Fact-checked by Grok 2 weeks ago

Climate change impacts

Climate change impacts encompass the observable consequences of alterations in Earth's , including a rise of approximately 1.2°C since the late baseline, regional shifts in and patterns, and a cumulative increase of 21–24 cm since 1880, with recent decadal rates averaging 3–4 mm per year. These changes, primarily driven by elevated atmospheric concentrations of and other greenhouse gases from combustion and land-use modifications, have led to documented effects such as retreat, thawing in high latitudes, and enhanced plant productivity through CO2 fertilization, which satellite observations attribute to 70% of global vegetation greening trends over the past four decades. While adverse outcomes include from and heat-related stresses on in vulnerable regions, empirical records reveal no consistent intensification of frequency or intensity, and drought patterns show regional variability rather than uniform worsening. Controversies persist regarding the attribution and magnitude of these impacts, as models have frequently overestimated warming rates—such as CMIP5 simulations projecting 16% faster surface air increases than observations since 1970—and projected catastrophic escalations in extremes that have not materialized in data. Such discrepancies underscore uncertainties in mechanisms like and highlight the need for causal analysis grounded in instrumental records over model-dependent scenarios.

Overview and Fundamentals

Definition and Causal Framework

Climate change impacts refer to the effects of observed and projected alterations in the Earth's —such as shifts in global mean temperature, regimes, and the frequency or intensity of extreme events—on natural ecosystems, human societies, and socioeconomic systems. These impacts arise from detectable changes in climatic variables that persist beyond internal variability, influencing phenomena like , , , and infrastructure . The causal framework originates from the physics of , where elevated atmospheric concentrations of greenhouse gases, primarily (CO₂) from combustion and land-use changes, absorb outgoing longwave infrared radiation emitted by the Earth's surface, thereby reducing the planet's effective thermal emission to space and creating a positive imbalance at the top of the atmosphere. This imbalance, quantified as approximately 0.5 to 1 W/m² in recent decades, drives an increase in global surface temperatures through enhanced downward infrared radiation and subsequent processes. Warming initiates a sequence of thermodynamic and dynamic responses: ocean heat uptake exceeds 90% of excess energy, causing that contributes to at rates of about 1.7 mm/year since 1900; land ice melt from and accelerates mass loss, adding roughly 0.4 mm/year to sea levels; and increased atmospheric moisture capacity (about 7% per 1°C warming per the Clausius-Clapeyron relation) alters evaporation, condensation, and storm dynamics, potentially intensifying heavy rainfall while shifting drought patterns in subtropical regions. Feedback loops, including water vapor amplification (doubling CO₂'s direct effect) and reduced surface from ice loss, further modulate the magnitude of these physical changes, propagating them into broader environmental and ecological consequences.

Attribution to Anthropogenic vs. Natural Factors

Detection and attribution studies employ climate models to simulate trends under scenarios of forcings alone (e.g., variations, volcanic aerosols, and internal variability from ocean-atmosphere oscillations like ENSO and AMO) versus combined and forcings (primarily gases and aerosols). Simulations with forcings only reproduce observed cooling from major volcanic eruptions, such as in 1991, but fail to capture the sustained warming trend since the mid-20th century, particularly post-1950, where global surface have risen approximately 0.8°C. The attributes the majority of observed warming since 1850-1900—about 1.1°C—to , estimating that human influence accounts for the net imbalance of +2.72 W/m² as of 2019, with natural factors contributing negligibly to the post-1950 trend. This conclusion relies on "fingerprinting" techniques matching observed patterns (e.g., stratospheric cooling and tropospheric warming) to model outputs from anthropogenic forcings, which differ from natural variability signatures. However, critiques highlight that global climate models systematically underestimate natural variability at multidecadal scales, potentially overattributing warming to human causes by underrepresenting internal ocean dynamics and solar influences. Early 20th-century warming (roughly 1910-1940, ~0.4°C) is often linked to a combination of natural factors, including increased solar activity during the early 20th-century solar maximum and reduced volcanic activity, alongside emerging anthropogenic influences and internal variability; some analyses suggest these natural components explain much of that period's rise without requiring dominant human forcing. Solar total irradiance reconstructions show a ~0.1-0.2% increase from the Maunder Minimum to modern levels, correlating with pre-1950 temperature recoveries, though post-1980 solar activity has declined slightly while temperatures rose, limiting its explanatory power for recent decades. Volcanic contributions are predominantly cooling via sulfate aerosols reflecting sunlight, with net effects estimated at -0.1 to -0.2°C for major eruptions, but insufficient to offset long-term trends. Oceanic cycles like the Atlantic Multidecadal Oscillation (AMO) and Pacific Decadal Oscillation (PDO) introduce variability of ±0.1-0.3°C over decades, potentially amplifying or masking forced trends; for instance, the positive AMO phase since the has contributed to North Atlantic warming, but studies indicate internal variability accounts for less than 20-30% of 20th-century total warming, with forced signals dominating the multi-decadal rise. Dissenting analyses using unadjusted observational data argue for a larger natural role, incorporating solar-oceanic feedbacks and questioning model reliance on homogenized datasets that may inflate trends; one reassessment claims human CO₂'s contribution is overstated, with natural forcings and variability explaining up to 50-70% of observed changes when and alternative proxies are prioritized. These critiques, often from non-mainstream sources, underscore uncertainties in attribution amid institutional pressures for , where peer-reviewed challenges to dominant narratives face publication barriers.

Measurement and Data Sources

Global surface air , a foundational metric for assessing climate change impacts, are derived from networks of land-based stations, buoys, and ships, with datasets such as NOAA's GlobalTemp incorporating land surface air records from the Global Historical Climatology Network-Monthly and sea surface from the Extended Reconstructed Sea Surface Temperature dataset. Independent analyses, like those from Berkeley Earth, utilize high-resolution gridded time series from over 39,000 land stations and measurements to produce global land- anomalies. Satellite-based measurements, such as those from microwave sounding units, capture lower tropospheric , offering coverage over remote areas including polar regions and where surface data are sparse. Sea level rise, critical for evaluating coastal impacts, is monitored through tide gauge networks for relative sea level changes at specific locations and satellite altimetry missions for global mean sea level trends. NASA's Sea Level Change Portal aggregates data from missions like TOPEX/Poseidon, Jason-1, -2, and -3, providing gridded sea surface height anomalies with uncertainties typically below 4 cm. The European Space Agency's Copernicus Climate Change Service offers daily and monthly gridded sea level anomaly datasets from satellite altimetry, enabling assessments of variability driven by thermal expansion and mass changes from land ice. Precipitation and extreme weather events, linked to hydrological impacts, rely on rain gauge networks and reanalysis products integrating observations with models. NOAA's Climate Data Online provides historical records from global weather stations, supporting analyses of heavy precipitation trends. For extremes, datasets from the National Centers for Environmental Information track metrics like one-day precipitation maxima, with U.S. data showing increases in the heaviest events since the mid-20th century based on over 1,000 stations. Satellite-derived products, such as those from NASA's Global Precipitation Climatology Project, complement ground data by estimating rainfall over oceans and data-poor regions. Broader monitoring incorporates satellite constellations from and NOAA, including instruments on platforms like Landsat and GOES for surface changes and cloud properties relevant to impact attribution. These sources undergo homogenization to account for non-climatic factors like station relocations and instrument changes, though independent datasets like HadCRUT from the UK's Hadley Centre and Climatic Research Unit provide cross-verification using distinct methodologies. Proxy records, such as ice cores and tree rings, extend baselines beyond instrumental eras but are secondary to direct measurements for recent impacts.

Observed Physical Impacts

Global surface air temperature anomalies, relative to the 1850-1900 pre-industrial baseline, have risen by approximately 1.2°C as of 2024, with independent datasets from , NOAA, and Berkeley Earth converging on this estimate despite methodological differences such as station adjustments and urban heat island corrections. The rate of increase has accelerated since the mid-20th century, averaging 0.06°C per decade from 1850 to present but reaching 0.18°C per decade since 1975, driven by land and ocean measurements from thousands of weather stations, buoys, and satellites. The year marked the warmest on record across major datasets, with a global anomaly of about 1.35°C above the 1850-1900 , surpassing 2023 by 0.1-0.2°C; preliminary 2025 through indicate the third-warmest year to date, with monthly anomalies exceeding 1.4°C in several months like and . Regional disparities persist, with amplification amplifying warming by 2-3 times the global average due to ice-albedo feedback, while has absorbed over 90% of excess energy, contributing to sustained surface trends. Heatwave frequency and intensity have increased globally since the late , with peer-reviewed analyses showing a rise in events exceeding three consecutive days above the 95th threshold, particularly in , , and . Compound heatwaves—simultaneous hot and dry conditions—have intensified by 3.32°C per decade from 1980-2022, correlating with expanded drought-prone areas and elevated mortality risks during peaks like the 2023 European events. These trends align across reanalysis datasets up to , though natural variability such as El Niño episodes in 2023-2024 amplified recent extremes by 0.2-0.4°C.

Sea Level Rise and Cryospheric Changes

Global mean sea level has risen by approximately 11.1 cm from 1993 to 2023, as measured by satellite altimetry, with an average rate of 3.3 mm per year over that period. This rate reflects contributions from thermal expansion of seawater due to warming oceans and mass loss from land-based ice, though the relative dominance of each factor varies by decade and region. Satellite data indicate a recent increase to around 4.5 mm per year, potentially influenced by short-term events like the 2023 El Niño, which caused a 0.76 cm jump from 2022 levels. However, longer-term tide gauge records, spanning over a century, often show a more consistent linear rise of 1.5–2 mm per year without clear evidence of acceleration when accounting for local vertical land motion and instrumental biases. Discrepancies between satellite and tide gauge trends highlight uncertainties in post-glacial rebound corrections and potential overestimation in altimetry data adjustments, underscoring the need for integrated analyses. The has contributed significantly to recent , with mass loss averaging 266 billion tons per year as of 2025, equivalent to about 0.7 mm of annual global increase. Inter-comparison exercises like IMBIE estimate cumulative losses of 4,890 Gt from 1992 to 2020, accelerating from near-zero in the to over 200 Gt per year post-2000, driven primarily by surface melt and glacier calving. In contrast, the exhibits regional variability: and the show net losses of around 144 Gt per year from 2011–2020 via satellite gravimetry, while has experienced mass gains from increased snowfall, resulting in a modest overall net loss of 135 billion tons annually. These dynamics complicate attribution, as dynamic ice discharge and patterns interact with oceanic and atmospheric forcing, with data subject to glacial isostatic adjustment uncertainties. Global glacier mass balance has been negative since systematic monitoring began, with the World Glacier Monitoring Service reporting an average annual loss of about -0.85 meters water equivalent for reference glaciers in recent years, based on over 130 monitored sites. Cumulative losses since the 1970s equate to roughly 10–15% of glacier volume worldwide, accelerating in the 21st century due to enhanced melt from rising air temperatures, though variability tied to regional precipitation tempers uniform retreat. Arctic sea ice extent at the September minimum has declined by 12.2% per decade relative to 1981–2010 averages, reaching 4.28 million square kilometers in 2024—the seventh lowest on record—primarily from reduced summer melt onset and thinner multi-year ice. This trend reflects thermodynamic forcing from warmer surface air and ocean temperatures, though natural oscillations like the Atlantic Multidecadal Oscillation influence interannual variability. Permafrost thaw in the has progressed with ground temperatures rising at 0.6°F per decade on average, leading to active layer deepening and talik formation in some regions. Observations indicate thaw rates of 0.3–0.9 cm per year vertically in sub- catchments over the past century, with accelerating to decadal scales exceeding 1 meter per year in vulnerable areas due to wave undercutting and thermodenudation. Pan- permafrost extent has decreased from 13.4 million km² in 2003–2013 to 12.51 million km² in 2014–2023, releasing stored carbon and contaminants, though thaw propagation slows in ice-rich sediments and is modulated by vegetation insulation and drainage. These changes amplify local feedbacks like lake expansion but do not directly contribute to , as permafrost is terrestrial.

Precipitation, Droughts, and Extreme Weather Events

Observations indicate a modest global increase in average annual over areas of approximately 0.03 inches per decade since 1901, with greater increases in the United States at 0.18 inches per decade, though regional patterns vary significantly, including drying in subtropical zones and wetting in higher latitudes and some tropical areas. from the Global Precipitation Climatology Centre, derived from station observations since 1891, confirm this slight upward trend without uniform intensification across all regions. Attribution studies link part of the observed increase in heavy events—defined as daily amounts exceeded once per decade in the pre-industrial baseline—to human-induced warming, with medium confidence for more frequent extremes over in model projections, though empirical detection remains regionally inconsistent due to natural variability and limitations. Drought trends exhibit substantial variability by metric and region, with no robust evidence of a global increase in meteorological (based on deficits alone) over the past century, as assessed in reviews of standardized indices like the ; however, inclusion of in indices like SPEI reveals stronger drying signals in arid and semi-arid zones due to elevated temperatures amplifying . Recent analyses, including high-resolution gridded SPEI data from 1981–2022, identify increases in drought frequency and severity over more than half of global land area attributable to land-use changes and warming, with forcing implicated in heightened maximum durations and intensities in some datasets. Critiques of these findings highlight methodological sensitivities, such as the choice of drought and baseline periods, which can exaggerate trends when models overestimate atmospheric demand, and note that greenery indices show no widespread expansion of drylands despite localized agricultural droughts in regions like the Mediterranean and . Extreme weather events display mixed attribution outcomes, with low confidence in global trends for tropical cyclone frequency but medium confidence for increases in their peak wind intensities and rainfall rates since the 1980s, potentially linked to warmer sea surface temperatures; however, normalized economic damages and U.S. landfall counts show no accelerating pattern when adjusted for improved detection and coastal development. River floods lack detected human influence on frequency globally, though compound events with heavy precipitation exhibit regional upticks in intensity. Attribution science for these events relies on probabilistic modeling, which some analyses criticize for underemphasizing natural oscillations like the Atlantic Multidecadal Oscillation and over-relying on climate models with historical biases in simulating extremes, leading to overstated causal links in media interpretations despite stagnant or declining trends in many non-thermal extremes. Overall, while warming contributes to thermodynamic enhancements (e.g., higher moisture convergence for storms), empirical records through 2024 reveal no universal surge in event counts or severities, underscoring the dominance of variability over linear anthropogenic signals in short observational periods.

Environmental and Ecological Impacts

Terrestrial Greening and Vegetation Response

Satellite observations from instruments such as the (MODIS) and (AVHRR) have documented a pronounced greening of Earth's terrestrial surfaces since the early , characterized by increases in the (LAI) and (NDVI). Between 1982 and 2015, vegetated lands experienced a net greening equivalent to an additional leaf area spanning approximately 18 million square kilometers, or twice the land area of the , with 25% to 50% of global vegetated regions showing statistically significant positive trends. This phenomenon is most evident in regions like , , and boreal forests, where LAI has increased by up to 0.05 units per decade in some biomes. The primary driver of this is the , whereby elevated atmospheric CO2 concentrations—rising from about 340 ppm in 1980 to over 420 ppm by 2023—enhance photosynthetic rates, particularly in C3 that dominate global vegetation. Empirical modeling integrated with attributes roughly 70% of the observed from 1982 to 2010 directly to CO2 fertilization, with supporting evidence from Free-Air CO2 Enrichment (FACE) experiments demonstrating 10-20% increases in above-ground biomass under CO2 levels of 550 ppm. Secondary contributors include deposition from agricultural and industrial sources, extended growing seasons due to warming in high latitudes (adding 8-10 days per decade in some areas), and land management practices such as in . These factors collectively improve water-use efficiency, allowing to maintain higher productivity under stress by reducing while sustaining carbon assimilation. Vegetation responses extend to enhanced global net primary productivity (NPP), estimated to have risen by 10-30% over the same period, bolstering terrestrial carbon sinks that absorbed approximately 25-30% of CO2 emissions annually in recent decades. In and temperate zones, has amplified biomass accumulation, with newly established forests contributing disproportionately to —up to 50% of total terrestrial uptake in some models. However, empirical assessments indicate spatial variability, with tropical regions showing muted responses due to nutrient limitations like scarcity, and arid zones where water constraints dominate. Recent analyses reveal signs of a slowdown in trends post-2000, particularly in biomes, with some studies reporting a decline in the (CFE) on from 1982 to 2015, linked to diminishing nutrient availability and deficits amid rising temperatures. For instance, global CFE correlated negatively with trends in foliar and concentrations, suggesting potential saturation where additional CO2 yields marginal gains without corresponding nutrient inputs. Despite this, aggregate persisted into the 2020s, with achieving record-high greenness in 2020 driven by and temperate expansions, underscoring that CO2-driven enhancements have not yet been fully offset by countervailing stressors. This dynamic highlights the causal primacy of CO2 as a growth stimulant, though long-term projections must account for biophysical feedbacks like altered from denser canopies, which can locally mitigate warming by 0.2-0.3°C in greened areas.

Ocean and Marine Ecosystem Changes

Oceans have absorbed more than 90% of the excess heat trapped by greenhouse gases since the mid-20th century, leading to increased . Observations indicate that heat content in the global ocean has been consistently above average since the mid-1990s, with the top 700 meters warming since 1955 across multiple data analyses. In 2023, reached record highs, with accelerated accumulation in upper layers compared to earlier rates. This warming has driven physiological stress in marine organisms, including elevated metabolic rates and shifts in distribution patterns. Many exhibit poleward migrations, with global analyses showing abundance increases at the poleward edges of ranges and declines at equatorward edges due to thermal tolerances. For instance, and other mobile species have redistributed at rates up to six times faster than terrestrial counterparts, altering structures and food webs. Such shifts reflect thermal niche constraints rather than uniform , though they disrupt fisheries in equatorial regions while potentially benefiting high-latitude productivity in some cases. Ocean , resulting from CO2 absorption, has lowered surface by approximately 0.1 units since the industrial era (from ~8.2 to 8.1), equivalent to a 30% increase in concentration. This trend affects in shell-forming organisms like pteropods and mollusks, reducing states for in surface waters by over 0.6 units from 1800 to 2014 in the top 100 meters. Impacts include impaired shell growth and increased dissolution risks, particularly in polar and upwelling regions, though via or behavioral changes occurs in some populations. Coral reefs have experienced recurrent bleaching from marine heatwaves, with frequency rising due to prolonged . NOAA data link bleaching to degree heating weeks exceeding thresholds, as seen in the 2014–2017 global event and the ongoing fourth global bleaching episode starting in 2023, affecting 84.4% of reef areas by September 2025. While bleaching expels symbiotic and can lead to mortality, recovery is possible under moderate stress, and some species demonstrate through symbiont shuffling or acclimation. Warmer conditions have also shortened incubation times and boosted growth rates for certain fish , enhancing in specific locales. Deoxygenation compounds these effects, with global oxygen inventories declining 1–2% since the 1950s due to warmer water holding less dissolved oxygen and reducing vertical mixing. This has expanded hypoxic zones, compressing habitable volumes for and altering metabolic scopes, though some mid-latitude exploit expanded ranges. Overall, these changes have contributed to a net 4% decline in sustainable fisheries catch potential since , with equatorial losses outweighing poleward gains.

Biodiversity Shifts and Habitat Alterations

Observed shifts in species distributions include poleward migrations in terrestrial and taxa, with meta-analyses documenting average latitudinal advances of 17.1 km per decade for terrestrial and 72.1 km per decade for from 1960 to 2005, though recent assessments indicate only 46.6% of range-shift observations align with expected directional changes toward higher latitudes, elevations, or depths, highlighting variability and non-climatic influences such as land-use changes. Altitudinal shifts average 11.0 meters per decade upward, but success depends on connectivity, with fragmented landscapes impeding dispersal for many . These movements disrupt ecological interactions, including trophic mismatches where phenological changes—such as earlier breeding in birds or flowering in plants—desynchronize with prey or availability, potentially reducing by up to 20% in affected populations. Habitat alterations manifest prominently in vulnerable ecosystems. Coral reefs, covering less than 0.1% of ocean area, have experienced recurrent bleaching events, with over 14% global cover lost between 2009 and 2018 due to marine heatwaves elevating sea surface temperatures by 0.88°C above seasonal norms, leading to symbiosis breakdown with and mortality rates exceeding 90% in severe cases like the 2014–2017 event. In boreal forests, warming-induced droughts and pest outbreaks, including infestations amplified by milder winters, have caused dieback affecting millions of hectares in and since the , altering canopy structure and composition. Tundra habitats face thaw, releasing 1.5–1.7 Gt of carbon annually from 2000–2017 and converting frozen soils to wetlands, which shifts vegetation from moss-lichen dominance to shrub encroachment, reducing habitat suitability for specialized arctic species like caribou. While climate-driven changes contribute, direct pressures like and land conversion remain the primary drivers of recent , accounting for over 70% of terrestrial declines since 1970, with climate acting as an exacerbating factor rather than the sole cause in most assessments. Some exhibit or range expansions into newly suitable areas, particularly generalists, but narrow-range endemics in montane or polar regions face heightened risks, with projections estimating 15–37% of potentially losing over half their climatic range by 2100 under moderate warming scenarios. Empirical data underscore the need to disentangle climatic signals from confounding variables, as models attributing shifts solely to often overestimate directional changes when incorporating dispersal barriers or interactions.

Human and Societal Impacts

Agricultural Productivity and Food Security

Global agricultural productivity has shown substantial increases over the past century, with major staple crop yields rising despite a 1.1°C global temperature increase since pre-industrial times. For example, global maize yields per hectare grew from approximately 1.8 metric tons in 1961 to over 5.8 metric tons in 2020, driven by technological advancements, improved crop varieties, and elevated atmospheric CO2 levels. Similarly, wheat and rice yields have more than doubled in the same period, outpacing population growth and contributing to a decline in global undernourishment from 23% of the population in 1990 to about 9% in 2019. These trends indicate that warming has not halted productivity gains, though regional variations exist. Elevated CO2 concentrations, now at around 420 ppm, provide a fertilization effect that enhances and water-use efficiency in C3 crops like , , and soybeans, accounting for a significant portion of historical yield improvements. Empirical analyses estimate that a 1 ppm rise in CO2 correlates with 0.4% higher corn yields, 0.6% for soybeans, and 1% for in U.S. field data from 1983 to 2020, implying that CO2 has boosted global yields by roughly 20-30% over the industrial era. This effect partially offsets warming-induced stresses, as free-air CO2 enrichment experiments demonstrate 10-20% yield increases for major crops at doubled CO2 levels under moderate temperatures. However, benefits diminish under nutrient limitations or extreme heat, and C4 crops like show smaller responses. Observed negative impacts include reduced yields in tropical regions from higher temperatures exceeding optimal thresholds for key crops. Statistical models applied to historical data find that each 1°C warming has decreased yields by 7.5±5.3%, by 6.0±3.3%, and by 3.3±2.9% globally, with stronger effects in low-latitude areas lacking . and floods, intensified in frequency in some areas, have caused localized losses; for instance, the U.S. drought reduced corn production by 13%, while floods in periodically damage paddies. Yet, these events have not reversed the upward global trend, as expansions and resilient varieties mitigate damages, and greening from CO2 has expanded suitability in higher latitudes. Food security, measured by caloric availability and undernourishment rates, has improved amid modest warming, with food rising 30% since per FAO data. Projections incorporating suggest net global yield declines of 3-12% by mid-century under moderate emissions scenarios, but underscores that technological progress and CO2 effects could sustain or exceed current growth if historical patterns hold. Challenges persist in vulnerable regions, where amplifies climate sensitivities, but attributing recent spikes primarily to warming overlooks dominant roles of conflicts and economic disruptions. Overall, causal factors like breeding innovations have outweighed climatic pressures to date, supporting in global systems.

Health Effects and Mortality Rates

Temperature-related mortality constitutes the primary direct health impact attributed to climate variability, with cold exposure causing substantially more deaths than heat worldwide. Globally, cold-related deaths outnumber heat-related ones by a factor of approximately 9:1, based on analyses of data across multiple regions. In , spanning 854 cities, the ratio reaches about 10:1. These disparities persist even in warmer climates, where still accounts for the majority of temperature-attributable fatalities. Heatwaves have driven spikes in mortality, particularly among vulnerable populations such as the elderly and those with pre-existing conditions, but adaptation measures—including widespread adoption, early warning systems, and urban heat mitigation—have reduced per-event death rates in developed regions. , heat-related mortality vulnerability has continued to decline over the past decade despite rising temperatures, reflecting improved and behavioral responses. From 1979 to 2022, U.S. heat exposure death rates remained low, fluctuating between 0.5 and 2 per million population annually. Globally, heat accounts for roughly 489,000 deaths yearly between 2000 and 2019, concentrated in (45%) and (36%), though these figures include indirect contributors like cardiovascular strain. Projections for future temperature-related mortality vary by scenario and adaptation assumptions, with some models estimating net increases driven by heat, potentially adding 41,850 to 96,072 annual deaths under moderate to high warming, largely due to and aging rather than alone. However, reductions in cold-related deaths—estimated at 8.3 times higher than heat currently—could offset heat gains in higher-latitude regions, leading to heterogeneous outcomes: net decreases in colder areas but increases in . Empirical trends underscore 's role, as U.S. winter death rates remain 8-12% higher than non-winter months, suggesting potential benefits from milder winters. Vector-borne diseases, such as and dengue, show sensitivity to and shifts, with enabling range expansions in some temperate and areas. These illnesses cause over 700,000 deaths annually, but attributable mortality increases from recent warming remain limited and confounded by socioeconomic factors, interventions like insecticides, and non-climatic drivers. For instance, global incidence has declined despite temperature rises, due to control measures, while evidence for climate-driven surges in mortality is correlative rather than causal in most peer-reviewed assessments. Over half of human pathogenic diseases have been aggravated by climatic hazards at some point, but net mortality impacts hinge on adaptive responses beyond temperature alone.

Water Resources and Infrastructure Vulnerabilities

Climate change has altered hydrological cycles, increasing variability globally by approximately 1.2% per decade since the mid-20th century, which elevates risks of both intensified floods and prolonged droughts in affected regions. This variability stems from enhanced atmospheric moisture and shifts in circulation patterns under warming conditions, resulting in more erratic daily rainfall rather than uniform increases or decreases. Empirical observations indicate heavier events over land areas like the , where extreme events have risen since 1996, linked to Atlantic warming. Conversely, some regions experience drying trends, exacerbating where evaporation rates outpace supply replenishment. Snowpack accumulation, a critical for seasonal in mountainous regions, has declined sharply since the due to rising temperatures shifting from to rain and accelerating melt. In the , reductions of 10-20% per decade have been observed in parts of the southwestern and , reducing spring and summer runoff for downstream and use. This has led to earlier peak streamflows, straining water management systems reliant on , as evidenced in western U.S. basins where historical patterns no longer align with current conditions. In , such changes have already contributed to disrupted operating seasons for and increased costs for reservoirs and facilities. Water , including , pipelines, and treatment plants, faces heightened vulnerabilities from these shifts, with extreme events causing physical damage and operational failures. Floods from intense can overload networks, leading to breaches or , while droughts increase on supplies and elevate risks of structural from reduced maintenance capacity. Coastal systems are particularly susceptible to saline intrusion from combined with droughts, allowing saltwater to infiltrate aquifers and estuaries farther inland, as observed in U.S. freshwater resources. Peer-reviewed assessments project that unadapted infrastructure could see damages multiply under escalating extremes, though historical data underscore that aging designs—often predating modern climate trends—amplify these risks beyond direct climatic forcing. In regions like and , concurrent heatwaves and droughts have increased infrastructure stress, with examples including failures from and flood-induced scour on foundations.

Economic and Policy Dimensions

Quantified Economic Damages and Costs

Integrated assessment models () such as , , and FUND quantify damages by linking emissions to rises, sectoral impacts, and GDP reductions, often estimating 1-3% global GDP losses for 2-3°C warming by 2100 under baseline scenarios without aggressive . Recent updates incorporating empirical damage functions have projected higher committed losses, including a 19% global income reduction by 2051 independent of future emissions pathways, driven by persistent effects on productivity in , labor, and . These models derive damage functions from econometric analyses of historical variations, but they vary significantly based on assumptions about , discount rates (typically 2-5%), and points. Criticisms of IAMs highlight their limitations in capturing dynamic human responses, such as and adaptive infrastructure, which may overstate net damages by assuming static economies and ignoring path-dependent growth. For instance, models often fail to fully account for induced technical change or regional heterogeneity, leading to exaggerated mitigation cost-benefit ratios and inconsistent (SCC) estimates across studies. Empirical calibrations suggest IAM damage projections exceed observed historical impacts, where normalized losses from temperature extremes show limited macroeconomic persistence beyond 1-2 years. The SCC, defined as the marginal global damage cost of emitting one additional of CO₂, serves as a key metric for , with peer-reviewed meta-analyses yielding central estimates of $50-70 per (2020 USD) at standard rates, though ranges span $44-413 due to uncertainties in and equity weighting. Higher values, such as $185 per , arise from lower rates (e.g., 2%) and inclusion of non-market damages like , but these are contested for lacking robust empirical validation of tail risks. Historical trends in economic losses from , normalized to GDP, reveal declining global vulnerability, with loss rates dropping markedly since the due to improved , infrastructure , and accumulation outpacing hazard intensity. Climate-attributable portions of these losses remain modest, averaging 0.05-0.82% of global GDP annually over recent decades, concentrated in sectors like and coastal property but offset by growth in unaffected regions. Severe events can reduce regional GDP by up to 2.2% short-term, with 1.7% persistence after five years, yet aggregate global impacts have not accelerated proportionally to warming. Overall, these observed costs represent a small of annual global growth (2-3%), underscoring adaptation's role in containing damages below model projections.

Adaptation Strategies and Resilience Measures

Adaptation strategies encompass a range of proactive and reactive measures designed to moderate the adverse effects of or to capitalize on opportunities arising from them, such as through enhanced practices or modifications. These include structural interventions like levees and reservoirs, as well as softer approaches such as diversification and early warning systems. Empirical evidence indicates that targeted can significantly reduce vulnerabilities; for instance, in , the adoption of drought-resistant varieties and improved has been shown to offset a portion of yield losses attributed to elevated temperatures, with global excess heat and linked to 9-10% net reductions in output absent such measures. measures, by contrast, emphasize systemic capacity to withstand and recover from disturbances, often integrating in and community-based planning to enhance overall durability against events like floods or heatwaves. In agricultural contexts, has demonstrated measurable efficacy through farmer-level interventions. Smallholder farmers employing strategies like adjusted planting schedules and practices have reported income improvements and better , as evidenced by surveys in regions prone to variable . A cost-benefit analysis of five such strategies— including and —revealed positive net returns, with benefits exceeding costs by factors of 1.5 to 3 in modeling for vulnerable farming communities. Psychological factors also play a role, with higher psychological capital among farmers correlating to greater uptake of these practices, thereby bolstering in empirical studies from . However, implementation barriers persist, including access to and , underscoring the need for evidence-based frameworks to scale interventions effectively. Coastal and urban resilience efforts provide further case studies of success. In flood-vulnerable areas, infrastructure enhancements such as elevated roadways and permeable surfaces have reduced , with policies prioritizing regulatory outperforming less structured approaches in vulnerability assessments across multiple municipalities. The U.S. Environmental Protection Agency documents instances where utilities integrated climate projections into planning via tools like the Climate Resilience Evaluation and Awareness Tool, yielding avoided losses from ; one evaluation projected substantial risk reductions for water facilities through modest retrofits costing millions but preventing billions in potential damages. European examples, including Dutch delta management and Italian flood barriers, illustrate long-term investments that have contained sea-level rise impacts, with benefit-cost ratios often exceeding 1.5 for efficient measures like and setback policies. Economically, adaptation investments generally yield high returns when focused on high-impact areas. A analysis estimates that each dollar spent on resilience generates over $10 in benefits within a decade, potentially unlocking $1.4 trillion globally through avoided disruptions in sectors like and . Nonetheless, comprehensive cost-benefit analyses reveal challenges, including uncertainties in projecting future hazards, which can inflate estimates; adaptation spending is projected to reach trillions by 2050, surpassing outlays but requiring prioritization to avoid inefficient allocations. Success hinges on integrating local empirical data over generalized models, as observational studies emphasize grounded strategies for regional efficacy.

Benefits from Warming and CO2 Effects

Increased atmospheric CO2 concentrations have enhanced global vegetation growth through the , contributing to a measurable "" of Earth's surface. observations from 1982 to 2015 indicate that CO2 fertilization accounts for approximately 70% of this greening, with the remainder attributed to factors like deposition and land management changes. This effect has boosted rates by an estimated 30% globally from 1900 to 2010 as CO2 levels rose from 296 to 389 . Empirical data from enclosure experiments and studies confirm that elevated CO2 improves plant water-use efficiency by reducing , allowing crops to maintain productivity under drier conditions. In , CO2 fertilization has directly increased of major crops like and by about 7.1% from 1961 to 2017, based on statistical analysis of historical data adjusted for other factors such as and . Free-air CO2 enrichment (FACE) experiments, conducted in open fields to mimic real-world conditions, demonstrate gains of 10-20% for crops like and under doubled CO2 levels, though benefits vary by availability and temperature. These gains have helped offset stagnation or declines in regions facing other stressors, preventing further losses in staple crop production. Mild warming has extended growing seasons in mid- and high-latitude regions, enabling additional crop cycles or longer-maturing varieties. In the and , frost-free periods have lengthened by 10-20 days since the mid-20th century, correlating with increased maize production in cooler areas due to warmer springs and localized cooling effects from . strategies leveraging these extended seasons, such as shifting planting dates, have projected yield boosts of up to 12% for and under moderate warming scenarios. NOAA assessments note short-term benefits for farmers in temperate zones from earlier springs and prolonged warm periods suitable for planting. Warming has also reduced cold-related mortality, which historically exceeds heat-related deaths by a factor of 8-10 globally. Analysis of 1368 small areas from 1991-2020 estimates 363,809 annual cold-related deaths versus 43,729 heat-related ones, with warming from 2000-2019 yielding a net decline in temperature-attributable due to fewer cold deaths. Short-term projections suggest could slightly lower net temperature-related deaths, as decreases in cold fatalities outpace heat increases in many regions, though long-term trends may reverse this without . Historical evidence from warmer intervals, such as the (circa 950-1250 CE), links elevated temperatures to expanded agricultural output and population growth in and , with stable warm phases correlating to higher food supplies and fertility rates. Proxy records indicate these periods facilitated settlement in and enhanced monsoon-driven productivity in some areas, contrasting with cooler eras marked by famines and societal stress. Such patterns underscore potential upsides of moderate warming for human societies adapted to cooler baselines, though modern contexts differ due to higher baseline temperatures and population densities.

Projections and Uncertainties

Model-Based Future Scenarios

Model-based future scenarios for climate change impacts rely on ensembles of global climate models, such as those from the Phase 6 (CMIP6), which simulate atmospheric, oceanic, and land surface responses to specified pathways. These projections integrate socioeconomic narratives via (SSPs), ranging from sustainability-focused SSP1 to fossil fuel-intensive SSP5, combined with concentration levels akin to Representative Concentration Pathways (RCPs). The IPCC's Sixth Assessment Report (AR6) employs five illustrative scenarios: SSP1-1.9 (very low emissions, net-zero by 2050), SSP1-2.6 (low emissions), SSP2-4.5 (middle-of-the-road), SSP3-7.0 (regional rivalry, high emissions), and SSP5-8.5 (very high emissions). These drive projections of physical changes like and , which are then fed into sectoral impact models for ecosystems, , and . Projected global mean surface air temperature increases by 2100 relative to 1850–1900 vary widely across scenarios, with likely ranges (66% probability) of 1.3–2.4°C for SSP1-2.6 and 3.3–5.7°C for SSP5-8.5. Sea level rise, dominated by thermal expansion and glacier melt, projects medians of 0.28–0.55 m under SSP1-2.6 and 0.63–1.01 m under SSP5-8.5, excluding rapid Antarctic ice sheet instability that could add decimeters. Precipitation changes amplify the water cycle, with models forecasting wetter tropics and high latitudes but drier subtropics, alongside increased intensity of heavy precipitation events by 5–20% per degree of warming. Extreme events, including heatwaves and droughts, are projected to intensify, though regional variability is high and model skill for tropical cyclones remains limited.
ScenarioKey AssumptionsMedian Warming (°C, 2081–2100)Median Sea Level Rise (m, 2100)
SSP1-1.9Strong mitigation, low ~1.6~0.37
SSP1-2.6, net-zero CO2 by 2050~1.8~0.44
SSP2-4.5Current trends, moderate challenges~2.7~0.55
SSP3-7.0High population, slow progress~3.6~0.71
SSP5-8.5High demand, fossil fuels~4.4~0.77
Impacts on include modeled poleward shifts in ranges and elevated risks, particularly in montane and ecosystems, with up to 20–30% of at risk under higher warming despite CO2 fertilization effects on . Agricultural projections show net global declines of 5–25% by 2100 under SSP2-4.5, varying by and —yields may rise in higher latitudes but fall in due to heat stress and —though and breeding are often underrepresented in models. Water resources face heightened risks in Mediterranean and , while coastal infrastructure contends with erosion and inundation. Uncertainties stem from equilibrium climate sensitivity (ECS), assessed at 2.5–4.0°C in AR6 but higher in many CMIP6 models (mean ~3.9°C), leading to potential overestimation of warming due to exaggerated feedbacks; paleoclimate and observational constraints suggest values toward the lower end. High-emission scenarios like SSP5-8.5, assuming continued rapid growth, are deemed low-likelihood given stagnant global use since 2013, emissions from GDP, and commitments, rendering them upper-bound rather than baseline cases. Impact models often underestimate compound extremes and adaptation potential, with sectoral projections sensitive to assumptions about technological progress and human responses. Academic consensus bodies like the IPCC, while synthesizing peer-reviewed literature, may amplify upper-tail risks due to institutional incentives favoring cautionary narratives.

Historical Prediction Accuracy and Failures

Numerous predictions of imminent and severe impacts, issued by scientists, policymakers, and international bodies since the , have not occurred as described, raising questions about the precision of impact forecasting amid complex interactions between climate variability, , and socioeconomic factors. These include expectations of rapid ice loss, diminished snowfall in temperate regions, accelerated submerging islands, and surges in events like hurricanes. Such misses often stem from overreliance on simplified model extrapolations that underweight natural variability, feedback uncertainties, and human responses, as critiqued in analyses of historical projections. A prominent example involves sea ice extent. In 2008, former U.S. Vice President referenced scientific models indicating a "75 percent chance" that the entire North Polar ice cap could be "completely ice-free" during summer months within five to seven years, pointing to a potential ice-free by 2013–2015. Similar timelines were suggested by researchers like Wieslaw Maslowski in 2008, who estimated a strong possibility of an ice-free summer as early as 2013 based on submarine data and modeling. In reality, the September minimum extent reached a low of 3.41 million square kilometers in 2012 but has since stabilized around 4–4.5 million square kilometers annually through 2024, with no ice-free conditions observed; multi-year ice persists, and recent years show slower decline rates than projected. These discrepancies highlight model sensitivities to cloud feedbacks, ocean currents, and ice dynamics, which have led to overestimations of melt rates in some ensemble projections. Predictions of regional weather pattern shifts have also faltered. In March 2000, Dr. David Viner, a climatologist at the University of East Anglia's Climatic Research Unit (a key contributor to IPCC assessments), forecasted that "within a few years winter snowfall will become 'a very rare and exciting event'" in the , with future generations of children unlikely to experience snow due to warming. Contrary to this, the UK has endured recurrent heavy snowfalls, including the widespread disruptions of the 2009–2010 winter (with over 500,000 stranded travelers), the 2013 cold snap, and the 2018 "" storm that dumped up to 50 cm of snow in parts of and , causing fatalities and economic losses exceeding £1 billion. Such events align more with decadal variability, including influences from the , than with a monotonic decline in snowfall probability as anticipated.
PredictionSource and DateProjected Impact/TimelineObserved Outcome
North Polar ice cap completely ice-free in summer, citing models (2008)By 2013–2015Persistent summer minima ~4 million km² through 2024; no ice-free state
Winter snowfall rare in UK; children unfamiliar with David Viner, CRU (2000)Within a few years (~2003–2005)Multiple major snow events (e.g., 2010, 2018); annual averages stable or variable but not eliminated
Coastal nations wiped out by Noel Brown, UNEP (1989)By 2000 if no actionLow-lying areas like experienced ~3 mm/yr rise; no submersion, with some atolls accreting land via /sediment; population grew 50% since
Early assessments overestimated other impacts as well. A 1989 report quoted UN environmental official Noel Brown warning that "entire nations could be wiped off the face of the Earth" by rising seas by 2000 without drastic measures, implying submersion of atolls like the . Observed global has averaged 3.3 mm per year since 1993, totaling about 10 cm from 1990–2020, insufficient to inundate nations; land area has expanded 10% since 1978 due to natural accretion and human engineering, accommodating . Likewise, 2005 UN estimates projected 50 million "climate refugees" by 2010 from rising seas and disasters, but actual environmentally displaced persons numbered under 25 million globally by 2010, with most attributed to conflict or development rather than alone, and no mass materialized. These cases illustrate how adaptive measures and geological processes can mitigate projected vulnerabilities, though proponents argue outliers do not invalidate broader trends.

Role of Natural Variability in Projections

Natural climate variability encompasses internal fluctuations in the Earth's arising from interactions among the atmosphere, oceans, land, and ice, without external forcings such as greenhouse gases or changes. Key modes include the El Niño-Southern Oscillation (ENSO), which operates on interannual timescales and drives global temperature swings of up to 0.2–0.3°C; the (PDO), a longer-term pattern in Pacific sea surface temperatures with phases lasting 20–30 years; and the Atlantic Multidecadal Oscillation (AMO), featuring warm and cool phases over 60–80 years that modulate North Atlantic temperatures by approximately 0.4°C. In climate projections from coupled general circulation models (e.g., those in the CMIP6 used by the IPCC), natural variability is simulated through large ensembles to generate probabilistic ranges, aiming to capture the nature of these modes. However, empirical comparisons reveal that models often underestimate the amplitude of observed year-to-year and decadal variability, particularly in patterns like those over the North Atlantic. This discrepancy arises partly because models struggle to replicate the full spectrum of ocean-atmosphere and teleconnections, leading to simulated internal variability that is 10–20% weaker than and reanalysis data indicate for ENSO and PDO influences on global temperatures. The underrepresentation of natural variability inflates confidence in projections by narrowing uncertainty bands, especially for near-term (2030–2050) outcomes where internal modes can dominate over forced trends. For example, the 1998–2012 global warming slowdown (or "hiatus"), during which surface temperatures rose only 0.05°C per decade despite rising CO2, was substantially influenced by a transition to negative PDO and declining AMO phases, masking warming estimated at 0.1–0.2°C per decade in models. Similarly, the accelerated warming from 1970–2000 aligned with positive AMO and PDO phases, contributing up to 0.1°C to multidecadal trends independent of forcings. Projections that initialize from current states without fully resolving these phase dependencies risk forecasting linear trends that overlook potential decadal coolings or hiatuses, as seen in historical analogs where PDO shifts reduced global temperature rises by 30–50% temporarily. On regional scales, natural variability exacerbates projection uncertainties; decadal oscillations can rival or exceed signals in modulating extremes and anomalies, with studies showing variability changes outweighing mean shifts in extreme event frequencies under certain scenarios. Attribution efforts attempting to isolate human influences must therefore employ detection methods that statistically separate modes like ENSO, but residual uncertainties persist due to incomplete observational baselines predating and model biases in simulating low-frequency variability. While forcings dominate long-term (century-scale) projections, incorporating empirical constraints from paleoclimate proxies and modern indices reveals that unforced multidecadal swings could alter effective warming rates by ±0.1–0.2°C over 20–30 years, underscoring the need for scenario ensembles that explicitly modulate internal variability amplitudes.

Controversies and Debates

Debates on Impact Magnitude and Alarmism

Critics of prevailing climate narratives argue that projections of catastrophic impacts have frequently overstated the magnitude of harms attributable to anthropogenic warming, fostering unnecessary alarmism that distorts policy priorities toward costly mitigation at the expense of adaptation and development. Bjorn Lomborg, in his 2020 book False Alarm, contends that exaggerated fears lead to inefficient spending, estimating that aggressive emission reduction policies could cost trillions while yielding negligible temperature reductions, whereas targeted investments in resilience yield higher returns. He substantiates this with data indicating a 98% decline in deaths from climate-related disasters like droughts, floods, and storms since the early 20th century, attributable primarily to technological and economic advancements rather than stabilized emissions. Lomborg's analysis draws on normalized disaster loss records, challenging claims of escalating human vulnerability. Climatologist emphasizes uncertainties in impact attribution, asserting that institutional incentives within academia and funding bodies amplify high-end scenarios while downplaying natural variability and adaptive capacities. She critiques the overreliance on equilibrium estimates that favor alarmist outcomes, noting peer-reviewed assessments showing observed warming rates below many model projections since the 1990s. For instance, evaluations of CMIP5 models reveal that a majority overestimated tropospheric warming by 10-20% compared to satellite observations through 2019. Curry argues this discrepancy underscores the need for probabilistic risk assessments over deterministic catastrophe narratives, particularly given stagnant trends in normalized economic damages from weather extremes. Debates intensify over attribution, where alarmist claims of surging hurricanes, floods, and heatwaves clash with observational data showing no statistically significant global increases. Pielke Jr.'s analyses of insured losses and EM-DAT records demonstrate that, after adjusting for , wealth, and exposure, disaster costs have not risen in tandem with . Proponents of heightened impacts, often citing IPCC summaries, invoke precautionary principles amid potential tipping points, yet critics highlight the absence of empirical validation for such nonlinear escalations and note systemic biases in media amplification of worst-case scenarios from sources with vested interests in sustained funding. This tension reflects broader concerns that alarmism, while mobilizing action, erodes credibility when predictions like rapid ice loss or submersion of island nations fail to materialize as forecasted in the early . Empirical focus on verifiable trends, such as modest sea-level rise accelerations (3-4 mm/year since 1993), supports arguments for measured responses over panic-driven overhauls.

Criticisms of Attribution Science

Criticisms of attribution science, which seeks to quantify the influence of factors on trends and events, primarily focus on statistical biases, model limitations, and inadequate handling of uncertainties that may inflate estimates of human causation. In detection and attribution analyses of , the prevalent optimal fingerprinting method applies (TLS) regression to align observed patterns with model-simulated "fingerprints" of forcings, but this approach introduces upward biases when and natural forcings exhibit negative correlations in models, leading to overstatements of impacts by up to several times the true effect. Ross McKitrick's analysis demonstrates that TLS overcorrects for errors-in-variables, a flaw exacerbated by the omission of instrumental variables methods more robust to such correlations. These issues render standard fingerprinting unreliable for isolating causal contributions, as it assumes model error structures that do not align with observational realities. Extreme event attribution (EEA), which estimates how human influence alters the probability or intensity of specific weather events, draws similar methodological critiques for relying on coarse-resolution climate models ill-suited to resolve localized and extremes. Studies often employ a small ensemble of 2 to 30 models, heightening the risk of where unrepresentative simulations miss the true signal or variability, potentially overstating effects. Proxy definitions in EEA, such as using multi-day averages for heatwaves or droughts, prioritize thermodynamic responses (e.g., Clausius-Clapeyron scaling for ) while sidelining circulatory like blocking highs, which can counteract or amplify risks independently of warming. For example, attribution of the 2013 Central European floods emphasized a 6% thermodynamic probability increase but overlooked dynamic reductions in overall risk, yielding misleading net assessments. Similarly, the analysis defined the event via seasonal mean temperatures, neglecting persistent anticyclonic patterns better explained by natural variability. Broader concerns label EEA as pseudoscientific for diverging from rigorous detection frameworks, assuming distributional shifts in extremes mirror global mean surface temperature trends without validating against multidecadal natural variability or local factors like aerosols and land use. Roger Pielke Jr. argues that EEA's probabilistic claims often stem from arbitrary data choices—such as truncating Nepal rainfall records from 1979 instead of 1950 or discarding Mediterranean datasets—yielding manipulable results that contradict IPCC findings of no detected anthropogenic trends in tropical cyclone frequency, drought persistence, or heavy precipitation indices through 2021. This event-centric focus invites selection bias by analyzing only realized disasters while ignoring non-events or warming-induced mitigations, such as reduced weaker storms, distorting perceptions of net climate impacts. Rapid, pre-peer-reviewed EEA outputs, as from the World Weather Attribution initiative, have propagated errors—like a 50% overestimate in Central European flood risk later revised to 5%—fueling media narratives and litigation without scrutiny. These flaws collectively undermine attribution's causal claims, as models' documented biases in simulating variability and extremes propagate into overconfident statements that prioritize thermodynamic signals over holistic assessments, potentially misleading policy on priorities. Critics emphasize that while attribution can inform thermodynamic contributions, its extension to probabilistic event blame requires fuller , including storyline approaches that integrate dynamics without assuming model perfection. Despite peer-reviewed advancements, the field's reliance on ensembles tuned to present-day observations may systematically understate internal variability, echoing longstanding debates over fingerprinting's sensitivity to forcing assumptions.

Comparative Historical Climate Contexts

The Earth's climate has exhibited substantial natural variability over the epoch (approximately 11,700 years ago to present), with periods of warming and cooling driven primarily by orbital forcings, solar activity, and volcanic influences rather than CO2. reconstructions from ice cores, tree rings, sediments, and corals indicate that global mean temperatures during the (roughly 9,000 to 5,000 years ) were 1 to 2°C warmer than mid-20th-century levels, particularly in the summers, coinciding with lower atmospheric CO2 concentrations around 260 ppm. This era featured expanded forests, a greener region due to intensified activity, and minimal ice cover in high latitudes, demonstrating to warmer conditions without industrial-era emissions. Earlier in the and , periods like the (approximately 130,000 to 115,000 years ago) saw global temperatures 1 to 2°C above pre-industrial levels, with sea levels 6 to 9 meters higher than today due to reduced and ice volumes. and sedimentary records from sites like confirm peak sea-level highs of up to 10 meters above present during this time, yet coastal ecosystems and early human populations adapted without the infrastructure vulnerabilities highlighted in modern impact assessments. These elevations resulted from natural dynamics, with rates of rise occasionally exceeding 1 meter per century in localized areas, underscoring that significant sea-level fluctuations are inherent to glacial-interglacial cycles. In the Common Era, the Roman Warm Period (circa 250 BCE to 400 CE) featured Mediterranean sea surface temperatures approximately 2°C higher than the late 20th-century average, facilitating agricultural expansion and urban growth across the Roman Empire through extended growing seasons. Tree-ring and speleothem data similarly reconstruct the Medieval Warm Period (circa 950 to 1250 CE) as regionally synchronous warmth in the North Atlantic, North America, and parts of Asia, with some proxy series indicating peak temperatures comparable to or exceeding those of the early 21st century in extratropical latitudes. Viking settlements in Greenland and viticulture in northern England during this interval reflect adaptive human responses to mild conditions, absent the high CO2 levels of today. Conversely, the Little Ice Age (circa 1300 to 1850 CE) brought Northern Hemisphere cooling of 0.5 to 1°C below the preceding baseline, marked by alpine glacier advances, frozen rivers like the Thames (last fully in 1814), and harvest failures contributing to social upheavals such as the 17th-century European crises. These historical episodes illustrate that climate shifts of 1 to 2°C magnitude—similar in scale to the observed 20th-21st century warming—have occurred without catastrophic disruptions to global biospheres or human societies, though regional impacts varied. Over the eon (past 485 million years), surface temperatures fluctuated between 11°C and 36°C, with biodiversity peaks during warmer intervals, challenging narratives of unprecedented modern sensitivity. Reconstructions emphasize natural forcings' dominance in pre-industrial variability, providing context for evaluating current changes against a backdrop of inherent climatic instability rather than equilibrium.

References

  1. [1]
    Global Temperature - Earth Indicator - NASA Science
    Sep 25, 2025 · NASA records show global temperature was 2.65 degrees Fahrenheit (1.47 degrees Celsius) warmer than the late-19th century (1850 - 1900) ...Missing: empirical | Show results with:empirical
  2. [2]
    Climate Change: Global Sea Level
    Global average sea level has risen 8–9 inches (21–24 centimeters) since 1880. In 2023, global average sea level set a new record high—101.4 mm (3.99 inches) ...
  3. [3]
    Carbon Dioxide Fertilization Greening Earth, Study Finds - NASA
    Apr 26, 2016 · Results showed that carbon dioxide fertilization explains 70 percent of the greening effect, said co-author Ranga Myneni, a professor in the ...
  4. [4]
    Global Warming and Hurricanes
    As far as Category 4-5 intensity storms, basin-wide unadjusted storm counts show a pronounced increase since the mid-1940s (Bender et al., 2010), but those ...Missing: empirical | Show results with:empirical<|separator|>
  5. [5]
    Analysis: How well have climate models projected global warming?
    Oct 5, 2017 · Global surface air temperatures in CMIP5 models have warmed about 16% faster than observations since 1970. About 40% of this difference is due ...<|control11|><|separator|>
  6. [6]
    Glossary — Global Warming of 1.5 ºC
    Climate change refers to a change in the state of the climate that can be identified (e.g., by using statistical tests) by changes in the mean and/or the ...
  7. [7]
    Climate Change 2022: Impacts, Adaptation and Vulnerability
    The report assesses climate change impacts on ecosystems, biodiversity, and human communities, and reviews adaptation capacities and limits.Summary for Policymakers · Chapter 14: North America · Chapter 5: Food, Fibre and
  8. [8]
    The Causes of Climate Change - NASA Science
    Oct 23, 2024 · Four Major Gases That Contribute to the Greenhouse Effect · Carbon Dioxide · Methane · Nitrous Oxide · Chlorofluorocarbons (CFCs).
  9. [9]
    What is the greenhouse effect? - NASA Science
    The greenhouse effect is the process through which heat is trapped near Earth's surface by substances known as 'greenhouse gases.'
  10. [10]
    Climate change widespread, rapid, and intensifying – IPCC
    Aug 9, 2021 · Climate change is intensifying the water cycle. This brings more intense rainfall and associated flooding, as well as more intense drought in ...
  11. [11]
    Thermal Expansion - NASA Sea Level Change Portal
    The warming of Earth is primarily due to accumulation of heat-trapping greenhouse gases, and more than 90 percent of this trapped heat is absorbed by the oceans ...
  12. [12]
    How Atmospheric Water Vapor Amplifies Earth's Greenhouse Effect
    Feb 8, 2022 · “As humans add carbon dioxide to the atmosphere, small changes in climate are amplified by changes in water vapor. This makes carbon dioxide a ...
  13. [13]
    Evidence - NASA Science
    Oct 23, 2024 · Global Temperature Is Rising. The planet's average surface temperature has risen about 2 degrees Fahrenheit (1 degrees Celsius) since the ...Missing: pre- | Show results with:pre-
  14. [14]
    The early 20th century warming: Anomalies, causes, and ...
    The results suggest that all factors discussed above contributed: anthropogenic forcing, natural forcing, and anomalous climate variability. Figure 5 ...
  15. [15]
    Detection, attribution, and modeling of climate change: Key open ...
    May 13, 2025 · This paper discusses a number of key open issues in climate science. It argues that global climate models still fail on natural variability at all scales.
  16. [16]
    Factors of natural climate variability contributing to the Early 20th ...
    The key issue is to assess the contribution of internal variability and external natural and human impacts.Missing: peer- | Show results with:peer-
  17. [17]
    Couldn't the Sun be the cause of global warming? | NOAA Climate.gov
    Oct 29, 2020 · Observations of sunspots and other indicators of solar activity indicate no changes that could have caused global warming over the past half ...
  18. [18]
    Long-term natural variability and 20th century climate change - NIH
    Observations suggest the warming of the 20th century global mean surface temperature has not been monotonic, even when smoothed by a 10–20 year low-pass filter.Missing: peer- reviewed
  19. [19]
    A limited role for unforced internal variability in 20th century warming
    May 20, 2019 · Currently, half of the observed warming during that time is attributed to internal ocean variability, which is a key reason why the estimate of ...
  20. [20]
    [PDF] A Critical Reassessment of the Anthropogenic CO₂-Global ...
    Mar 18, 2025 · This analysis integrates unadjusted observational data and recent peer-reviewed studies to demonstrate that the assertion of human CO₂ ...<|separator|>
  21. [21]
    Global Surface Temperature Dataset Updated | News
    Feb 14, 2024 · NOAAGlobalTemp consists of land surface air temperature (LSAT) records from the Global Historical Climatology Network-Monthly, and sea surface ...Missing: sources | Show results with:sources
  22. [22]
    Data Overview - Berkeley Earth
    Berkeley Earth provides high-resolution land and ocean time series data and gridded temperature data.
  23. [23]
    Climate Change Indicators: U.S. and Global Temperature - EPA
    Surface data come from land-based weather stations. Satellite measurements cover the lower troposphere, which is the lowest level of the Earth's atmosphere. “ ...
  24. [24]
    Climate Change Indicators: Sea Level | US EPA
    These data are based on measurements collected by satellites and tide gauges. Relative sea level data are available from the National Oceanic and Atmospheric ...
  25. [25]
    NASA Sea Level Change Portal
    Visualize and access information and data relevant to understanding and planning for sea level rise in response to ongoing climate change.
  26. [26]
    Sea level gridded data from satellite observations for the global ...
    Jun 14, 2018 · This dataset provides gridded daily and monthly mean global estimates of sea level anomaly based on satellite altimetry measurements.
  27. [27]
    Climate Data Online (CDO)
    Climate Data Online (CDO) provides free access to NCDC's archive of global historical weather and climate data in addition to station history information.Data Tools · NOAA CDO Web Services · Climate Data Records · NCEI Web ServicesMissing: impacts | Show results with:impacts
  28. [28]
    Climate Change Indicators: Heavy Precipitation | US EPA
    Data Sources. The data used for this indicator come from a large national network of weather stations and were provided by the National Oceanic and ...
  29. [29]
    Extreme Weather and Climate Change - NASA Science
    Oct 23, 2024 · As carbon dioxide, methane, and other gases increase, they act as a blanket, trapping heat and warming the planet. In response, Earth's air and ...
  30. [30]
    NASA Earthdata: Your Gateway to NASA Earth Observation Data
    The Earth Science Data Systems (ESDS) Program provides open access to NASA's archive of Earth science data, empowering researchers and decision makers.About the ESDS Program · Find Data · Worldview · Lance
  31. [31]
    Satellites | National Oceanic and Atmospheric Administration
    Aug 29, 2025 · The NOAA Satellite and Information Service provides timely access to global environmental data from satellites and other sources to monitor and understand our ...
  32. [32]
    CO₂ and Greenhouse Gas Emissions - Our World in Data
    It is called the “HadCRUT” (Hadley Centre/Climatic Research Unit Temperature) dataset. It measures temperature anomalies across the world at high resolutions.Climate Change · Global aviation · The safest energy sources · Fossil fuel subsidies
  33. [33]
    Global Temperature Report for 2024 - Berkeley Earth
    Jan 10, 2025 · We conclude that 2024 was the warmest year on Earth since 1850, exceeding the previous record just set in 2023 by a clear and definitive margin.
  34. [34]
    Climate change: global temperature
    May 29, 2025 · Earth's temperature has risen by an average of 0.11° Fahrenheit (0.06° Celsius) per decade since 1850, or about 2° F in total.
  35. [35]
    GISS Surface Temperature Analysis (GISTEMP v4) - Data.GISS
    The GISS Surface Temperature Analysis version 4 (GISTEMP v4) is an estimate of global surface temperature change. Graphs and tables are updated about the 10th ...Global Maps · History · FAQs · News, Updates, and Features
  36. [36]
  37. [37]
    Monthly Climate Reports | Global Climate Report | July 2025
    July 2025 recorded a global surface temperature 1.00°C (1.80°F) higher than the 20th-century average, making it the third-warmest July since records began in ...
  38. [38]
    Increasing trends in regional heatwaves | Nature Communications
    Jul 3, 2020 · Heatwaves have increased in intensity, frequency and duration, with these trends projected to worsen under enhanced global warming.
  39. [39]
    Heat waves: a hot topic in climate change research - PubMed Central
    It is likely that the frequency of heat waves has increased in large parts of Europe, Asia and Australia. It is very likely that human influence has contributed ...
  40. [40]
    Global prevalence of compound heatwaves from 1980 to 2022
    Our results demonstrate a pronounced increase in global CHW, with an annual cumulative intensity rising by 3.32 °C per decade(p<0.001).
  41. [41]
    A new understanding of global heatwave characteristics: trends on ...
    May 29, 2025 · This study provides an updated analysis of global heatwave characteristics using reanalysis data up to 2023.
  42. [42]
    August 2025 Temperature Update - Berkeley Earth
    Sep 17, 2025 · Furthermore, there is a 1.2% chance that global average temperature anomalies for 2025 exceed the 1.5 C (2.7 °F) benchmark above the 1850-1900 ...
  43. [43]
    Data in Action: The rate of global sea level rise doubled ... - PO.DAAC
    Feb 5, 2025 · The total sea level rise from 1993 to 2023 was 11.1 cm (bottom figure). Overall, the rate of global mean sea level is about 3.3 mm/year +/- 0.3 ...
  44. [44]
    El Niño causes spike in 2023 global sea level - EarthSky
    Mar 25, 2024 · Global average sea level rose by about 0.3 inches (0.76 centimeters) from 2022 to 2023, a relatively large jump due mostly to a warming climate ...
  45. [45]
    The Clear Evidence for Accelerating Sea Level Rise
    Sep 10, 2025 · A 2024 study in Communications Earth & Environment confirmed that global mean sea level rise has reached 4.5 mm per year due to warming oceans ...
  46. [46]
    New Evidence Confirms That Sea Level Rise Is Not Accelerating
    Oct 13, 2025 · The observations come from tide gauge records, which provide a more complete dataset than satellite altimetry that dates from only 1993. It's ...
  47. [47]
    Ice Sheets - Earth Indicator - NASA Science
    Sep 25, 2025 · Antarctica is losing ice mass at an average rate of about 135 billion tons per year, and Greenland is losing about 266 billion tons per year.
  48. [48]
    Mass balance of the Greenland and Antarctic ice sheets from 1992 ...
    Apr 20, 2023 · We present a new 29-year record of ice sheet mass balance from 1992 to 2020 from the Ice Sheet Mass Balance Inter-comparison Exercise (IMBIE).Peer review · Related articles · Assets · Metrics
  49. [49]
    Globally consistent estimates of high-resolution Antarctic ice mass ...
    Feb 20, 2024 · The Antarctic ice sheet mass balance is -144 ± 27 Gt a-1 from 2011-2020, with a GIA effect of 86 ± 21 Gt a-1.
  50. [50]
    Antarctic Ice Mass Loss 2002-2023 - NASA SVS
    These images, created from GRACE and GRACE-FO data, show changes in Antarctic ice mass since 2002. Orange and red shades indicate areas that lost ice mass ...
  51. [51]
    latest glacier mass balance data - World Glacier Monitoring Service
    Oct 8, 2024 · Mass balance values for the observation periods will be reported from more than 130 glaciers worldwide. The mass balance statistics (Table 1) ...
  52. [52]
    global glacier state - World Glacier Monitoring Service
    Feb 20, 2025 · The mass balance estimates considered here are based on a set of global reference glaciers with more than 30 continued observation years for the ...
  53. [53]
    Arctic Sea Ice Minimum Extent - Earth Indicator - NASA Science
    The measurements have shown that September Arctic sea ice is shrinking at a rate of 12.2% per decade, compared to its average extent during the period from 1981 ...
  54. [54]
    Arctic sea ice extent levels off; 2024 minimum set
    Sep 24, 2024 · On September 11, Arctic sea ice likely reached its annual minimum extent of 4.28 million square kilometers (1.65 million square miles).
  55. [55]
    Sea Ice Today | National Snow and Ice Data Center
    In the Arctic, the annual minimum sea ice extent occurs in September and the annual sea ice maximum extent occurs in March. In the Antarctic, the minimum ...Charctic Interactive Sea Ice... · Data Tools · Studying sea ice · Sea Ice Analysis Tool
  56. [56]
    Climate Change Indicators: Permafrost | US EPA
    Overall, permafrost temperatures have increased at an average rate of 0.6°F per decade. Changes in permafrost temperature shown in this indicator are consistent ...
  57. [57]
    Estimation of permafrost thawing rates in a sub-arctic catchment ...
    May 12, 2009 · Based on recession flow analysis, we estimate that permafrost in this catchment may be thawing at an average rate of about 0.9 cm/yr during the past 90 years.
  58. [58]
    Coastal Permafrost Erosion - NOAA Arctic
    Nov 9, 2020 · 2). Observations from all but 1 of the 14 coastal permafrost sites around the Arctic indicate that decadal-scale erosion rates are increasing.Arctic Report Card 2020 · Introduction · Historic And Contemporary...
  59. [59]
    The freezing‒thawing index and permafrost extent in pan-Arctic ...
    The permafrost area in pan-Arctic experienced a decline from 13.4 × 106 km2 in 2003–2013 to 12.51 × 106 km2 in 2014–2023, representing a decadal reduction of ...2. Materials And Methods · 3. Results · 4. Discussion
  60. [60]
    Permafrost Beneath Arctic Ocean Margins
    Nov 14, 2023 · The longer the period of inundation, the deeper the top of ice-bearing permafrost thaws, but the slower its thaw rate. Thaw depths observed ...
  61. [61]
    Climate Change Indicators: U.S. and Global Precipitation - EPA
    Globally, precipitation has increased by 0.03 inches per decade since 1901, while the US has increased by 0.18 inches per decade. Some US areas have seen ...
  62. [62]
    Global Precipitation Climatology Centre (GPCC)
    GPCC provides a monthly precipitation dataset from 1891-present, calculated from global station data, with a full data product from 1891-2019.
  63. [63]
    Chapter 11: Weather and Climate Extreme Events in a Changing ...
    This chapter assesses changes in weather and climate extremes on regional and global scales, including observed changes and their attribution, as well as ...
  64. [64]
    Global drought trends and future projections - Journals
    Oct 24, 2022 · This opinion piece includes a review of the recent scientific literature on the topic and analyses trends in meteorological droughts.Missing: peer | Show results with:peer
  65. [65]
    Comparisons of SPI and SPEI in capturing drought dynamics
    The SPEI shows stronger droughts trend than the SPI particularly in arid and semi-arid regions. · The SPI-SPEI differences is highest in hyper-arid/arid zones, ...
  66. [66]
    Global high-resolution drought indices for 1981–2022
    Dec 4, 2023 · We developed four high-resolution (5 km) gridded drought records based on the standardized precipitation evaporation index (SPEI) covering the period 1981–2022.
  67. [67]
    Evidence of anthropogenic impacts on global drought frequency ...
    May 12, 2021 · We show that the presence of anthropogenic forcing has increased the drought frequency, maximum drought duration, and maximum drought intensity experienced in ...
  68. [68]
    Drought frequency, intensity, and exposure have increased due to ...
    May 22, 2025 · Our results show that historical LULCC has increased the frequency, duration, severity, and intensity of drought over half of the global land, ...
  69. [69]
    Did Climate Change Do That? | Cato Institute
    He argues that, given the reality of anthropogenic climate change, “Should not the burden of proof be on showing that there is no human influence?” He critiques ...
  70. [70]
    Overstating the effects of anthropogenic climate change? A critical ...
    Feb 8, 2023 · For example, here the complaint could be that the storyline approach affirms that ACC caused extreme temperature happening in a particular ...
  71. [71]
    Monthly Climate Reports | Global Climate Report | Annual 2022
    The year 2022 was the sixth warmest year since global records began in 1880 at 0.86°C (1.55°F) above the 20th century average of 13.9°C (57.0°F).
  72. [72]
    Greening of the Earth and its drivers | Nature Climate Change
    Apr 25, 2016 · Factorial simulations with multiple global ecosystem models suggest that CO2 fertilization effects explain 70% of the observed greening trend, ...Missing: benefits satellite
  73. [73]
    Newly established forests dominated global carbon sequestration ...
    Jul 17, 2025 · As the world's primary carbon sink, terrestrial ecosystems have captured about 20% of carbon dioxide (CO2) from anthropogenic emissions during ...
  74. [74]
    Recent global decline of CO2 fertilization effects on vegetation ...
    We showed that global CFE has declined across most terrestrial regions of the globe from 1982 to 2015, correlating well with changing nutrient concentrations ...
  75. [75]
    Slowdown of the greening trend in natural vegetation with further ...
    Sep 13, 2021 · The probabilistic attribution method clearly identifies the CO2 fertilization effect as the dominant driver in only two biomes, the temperate ...
  76. [76]
    Earth's record-high greenness and its attributions in 2020
    Jan 1, 2025 · Earth recorded its highest level of vegetation greenness in 2020 since early 2000s. This greening largely linked to continuous growth in boreal and temperate ...
  77. [77]
    Biophysical impacts of earth greening can substantially mitigate ...
    Jan 9, 2023 · When considering those pixels with statistically significant LAI trends (Mann-Kendall test, P < 0.05), the cooling trend can increase to −0.029 ...
  78. [78]
    Climate Change: Ocean Heat Content
    Heat content in the global ocean has been consistently above-average (red bars) since the mid-1990s. More than 90 percent of the excess heat trapped in the ...Missing: peer- reviewed
  79. [79]
    Climate Change Indicators: Ocean Heat | US EPA
    Aug 12, 2025 · In four different data analyses, the long-term trend shows that the top 700 meters of the oceans have become warmer since 1955 (see Figure 1).Missing: peer- reviewed
  80. [80]
    Ocean Heat Content Rises | News
    Jan 22, 2020 · Based on data collected worldwide, ocean heat content continues to rise, outpacing its early rate of warming by a large percentage.Missing: 1955-2023 peer- reviewed
  81. [81]
    Climate Change Drives Poleward Increases and Equatorward ...
    Apr 20, 2020 · Global-scale analysis of marine species shows abundance changes linked to warming. Increases at poleward sides of species ranges reflect new ecological ...
  82. [82]
    Climate-driven global redistribution of an ocean giant predicts ...
    Oct 7, 2024 · As a result, marine species are moving poleward as much as six times faster, with global redistributions projected for over 12,000 species.
  83. [83]
    Climate-Driven Shifts in Marine Species Ranges - Annual Reviews
    Jan 3, 2020 · Climate change alters the structure of arctic marine food webs due to poleward shifts of boreal generalists. Proc. Biol. Sci. 282:20151546.
  84. [84]
    Ocean acidification | National Oceanic and Atmospheric Administration
    Sep 25, 2025 · During this time, the pH of surface ocean waters has fallen by 0.1 pH units. This might not sound like much, but the pH scale is logarithmic, so ...
  85. [85]
    Global Surface Ocean Acidification Indicators From 1750 to 2100
    Mar 23, 2023 · From 1750 to 2000, the average pH of the global surface ocean decreased by ∼0.11 units, equivalent to an acidity increase of ∼30% (Caldeira & ...
  86. [86]
    Progression of ocean interior acidification over the industrial era
    Nov 27, 2024 · Across the top 100 meters and from 1800 to 2014, the saturation state of aragonite (Ωarag) and pH = −log[H+] decreased by more than 0.6 and 0.1, ...
  87. [87]
    Current Global Bleaching: Status Update & Data Submission
    Sep 12, 2025 · From 1 January 2023 to 11 September 2025, bleaching-level heat stress has impacted 84.4% of the world's coral reef area and mass coral bleaching ...
  88. [88]
    What is coral bleaching? - NOAA's National Ocean Service
    Jun 16, 2024 · Warmer water temperatures can result in coral bleaching. When water is too warm, corals will expel the algae (zooxanthellae) living in their ...
  89. [89]
    A global analysis of coral bleaching over the past two decades
    Mar 20, 2019 · Thermal-stress events associated with climate change cause coral bleaching and mortality that threatens coral reefs globally.<|separator|>
  90. [90]
    Ocean Deoxygenation: A Primer - ScienceDirect.com
    Jan 24, 2020 · Earth's ocean is losing oxygen; since the mid-20 th century, 1%–2% of the global ocean oxygen inventory has been lost.Effects On Marine Ecosystems · Effects On Fisheries · Effect Of Hypoxia On...
  91. [91]
    Ocean Deoxygenation | CMEMS - Copernicus Marine Service
    Globally, the ocean has lost around 2% of dissolved oxygen since the 1950s and is expected to lose a further 1–7% by 2100. Although oxygen levels vary on a ...<|separator|>
  92. [92]
    Ocean warming has already changed marine fisheries - emLab, UCSB
    Apr 17, 2019 · Although warming has benefited some fisheries, it has hurt others. The losers outweighed the winners, resulting in a net 4% decline in sustainable catch ...
  93. [93]
    Climate change and the global redistribution of biodiversity
    Apr 11, 2023 · We found that less than half of all range-shift observations (46.60%) documented shifts towards higher latitudes, higher elevations, and greater marine depths.
  94. [94]
    Meta‐analyzing the likely cross‐species responses to climate change
    Aug 29, 2019 · Species' traits as predictors of range shifts under contemporary climate change: A review and meta-analysis. Global Change Biology, 23, 4094– ...
  95. [95]
    Where and why are species' range shifts hampered by unsuitable ...
    May 19, 2022 · There is widespread concern that species will fail to track climate change if habitat is too scarce or insufficiently connected.
  96. [96]
    Review Climate change effects on biodiversity, ecosystems ...
    Sep 1, 2020 · We find that species are responding to climate change through changes in morphology and behavior, phenology, and geographic range shifts, and ...Review · 2. Individuals, Populations... · 6. Implications For Natural...<|separator|>
  97. [97]
    Special Report on the Ocean and Cryosphere in a Changing ...
    Almost all coral reefs will degrade from their current state, even if global warming remains below 2ᴼC (very high confidence), and the remaining shallow coral ...
  98. [98]
    Overcoming the coupled climate and biodiversity crises and their ...
    Apr 21, 2023 · Climate change is causing species shifts and biodiversity loss globally because species are confined to limited thermal performance ranges ...
  99. [99]
    The direct drivers of recent global anthropogenic biodiversity loss
    Nov 9, 2022 · We show that land/sea use change has been the dominant direct driver of recent biodiversity loss worldwide.Missing: alterations | Show results with:alterations
  100. [100]
    Climate change reshuffles northern species within their niches
    Jun 2, 2022 · Indeed, climate change has prompted extensive shifts in species abundance and distributions, potentially leading to increased homogenization of ...
  101. [101]
    How well does climate change explain species' range shifts?
    Jul 20, 2023 · Our study highlights how complex species' range responses to climate change can be, and how they can be moderated by other factors such as land use change, ...
  102. [102]
    Global food security under climate change - PNAS
    This article reviews the potential impacts of climate change on food security. It is found that of the four main elements of food security, ie, availability, ...
  103. [103]
    The Influence of Climate Change on Global Crop Productivity - PMC
    Over the next few decades, CO2 trends will likely increase global yields by roughly 1.8% per decade. At the same time, warming trends are likely to reduce ...
  104. [104]
    Rising temperatures can negate CO2 fertilization effects on global ...
    Nov 15, 2023 · Many meta-analyses have reported increased seed yields in response to rising carbon dioxide concentrations ([CO2]) or decreased yields in ...
  105. [105]
    Climate change impacts on crop yields: A review of empirical ...
    Evidence was strong that increasing temperatures reduce crop yields. A 1 °C warming decreased the yields by 7.5±5.3% (maize), 6.0±3.3% (wheat), 6.8±5.9% ( ...
  106. [106]
    Climate Change Impacts on Agriculture and Food Supply | US EPA
    Aug 11, 2025 · Climate change can make conditions better or worse for growing crops in different regions. For example, changes in temperature, rainfall, and ...
  107. [107]
    Crop yields have increased dramatically in recent decades, but ...
    Sep 30, 2024 · Climate change has slowed the productivity of key crops such as maize and soybeans, but might have had small positive impacts on wheat.
  108. [108]
    Global vulnerability of crop yields to climate change - ScienceDirect
    The climate change could reduce global crop yields by 3–12% by mid-century and 11–25% by century's end, under a vigorous warming scenario.
  109. [109]
    The State of Food Security and Nutrition in the World
    SOFI 2025 presents the latest data and analysis on hunger, food security and nutrition worldwide, including updated estimates on the cost and affordability of ...
  110. [110]
    [PDF] CO₂ FERTILIZATION OF US FIELD CROPS Charl
    Second, our finding that CO2 fertilization has driven a large portion of the historical increase in crop yields has implications for how we think about the ...
  111. [111]
    How many people die from extreme temperatures, and how this ...
    Jul 1, 2024 · Globally, cold deaths are 9 times higher than heat-related ones. In no region is this ratio less than 3, and in many, it's over 10 times higher.
  112. [112]
    Do more people die from heat or cold? How will this change in the ...
    Oct 13, 2024 · A study across 854 cities in Europe found that cold-related deaths were around ten times higher than heat-related ones.
  113. [113]
    Estimating future heat-related and cold-related mortality under ...
    Jan 27, 2025 · Estimates generally report that there are roughly ten cold-related deaths for each heat-related death; some studies suggested that temperature- ...
  114. [114]
    Recent Trends in Heat-Related Mortality in the United States
    Our results show that in the past decade there is heterogeneity in the trends of heat-related human mortality. The decrease in heat vulnerability continues ...
  115. [115]
    Climate Change Indicators: Heat-Related Deaths | US EPA
    Between 1979 and 2022, the death rate as a direct result of exposure to heat (underlying cause of death) generally hovered between 0.5 and 2 deaths per million ...
  116. [116]
    Heat and health - World Health Organization (WHO)
    May 28, 2024 · Between 2000–2019 studies show approximately 489 000 heat-related deaths occur each year, with 45% of these in Asia and 36% in Europe (2). In ...Heat-health action plans · Biodiversity · El Niño Southern OscillationMissing: trends | Show results with:trends
  117. [117]
    Temperature-related mortality burden and projected change in 1368 ...
    Aug 21, 2024 · Temperature-related deaths are projected to increase by 41 850–96 072 additional deaths annually depending on the warming scenario, driven by ...
  118. [118]
    The future of the temperature–mortality relationship - The Lancet
    Aug 21, 2024 · The number of cold-related deaths is 8·3 times higher than heat-related deaths, but in a scenario in which global temperature increases by 3°C, ...
  119. [119]
    Climate Change Indicators: Cold-Related Deaths | US EPA
    In recent years, U.S. death rates in winter months have been 8 to 12 percent higher than in non-winter months. Much of this increase relates to seasonal changes ...
  120. [120]
    Impact of recent and future climate change on vector‐borne diseases
    There is a wealth of evidence that recent climate change has already affected pathogen–vector–host systems, in particular over temperate, peri‐Arctic and Artic ...
  121. [121]
    Vector-borne diseases - World Health Organization (WHO)
    Sep 26, 2024 · Vector-borne diseases account for more than 17% of all infectious diseases, causing more than 700 000 deaths annually.
  122. [122]
    Climate Change and Vectorborne Diseases
    Nov 23, 2022 · Warming temperatures and changes in precipitation associated with climate change are affecting the occurrence of vectorborne diseases.Missing: mortality | Show results with:mortality
  123. [123]
    Over half of known human pathogenic diseases can be aggravated ...
    Aug 8, 2022 · We found that 58% (that is, 218 out of 375) of infectious diseases confronted by humanity worldwide have been at some point aggravated by climatic hazards.
  124. [124]
    Anthropogenic amplification of precipitation variability over the past ...
    Jul 25, 2024 · The projected increase in precipitation variability results from a combination of increased atmospheric moisture and weakened circulation ...
  125. [125]
    Why are extreme precipitation events becoming more frequent in a ...
    Nov 20, 2024 · In addition, the daily precipitation variability increases by 1.2% per 10 years globally, especially in Europe, Australia, and eastern North ...
  126. [126]
    Rise in Northeast US extreme precipitation caused by Atlantic ...
    Extreme precipitation (EP) in the Northeastern United States increased abruptly after 1996, coinciding with warming Atlantic sea surface temperatures (SSTs).
  127. [127]
    How climate change worsens heatwaves, droughts, wildfires and ...
    Nov 14, 2024 · But climate change is shifting global rainfall patterns. While some of the world is getting wetter, other parts are becoming drier, which can ...
  128. [128]
    Climate Change Behind Sharp Drop in Snowpack Since 1980s
    Jan 10, 2024 · The sharpest global warming-related reductions in snowpack—between 10% to 20% per decade—are in the Southwestern and Northeastern United States, ...
  129. [129]
    Changing Climate Behind Sharp Drop in Snowpack Since 1980s
    Jan 24, 2024 · The sharpest global warming-related reductions in snowpack—between 10% to 20% per decade—are in the Southwestern and Northeastern United States, ...
  130. [130]
    Highlighting Changes to Snow and Water in the West - USGS.gov
    Apr 26, 2024 · One certainty is that the warming climate has already decreased snowpack and will continue to do so over the coming decades, with serious ...
  131. [131]
    [PDF] Water Resources - Canada in a Changing Climate
    Climate change impacts to the water cycle have already occurred, leading to damaged infrastructure, increased operating costs, disrupted operating seasons, and ...
  132. [132]
    Climate change's ripple effect on water supply systems and the ...
    Rising temperatures, precipitation fluctuations, and extreme weather events increase the risk of water shortages or network overloads [[68], [69]]. Despite ...
  133. [133]
    Vulnerability of urban water infrastructures to climate change at city ...
    Urban water infrastructures are facing substantial risks from floods and droughts aggravated by global climate change, while the adaptation is highly dependent ...
  134. [134]
    Climate Change Impacts on Freshwater Resources | US EPA
    Aug 12, 2025 · Rising sea level and increased drought can enable saline water to advance farther upstream and inland in estuaries, wetlands, and aquifers.Missing: empirical | Show results with:empirical
  135. [135]
    Escalating impacts of climate extremes on critical infrastructures in ...
    Projections of multiple climate risks to critical infrastructures are assessed. Impacts could rise up to 10 times present damages by 2100 due to global ...
  136. [136]
    Droughts and floods in a changing climate and implications for multi ...
    Rising atmospheric temperatures disrupt hydrological processes, leading to increased evaporation, which in turn can result in either floods or droughts ...
  137. [137]
    Quantifying Economic Damages from Climate Change
    Oct 31, 2018 · The paper seeks to provide both an accessible and comprehensive overview of how econ- omists think about parameterizing damage functions and ...
  138. [138]
    The economic commitment of climate change | Nature
    Apr 17, 2024 · We find that the world economy is committed to an income reduction of 19% within the next 26 years independent of future emission choices.
  139. [139]
    Climate Damage Functions for Estimating the Economic Impacts of ...
    This article shows how damage functions can be developed from the results of detailed modeling studies and then used to estimate future economic impacts.
  140. [140]
    Limitations of integrated assessment models of climate change
    Apr 2, 2009 · In addition, IAMs may exaggerate mitigation costs by failing to reflect the socially determined, path-dependent nature of technical change and ...
  141. [141]
    The failure of Integrated Assessment Models as a response to ...
    Dec 14, 2020 · They are unable to adequately represent human responses to changing climate and ecological processes. This also affects complementary scenario ...
  142. [142]
    [PDF] On the GDP Effects of Severe Physical Hazards
    We assess the impacts from physical hazards (or severe weather events) on economic activity in a panel of 98 countries using local projection methods. Prox-.
  143. [143]
    Comprehensive evidence implies a higher social cost of CO2 - Nature
    Sep 1, 2022 · Our preferred mean SC-CO 2 estimate is $185 per tonne of CO 2 ($44–$413 per tCO 2 : 5%–95% range, 2020 US dollars) at a near-term risk-free discount rate of 2%.
  144. [144]
    Trends and biases in the social cost of carbon - Tol - 2025
    May 7, 2025 · An updated and extended meta-analysis confirms that the central estimate of the social cost of carbon is around $200–250/tC ($700–900/tCO2) ...
  145. [145]
    Synthesis of evidence yields high social cost of carbon due to ...
    Dec 17, 2024 · This paper synthesizes results from the published literature over the last 20 y, revealing a wide range of SCC estimates.
  146. [146]
    Empirical evidence of declining global vulnerability to climate ... - NIH
    Results show a clear decreasing trend in both human and economic vulnerability, with global average mortality and economic loss rates that have dropped by 6.5 ...
  147. [147]
    The global costs of extreme weather that are attributable to climate ...
    Climate change-attributed economic costs from extreme weather events vary between 0.05% to 0.82% of global GDP annually over the study period.
  148. [148]
    [PDF] The macroeconomic implications of extreme weather events - OECD
    We find that the most severe events reduce regional GDP by 2.2%, with losses of around 1.7% persisting after five years.
  149. [149]
    [PDF] Climate Change, Adaptation, and Agricultural Output Patrick Regan ...
    A recent study estimated a 9-10% net crop yield reduction due to the excess temperature and drought at the global level (Lesk et al. 2016) and the ...
  150. [150]
    [PDF] Adaptation to Climate Change - MIT Economics
    In discussing the empirical literature on the adoption and efficacy of adaptation strategies in Section 3.2, we use E to indicate an intervention that ...
  151. [151]
    The Impact of Climate Change Adaptation Strategies on Income and ...
    This study aims to estimate the impact of adaptation strategies on fishers' income and their household's food security. Data were collected from small-scale ...<|separator|>
  152. [152]
    Cost and benefit analysis of adopting climate adaptation practices ...
    This study conducted an ex-ante in-depth empirical analysis of the costs and benefits of implementing five climate adaptation strategies identified among ...<|control11|><|separator|>
  153. [153]
    Psychological capital and climate change adaptation: Empirical ...
    Jun 28, 2021 · This study is the first attempt to assess how psychological capital affects climate change adaptation amongst smallholder farmers.
  154. [154]
    An Evidence-Based Public Health Approach to Climate Change ...
    The current approach to EBPH can, with modifications, support climate change adaptation activities, but there is little evidence regarding interventions and ...
  155. [155]
    Resilient by design: isolating impactful climate adaptation measures ...
    Aug 26, 2025 · Results show that policies focused on infrastructure enhancements and regulatory measures are most effective in reducing vulnerability. In ...
  156. [156]
    [PDF] Case Studies for Climate Change Adaptation - US EPA
    The Climate Resilience Evaluation and Awareness Tool (CREAT) helps understand expected utility facility vulnerabilities from a changing climate.
  157. [157]
    [PDF] 10 case studies Climate-ADAPT
    This unique collection of 10 European case studies showcases measures that are already being carried out in Europe to increase resilience to extreme weather.
  158. [158]
    Assessing the costs and benefits of climate change adaptation
    Mar 3, 2023 · Adaptation actions are cost-efficient when the benefit-cost ratio exceeds 1.5. Measures resulting in a lower ratio require careful consideration ...
  159. [159]
    RELEASE: WRI Study Finds Climate Adaptation Investments Yield ...
    Jun 3, 2025 · Every $1 invested in adaptation and resilience generates more than $10 in benefits over ten years. This translates to potential returns of over $1.4 trillion, ...
  160. [160]
    Challenges in the cost-benefit analysis of climate change adaptation ...
    Nov 20, 2024 · This perspective explores the complexities and challenges of cost-benefit analyses in designing effective climate change adaptation initiatives.
  161. [161]
    Climate adaptation may cost trillions - Wellington Management
    Jan 1, 2025 · A recent study projects that by 2050, spending on building resilience to damaging weather extremes will exceed transition spending sixfold.
  162. [162]
    An empirical perspective for understanding climate change impacts ...
    Jul 17, 2017 · Observational approaches provide the grounding in evidence that is needed to develop local to regional climate adaptation strategies. Our ...
  163. [163]
    Higher than expected CO2 fertilization inferred from leaf to global ...
    We estimate a historic global CO2 fertilization effect on photosynthesis of 30% (1900–2010; 296–389 ppm ca) that is significantly higher than current TBM ( ...
  164. [164]
    NASA Study: Rising Carbon Dioxide Levels Will Help and Hurt Crops
    May 3, 2016 · Elevated carbon dioxide concentrations in the atmosphere may increase water-use efficiency in crops and considerably mitigate yield losses ...
  165. [165]
    Crops and rising atmospheric CO2: friends or foes? - PMC
    May 29, 2025 · In regions where crop yield gains have stalled, studies suggest that recent increases in [CO2] have prevented further crop yield losses, ...
  166. [166]
    Changing temperatures are helping corn production in U.S. — for now
    Nov 5, 2018 · Research links warming temperatures and localized cooling to increased maize production.<|separator|>
  167. [167]
    Adapting growing seasons to climate change can boost yields of ...
    Jan 24, 2023 · The researchers found that adaptation could increase crop yields by 12 percent, compared to non-adaptation, with maize and rice showing the ...
  168. [168]
    Are there positive benefits from global warming? | NOAA Climate.gov
    Oct 29, 2020 · In the short term, farmers in some regions may benefit from the earlier onset of spring and from a longer warm season that is suitable for ...
  169. [169]
    Temperature-related mortality burden and projected change in 1368 ...
    Between 1991 and 2020, we estimated 363 809 cold-related deaths (empirical 95% CI 362 493–365 310) and 43 729 heat-related deaths (39 880–45 921), subject to ...
  170. [170]
    Human Deaths from Hot and Cold… - The Breakthrough Institute
    Dec 1, 2022 · It has been estimated that warming from 2000 to 2019 has resulted in a net decline in excess deaths globally (a larger decrease in cold deaths ...
  171. [171]
    Global, regional, and national burden of mortality associated with ...
    At a global level, the results indicate that global warming might slightly reduce net temperature-related deaths in the short term, although, in the long run, ...
  172. [172]
    Insights from past millennia into climatic impacts on human health ...
    However, food supplies, fertility, and population growth typically increased during longer-term stable warmer periods. Written records from up to 5 ...
  173. [173]
    What was the Medieval Warm Period (MWP)? - Greenly
    Jan 17, 2025 · The Medieval Warm Period (MWP) had an overall beneficial impact on the world, as crops were plentiful and population doubled in several cities ...
  174. [174]
    [PDF] Future Global Climate: Scenario-based Projections and Near-term ...
    in climate response to CDR causes hysteresis between key climate variables such as temperature, precipitation, AMOC and sea level, and atmosphere CO2 with ...
  175. [175]
    [PDF] CLIMATE CHANGE 2023
    The IPCC's Climate Change 2023 report summarizes climate change knowledge, impacts, risks, mitigation, and adaptation, integrating the Sixth Assessment Report.
  176. [176]
  177. [177]
    State-of-the-art global models underestimate impacts from climate ...
    Mar 1, 2019 · We find that a majority of models underestimate the extremeness of impacts in important sectors such as agriculture, terrestrial ecosystems, and heat-related ...Results · River Flow And Water... · Climate Forcing Data<|control11|><|separator|>
  178. [178]
    Biased Estimates of Equilibrium Climate Sensitivity and Transient ...
    Dec 9, 2021 · This study assesses the effective climate sensitivity (EffCS) and transient climate response (TCR) derived from global energy budget constraints within ...
  179. [179]
    High radiative forcing climate scenario relevance analyzed with a ...
    Sep 18, 2024 · Additionally, the IPCC AR6 report stated that “the likelihood of high-emissions scenarios such as RCP8.5 or SSP5-8.5 is considered low in light ...
  180. [180]
    Wrong Again: 50 Years of Failed Eco-pocalyptic Predictions
    Sep 18, 2019 · None of the apocalyptic predictions with due dates as of today have come true. What follows is a collection of notably wild predictions from notable people in ...
  181. [181]
    Flawed Climate Models - Hoover Institution
    Apr 5, 2017 · The ultimate test for a climate model is the accuracy of its predictions. But the models predicted that there would be much greater warming ...
  182. [182]
    Al Gore has history of climate predictions, statements proven false
    Jan 22, 2023 · Al Gore has history of climate predictions, statements proven false. Gore said there was a 75% chance the entire north polar ice cap would likely be gone by ...
  183. [183]
    How Climate Scenarios Lost Touch With Reality
    A failure of self-correction in science has compromised climate science's ability to provide plausible views of our collective future.
  184. [184]
    Chapter 4 | Climate Change 2021: The Physical Science Basis
    Climate models seem to underestimate the forced component of the year-to-year variability in the atmospheric circulation, particularly in the North Atlantic ...
  185. [185]
    PDO and AMO Modulation of the ENSO–Asian Summer Monsoon ...
    Dec 28, 2023 · Our analyses reveal that the PDO exerts a more pronounced impact on ASM variability than the AMO. By comparing different linear regression ...
  186. [186]
    Influence of the Atlantic Multidecadal Oscillation and the Pacific ...
    For the periods 1950-1976 and 1998-2012, global temperature increased little, with positive AMO and negative PDO indices; subsequently, the rate of temperature ...Missing: projections | Show results with:projections
  187. [187]
    Changes in global teleconnection patterns under global warming ...
    May 26, 2023 · Global warming increases variance in AMO and PDO, decreases in ENSO. SAI reverses these changes, but does not suppress decadal variability ...
  188. [188]
    and Naturally‐Caused Temperature Trends: A Systematic Approach ...
    May 1, 2023 · A data-driven approach is proposed to decompose the observed trends from global and regional scales into anthropogenically- and naturally-caused trends.
  189. [189]
    [PDF] Distinguishing the roles of natural and anthropogenically forced ...
    Capsule: In decadal forecasts, the magnitude of natural decadal variations may rival that of anthropogenically forced climate change on regional scales.
  190. [190]
    Role of Natural Climate Variability in the Detection of Anthropogenic ...
    The goal of the present study is to assess how natural climate variability affects the ability to detect the climate change signal for mean and extreme ...
  191. [191]
    Climate variability can outweigh the influence of climate mean ... - ACP
    Feb 5, 2025 · In this study, we show that changes in variability have a stronger influence on the number of extreme precipitation days than the change in the mean state in ...
  192. [192]
    New insights into natural variability and anthropogenic forcing of ...
    Jun 11, 2019 · As the GHGs increase further, our study shows a declined influence of the PDV and AMV to global and regional climate relative to that of GHGs.
  193. [193]
    False Alarm: How Climate Change Panic Costs Us Trillions, Hurts ...
    Children panic about their future, and adults wonder if it is even ethical to bring new life into the world. Enough, argues bestselling author Bjorn Lomborg.
  194. [194]
    Bjorn Lomborg: 'Climate alarm' is as big a threat as climate change
    Sep 10, 2020 · Bjorn Lomborg says climate alarm causes nothing but anxiety and bad policies, arguing we can do better with smarter solutions to the problem.Missing: critique | Show results with:critique
  195. [195]
    More misinformation and nonsense on climate from Lomborg and Tol
    Mar 22, 2024 · Dr Lomborg next makes the following claim: “The data show that climate-related deaths from droughts, storms, floods and fires have declined by ...Missing: alarmism | Show results with:alarmism
  196. [196]
    [PDF] State of the climate debate
    State of the climate debate. Judith Curry. Climate Forecast Applications Network. Page 2. Agreement: • Surface temperatures have increased since 1880. • Humans ...
  197. [197]
    Evaluating the Performance of Past Climate Model Projections
    Dec 4, 2019 · Retrospectively comparing future model projections to observations provides a robust and independent test of model skill.
  198. [198]
    Current Wisdom: Observations Now Inconsistent with Climate Model ...
    Nov 5, 2013 · Everyone has read that over the past 10–15 years, most climate models' forecasts of the rate of global warming have been wrong. Most predicted a ...
  199. [199]
    [PDF] Climate Debate
    Climate Debate. Judith Curry. Page 2. President's Climate Action Plan. Hillary Clinton: climate change is “the most consequential, urgent, sweeping collection ...
  200. [200]
    Climate Change: The Science Doesn't Support the Heated Rhetoric
    Oct 16, 2024 · Other than modest warming, there has been little change in any kind of severe weather that can be attributed to global greenhouse gas emissions.
  201. [201]
    Alarmism and Accountability in Climate Communication During ...
    Jun 7, 2023 · Pielke (2014) claimed that discussions linking catastrophic loss to climate change were an effect of alarmist content (“climate porn”) ...
  202. [202]
    Beware of exaggerated claims of climate harm | Fraser Institute
    May 24, 2023 · There is, in fact, good evidence showing the effect of climate change on economic growth will likely be insignificant and may even be positive.
  203. [203]
  204. [204]
    Climate attribution method overstates “fingerprints” of external forcing
    Dec 18, 2023 · This is the third in my series of papers on flaws in standard fingerprinting methods: blog posts on the first two are here and here.
  205. [205]
    Behind the Curtain - by Roger Pielke Jr. - The Honest Broker
    Apr 5, 2025 · Rather, my critique is that extreme event attribution is pseudoscience, seemingly created to undermine the scientifically robust detection and ...
  206. [206]
    Revisiting the Holocene global temperature conundrum - Nature
    Feb 15, 2023 · Another recent reconstruction focused on a globally distributed pollen-only dataset shows maximum temperatures extending from around 8 to 5 ka ...
  207. [207]
  208. [208]
    Ice Volume and Sea Level During the Last Interglacial - Science
    Jul 13, 2012 · During the last interglacial period, ~125000 years ago, sea level was at least several meters higher than at present, with substantial ...
  209. [209]
    Sea level variations during the last interglacial - Past Global Changes
    The last interglacial sea level was 6 to 9 m above present, with a possible fluctuation of 6.5-10.5m, and a range of 5.5 to 9m.
  210. [210]
    Fact Sheet fs002-00: Sea Level and Climate
    Sea levels during several previous interglacials were about 3 to as much as 20 meters higher than current sea level. The evidence comes from two different ...
  211. [211]
    Roman Warm Period and Late Antique Little Ice Age in an Earth ...
    Aug 9, 2022 · Climate changes during the Roman Warm Period (RWP) and Late Antique Little Ice Age (LALIA) had large impacts on human migration, agricultural ...
  212. [212]
    Was the Medieval Era Warmer Than Now? New Tree Ring Study ...
    Aug 4, 2023 · Past studies of tree rings had indicated that the Medieval period was as warm or warmer than today, but climate models found otherwise. Climate ...
  213. [213]
    How the Little Ice Age Changed History | The New Yorker
    Mar 25, 2019 · During this epoch, often known as the Little Ice Age, temperatures dropped by as much as two degrees Celsius, or 3.6 degrees Fahrenheit. ...
  214. [214]
    A 485-million-year history of Earth's surface temperature | Science
    Sep 20, 2024 · PhanDA indicates that Earth's temperature has varied between 11° and 36°C over the past 485 million years. This range is larger than previous reconstructions.
  215. [215]
    Challenges and perspectives for large‐scale temperature ...
    Dec 13, 2016 · Large-scale temperature reconstructions of the last one to two millennia and the underlying methods are reviewed It is studied how different ...