement for Multinational Fisheries Governance in the Northeastern Pacific; Stanford University: Stanford, CA, USA, 2021; Available online: https://purl.stanford.edu/zr870js8104 (accessed on 13 December 2024).
- Coote, M.L. Environmental Decision-Making in the Arctic Council: What is the Role of Indigenous Peoples? Ph.D. Thesis, University of Iceland, Reykjavik, Iceland, 2016. [Google Scholar]
- Malik, I.H.; Ford, J.D. Monitoring climate change vulnerability in the Himalayas. Ambio 2024, 54, 1–19. [Google Scholar] [CrossRef] [PubMed]
- Malik, I.H.; Ford, J.D. Addressing the climate change adaptation gap: Key themes and future directions. Climate 2024, 12, 24. [Google Scholar] [CrossRef]
- Konnov, A.; Khmelnitskaya, Y.; Dugina, M.; Borzenko, T.; Tysiachniouk, M.S. Traditional livelihood, unstable environment: Adaptation of traditional fishing and reindeer herding to environmental change in the Russian Arctic. Sustainability 2022, 14, 12640. [Google Scholar] [CrossRef]
- Kirchner, S.; Mazzullo, N.; Nebasifu, A.A.; Lesser, P.; Tulppo, P.; Kyllönen, K.M.; Heinrich, K. Towards a holistic cross-border environmental governance in the European Arctic. J. Territ. Marit. Stud. 2022, 9, 31–46. [Google Scholar]
- Hjort, J.; Streletskiy, D.; Doré, G.; Wu, Q.; Bjella, K.; Luoto, M. Impacts of permafrost degradation on infrastructure. Nat. Rev. Earth Environ. 2022, 3, 24–38. [Google Scholar] [CrossRef]
- McCarty, J.L.; Aalto, J.; Paunu, V.-V.; Arnold, S.R.; Eckhardt, S.; Klimont, Z.; Fain, J.J.; Evangeliou, N.; Venäläinen, A.; Tchebakova, N.M.; et al. Reviews & syntheses: Arctic fire regimes and emissions in the 21st century. Biogeosci. Discuss. 2021, 18, 5053–5083. [Google Scholar]
- Paul, M.; Swistek, G. Russia in the Arctic: Development Plans, Military Potential, and Conflict Prevention; SWP: Berlin, Germany, 2022. [Google Scholar] [CrossRef]
- Groisman, P.Y.; Gutman, G. (Eds.) Regional Environmental Changes in Siberia and Their Global Consequences; Springer Science & Business Media: Berlin, Germany, 2012. [Google Scholar]
- Swanson, D.; Murphy, D.; Temmer, J.; Scaletta, T. Advancing the climate resilience of Canadian infrastructure. Int. Inst. Sustain. Dev. 2021, 118, 2021-07. [Google Scholar]
- Crépin, A.S.; Karcher, M.; Gascard, J.C. Arctic climate change, economy and society (ACCESS): Integrated perspectives. Ambio 2017, 46, 341–354. [Google Scholar] [CrossRef] [PubMed]
- Wamsler, C.; Niven, L.; Beery, T.H.; Bramryd, T.; Ekelund, N.; Jönsson, K.I.; Osmani, A.; Palo, T.; Stålhammar, S. Operationalizing ecosystem-based adaptation: Harnessing ecosystem services to buffer communities against climate change. Ecol. Soc. 2016, 21, 31. [Google Scholar] [CrossRef]
- Flynn, M.; Ford, J.D.; Pearce, T.; Harper, S.L.; IHACC Research Team. Participatory scenario planning and climate change impacts, adaptation and vulnerability research in the Arctic. Environ. Sci. Policy 2018, 79, 45–53. [Google Scholar] [CrossRef]
- Bhavani, K.; Gajendra, N. Smart data-driven sensing: New opportunities to combat environmental problems. In Bio-Inspired Data-Driven Distributed Energy in Robotics and Enabling Technologies; CRC Press: Boca Raton, FL, USA, 2024; pp. 17–47. [Google Scholar]
- Pigford, A.A. Exploring the Public Value of Networked Science in the Canadian Arctic. Ph.D. Thesis, McGill University, Montreal, QC, Canada, 2021. [Google Scholar]
- Buschman, V.Q. Indigenous Contributions to Arctic Biodiversity Conservation. Ph.D. Thesis, University of Washington, Seattle, WA, USA, 2021. [Google Scholar]
| Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).