Fact-checked by Grok 2 weeks ago

Climate change

Climate change denotes long-term shifts in the average patterns of temperature, precipitation, and other climatic variables on regional and global scales, with the contemporary focus on the observed rise in Earth's global surface air temperature by approximately 1.2°C since the pre-industrial baseline of 1850–1900, accelerating in recent decades to reach 1.43°C above that level in 2023. This warming manifests in phenomena such as reduced Arctic sea ice extent (with natural variability contributing 30-50% to the observed decline), altered precipitation regimes, and more frequent heatwaves and heavy precipitation events, though attribution of specific events remains challenging due to natural variability. The primary drivers of post-1950 warming are widely attributed to anthropogenic emissions of greenhouse gases like carbon dioxide and methane from fossil fuel combustion, deforestation, and industrial processes, which have elevated atmospheric CO₂ concentrations from ~280 ppm pre-industrially to over 420 ppm today, enhancing the greenhouse effect. Natural factors, including solar irradiance variations, volcanic aerosols, and internal climate oscillations like El Niño-Southern Oscillation, explain much of the early 20th-century warming and interdecadal fluctuations but insufficiently account for the late-20th-century acceleration without human influence. Scientific literature reflects a strong consensus, with over 99% of peer-reviewed studies endorsing a substantial human role in recent warming, though methodological critiques of consensus surveys, including keyword-based searches for skeptical papers in studies like Lynas et al., highlight potential overestimation by including implicit endorsements, conflating natural forcings discussions with denial, and underrepresenting dissenting analyses of model discrepancies or natural forcings. Controversies center on the degree of future warming projected by general circulation models—which have historically overestimated tropospheric warming rates compared to satellite observations—and the net impacts, including CO₂ fertilization effects boosting global greening alongside risks like sea-level rise of ~20 cm since 1900. Policy responses, from the Paris Agreement's emissions targets to adaptation measures, underscore ongoing debates over causal attribution, economic costs, and the balance between mitigation and natural adaptation capacities.

Definitions and Basic Concepts

Core Definitions

Climate refers to the average and variability of weather conditions, such as temperature, precipitation, humidity, wind, and atmospheric pressure, over an extended period, typically at least 30 years, for a specific region or the globe. In contrast, weather describes short-term atmospheric conditions at a particular time and location, including immediate phenomena like rain, storms, or temperature fluctuations, which can vary day-to-day or seasonally. Climate change denotes a statistically significant alteration in the mean state or variability of the climate system, detectable through changes in long-term averages of temperature, precipitation, or other elements, persisting over decades or longer. This encompasses shifts driven by both natural factors, such as volcanic eruptions or solar variability, and human activities, particularly the emission of greenhouse gases from fossil fuel combustion, deforestation, and industrial processes since the Industrial Revolution. Global warming specifically describes the observed rise in Earth's average surface temperature, approximately 1.1°C since the late 19th century, largely attributed to increased concentrations of greenhouse gases like carbon dioxide (now at about 420 parts per million, up from pre-industrial levels of 280 ppm) and methane. Greenhouse gases are atmospheric constituents that absorb and re-emit infrared radiation, trapping heat near the surface in a process known as the greenhouse effect, which is essential for maintaining habitable temperatures but intensified by anthropogenic emissions. Principal greenhouse gases include water vapor, carbon dioxide, methane, nitrous oxide, and fluorinated gases, with their radiative forcing contributing to net warming when concentrations rise.

Greenhouse Effect and Radiative Forcing

The greenhouse effect is the process by which certain atmospheric gases absorb outgoing longwave infrared radiation emitted from Earth's surface and re-emit it in all directions, including downward, thereby reducing the net loss of heat to space and warming the lower atmosphere and surface. This natural phenomenon arises from the differential absorption properties of gases like water vapor, carbon dioxide (CO₂), methane (CH₄), and nitrous oxide (N₂O), which are opaque to infrared wavelengths but transparent to most incoming shortwave solar radiation. Without the greenhouse effect, Earth's global mean surface temperature would be approximately -18°C (0°F), based on the planet's effective radiating temperature derived from absorbed solar flux and planetary albedo, rather than the observed 15°C (59°F). Water vapor dominates the natural effect due to its high concentration and strong absorption bands, while CO₂ contributes roughly 80% of the total radiative forcing in equilibrium calculations excluding water vapor feedbacks. Anthropogenic emissions have elevated concentrations of long-lived greenhouse gases—CO₂ from 280 ppm pre-industrially to over 420 ppm by 2023, CH₄ from 722 ppb to 1,900 ppb, and N₂O from 270 ppb to 336 ppb—enhancing the greenhouse effect beyond natural variability. This increase amplifies trapping of infrared radiation, with water vapor acting as a feedback that further intensifies warming as temperatures rise, though its concentration is thermodynamically limited by saturation processes. Observational spectra from satellites confirm the fingerprint of anthropogenic CO₂ in reduced outgoing longwave radiation at specific wavelengths (e.g., 15 μm band), distinct from natural solar or volcanic influences. Radiative forcing quantifies perturbations to Earth's top-of-atmosphere energy balance, defined as the change in net downward irradiance (shortwave plus longwave) at the tropopause following a climate driver change, after allowing for stratospheric temperature adjustment but before surface or tropospheric responses. Top-of-atmosphere radiation from NASA’s CERES satellites and in-situ ocean measurements from the global Argo array show that Earth’s energy imbalance (EEI) approximately doubled from mid-2005 to mid-2019, implying a faster net energy gain by the climate system; the rise reflects increased absorbed solar radiation (reduced cloud/sea-ice reflectivity) alongside strengthened greenhouse trapping. The Earth Heat Inventory synthesizing observations indicates the system accumulated 381 ± 61 ZJ of heat (1971–2020)—an average 0.48 ± 0.10 W m⁻²—with about 89–90% stored in the ocean, making ocean heat content (OHC) the most robust constraint on EEI. Because a positive and increasing EEI guarantees further warming until the energy budget is rebalanced, the recent EEI increase is consistent with the observed post-2010 acceleration in warming rates and supports expectations of continued near-term surface warming as aerosol cooling diminishes and ocean heat uptake rises. Effective radiative forcing (ERF) extends this to include rapid tropospheric and surface adjustments, such as cloud or lapse rate changes, providing a better predictor of eventual temperature response via the relationship ΔT ≈ λ × ERF, where λ is climate sensitivity. Key components include positive forcing from well-mixed greenhouse gases (approximately 3.2 W/m² anthropogenic since pre-industrial era as of recent assessments) and negative forcing from aerosols (around -1.0 to -1.5 W/m²), with net anthropogenic ERF estimated at about 2.7 W/m² by 2020, reflecting a 51% rise in greenhouse gas forcing alone since 1990 per the Annual Greenhouse Gas Index. Uncertainties persist in aerosol-cloud interactions and black carbon effects, with peer-reviewed analyses bounding total aerosol ERF at -2.0 to -0.4 W/m² for climate change attribution. Natural forcings, such as solar irradiance variations (±0.1 W/m² over decades) and volcanic eruptions (short-term negative spikes up to -3 W/m²), are smaller and better constrained observationally than anthropogenic drivers.

Distinction from Weather and Natural Variability

Weather refers to the state of the atmosphere at a specific location and time, encompassing short-term conditions such as temperature, precipitation, wind, and humidity over hours to weeks. Climate, by contrast, describes the long-term average of these weather elements, including their variability and extremes, typically assessed over periods of 30 years or more to capture representative patterns. This temporal distinction ensures that transient events, like a single heatwave or cold snap, do not equate to a climatic shift, as they fall within the bounds of daily or seasonal fluctuations rather than sustained alterations in statistical norms. Natural climate variability arises from internal oscillations within the Earth's climate system, such as the El Niño-Southern Oscillation (ENSO), which operates on interannual timescales of 2–7 years and influences global temperature and precipitation distributions through ocean-atmosphere interactions, and the North Atlantic Oscillation (NAO), affecting regional weather patterns over months to decades. External drivers of variability include solar irradiance variations, which fluctuate on 11-year cycles with amplitudes of about 0.1% but have contributed minimally to 20th-century warming (less than 0.1°C), and volcanic eruptions, which inject aerosols into the stratosphere causing temporary global cooling of 0.1–0.5°C for 1–3 years, as observed after the 1991 Mount Pinatubo eruption. These mechanisms produce oscillations superimposed on longer-term trends, but paleoclimate reconstructions from proxies like ice cores and tree rings indicate that pre-industrial variability, such as the Medieval Warm Period (circa 950–1250 CE) or Little Ice Age (circa 1450–1850 CE), occurred within a range of ±0.5°C globally without exceeding recent rates of change. Distinguishing anthropogenic climate change from natural variability relies on detection-attribution methods, which compare observed trends against model simulations isolating forced responses (e.g., greenhouse gases) from unforced internal variability. For instance, post-1950 global temperature increases of approximately 0.8°C exceed the amplitude of multidecadal natural oscillations like the Atlantic Multidecadal Oscillation (AMO), which varies by 0.3–0.5°C over 60–80 years, as evidenced by fingerprinting techniques matching spatial patterns of warming (e.g., tropospheric amplification and stratospheric cooling) unique to radiative forcing rather than solar or oceanic cycles. However, on regional scales and shorter decadal periods, natural variability can mask or amplify signals, complicating attribution; studies show that internal atmospheric dynamics account for up to 50% of year-to-year temperature variance, underscoring the need for multi-decadal datasets to achieve statistical confidence. While consensus attributes dominant recent warming to human forcings, debates persist over model fidelity in simulating variability, with some analyses indicating that unforced simulations alone fail to replicate observed low-frequency trends without adjustments.

Historical Context

Paleoclimate Records and Long-Term Variations

Paleoclimate records, derived from proxies such as ice cores, tree rings, sediment layers, and coral reefs, provide evidence of Earth's climate variations over millennia to millions of years. These archives reveal recurrent glacial-interglacial cycles during the Pleistocene epoch, with global temperatures fluctuating by approximately 4–7°C between cold glacial maxima and warmer interglacials. Ice cores from Antarctica, including Vostok and EPICA Dome C, document these cycles over the past 800,000 years, showing temperature reconstructions based on deuterium and oxygen isotopes that correlate with orbital forcings. Milankovitch cycles—variations in Earth's orbital eccentricity (period ~100,000 years), axial tilt (41,000 years), and precession (23,000 years)—drive these long-term changes by altering seasonal insolation, particularly at high northern latitudes, initiating ice sheet growth or retreat. Evidence from marine sediment cores and speleothems confirms that these orbital parameters paced the timing of glacial advances and retreats, with ice volume proxies like benthic foraminifera δ¹⁸O isotopes aligning closely with insolation minima. Solar variability, including grand solar minima like the Maunder Minimum (1645–1715 CE), and volcanic eruptions contributing stratospheric aerosols have superimposed shorter-term cooling episodes on these orbital trends, as seen in tree-ring width and density records spanning centuries. During the Holocene epoch (last ~11,700 years), proxy data indicate a generally warmer early period, the Holocene Climatic Optimum (~9,000–5,000 years before present), followed by gradual cooling toward the neoglaciation, with regional multi-centennial oscillations. Multi-proxy reconstructions from 1,319 records, including pollen, chironomids, and Mg/Ca ratios, show Northern Hemisphere summer temperatures peaking 1–2°C above late 20th-century levels in some mid-latitude sites, influenced by orbital-induced insolation decline and amplified by feedback from vegetation and ocean circulation. Atmospheric CO₂ levels, measured from Antarctic ice cores, remained stable between 260–280 ppm throughout most of the Holocene, rising only modestly to ~284 ppm pre-industrially, contrasting with tighter correlations to temperature in deeper glacial records where CO₂ lagged orbital-driven warming by centuries. These records underscore that natural forcings, rather than anthropogenic factors, dominated long-term variations prior to the industrial era, with no precedent in the instrumental record for the sustained Holocene cooling trend until recent reversals. Tree-ring δ¹⁸O from the European Alps reveals a significant drying and cooling trend over the Holocene, aligned with decreasing summer insolation from orbital precession. Volcanic impacts, such as the ~75,000-year-old Toba eruption, demonstrate episodic global cooling of ~3–5°C lasting years to decades, but such events are infrequent compared to orbital pacing.

Pre-20th Century Climate Shifts

Paleoclimate reconstructions derived from proxies including tree rings, ice cores, lake sediments, and historical documents reveal multiple episodes of natural climate variability prior to 1900, driven primarily by solar irradiance fluctuations, volcanic eruptions, and orbital forcings rather than anthropogenic greenhouse gases. These shifts occurred on timescales of centuries to millennia, with regional temperature anomalies often exceeding 1°C, demonstrating the Earth's climate system's sensitivity to natural forcings. The Roman Warm Period, approximately 250 BCE to 400 CE, featured elevated temperatures in the Mediterranean basin, where alkenone-based sea surface temperature proxies indicate values around 2°C warmer than the subsequent Byzantine average, facilitating agricultural expansion and maritime activity. This warmth appears regionally pronounced rather than globally uniform, with pollen and speleothem records from Europe corroborating milder conditions conducive to viticulture in Britain. A transitional cool phase, sometimes termed the Late Antique Little Ice Age from circa 536 to 660 CE, followed, linked to massive volcanic eruptions that injected sulfate aerosols into the stratosphere, causing summer temperature drops of up to 2.5°C in the Northern Hemisphere as recorded in tree-ring oxygen isotopes. The Medieval Warm Period, spanning roughly 950 to 1250 CE, succeeded this, with multi-proxy hemispheric reconstructions showing North Atlantic and European temperatures 0.2–0.5°C above the subsequent Little Ice Age baseline, enabling Norse colonization of Greenland and reduced sea ice extent. While not synchronously global, borehole and documentary evidence from China and South America indicate concurrent warm anomalies in select regions. The Little Ice Age, from approximately 1300 to 1850 CE, marked a pronounced cooling of about 0.6°C below the 19th-century mean in global multi-proxy averages, manifested in alpine glacier advances, frozen harbors like the Thames in London (last in 1814), and crop failures across Europe. Contributing factors included the Spörer and Maunder solar minima, which reduced total solar irradiance by up to 0.25%, compounded by frequent explosive volcanism that enhanced radiative forcing deficits. These events underscore pre-industrial climate dynamism, with recovery toward 1900 aligning with solar rebound and diminished volcanism.

20th Century Observations

Instrumental records of near-surface air temperatures over land and sea, compiled from thousands of weather stations and ship measurements, indicate an overall global warming of approximately 0.6°C from 1900 to 2000, with regional variations including stronger warming over land than oceans. This trend featured early 20th-century warming peaking around 1940, a mid-century interval of stability or slight decline amid incomplete global coverage and potential influences like aerosol emissions, and renewed warming post-1975 coinciding with improved observational networks. Analyses such as those from NOAA's GlobalTemp dataset confirm this pattern, though adjustments for station relocations, time-of-observation biases, and urban heat effects have been applied, with some critiques noting that unadjusted records show less pronounced mid-century cooling. Recent reevaluations, including corrections for early-20th-century sea surface temperature under-sampling, suggest the instrumental record may underestimate warming in that period due to cold biases in bucket measurements. Tide gauge measurements from over 900 stations worldwide record an average global mean sea level rise of 1.6 to 1.8 mm per year over the 20th century, totaling roughly 16-18 cm from 1900 to 2000, consistent with thermal expansion and land ice melt contributions. These relative sea level changes varied regionally, with accelerations emerging after the 1920s in some reconstructions but remaining within historical variability rates of 1-2 mm/year until mid-century, after which rates increased to about 2 mm/year by the 1990s. Uncertainties arise from vertical land motion, isostatic rebound, and sparse coverage in the Southern Hemisphere, but ensemble analyses of tide gauge data affirm no significant deviation from linear trends until recent decades. Glacier observations, tracked by networks like the World Glacier Monitoring Service since the early 1900s, document widespread retreat and negative mass balances across mountain ranges, with cumulative ice loss equivalent to several millimeters of sea level equivalent rise. For instance, Alpine glaciers lost volume at rates accelerating from the 1920s onward, while small glaciers in temperate regions provided early evidence of atmospheric warming through terminus recession and thinning, though some advances occurred during cooler mid-century episodes. Arctic sea ice reconstructions from ship logs and proxies indicate variable extent, with summer minima potentially lower in the early 1930s than mid-century but declining sharply post-1970 amid limited pre-satellite data. Global precipitation patterns showed modest increases over land areas, with a net rise of about 9 mm (roughly 2%) from 1900 to 2000, concentrated in mid-to-high latitudes and influenced by enhanced moisture convergence in a warming atmosphere. Oceanic trends were smaller at 0.13 mm per day per century, while regional extremes varied, including wetter conditions in the Northern Hemisphere extratropics but drier subtropics in parts of the Southern Hemisphere. These changes align with thermodynamic expectations of higher evaporation rates but remain within natural decadal oscillations, with data sparsity complicating attribution before widespread rain gauge networks expanded post-1950.

Observed Changes

Instrumental measurements of global surface air temperature began in the mid-19th century, with reliable records from land stations, sea surface temperatures from ships and buoys, and later Arctic infilling via statistical methods. Major datasets, including NASA's GISTEMP, the UK Met Office's HadCRUT5, NOAA's GlobalTemp, and Berkeley Earth's surface temperature series, reconstruct global mean anomalies relative to baselines like 1850-1900 or 1961-1990. These independent analyses converge on a net warming of approximately 1.1°C from 1880 to 2023, with acceleration since the mid-20th century. Linear trends across datasets indicate an overall global surface warming rate of about 0.08°C per decade from 1880 to 2024, rising to 0.18°C per decade since 1975. The year 2024 marked the warmest on record in all major series, with anomalies of 1.28°C above the 1951-1980 baseline in GISTEMP and 1.62°C above 1850-1900 in Berkeley Earth, surpassing 2023's prior record by 0.10-0.15°C. This recent spike correlates with a strong El Niño event dissipating in mid-2024, though underlying multidecadal trends persist amid natural oscillations like the Atlantic Multidecadal Oscillation. Land areas have warmed faster than oceans, at roughly 1.5 times the global average since 1970, amplifying continental trends. Satellite-based microwave sounding unit (MSU) records since December 1978 provide lower tropospheric temperature trends, less prone to surface biases like urban heat islands (UHI). The University of Alabama in Huntsville (UAH) dataset reports a +0.15°C per decade trend through November 2024, lower than surface estimates, while Remote Sensing Systems (RSS) shows ~0.21°C per decade; discrepancies arise from orbital decay corrections and tropical tropospheric amplification debates. Periods of slower warming, such as 1998-2012 (+0.05°C per decade in HadCRUT), highlight internal variability's role, with recoveries tied to reduced volcanic and solar forcing minima. Data adjustments for station relocations, time-of-observation biases, and UHI—where urbanization adds local heating of 0.1-1°C in cities—aim to isolate climatic signals, but residual effects may inflate land trends by 10-30% in some analyses. Rural-only subsets from Berkeley Earth and others yield trends ~0.9-1.0°C since 1950, close to adjusted global figures, though critics argue UHI contamination persists due to population growth near stations, potentially overstating recent decadal rates by 0.05°C or more. Coverage gaps in the early record (pre-1900) and Southern Hemisphere introduce uncertainties of ±0.05-0.1°C in century-scale trends, with homogeneity tests confirming robustness but not eliminating debates over methodological choices.

Sea Levels and Ice Cover

Global mean sea level has risen by approximately 21-24 centimeters since 1880, with tide gauge records indicating an average rate of about 1.7 millimeters per year from 1900 to 1990, accelerating to around 3.7 millimeters per year from satellite altimetry data starting in 1993. Recent observations show variability, including a 0.76-centimeter increase from 2022 to 2023 attributed primarily to El Niño-driven thermal expansion and land water storage changes. Reconstructions from 945 tide gauges from 1900 to 2022 confirm ongoing rise but highlight regional differences due to vertical land motion and local subsidence. Arctic sea ice extent has declined markedly since satellite records began in 1979, with summer minima decreasing at an average rate of about 13% per decade and winter maxima showing a slower decline of around 3% per decade. The 2025 winter maximum reached a record low of the 47-year record on March 22, while the September 2025 minimum was 4.60 million square kilometers, ranking among the ten lowest. In contrast, Antarctic sea ice extent showed slight increases through much of the satellite era until abrupt declines, with record lows from 2022 to 2025 driven by subsurface ocean warming and reduced stratification; the 2024 winter maximum was the second lowest on record. Land-based ice contributes significantly to sea level rise through mass loss. Greenland's ice sheet has lost mass at an accelerating rate, averaging 266 billion tons per year from GRACE satellite gravimetry data through recent years, primarily from surface melt and iceberg calving. Antarctica's ice sheet shows net loss of about 135 billion tons per year over the same period, though with regional gains in East Antarctica offsetting faster losses in West Antarctica and the peninsula. Global glaciers, excluding ice sheets, have retreated since the 1970s, with mass loss accelerating; from 2000 to present, they have shed about 5% of their ice volume regionally variable, contributing roughly 18% more to sea level than previously estimated, per community assessments. The World Glacier Monitoring Service reports the 2022-2024 period as the largest three-year mass loss on record for monitored glaciers.

Precipitation and Extreme Events

Global precipitation has shown a modest upward trend since the early 20th century, with land areas experiencing an average increase of approximately 1-2% per decade from 1901 to 2020, though this varies regionally and is influenced by natural variability such as the El Niño-Southern Oscillation. Observations indicate wetter conditions in high latitudes and parts of the tropics, contrasted by drying in subtropical regions like the Mediterranean and southern Africa, but these shifts align partly with multidecadal oscillations rather than a uniform anthropogenic signal. Day-to-day precipitation variability has increased over most land regions since 1900, potentially amplifying flood risks in some areas, though long-term totals remain stable or slightly elevated without exceeding historical precedents adjusted for data quality. Heavy precipitation events, defined as the upper percentiles of daily or sub-daily rainfall, have intensified in frequency and magnitude over many mid-latitude and tropical land regions since the mid-20th century, with detection and attribution studies linking this to human-induced warming through enhanced atmospheric moisture capacity under the Clausius-Clapeyron relation, which predicts about 7% more water vapor per 1°C temperature rise. However, these changes are not globally coherent, and event attribution remains probabilistic, with natural variability confounding signals; for instance, no significant increase in extreme rainfall has been observed in all regions, and claims of dramatic escalation often overlook sparse historical records pre-1950. Drought trends exhibit regional divergence rather than a global intensification: meteorological droughts (precipitation deficits) show no widespread increase over land areas from 1900 to 2020, while agricultural and hydrological droughts vary due to land use, irrigation, and evaporation changes rather than precipitation alone. In the United States, for example, drought frequency has not trended upward when normalized for natural cycles, challenging narratives of climate-driven aridification. Flood occurrences lack a clear global trend attributable to climate change, as riverine floods are influenced more by upstream precipitation, land management, and urbanization than by overall precipitation volume; pluvial (flash) floods may rise with intense downpours, but economic damages reflect exposure growth rather than event frequency. Normalized flood losses have not increased disproportionately to GDP or population since 1900, per analyses critiquing unadjusted disaster cost tallies. Tropical cyclone frequency and intensity show no robust increase globally or in landfall rates through 2020, with accumulated cyclone energy metrics stable or declining in some basins despite warmer seas; attribution to greenhouse gases is limited to potential modest intensification of the strongest storms, but observed data do not support claims of more frequent major hurricanes. U.S. hurricane landfalls, for instance, exhibit no upward trend since reliable records began in the late 19th century, underscoring the dominance of natural decadal variability over anthropogenic forcing in cyclone metrics. Overall, while thermodynamic principles suggest potential for altered extremes under warming, empirical records reveal mixed signals dominated by regionality and variability, with many alarmist attributions criticized for conflating correlation with causation and ignoring socioeconomic drivers of impacts. Peer-reviewed critiques emphasize that disaster cost escalations, such as U.S. billion-dollar events rising to 28 in 2023, stem primarily from expanded asset values and vulnerability rather than climatological shifts.

Causes and Attribution

Natural Drivers

Natural drivers of climate encompass solar variability, volcanic activity, orbital parameters, and internal oscillations within the Earth system, which have historically influenced global temperatures over diverse timescales. These factors operate independently of human emissions and can produce both warming and cooling effects, though their net contribution to the observed warming since the late 19th century has been assessed as minimal in magnitude compared to the sustained trend. Attribution analyses, drawing on radiative forcing estimates and paleoclimate proxies, indicate that natural forcings alone cannot account for the rapid post-1950 temperature rise, as solar output has stagnated or declined while volcanic influences have been episodic and short-lived. Changes in Earth's orbital geometry, known as Milankovitch cycles, modulate the seasonal and latitudinal distribution of solar insolation through variations in eccentricity (cycle ~100,000 years), obliquity (tilt, ~41,000 years), and precession (~23,000 years). These cycles drive glacial-interglacial transitions by altering summer insolation in the Northern Hemisphere, with peak effects on the order of 100 W/m² regionally but averaging near zero globally. Over the Holocene and recent centuries, orbital forcing has trended toward gradual cooling, exerting a negative influence of approximately -0.1 W/m² since 1850, insufficient to explain contemporary warming and operating too slowly for decadal-scale detection. Solar total irradiance (TSI) fluctuates by ~1 W/m² (0.1%) over the 11-year Schwabe cycle, driven by sunspot activity, with longer-term modulations linked to grand solar minima like the Maunder Minimum (1645–1715), which coincided with the Little Ice Age cooling of ~0.5–1°C in parts of Europe. Reconstructing TSI via proxies such as cosmogenic isotopes shows a rise of ~0.3–0.4 W/m² from the 17th to mid-20th century, potentially contributing 0.02–0.1°C to early 20th-century warming, but no net increase since the 1950s despite accelerating global temperatures; the implied forcing per W/m² is ~0.06–0.07°C globally after accounting for climate feedbacks. Recent satellite measurements confirm TSI stability or slight decline post-1980, underscoring limited explanatory power for late-20th-century trends. Volcanic eruptions release sulfur dioxide that forms stratospheric sulfate aerosols, reflecting ~1–2% of incoming solar radiation and inducing temporary global cooling of 0.1–0.5°C lasting 1–3 years; the 1815 Tambora eruption, for instance, triggered the "Year Without a Summer" with Northern Hemisphere temperature drops of up to 3°C regionally. Major 20th-century events like El Chichón (1982) and Pinatubo (1991) contributed transient forcings of -2 to -3 W/m², offsetting warming briefly but yielding a net volcanic forcing near zero over decades due to clustering of eruptions. Paleoclimate records from ice cores reveal that explosive volcanism has driven multiyear cools in the past, but reconstructions show no long-term trend amplifying recent warming; underestimation of aerosol lifetime in models may amplify projected cooling from future events by a factor of two. Internal variability arises from coupled ocean-atmosphere dynamics, notably the El Niño-Southern Oscillation (ENSO) on interannual scales, the Pacific Decadal Oscillation (PDO), and Atlantic Multidecadal Oscillation (AMO) on 20–70-year scales, redistributing heat without net energy addition to the system. ENSO phases modulate global temperatures by ±0.1–0.2°C, with El Niño events enhancing short-term warming. The AMO's positive phase from ~1925–1965 and PDO's warm regime in the mid-20th century amplified early-century Arctic and global anomalies, contributing up to 0.2–0.3°C to hemispheric trends through altered circulation and heat release from ocean depths. These modes explain ~30–50% of early 20th-century variance but transitioned to neutral or negative phases post-1970 (e.g., AMO peak ~1995–2010), correlating with slower surface warming rates in some regions despite overall trend continuation, highlighting their role in fluctuations rather than secular change. One analysis attributes roughly half of early 20th-century warming (1910–1940) to well-mixed greenhouse gases alongside reductions in shortwave-absorbing aerosols and solar factors, suggesting over-reliance on anthropogenic dominance.

Anthropogenic Influences

Human activities have significantly altered the composition of Earth's atmosphere, primarily through emissions of greenhouse gases (GHGs) and aerosols, exerting a net positive radiative forcing that contributes to global warming. The most prominent anthropogenic influence is the increase in atmospheric carbon dioxide (CO₂) concentration, which rose from approximately 280 parts per million (ppm) in the pre-industrial era (prior to the mid-19th century) to over 420 ppm by 2023, representing more than a 50% increase driven largely by fossil fuel combustion and cement production. Isotopic analysis of atmospheric CO₂ confirms the fossil fuel origin of this rise, as evidenced by the decline in the ¹³C/¹²C ratio and the near-absence of radiocarbon (¹⁴C), signatures unique to "old" carbon from ancient biomass rather than recent biogenic or oceanic sources. Methane (CH₄), the second most important anthropogenic GHG after CO₂, has seen concentrations increase by about 150% since pre-industrial levels, with over 60% of current emissions attributable to human sources such as agriculture (including livestock enteric fermentation and rice cultivation, accounting for around 40%), fossil fuel extraction and distribution (leaks from natural gas systems), and waste management (landfills). Nitrous oxide (N₂O) emissions, primarily from agricultural fertilizer use and industrial processes, have risen by roughly 20%, contributing additional forcing despite lower concentrations. Land-use changes, particularly deforestation for agriculture and logging, release stored carbon and reduce terrestrial sinks, accounting for 6-12% of annual global CO₂ emissions; for instance, tropical forest loss alone emitted over 5.6 billion tonnes of CO₂-equivalent gases yearly in recent decades. Anthropogenic aerosols, including sulfates from fossil fuel burning and black carbon from incomplete combustion, introduce a countervailing cooling effect by scattering sunlight and enhancing cloud reflectivity, estimated to offset 0.5-1.1°C of GHG-induced warming globally. This negative forcing masks some warming but varies regionally and temporally, with reductions in aerosol emissions (e.g., due to air quality regulations) potentially accelerating temperature rises in coming decades. Overall, peer-reviewed assessments attribute the net anthropogenic radiative forcing since 1750 at approximately +2.0 to +2.5 W/m², dominated by well-mixed GHGs, though uncertainties persist in aerosol-cloud interactions and historical emission inventories.

Evidence and Attribution Methods

Detection and attribution studies aim to determine whether observed climate changes exceed expected natural variability and to quantify the contributions from specific forcings, such as greenhouse gases, aerosols, solar irradiance, and volcanic activity. These methods rely on statistical analyses and climate model simulations to compare observed data against ensembles of model runs that isolate individual or combined forcings. Detection identifies a significant signal in observations inconsistent with internal variability alone, while attribution assesses the best explanation among possible external drivers by evaluating goodness-of-fit metrics like scaling factors in optimal fingerprinting techniques. Primary methods include process-based modeling using global climate models from projects like CMIP6, where simulations with natural forcings only (e.g., solar cycles and volcanic eruptions) are contrasted with those incorporating anthropogenic forcings (e.g., CO2 and methane increases since the Industrial Revolution). Optimal fingerprinting applies generalized least squares regression to match spatial or temporal patterns ("fingerprints") in observations to model-predicted responses, estimating the amplitude of each forcing's influence. Evidence-based approaches supplement models by directly comparing physical indicators, such as the tropospheric warming and stratospheric cooling pattern, which aligns with radiative forcing from well-mixed greenhouse gases rather than natural solar variations. Key evidence supporting anthropogenic attribution includes the inability of natural-forcing-only simulations to reproduce post-1950 global surface warming, which models match only when including rising greenhouse gas concentrations; for instance, CMIP6 ensembles show natural forcings alone projecting near-zero or slight cooling trends from 1850–2020 due to volcanic influences, contrasting observed warming of approximately 1.1°C. Ocean heat content increases in the upper 2000 meters, measured by Argo floats since 2004, exhibit trends attributable to anthropogenic forcing with high confidence, as natural variability alone cannot explain the sustained energy uptake exceeding 90% of Earth's excess radiative imbalance. Spatial patterns, such as amplified warming over land and Arctic regions, further fingerprint human influence, with regression analyses yielding scaling factors near unity (indicating no need for model adjustments to fit observations). Critiques of these methods highlight uncertainties in model representation of natural variability, including multidecadal oscillations like the Atlantic Multidecadal Variability, which some analyses suggest are underrepresented in CMIP ensembles, potentially leading to overestimation of anthropogenic signals. Optimal fingerprinting assumes linear responses and Gaussian statistics, assumptions challenged by nonlinear feedbacks and regime shifts in paleoclimate records, raising questions about the robustness of attribution statements. Independent assessments argue that reliance on equilibrium climate sensitivity estimates from models, rather than emergent constraints from observations, introduces circularity, as models tuned to historical data may amplify confirmation of their own forcings. Despite these limitations, multi-method convergence—combining instrumental records, proxies, and reanalyses—provides medium to high confidence in dominant human causation for global warming since the mid-20th century, though precise quantification of contributions (e.g., 100% vs. partial) remains debated due to unforced variability estimates varying by 0.1–0.3°C in recent decades.

Modeling and Predictions

Development of Climate Models

The foundations of climate modeling trace back to 19th-century theoretical work on atmospheric heat transfer and radiative forcing, including Joseph Fourier's 1824 recognition of the greenhouse effect and Svante Arrhenius's 1896 calculation estimating that doubling atmospheric CO2 could raise global temperatures by 5–6°C. These early efforts were analytical rather than numerical, relying on simplified energy balance equations without computational simulation of dynamic processes. The transition to numerical climate models began in the mid-20th century, building on advances in numerical weather prediction (NWP). In 1956, Norman Phillips produced the first general circulation model (GCM) of the atmosphere, simulating global circulation patterns using a two-dimensional grid on an early computer, which demonstrated realistic zonal wind structures despite coarse resolution and simplified physics. This marked the shift from static calculations to dynamic simulations governed by the primitive equations of fluid motion, though initial runs required manual adjustments and were limited by computational constraints to short integrations. By the 1960s, GCMs evolved into three-dimensional frameworks incorporating radiative transfer and moist convection. Joseph Smagorinsky's group at NOAA developed operational GCMs starting in 1963, emphasizing realistic simulations of large-scale circulation driven by solar heating gradients. A pivotal advancement came in 1967 with Syukuro Manabe and Richard Wetherald's one-dimensional radiative-convective model at GFDL, which quantified the surface warming from CO2 doubling as approximately 2.3°C after accounting for water vapor feedback, laying groundwork for integrating greenhouse gas forcings into multi-level atmospheric models. The 1970s saw the maturation of coupled atmosphere-ocean GCMs (AOGCMs), addressing the limitations of atmosphere-only models that prescribed sea surface temperatures. The first such coupled model appeared in 1969, simulating air-sea interactions, followed by refinements in the 1975 GFDL model that included oceanic heat diffusion and salinity effects for decadal simulations. These developments incorporated parameterizations for sub-grid processes like clouds and turbulence, as direct resolution remained infeasible due to computing power—early GCMs operated at resolutions of hundreds of kilometers horizontally. Subsequent decades brought increased complexity through Earth system models (ESMs), integrating biogeochemical cycles such as carbon and aerosols. The 1980s introduced interactive vegetation and land surface schemes, while the 1990s Coupling Model Intercomparison Project (CMIP) standardized multi-model ensembles, enabling systematic evaluation and refinement across institutions like Hadley Centre and NCAR. Resolution improved from ~300 km in early GCMs to ~10–100 km by the 2010s, facilitated by supercomputing advances, though parameterizations for unresolved physics persist as sources of uncertainty. International coordination via the World Climate Research Programme has driven phases like CMIP6 (circa 2016), incorporating higher-fidelity ocean eddies and ice sheets for millennial-scale projections.

Evaluation of Model Accuracy

Climate models are evaluated for accuracy through hindcasting—simulating historical climate conditions and comparing outputs to observational data—and forecasting, where past projections are assessed against subsequent real-world measurements. Hindcast performance assesses how well models reproduce known past trends in variables like global surface air temperature (SAT), precipitation, and sea levels, while forecast skill examines out-of-sample predictions, such as post-publication warming rates. Evaluations often use multi-model ensembles like those from the Coupled Model Intercomparison Project (CMIP) phases 3 through 6, comparing ensemble means and individual model runs to datasets from sources including satellites, weather stations, and buoys. Discrepancies arise due to model assumptions about feedbacks, such as cloud responses and aerosol effects, which amplify or dampen warming. Global SAT hindcasts in CMIP3 and CMIP5 ensembles show reasonable skill for 20th-century trends, with improvements in regional and decadal variability from CMIP3 to CMIP5, though models exhibit biases in polar amplification and tropical patterns. For forecasts, a 2019 analysis of models published from 1970 to 2007 found that, after adjusting for differences in radiative forcing scenarios, projected warming aligned closely with observations through 2017, with no systematic over- or underestimation in the ensemble median. However, CMIP5 models simulated SAT increases about 16% faster than observed global averages since 1970, with roughly 40% of the divergence attributable to excessive tropical warming in models; CMIP6 exhibits even larger discrepancies, overestimating warming over 63% of Earth's surface area in recent decades. Upper-air temperature trends, measured by satellites since 1979, reveal persistent model overestimation of mid-tropospheric warming rates, exceeding observations by factors of 2-3 in the tropics. Beyond temperature, model accuracy varies by variable. Precipitation hindcasts in CMIP ensembles capture broad trends but underestimate extremes and regional variability, with CMIP6 showing mixed improvements over CMIP5 yet persistent dry biases in subtropical zones. Sea ice extent simulations overestimate Arctic summer minima in recent decades compared to satellite observations, while Antarctic trends are better matched but still diverge in multi-year forecasts. Equilibrium climate sensitivity (ECS), the long-term warming from doubled CO2, implied by CMIP6 models averages around 3.7°C, higher than many observationally derived estimates of 1.5-2.5°C from instrumental records and paleoclimate proxies, suggesting models may amplify positive feedbacks like water vapor while underrepresenting negative cloud feedbacks. Evaluations indicate that while early single-model projections were often skillful for global SAT, modern ensembles tuned to historical data tend to run "hot" for post-2000 forecasts, prompting calls for weighting schemes favoring lower-sensitivity models in projections.

Projections and Equilibrium Climate Sensitivity

Equilibrium climate sensitivity (ECS) represents the long-term global surface air temperature response to a doubling of atmospheric CO₂ concentration from pre-industrial levels (approximately 280 ppm), after the climate system reaches a new equilibrium, incorporating slow feedbacks such as ice sheet changes. ECS estimates derive from three primary approaches: process-based general circulation models (GCMs), instrumental records using energy budget constraints, and paleoclimate proxies like ice ages or volcanic eruptions. Model-based estimates from CMIP6 GCMs range from 1.8°C to 5.6°C, with a multimodel mean of about 3.9°C, reflecting diverse representations of cloud feedbacks and aerosol effects. However, these higher sensitivities in CMIP6 have been critiqued for inconsistency with observed historical warming patterns, potentially biasing effective climate sensitivity upward due to flawed spatial patterns in simulated surface temperatures. Instrumental estimates, which constrain ECS using observed 20th-century warming, radiative forcing, and ocean heat uptake, typically yield lower values. A 2021 analysis of energy budget methods found a median ECS of 2.16°C (5–95% range: 1.1–3.9°C), lower than many GCMs, attributing discrepancies to overestimated forcing or underestimated historical aerosol cooling. Paleoclimate-based assessments, such as those from the Last Glacial Maximum (around 21,000 years ago), support ECS values around 2.5–2.7°C when accounting for revised estimates of polar amplification and dust forcings. The Intergovernmental Panel on Climate Change's Sixth Assessment Report (AR6, 2021) synthesizes these methods to assess ECS as likely (66–100% probability) between 2.5°C and 4.0°C, with a best estimate of 3.0°C, narrowing the prior AR5 range of 1.5–4.5°C but retaining substantial uncertainty due to cloud feedback ambiguities. This assessment has faced scrutiny for overweighting model ensembles over observationally derived bounds, amid evidence that low-ECS models better match recent Earth energy imbalance trends when adjusted for shortwave and longwave components. Climate projections for global temperature rise integrate ECS alongside transient climate response (TCR, the warming during gradual CO₂ increase) and socioeconomic pathways (Shared Socioeconomic Pathways, SSPs) in ensembles like CMIP6. Under SSP1-1.9 (very low emissions aligning with 1.5°C Paris goals), projected median warming by 2081–2100 is 1.4°C (likely range: 1.0–1.8°C) relative to 1850–1900. The intermediate SSP2-4.5 scenario forecasts 2.7°C (range: 2.1–3.5°C), while high-emissions SSP5-8.5 anticipates 4.4°C (3.3–5.7°C), driven by cumulative CO₂ emissions and non-CO₂ forcings. These projections assume continued historical trends in emissions and land use but exhibit wide intermodel spread, partly from ECS variability; subsets with ECS below 3°C align more closely with observed 1970–2020 warming rates of about 0.2°C per decade. Critiques highlight that full CMIP6 projections warm faster than AR6-assessed likely ranges, potentially overstating near-term risks due to inflated TCR (CMIP6 mean 2.1°C vs. AR6 1.8°C). Near-term projections (to 2050) show less scenario divergence, with global mean temperature likely exceeding 1.5°C by the early 2030s across SSPs, though natural variability could delay this by up to a decade. Uncertainties in projections stem from ECS estimation challenges, including incomplete treatment of tipping elements like permafrost thaw or Amazon dieback, which could amplify warming beyond linear responses. Recent studies using emergent constraints—linking model ECS to observable variables like tropical cloud regimes—suggest the high end (>4°C) is less probable, with observationally calibrated ECS favoring 2–3°C and reducing projected 2100 warming by 0.5–1°C under high-emission paths. Conversely, arguments against narrowing ECS further emphasize persistent gaps in process understanding, such as low-cloud feedback strength, validated by GCM-paleoclimate mismatches during the Last Glacial Maximum. Overall, while AR6 projections underscore risks of exceeding 2°C without rapid mitigation, empirical constraints imply milder long-term sensitivities, tempering the upper bounds of catastrophe narratives from uncalibrated models.

Potential Impacts

Environmental Consequences

Global temperatures have risen by approximately 1.1°C since pre-industrial times, leading to observable shifts in ecosystems, including poleward migrations of species ranges in terrestrial and marine environments. Empirical studies document that about 46.6% of range-shift observations align with expected poleward, upslope, or deeper-water movements, though many species exhibit stasis or idiosyncratic responses due to dispersal limitations, habitat fragmentation, and non-climatic stressors. Land-use changes remain the primary driver of recent global biodiversity loss, with climate warming contributing through altered phenology, extinction risks for endemic species, and intensified interactions with other pressures like habitat destruction. In marine systems, ocean warming and acidification—driven by CO2 absorption reducing surface pH by 0.1 units since the Industrial Revolution—have impaired calcification in shell-forming organisms such as pteropods and corals. Laboratory and field experiments indicate negative effects on survival, growth, and reproduction in various taxa, including reduced larval development in shellfish and disrupted symbiosis in corals leading to bleaching. The fourth global coral bleaching event, confirmed in 2023-2024, affected reefs across the Atlantic, Pacific, and Indian Oceans, with heat stress exceeding thresholds for prolonged periods; attribution analyses link increased frequency to anthropogenic warming, though local factors like pollution exacerbate vulnerability. Terrestrial ecosystems show amplified responses in high-latitude and high-elevation regions, where warming exceeds the global average. Permafrost thaw, affecting up to 24% of Northern Hemisphere permafrost by 2100 under moderate emissions scenarios, releases stored organic carbon as CO2 and methane, potentially adding 0.13-0.27 GtC/year, while destabilizing landscapes through thermokarst formation and altering hydrology. Glacier retreat has accelerated, with mass loss from Greenland and Antarctic ice sheets contributing roughly 50% of observed sea-level rise, which totals 21-24 cm globally since 1880, accelerating to 3.7 mm/year in recent satellite records. These changes disrupt alpine and Arctic habitats, promoting shrub encroachment over tundra and increasing wildfire risk in boreal forests via drier conditions. Projections indicate further ecosystem reorganization, with meta-analyses forecasting up to 16% of species at high extinction risk under 2°C warming, concentrated in biodiversity hotspots like tropical mountains and islands; however, adaptation through genetic variation and dispersal may mitigate losses for some taxa, and CO2 fertilization has enhanced vegetation productivity in mid-latitudes, countering some desiccation effects. Uncertainties persist in attributing specific biodiversity declines solely to climate, as synergistic drivers like invasive species and overexploitation often dominate empirical datasets.

Socioeconomic Effects

Climate change has been associated with economic damages primarily through intensified extreme weather events, altered agricultural productivity, and disruptions to human settlements, though the magnitude of these effects remains debated among economists due to uncertainties in attribution, adaptation potential, and baseline comparisons. Empirical estimates of global GDP impacts vary widely; a meta-analysis of studies through 2023 found central projections of 1.9% income loss for 2.5°C warming and 7.9% for 5°C, after accounting for publication biases that tend to inflate pessimistic outcomes. Observed damages from weather extremes, partially attributable to anthropogenic warming, reached approximately $143 billion annually in recent years, with human losses comprising 63% of the total. In the United States, climate-related events from 1980 to early 2025 inflicted over $2.9 trillion in costs, predominantly from hurricanes, floods, and wildfires. These figures, however, often include uninsured losses and do not fully isolate climate signals from natural variability or socioeconomic factors like population density in vulnerable areas. Agricultural sectors face uneven impacts, with empirical models indicating yield reductions in tropical and subtropical regions due to heat stress and water scarcity, while higher latitudes may see modest gains from longer growing seasons. A 2025 analysis projected global crop production declines concentrated in modern breadbaskets like the U.S. Midwest and parts of Europe and Asia, exacerbating food price volatility and contributing to undernutrition in low-income populations. In developing economies reliant on rain-fed agriculture, smallholder farmers experience compounded vulnerabilities, including health effects from reduced nutritional quality and labor productivity losses during heatwaves, potentially driving rural-to-urban migration. Observed data from IPCC assessments link these changes to accelerated shifts from farm-based livelihoods to urban employment, particularly in Asia and Africa, where climate variability has intersected with poverty to hinder sustainable development. Human health burdens include excess mortality from heat extremes and expanded ranges for vector-borne diseases like malaria and dengue, with the World Health Organization estimating that climate change could cause hundreds of thousands of additional deaths annually by mid-century through direct and indirect pathways such as malnutrition and injury. Conversely, fewer cold-related deaths in temperate zones partially offset these, though net global health costs are projected positive in most models. Migration patterns are influenced by climate stressors, with evidence showing intensified flows from low-latitude, low-income areas to higher-latitude urban centers, amplifying inequality as wealthier nations absorb migrants while origin regions lose labor. A 2021 study quantified climate's role in boosting such movements, projecting up to 200 million additional climate-displaced individuals by 2050 under high-emission scenarios, often triggered by droughts, floods, and sea-level encroachment in coastal deltas. These displacements strain public resources in receiving areas, including infrastructure and social services, while empirical critiques highlight that many projections overlook adaptive migration benefits, such as remittances and knowledge transfers. Regional disparities underscore that developing nations bear disproportionate burdens relative to their emissions, with OECD projections estimating global output reductions rising from 1.75% of GDP currently to nearly 9% by 2100, heavily weighted toward tropical economies lacking adaptive capacity. Some empirical research cautions against overreliance on integrated assessment models, which may underestimate damages by ignoring spatial convergence in economic growth or overestimate by downplaying historical adaptation rates observed during past warming periods. Overall, while projections like a committed 19% global income drop by 2049 from past emissions reflect panel data analyses, they contrast with lower-end estimates from meta-reviews, reflecting ongoing debates over model assumptions and the inclusion of unquantified benefits such as reduced heating demands in colder regions.

Countervailing Benefits

Satellite observations indicate that Earth has experienced significant greening over the past several decades, with increased leaf area index across global vegetation. A 2016 NASA study attributed approximately 70% of this greening to carbon dioxide fertilization, where elevated atmospheric CO2 enhances photosynthesis and plant growth, particularly in regions like China and India. This effect has contributed to a net increase in global vegetation cover, potentially boosting biomass production and carbon sequestration, though long-term sustainability remains uncertain due to nutrient limitations. In agriculture, CO2 fertilization has demonstrably increased crop yields. Analysis of historical data from 1961 to 2017 shows that elevated CO2 raised yields of C3 crops such as rice and wheat by 7.1%, offsetting some negative impacts from warming. Experimental evidence confirms that higher CO2 levels improve water-use efficiency in crops, mitigating drought stress and supporting higher productivity under elevated temperatures. Longer growing seasons in higher latitudes and altitudes, driven by warming, have enabled expanded cultivation of crops like potatoes in northern Europe and shifted viable farmland to elevated areas. Mortality patterns reveal that cold-related deaths substantially outnumber heat-related ones globally. A comprehensive analysis found excess cold-attributable deaths at roughly 9 times the rate of heat-related deaths, totaling about 4.6 million versus 489,000 annually. From 2000–2019, cold temperatures were linked to 8.52% of excess deaths worldwide, compared to 0.91% for hot temperatures, suggesting that moderate warming could yield a net reduction in temperature-related mortality by decreasing cold fatalities. Empirical trends indicate rising temperatures have already averted approximately 166,000 deaths net globally through reduced cold exposure. Economic opportunities arise from the opening of Arctic shipping routes due to ice melt. Projections suggest that by 2050, up to 5% of global shipping could utilize these shorter paths, reducing transit times and fuel costs between Europe and Asia. The Northern Sea Route, for instance, offers distances 40% shorter than traditional Suez or Panama alternatives, facilitating trade and resource extraction while benefiting northern communities through enhanced access. Historically, warmer periods such as the Medieval Warm Period correlated with agricultural expansions and Norse settlements in Greenland, underscoring potential societal adaptations to benign warming. These benefits, while regionally variable, counterbalance some projected adverse effects, emphasizing the need for nuanced assessment beyond predominant negative narratives.

Policy Responses

Mitigation Approaches and Costs

Mitigation of climate change primarily involves strategies to reduce anthropogenic greenhouse gas emissions, particularly carbon dioxide from fossil fuel combustion, through technological innovation, policy interventions, and land-use changes. Key approaches in the energy sector include transitioning to low-carbon electricity generation via solar photovoltaic, onshore wind, nuclear fission, and, to a lesser extent, carbon capture and storage applied to fossil fuels. In transportation, electrification of vehicles powered by low-emission grids and shifts to biofuels or hydrogen are emphasized, while industrial processes target efficiency gains and electrification or hydrogen substitution. Agriculture and forestry mitigation focuses on reducing methane from livestock, enhancing soil carbon sequestration, and curbing deforestation. These strategies aim to achieve net-zero emissions by mid-century in ambitious scenarios, though their feasibility depends on scalability and integration challenges like grid reliability for intermittent renewables. Economic costs of mitigation vary by approach and are often assessed using integrated assessment models (IAMs) or levelized cost of energy (LCOE) metrics, which calculate lifetime costs per unit of electricity generated. Unsubsidized LCOE for utility-scale solar PV in 2024 ranges from $24 to $96 per MWh, onshore wind from $24 to $75 per MWh, combined-cycle natural gas from $39 to $101 per MWh, and new nuclear from $141 to $221 per MWh, excluding system-level integration costs such as storage for renewables' intermittency or backup capacity. Critics note that standard LCOE understates total system expenses for renewables, which require overbuild, firming capacity, and transmission upgrades, potentially doubling effective costs in high-penetration scenarios; nuclear, despite higher upfront capital, provides dispatchable baseload power with lower long-term fuel and operational variability. Empirical analyses of implemented policies, such as the European Union's Emissions Trading System, indicate emission reductions of 1-2% annually attributable to carbon pricing, but at marginal abatement costs exceeding $100 per ton of CO2 in some sectors, contributing to elevated energy prices and industrial leakage.
TechnologyUnsubsidized LCOE Range (USD/MWh, 2024)Key Limitations
Utility-Scale Solar PV24–96Intermittency requires storage; land use
Onshore Wind24–75Variability; curtailment in oversupply
Gas Combined Cycle39–101Ongoing emissions; fuel price volatility
Nuclear (New Build)141–221High capital; regulatory delays
Global mitigation to limit warming to 2°C by 2100 is projected by IAMs to reduce GDP by 1-4% cumulatively compared to baseline scenarios, with costs rising nonlinearly for deeper cuts due to diminishing marginal returns on low-cost abatement options like efficiency improvements. Net-zero pathways by 2050 could impose annual GDP losses of 2-10% in high-income economies, driven by energy sector transformations, though some models offset portions via co-benefits such as reduced air pollution health costs estimated at $50-200 per ton of CO2 abated. Cost-benefit analyses reveal tensions: benefits hinge on assumptions of high climate damages (often 2-4% GDP per degree Celsius), yet empirical damage functions from observed warming suggest lower sensitivities, rendering aggressive mitigation uneconomic if equilibrium climate sensitivity is below 3°C. Policies like subsidies for renewables have accelerated deployment but distorted markets, with U.S. Production Tax Credits totaling over $15 billion annually by 2023, while suppressing nuclear development through regulatory hurdles that inflate costs by 20-50% relative to historical builds. Ex post evaluations of three decades of mitigation efforts show discernible emission drivers curbed, such as coal phase-outs, but total global emissions rose 60% since 1990 due to developing economy growth outpacing reductions in the West, questioning the efficacy of unilateral high-cost actions.

Adaptation Strategies

Adaptation strategies encompass a range of measures designed to reduce vulnerability to climate variability and change, including infrastructure hardening, behavioral adjustments, and technological innovations. These differ from mitigation by addressing impacts rather than causes, often yielding high returns on investment; for instance, the Global Commission on Adaptation estimated that every United States dollar invested in adaptation could generate up to ten dollars in net benefits through avoided damages and enhanced resilience. Empirical reviews of over 1,600 studies document implemented adaptations worldwide, spanning sectors like agriculture, water, and health, with many actions proving effective against observed weather extremes rather than solely long-term trends. In coastal regions, structural defenses have demonstrated longevity and efficacy. The Thames Barrier in London, operational since 1982, has been raised over 200 times to protect against tidal surges exacerbated by storm events, averting billions in potential flood damages. Similarly, the Netherlands' Delta Works, initiated after the 1953 North Sea flood that killed over 1,800 people, integrate dikes, sluices, and storm surge barriers, reducing flood risk for 60% of the population living below sea level. Autonomous adaptations, such as farmers altering planting schedules or adopting drought-tolerant crops, occur without formal policy; econometric analyses in agriculture show these responses mitigate yield losses from temperature and precipitation shifts by 10-30% in regions like sub-Saharan Africa. Water management strategies include enhanced storage and efficiency measures. In California, expanded reservoir capacities and groundwater recharge programs implemented since the 2012-2016 drought have buffered against hydrologic variability, sustaining urban and agricultural supplies amid reduced Sierra Nevada snowpack. Nature-based approaches, like wetland restoration and afforestation, offer co-benefits such as biodiversity support; peer-reviewed case studies in South Africa and the Netherlands highlight their role in flood attenuation and erosion control, with cost-benefit ratios often surpassing 4:1. Health adaptations focus on heat and vector-borne risks, including early warning systems that have contributed to a 90% decline in weather-related disaster deaths globally since 1920, per United Nations data, through improved forecasting and sheltering. Economic evaluations underscore adaptation's feasibility over inaction. A European Environment Agency assessment found adaptation measures cost-efficient when benefit-cost ratios exceed 1.5, though quantifying avoided damages remains challenging due to baseline uncertainties. In developing contexts, risk-pooling insurance mechanisms in African nations have enabled rapid post-disaster recovery, covering livestock losses from droughts since 2008. Historical precedents, such as U.S. Great Plains farmers introducing irrigation and hybrid crops during 1930s Dust Bowl conditions, illustrate adaptive capacity driven by market incentives rather than centralized planning. Constraints persist in quantifying limits, with some studies noting diminishing returns in highly exposed low-lying areas, yet empirical evidence favors proactive, localized strategies over broad assumptions of uniform vulnerability.

International Frameworks and Outcomes

The United Nations Framework Convention on Climate Change (UNFCCC), adopted on May 9, 1992, and entering into force on March 21, 1994, established a framework for international cooperation to stabilize atmospheric greenhouse gas concentrations at levels preventing dangerous anthropogenic interference with the climate system. Ratified by 198 parties, it categorized countries into Annex I (developed nations with binding reporting) and non-Annex I (developing nations with fewer obligations), laying the groundwork for subsequent protocols but imposing no immediate emission reduction targets. The Kyoto Protocol, adopted in 1997 and entering into force in 2005, required 37 industrialized countries and the European Union to reduce greenhouse gas emissions by an average of 5% below 1990 levels during 2008–2012. Empirical analyses indicate that ratifying Annex I countries achieved approximately 7% lower emissions than projected under a no-protocol baseline, attributed to policy implementations like emissions trading. However, the United States never ratified, Canada withdrew in 2011, and major developing emitters like China faced no binding cuts, resulting in global emissions rising 32% from 1990 to 2010 despite the protocol's efforts. The Paris Agreement, adopted on December 12, 2015, and entering into force on November 4, 2016, shifted to a universal framework where all 196 parties submit nationally determined contributions (NDCs) for emission reductions and adaptation, aiming to limit warming to well below 2°C above pre-industrial levels while pursuing 1.5°C. Current NDCs, if fully implemented, project emissions insufficient to meet these goals, with global greenhouse gases needing to peak before 2025 and decline 43% by 2030 for 1.5°C compatibility; instead, trends forecast 2.5–3°C warming by 2100. Post-2015, global CO2 emissions have continued increasing annually, driven by growth in developing economies, underscoring the voluntary nature's limitations in enforcing absolute reductions. Annual Conference of the Parties (COP) meetings under the UNFCCC have produced incremental outcomes, such as the 2009 Copenhagen Accord's voluntary pledges and the 2021 Glasgow Climate Pact's coal phase-down language, but compliance remains uneven, with many NDCs lacking enforcement mechanisms. Overall, these frameworks have facilitated technology transfer and reporting but failed to reverse the upward trajectory of global emissions, as empirical data from 1992 onward shows concentrations rising from ~355 ppm to over 420 ppm CO2 equivalent, reflecting causal drivers like economic development outweighing mitigation pledges. Critiques from economic analyses highlight that binding targets on developed nations alone shifted emissions to unregulated sectors, yielding negligible net global impact.

Controversies and Alternative Views

Challenges to Consensus Narratives

Critics contend that claims of near-unanimous scientific consensus on anthropogenic causes dominating recent climate change overstate agreement due to selective categorization in literature reviews. A peer-reviewed reanalysis of a widely cited study purporting 97% endorsement found methodological inconsistencies, including the improper classification of neutral papers and low endorsement rates among those taking a position, rendering the figure unreliable. Disparities between surface-based temperature datasets and satellite-derived lower tropospheric measurements challenge the uniformity of warming trends. Satellite records from the University of Alabama in Huntsville (UAH) report a global lower troposphere warming rate of approximately 0.14°C per decade from 1979 to 2023, lower than surface estimates around 0.18°C per decade, potentially indicating overestimation in ground records. Urban heat island (UHI) effects, arising from land-use changes in populated areas, contribute significantly to local warming biases. One study quantified UHI as responsible for 22% of observed U.S. summer surface temperature increases since 1895. Homogenization adjustments to correct for station moves and instrument changes have been shown to inadvertently blend urban influences into rural records, amplifying apparent trends. General circulation models (GCMs) underpinning IPCC projections exhibit systematic errors, including overprediction of historical warming in key regions like the tropical troposphere. Many CMIP5 and CMIP6 models simulate warming rates exceeding observations, with discrepancies attributed to overstated cloud feedbacks and equilibrium climate sensitivity (ECS) values—CMIP6 averages around 3.7°C per CO2 doubling, higher than some instrumental estimates below 3°C. These models failed to anticipate the 1998–2013 warming hiatus, during which natural variability dominated despite rising CO2, highlighting limitations in capturing internal ocean-atmosphere oscillations. Specific forecasts tied to consensus narratives have diverged from outcomes, eroding predictive credibility. Predictions of Arctic sea ice vanishing by 2013 or widespread submersion of low-lying nations like the Maldives by 2020 did not occur, with Arctic summer ice persisting at multi-million square kilometer levels and Maldives land area expanding due to accretion. James Hansen's 1988 congressional testimony projected U.S. warming of 0.45°C by 2010 under moderate emissions, yet observed increases were about 0.3°C. Attribution studies emphasizing near-total human forcing overlook amplified roles for natural drivers in recent decades. The Pacific Decadal Oscillation (PDO) positive phase since the 1970s correlates with enhanced Pacific warming, while solar irradiance variations and Atlantic Multidecadal Oscillation contribute to multidecadal patterns not fully replicated in models. Such critiques, often from peer-reviewed sources outside dominant institutional frameworks, underscore uncertainties amplified by systemic biases in funding and publication favoring alarmist interpretations.

Economic and Policy Critiques

Critics argue that aggressive climate mitigation policies, such as those pursuing net-zero emissions by 2050, impose substantial economic burdens that often outweigh projected benefits. Estimates indicate that global implementation could cost over $200 trillion by mid-century and up to $2 quadrillion by 2100, primarily through subsidies, regulatory mandates, and infrastructure overhauls for renewables, with returns yielding approximately 17 cents in avoided damages per dollar expended. These figures derive from integrated assessment models incorporating discounting and uncertainty, highlighting how front-loaded costs in developing economies exacerbate energy poverty and hinder growth. Empirical evidence underscores the limited efficacy of such policies in curbing emissions. Global climate finance reached $1.9 trillion in 2023, with mitigation efforts doubling since 2018, yet CO2 emissions rose 5.6% from 2015 to 2024, outpacing GDP growth and undermining Paris Agreement goals. Official development assistance for mitigation shows no discernible correlation with emissions reductions in recipient countries, suggesting misallocation toward low-impact projects like inefficient subsidies rather than innovation. Bjorn Lomborg, drawing on Copenhagen Consensus analyses by Nobel laureates, contends that the Paris Agreement's commitments, costing $819–$1,890 billion annually by 2030, would achieve less than 1% of required reductions, rendering it a high-cost, low-yield endeavor. Opportunity costs further amplify these concerns, as funds diverted to mitigation forego interventions with higher returns in human welfare. For instance, $1.1 trillion spent globally on green technologies in 2022 yielded marginal temperature reductions, while reallocating a fraction to research and development—currently under 4% of GDP in relevant sectors—could accelerate breakthroughs like advanced nuclear or carbon capture at lower long-term expense. Economists associated with the Copenhagen Consensus prioritize adaptation measures, such as resilient infrastructure, over emission cuts, estimating that optimal policies could limit warming to 3.75°C at 2.6% of GDP versus unchecked damages of 3%, but current regulatory approaches ignore such trade-offs. Policy frameworks like the Paris Agreement lack enforceable mechanisms, leading to non-compliance and emissions leakage, where reductions in one nation shift production to unregulated areas, netting minimal global impact. Alternative strategies emphasize innovation-driven growth over degrowth mandates. Lomborg's assessments reveal that historical warming since 1900 has correlated with net GDP gains of 1.5% annually through agricultural and health benefits, suggesting that prioritizing R&D in energy abundance—rather than rationing—avoids stifling development in low-income regions. Critics of net-zero timelines, including peer-reviewed modeling, warn of 8–18% annual GDP losses by 2050 from sector disruptions like offshoring emissions-intensive industries, which erode domestic value added without commensurate environmental gains. In sum, these critiques posit that policies should integrate cost-benefit rigor, favoring targeted investments over blanket restrictions to align economic realities with causal climate drivers.

Role of Media and Advocacy

Media coverage of climate change has often emphasized catastrophic scenarios and extreme weather attribution, contributing to heightened public alarm. A 2022 study analyzing English-language news media found a significant reporting bias favoring storms and wildfires, with these events receiving disproportionate attention compared to droughts or other impacts, potentially skewing perceptions of risk materiality. This selective focus aligns with broader trends where outlets attribute unusual weather—such as heat waves, floods, and hurricanes—to anthropogenic climate change, even when causal links remain uncertain or unproven. Such framing has fostered a moral panic, as mainstream reporting routinely overlooks empirical inconsistencies in alarmist predictions, including historical forecast failures like exaggerated sea-level rise or Arctic ice loss timelines. Advocacy organizations, including non-governmental groups like Greenpeace, Sierra Club, and Extinction Rebellion, have played a pivotal role in amplifying these narratives through campaigns, protests, and public shaming tactics. These entities mobilize grassroots pressure and influence policy by correlating their activities with shifts in public opinion on global warming, with studies showing moderate positive associations between environmental NGO presence and supportive attitudes. For instance, Extinction Rebellion's 2019 actions, including street blockades in major cities, aimed to disrupt normalcy and demand immediate emissions cuts, drawing global media attention and pressuring governments toward net-zero commitments. International NGOs further exert leverage via reputational costs, using public opinion channels and transnational advocacy to shape national laws, as evidenced in analyses of "climate shaming" campaigns. However, this advocacy-media synergy has drawn criticism for promoting alarmism that outpaces verifiable data, such as overstating near-term existential threats while underemphasizing adaptation successes or natural variability. Research indicates that sensationalist coverage and NGO-driven narratives can erode trust when predictions falter, with surveys revealing public irritation toward exaggerated claims that deter engagement rather than foster informed action. Mainstream outlets, influenced by institutional biases favoring consensus enforcement, often marginalize empirical critiques, granting limited space to analyses questioning model reliability or policy efficacy. Advocacy efforts, while effective in building political will, sometimes prioritize narrative persuasion over nuanced causal assessment, as seen in storytelling techniques that simplify complex dynamics to spur urgency. The interplay between media and advocacy has accelerated the adoption of crisis-oriented language, with Google Trends data showing sharp rises in searches for terms like "climate crisis" and "climate emergency" post-2018, coinciding with intensified NGO campaigns and editorial shifts. This linguistic evolution reflects coordinated efforts to frame the issue as an existential imperative, influencing public discourse and policy without always grounding in proportionate empirical evidence.

References

  1. [1]
    Indicators of Global Climate Change 2024: annual update of key ...
    Jun 19, 2025 · The 2024-observed best estimate of global surface temperature (1.52 °C) is well above the best estimate of human-caused warming (1.36 °C).
  2. [2]
    A year above 1.5 °C signals that Earth is most probably within the 20 ...
    Feb 10, 2025 · In 2023, global mean surface air temperature reached 1.43 °C above pre-industrial level (1.32–1.53 °C, likely range). This exceeded the ...
  3. [3]
    Evidence - NASA Science
    Oct 23, 2024 · There is unequivocal evidence that Earth is warming at an unprecedented rate. Human activity is the principal cause.
  4. [4]
    Causes of Climate Change | US EPA
    Aug 25, 2025 · Human activities have released large amounts of carbon dioxide and other greenhouse gases into the atmosphere, which has changed the earth's climate.
  5. [5]
    Long-term natural variability and 20th century climate change - PNAS
    Delineating the relative role of anthropogenic forcing, natural forcing, and long-term natural variability in 20th century climate change presents a significant ...
  6. [6]
    The early 20th century warming: Anomalies, causes, and ...
    The mid‐20th century end to the warming signaled the first period in that century that shows a detectable temperature change outside internal variability over 2 ...
  7. [7]
    Greater than 99% consensus on human caused climate change in ...
    Oct 19, 2021 · Greater than 99% consensus on human caused climate change in the peer-reviewed scientific literature, Lynas, Mark, Houlton, Benjamin Z, ...Abstract · Introduction · Method · Discussion
  8. [8]
    More than 99.9% of studies agree: Humans caused climate change
    Oct 19, 2021 · More than 99.9% of peer-reviewed scientific papers agree that climate change is mainly caused by humans, according to a new survey of 88,125 ...
  9. [9]
    Scientific Consensus - NASA Science
    Oct 21, 2024 · Based on well-established evidence, about 97% of climate scientists have concluded that human-caused climate change is happening.
  10. [10]
    What's the Difference Between Weather and Climate? | News
    Mar 23, 2018 · Whereas weather refers to short-term changes in the atmosphere, climate describes what the weather is like over a long period of time in a ...
  11. [11]
    Weather or Climate ... What's the Difference?
    Oct 19, 2023 · While weather refers to short-term changes in the atmosphere, climate refers to atmospheric changes over longer periods of time, usually 30 years or more.
  12. [12]
    What is the difference between weather and climate?
    Jun 16, 2024 · Weather reflects short-term conditions of the atmosphere while climate is the average daily weather for an extended period of time at a certain location.
  13. [13]
    Frequently Asked Questions — IPCC
    Climate change is a change in the state of the climate that can be identified (e.g., by using statistical tests) by changes in the mean and/ or the variability ...
  14. [14]
    What Is Climate Change? - NASA Science
    Oct 21, 2024 · Climate change is a long-term change in the average weather patterns that have come to define Earth's local, regional and global climates.
  15. [15]
    Basics of Climate Change | US EPA
    Aug 12, 2025 · These changes are due to a buildup of greenhouse gases in our atmosphere and the warming of the planet due to the greenhouse effect.
  16. [16]
    What is the greenhouse effect? - NASA Science
    Greenhouse gases consist of carbon dioxide, methane, ozone, nitrous oxide, chlorofluorocarbons, and water vapor. Water vapor, which reacts to temperature ...
  17. [17]
    Overview of Greenhouse Gases | US EPA
    Jan 16, 2025 · Gases that trap heat in the atmosphere are called greenhouse gases. This section provides information on emissions and removals of the main greenhouse gases.
  18. [18]
    What Are Greenhouse Gases and Why Do They Matter - Climate
    Greenhouse gases (GHGs) are a category of gases that absorb heat energy emitted from the planet's surface and they remain in Earth's atmosphere for a long time.
  19. [19]
    What Are Greenhouse Gases? - NASA Science
    Greenhouse gases are gases that can trap the heat from the sun near Earth's surface. They do this through a process known as the greenhouse effect.
  20. [20]
    The Causes of Climate Change - NASA Science
    Oct 23, 2024 · The greenhouse effect is essential to life on Earth, but human-made emissions in the atmosphere are trapping and slowing heat loss to space.
  21. [21]
    Climate and Earth's Energy Budget - NASA Earth Observatory
    Jan 14, 2009 · The natural greenhouse effect raises the Earth's surface temperature to about 15 degrees Celsius on average—more than 30 degrees warmer than it ...
  22. [22]
    How Carbon Dioxide Controls Earth's Temperature - NASA GISS
    Oct 14, 2010 · By this accounting, carbon dioxide is responsible for 80 percent of the radiative forcing that sustains the Earth's greenhouse effect. The ...
  23. [23]
    Annual Greenhouse Gas Index (AGGI) - Global Monitoring Laboratory
    For 2023, the AGGI was 1.51, which represents a 51% increase in effective radiative forcing from human-derived emissions of these gases since 1990. Changes in ...
  24. [24]
    How Atmospheric Water Vapor Amplifies Earth's Greenhouse Effect
    the process that occurs when gases in Earth's atmosphere trap the Sun's heat.
  25. [25]
    [PDF] Anthropogenic and Natural Radiative Forcing
    Radiative forcing includes present-day anthropogenic forcing from greenhouse gases, aerosols, and land changes, and natural forcing from solar and volcanic ...
  26. [26]
    Effective radiative forcing and adjustments in CMIP6 models - ACP
    Aug 17, 2020 · The effective radiative forcing, which includes the instantaneous forcing plus adjustments from the atmosphere and surface, has emerged as the ...
  27. [27]
    Bounding Global Aerosol Radiative Forcing of Climate Change
    Nov 1, 2019 · An assessment of multiple lines of evidence supported by a conceptual model provides ranges for aerosol radiative forcing of climate change ...
  28. [28]
    What's the difference between climate and weather? - NOAA
    Mar 9, 2016 · Climate is the average of the weather patterns in a location over a longer period of time, usually 30 years or more.Missing: variability | Show results with:variability
  29. [29]
    1.2.2 Natural Variability of Climate
    A distinction is made between physical feedbacks involving physical climate processes, and biogeochemical feedbacks often involving coupled biological ...
  30. [30]
    Climate change: evidence and causes | Royal Society
    Scientists have determined that, when all human and natural factors are considered, Earth's climate balance has been altered towards warming.
  31. [31]
    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 ...
  32. [32]
    and Naturally‐Caused Temperature Trends: A Systematic Approach ...
    May 1, 2023 · In the climate system, temperature change is jointly determined by internal natural variabilities, natural forcings, and anthropogenic forcings.<|control11|><|separator|>
  33. [33]
    [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.
  34. [34]
    Variable Walks In Our Climate Forest
    Oct 28, 2021 · Why doesn't the climate behave like we expect? The answer often lies in the internal variability of our atmosphere.
  35. [35]
    NOAA/WDS Paleoclimatology - EPICA Dome C - 800KYr CO2 Data
    So far, the Antarctic Vostok and EPICA Dome C ice cores have provided a composite record of atmospheric carbon dioxide levels over the past 650,000 years.
  36. [36]
    Milankovitch (Orbital) Cycles and Their Role in Earth's Climate
    Feb 27, 2020 · The small changes set in motion by Milankovitch cycles operate separately and together to influence Earth's climate over very long timespans, ...
  37. [37]
    Orbital Influences on Conditions Favorable for Glacial Inception
    Oct 18, 2021 · Milankovitch suggested that glacial cycles are driven by insolation integrated over the summer half-year at 65°N, which we will refer to as ...1 Introduction · 2 Data And Methods · 4 Conclusions
  38. [38]
    Holocene global mean surface temperature, a multi-method ... - Nature
    Jun 30, 2020 · The multi-proxy database includes a total of 1319 paleo-temperature records from 470 terrestrial and 209 marine sites where ecological, ...
  39. [39]
    A global database of Holocene paleotemperature records - Nature
    Apr 14, 2020 · A pre-deforestation pollen-climate calibration model for New Zealand and quantitative temperature reconstructions for the past 18 000 years BP.
  40. [40]
    Tree-ring stable isotopes from the European Alps reveal long-term ...
    Apr 4, 2025 · Our continuous tree-ring δ 18 O record reveals a significant long-term drying trend over much of the Holocene (P < 0.001), which is in line with orbital ...<|separator|>
  41. [41]
    Natural Causes of Climate Change - Florida Atlantic University
    Natural causes of climate change include orbital changes, volcanic eruptions, solar radiation variation, movement of crustal plates, and El Niño-Southern ...
  42. [42]
    Climate over past millennia - Jones - 2004 - AGU Journals - Wiley
    May 6, 2004 · We review evidence for climate change over the past several millennia from instrumental and high-resolution climate “proxy” data sources and climate modeling ...
  43. [43]
    Persistent warm Mediterranean surface waters during the Roman ...
    Jun 26, 2020 · This record comparison consistently shows the Roman as the warmest period of the last 2 kyr, about 2 °C warmer than average values for the late centuries.
  44. [44]
    Roman Warm Period and Late Antique Little Ice Age in an Earth ...
    Aug 9, 2022 · We analyze the climate differences during the two periods using large ensemble climate model simulations and the latest proxy reconstructions.
  45. [45]
    The Maunder minimum and the Little Ice Age: an update from recent ...
    Dec 4, 2017 · More striking for the LIA periods are the simulated temperature changes due to explosive volcanic activity, which produce sporadic large cooling ...
  46. [46]
    2 The Instrumental Record | Surface Temperature Reconstructions ...
    Global average temperature estimates based on the instrumental record indicate a near-level trend from 1856 to about 1910, a rise to 1945, a slight decline ...
  47. [47]
    Climate at a Glance | Global Time Series
    Coordinate temperature anomalies are with respect to the 1991-2020 average. All other regional temperature anomalies are with respect to the 1910-2000 average.
  48. [48]
    Early-twentieth-century cold bias in ocean surface ... - Nature
    Nov 20, 2024 · Instrumental GMST datasets, which blend sea surface temperature (SST) with surface air temperature over land and sea ice (LSAT), agree broadly ...
  49. [49]
  50. [50]
    Twentieth century sea level: An enigma - PNAS
    After the little ice age early in the 19th century, sea level rose at 18 cm/cy (the historic rate) with no measurable acceleration until the mid-20th century, ...Abstract · Sea Level During The Late... · Astronomic Constraints
  51. [51]
    Observation‐Driven Estimation of the Spatial Variability of 20th ...
    Mar 15, 2018 · The resulting map of 20th century regional sea level rise is optimized to agree with the tide gauge-measured trends, and provides an indication ...
  52. [52]
    Twentieth century climate change: Evidence from small glaciers
    Before the International Geophysical Year (1957–1959), fewer than 10 glaciers were observed. The number of measured glaciers has grown rapidly since the ...
  53. [53]
    More than 100 years of Arctic sea ice volume reconstructed with ...
    Aug 8, 2019 · A new paper now extends the estimate of Arctic sea ice volume back more than a century, to 1901. To do so it used both modern-day computer simulations and ...
  54. [54]
    Precipitation measurements and trends in the twentieth century - New
    Jan 28, 2002 · Precipitation gauge data indicate that global land precipitation (excluding Antarctica) has increased by about 9 mm over the twentieth century (a trend of 0.89 ...
  55. [55]
    Global precipitation trends in 1900–2005 from a reconstruction and ...
    Feb 15, 2013 · The mean trend is 0.13 mm day−1 per 100a over the oceans, 0.10 mm day−1 per 100a over the lands, and 0.13 mm day−1 per 100a globally. The global ...Introduction · Data · Trends in 20th Century · Future Projections: Trends in...
  56. [56]
    Science Briefs: Precipitation Trends in the 20th Century - NASA GISS
    We find that in much of the middle and high latitudes, precipitation has systematically increased over the 20th Century.
  57. [57]
    Data.GISS: GISS Surface Temperature Analysis (GISTEMP v4) - NASA
    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
  58. [58]
    HadCRUT5 - Met Office Hadley Centre observations datasets
    Sep 17, 2025 · HadCRUT5 is a gridded dataset of global historical surface temperature anomalies relative to a 1961-1990 reference period.
  59. [59]
    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.Annual Temperature Anomaly · Causes Of Warmth In 2024 · Comparisons With Other...Missing: 1900-2000 | Show results with:1900-2000
  60. [60]
    Data Overview - Berkeley Earth
    Berkeley Earth provides high-resolution land and ocean time series data, gridded temperature data, global monthly averages, and land-only data.Temperature City List · Temperature Country List · Asia · Results by RegionMissing: 20th | Show results with:20th
  61. [61]
    Global Temperature Anomalies from 1880 to 2024 - NASA SVS
    Jan 10, 2025 · In 2024, Earth's average surface temperature was the warmest on record, 2.30 degrees Fahrenheit (1.28 degrees Celsius) above the 1951-1980 ...
  62. [62]
    Global Temperature Report :: The University of Alabama in Huntsville
    GLobal Temperature Report. Earth System science Center THe University of Alabama in Huntsville. SEPTEMBER 2025 :: MAPS AND GRAPHS.Missing: 1979-2024 | Show results with:1979-2024
  63. [63]
    UAH v6.1 Global Temperature Update for November, 2024
    Dec 3, 2024 · The Version 6.1 global area-averaged temperature trend (January 1979 through November 2024) remains at +0.15 deg/ C/decade (+0.21 C/decade over ...
  64. [64]
    Temperature data (HadCRUT, CRUTEM, HadCRUT5, CRUTEM5 ...
    HadCRUT is a global temperature dataset, providing gridded temperature anomalies across the world as well as averages for the hemispheres and the globe as a ...
  65. [65]
    Can you explain the urban heat island effect? - NASA Science
    Oct 23, 2024 · While urban areas are typically warmer than the surrounding rural areas, the urban heat island effect doesn't significantly impact overall global warming.
  66. [66]
    Urban Heat Island warming effects related to population density are ...
    May 5, 2025 · A new research study from The University of Alabama in Huntsville addresses the question of how much urban areas have warmed from the Urban Heat Island (UHI) ...
  67. [67]
    An Updated Assessment of Near‐Surface Temperature Change ...
    Dec 15, 2020 · HadCRUT5 presents monthly average near-surface temperature anomalies, relative to the 1961–1990 period, on a regular 5° latitude by 5° longitude grid from 1850 ...1 Introduction · 3 Methods · 4 Results
  68. [68]
    GISTEMP: NASA Goddard Institute for Space Studies (GISS) Surface ...
    A comprehensive global surface temperature data set spanning 1880 to the present at monthly resolution, on a 2x2 degree latitude-longitude grid.
  69. [69]
  70. [70]
    [PDF] Reconstructing sea level rise at global 945 tide gauges since 1900
    Jun 24, 2025 · The average of reconstructed sea level across 945 tide gauges reveals a GMSL rise of 1.75±0.05 mm/yr over 1900-2020, and shows strong agreements ...
  71. [71]
    Sea Ice | National Snow and Ice Data Center
    However, it is important to note that the Arctic sea ice minimum extent has been consistently below 6.5 million square kilometers (2.5 million square miles) ...Science · Why it Matters · Quick Facts
  72. [72]
    Arctic sea ice sets a record low maximum in 2025
    Arctic sea ice extent appears to have reached its annual maximum on March 22, 2025. This is the lowest maximum in the 47-year satellite record.
  73. [73]
    2025 Arctic sea ice minimum squeezes into the ten lowest minimums
    Sep 17, 2025 · On September 10, Arctic sea ice likely reached its annual minimum extent of 4.60 million square kilometers (1.78 million square miles).
  74. [74]
    Antarctic Sea Ice #4: Record lows between 2022 and 2025
    Jun 19, 2025 · The most recent extremes in sea ice coverage were driven by ocean changes, most notably ongoing warming at around 100-200 m depth, combined ...
  75. [75]
    2024 Antarctic sea ice winter maximum second lowest on record
    Oct 8, 2024 · The sea ice in the Southern Ocean around Antarctica sea ice probably reached its winter maximum extent on September 19, 2024.
  76. [76]
    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.
  77. [77]
    Community estimate of global glacier mass changes from 2000 to ...
    Feb 19, 2025 · Since 2000, glaciers have lost between 2% and 39% of their ice regionally and about 5% globally. Glacier mass loss is about 18% larger than the ...Missing: 20th | Show results with:20th
  78. [78]
    Climate Change Indicators: Glaciers | US EPA
    On average, glaciers worldwide have been losing mass since at least the 1970s (see Figure 1), which in turn has contributed to observed changes in sea level ( ...
  79. [79]
    Glacier melt will unleash avalanche of cascading impacts
    Mar 21, 2025 · Five of the past six years have witnessed the most rapid glacier retreat on record. 2022-2024 witnessed the largest three-year loss of glacier ...
  80. [80]
    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 ...
  81. [81]
    Observed rainfall changes in the past century (1901–2019) over the ...
    Jan 21, 2021 · Changes in rainfall affect drinking water, river and surface runoff, soil moisture, groundwater reserve, electricity generation, agriculture ...
  82. [82]
    Anthropogenic amplification of precipitation variability over the past ...
    Jul 25, 2024 · From 1900 to 2020, day-to-day precipitation variability has increased over most land regions where observations are available according to the ...
  83. [83]
    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 ...
  84. [84]
    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.
  85. [85]
    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 ...
  86. [86]
    Don't Believe the Hype - by Roger Pielke Jr. - The Honest Broker
    Dec 19, 2022 · But to be sure, the prospect of human-caused climate change holds the potential for making extreme events more common or worse. For instance ...
  87. [87]
    Did Climate Change Do That? | Cato Institute
    The AR6 finds that extreme heat events like heatwaves are becoming more frequent and intense, and these trends can be attributed to human activities (likewise, ...
  88. [88]
    Billion-Dollar Weather and Climate Disasters | Time Series
    Visualize the frequency and cost of billion-dollar weather and climate events using the interactive time series. All Disasters. Drought. Flooding. Freeze.
  89. [89]
    Milankovitch cycles: What are they and how do they affect Earth?
    Jun 14, 2022 · Although Milankovitch cycles have nothing to do with the current climate change, they are believed to have dictated Earth's climate for millions ...What drives the Milankovitch... · Do Milankovitch cycles have...
  90. [90]
    Solar influence on climate during the past millennium - PNAS
    Overall, the magnitude of temperature response to solar irradiance changes is relatively small. Global average temperatures change by ≈0.066 ± 0.005°C for each ...
  91. [91]
    Changes in the Total Solar Irradiance and climatic effects
    Jul 22, 2021 · This would yield an increase of the terrestrial global temperature on ∆T⊕ = 0.02–0.05 °C, depending on the TSI amplitude of a particular solar ...
  92. [92]
    Volcanoes Can Affect Climate | U.S. Geological Survey - USGS.gov
    Volcanic gases like sulfur dioxide can cause global cooling, while volcanic carbon dioxide, a greenhouse gas, has the potential to promote global warming.
  93. [93]
    Effect of volcanic eruptions significantly underestimated in climate ...
    Jun 23, 2023 · Researchers have found that the cooling effect that volcanic eruptions have on Earth's surface temperature is likely underestimated by a factor of two.
  94. [94]
    Contribution of Atlantic and Pacific Multidecadal Variability to ...
    During most of the twentieth century, AMV dominates over PMV for the multidecadal internal variability imprint on global and Northern Hemisphere temperatures.
  95. [95]
    Imprint of the Atlantic multi-decadal oscillation and Pacific decadal ...
    Sep 5, 2013 · We find that in the early twentieth century the warming was dominated by a positive phase of the Atlantic multi-decadal oscillation (AMO) with minor ...
  96. [96]
    Carbon dioxide now more than 50% higher than pre-industrial levels
    Jun 3, 2022 · Prior to the Industrial Revolution, CO2 levels were consistently around 280 ppm for almost 6,000 years of human civilization. Since then, humans ...
  97. [97]
    Trends in CO2 - NOAA Global Monitoring Laboratory
    Sep 5, 2025 · The graphs show monthly mean carbon dioxide measured at Mauna Loa Observatory, Hawaii. The carbon dioxide data on Mauna Loa constitute the longest record of ...Data · Last 1 Year · Last Month · GlobalMissing: pre- industrial
  98. [98]
    How do we know the build-up of carbon dioxide in the atmosphere is ...
    Oct 12, 2022 · Fossil fuels are the only source large enough to cause the rapid CO2 increase, with a chemical fingerprint consistent with terrestrial plant ...
  99. [99]
    Changes to Carbon Isotopes in Atmospheric CO2 Over the Industrial ...
    Oct 23, 2020 · As fossil fuels are slightly depleted in 13C and entirely depleted in 14C, the burning of fossil fuels increases 12CO2 at a faster relative rate ...
  100. [100]
    Methane Emissions | US EPA
    Mar 31, 2025 · Human activities emitting methane include leaks from natural gas systems and the raising of livestock. Methane is also emitted by natural ...
  101. [101]
    Methane | Climate & Clean Air Coalition
    Over 60% of methane emissions come from human activity. ... Agriculture is the largest human source of methane emissions, responsible for 40%.
  102. [102]
    Deforestation and Climate Change
    Dec 10, 2024 · This forest loss produced roughly six percent of estimated global carbon dioxide emissions in 2023.
  103. [103]
    How halting deforestation can help counter the climate crisis - UNEP
    Jul 18, 2024 · The felling of trees in tropical areas alone releases more than 5.6 billion tonnes of planet-warming greenhouse gasses every year.
  104. [104]
    Aerosols: Small Particles with Big Climate Effects - NASA Science
    Jun 12, 2023 · Aerosols are small particles in the air that can either cool or warm the climate, depending on the type and color of the particle.Volcanoes Make Natural... · Air-Pollution Aerosols Also... · Soot Is a Dark-Colored...
  105. [105]
    Climate Impacts From a Removal of Anthropogenic Aerosol Emissions
    Removing aerosols induces a global mean surface heating of 0.5–1.1°C, and precipitation increase of 2.0–4.6%. Extreme weather indices also increase.
  106. [106]
    Explainer: How human-caused aerosols are 'masking' global warming
    Jun 10, 2025 · Aerosols affect the climate by absorbing or reflecting incoming sunlight, or by influencing the formation and brightness of clouds. Most ...
  107. [107]
    Contributions of Natural and Anthropogenic Forcing Agents to the ...
    We found that about half of the global warming is caused by the increase of WMGHGs (CO2, CH4, and N2O), while the decrease of the weakly absorbed (VIS) solar ...
  108. [108]
    Robust evidence for reversal of the trend in aerosol effective climate ...
    Sep 21, 2022 · Anthropogenic pollution particles, aerosols, exert an effective radiative forcing (ERF) on climate due to aerosol–radiation interactions ( ...
  109. [109]
    Chapter 3: Human Influence on the Climate System
    This chapter assesses the extent to which the climate system has been affected by human influence and to what extent climate models are able to simulate ...
  110. [110]
    Detection and attribution of climate change
    Detection and attribution studies can help evaluate whether model simulations are consistent with observed trends or other changes in the climate system.
  111. [111]
    Using Detection And Attribution To Quantify How Climate Change Is ...
    Dec 7, 2020 · We describe a set of formal methods, termed detection and attribution, used by climatologists to determine whether a climate trend or extreme event has changed.
  112. [112]
    Figure AR6 WG1 | Climate Change 2021: The Physical Science Basis
    Methods and systems used to test the attribution hypothesis or theory include: model-based fingerprinting; other model-based methods; evidence-based ...
  113. [113]
    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.
  114. [114]
    Overstating the effects of anthropogenic climate change? A critical ...
    Feb 8, 2023 · A critical assessment of attribution methods in climate science ... Climate change attribution: when is it appropriate to accept new methods?
  115. [115]
    The IPCC's attribution methodology is fundamentally flawed
    Aug 18, 2021 · Their “Optimal Fingerprinting” methodology on which they have long relied for attributing climate change to greenhouse gases is seriously flawed.
  116. [116]
    [PDF] History of climate modeling - University of Michigan Library
    The history of climate modeling begins with conceptual models, followed in the. 19th century by mathematical models of energy balance and radiative transfer ...
  117. [117]
    Timeline: The history of climate modelling - Carbon Brief
    Jan 16, 2018 · In the interactive timeline above, Carbon Brief charts more than 50 key moments in the history of climate modelling.
  118. [118]
    General Circulation Models of the Atmosphere
    Modeling techniques and entire GCMs spread by a variety of means. In the early days, as Phillips recalled, modelers had been like "a secret code society ...
  119. [119]
    [PDF] Climate Models An Assessment of Strengths and Limitations
    demonstrated by Norman Phillips in 1956. About that time, Joseph Smagorinsky started a program in climate modeling that ultimately be- came one of the most ...
  120. [120]
    [PDF] Thermal equilibrium of the atmosphere with a given distribution of ...
    As we explained in the introduction, the deviation from the fourth-power law is mainly due to the de- pendence of the effective source of outgoing radiation.
  121. [121]
    Weather forecasting and climate modelling: a short history
    Jun 19, 2024 · The first 'coupled' model, simulating how the atmosphere and ocean interact, was developed in 1969. Terrestrial globe with marked ocean currents ...Why future climate matters · Modelling global climate · What can we learn from...
  122. [122]
    The Fall and Rise of the Global Climate Model - AGU Journals - Wiley
    Aug 29, 2021 · Why do general circulation models overestimate the aerosol cloud lifetime effect?: A case study comparing CAM5 and a CRM. Atmospheric ...<|separator|>
  123. [123]
    Evaluating the Performance of Past Climate Model Projections
    Dec 4, 2019 · We find that climate models published over the past five decades were generally quite accurate in predicting global warming in the years after publication.Missing: milestones | Show results with:milestones
  124. [124]
    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 ...
  125. [125]
    The hindcast skill of the CMIP ensembles for the surface air ...
    Aug 28, 2012 · Surface air temperature trends from CMIP3 and 5 historical experiments are evaluated Some improvements in CMIP5 for regional and decadal ...
  126. [126]
    Comparison of Monthly Temperature Extremes Simulated by CMIP3 ...
    Comparisons with observations and reanalyses indicate that the models from both CMIP3 and CMIP5 perform well in simulating temperature extremes, which are ...
  127. [127]
    Performance-based sub-selection of CMIP6 models for impact ... - ESD
    Apr 21, 2023 · We have created a performance-based assessment of CMIP6 models for Europe that can be used to inform the sub-selection of models for this region.
  128. [128]
    Equilibrium Climate Sensitivity Estimated by Equilibrating Climate ...
    Nov 19, 2019 · 27 simulations with 15 climate models forced with a range of CO 2 concentrations show a median 17% larger equilibrium warming than estimated from the first 150 ...Estimating Equilibrium Climate... · Global Feedback Evolution · Implications
  129. [129]
    Chapter 7: The Earth's Energy Budget, Climate Feedbacks, and ...
    The equilibrium climate sensitivity, ECS (units: °C), is defined as the equilibrium value of ΔT in response to a sustained doubling of atmospheric CO 2 ...
  130. [130]
    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 ...
  131. [131]
    Revisiting a Constraint on Equilibrium Climate Sensitivity From a ...
    Oct 26, 2023 · Our central estimates of equilibrium climate sensitivity are 2.5–2.7 K, consistent with prior results and the IPCC's consensus estimate. 1 ...
  132. [132]
    [PDF] Climate Change 2021
    ... range of equilibrium climate sensitivity is between 2°C (high confidence) and 5°C (medium confidence). The AR6 assessed best estimate is 3°C with a likely ...
  133. [133]
    Observed trend in Earth energy imbalance may provide a ... - Science
    Jun 12, 2025 · We show that low climate sensitivity models fail to reproduce the trend in Earth energy imbalance, particularly in the individual longwave and shortwave ...
  134. [134]
    [PDF] Future Global Climate: Scenario-based Projections and Near-term ...
    This chapter covers future global climate projections, near-term changes, and projected changes in global climate indices in the 21st century.
  135. [135]
    How well have CMIP3, CMIP5 and CMIP6 future climate projections ...
    Jul 14, 2022 · The results show that CMIPs future temperature projections here investigated portray a future temperature increase over land areas well in line, ...
  136. [136]
    Factcheck: Why the recent 'acceleration' in global warming is what ...
    Apr 4, 2024 · The full CMIP6 ensemble of models has notably more warming than the IPCC-assessed warming projections. CMIP6 models, on average, warm by 0.29C ...
  137. [137]
    Opinion: Can uncertainty in climate sensitivity be narrowed further?
    Feb 29, 2024 · The distributions could nonetheless be narrowed in the future, particularly through better understanding of certain climate processes and paleoclimate proxies.
  138. [138]
    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.
  139. [139]
    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.
  140. [140]
    Impacts of ocean acidification on marine fauna and ecosystem ...
    We conclude that ocean acidification and the synergistic impacts of other anthropogenic stressors provide great potential for widespread changes to marine ...Abstract · Introduction · Effects of elevated pCO2 on... · Increased pCO2 and other...
  141. [141]
    NOAA confirms 4th global coral bleaching event
    Apr 15, 2024 · “Climate model predictions for coral reefs have been suggesting for years that bleaching impacts would increase in frequency and magnitude as ...Missing: attribution | Show results with:attribution
  142. [142]
    Coral‐bleaching responses to climate change across biological scales
    Climate change is increasing the frequency and intensity of coral‐bleaching events and is changing the composition, architectural complexity, and ...
  143. [143]
    Thawing permafrost poses environmental threat to thousands of ...
    Mar 28, 2023 · The effects of thawing permafrost, with all its consequences such as loss of hydrological barriers, improved hydrological connectivity, reduced ...
  144. [144]
    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) ...Missing: observations | Show results with:observations
  145. [145]
    Sea level rise - Project background - noaa/nesdis/star
    Current satellite altimeter observations show sea level rising at ~3 mm/yr, nearly twice as fast as the tide gauge measured rate over the past 100 years.
  146. [146]
    Climate Change Impacts on Ecosystems | US EPA
    Aug 8, 2025 · Climate change affects ecosystems in many ways. Climate controls how plants grow, how animals behave, which organisms thrive, and how they all interact with ...
  147. [147]
  148. [148]
    Review Climate change effects on biodiversity, ecosystems ...
    Sep 1, 2020 · Climate change is driving large-scale shifts in species distribution, abundance, and reorganization of terrestrial and aquatic ecosystems ( ...
  149. [149]
    A meta-analysis of the total economic impact of climate change
    The central estimate is more pessimistic. For 2.5 °C warming, the impact falls from −1.4% to −1.9% of income; for 5.0 °C, from −4.2% to ...
  150. [150]
    The global costs of extreme weather that are attributable to climate ...
    We find that US $ 143 billion per year of the costs of extreme events is attributable to climatic change. The majority (63%), of this is due to human loss of ...Missing: critiques | Show results with:critiques
  151. [151]
    [PDF] CHANGE - UC Berkeley Law
    2 Between 1980 and early 2025, climate-driven extreme events cost over $2.9 trillion nationwide.
  152. [152]
    Impacts of climate change on global agriculture accounting ... - Nature
    Jun 18, 2025 · We find that global impacts are dominated by losses to modern-day breadbaskets with favourable climates and limited present adaptation.
  153. [153]
    Health impacts of climate change on smallholder farmers
    Climate change impacts on smallholder farmers' health and agriculture may increase the migration of smallholder farmers [84,85].
  154. [154]
    Chapter 8: Poverty, Livelihoods and Sustainable Development
    Scientific evidence shows that climate change is accelerating livelihood transitions from rural agricultural production to urban wages (Cai et al., 2016; ...<|separator|>
  155. [155]
    Climate change - World Health Organization (WHO)
    Oct 12, 2023 · Climate change presents a fundamental threat to human health. It affects the physical environment as well as all aspects of both natural and human systems.Fast Facts on Climate Change... · The climate crisis · Alliance for TransformativeMissing: socioeconomic | Show results with:socioeconomic
  156. [156]
    Climate Change, Inequality, and Human Migration - Oxford Academic
    Dec 13, 2021 · We find that climate change strongly intensifies global inequality and poverty, reinforces urbanization, and boosts migration from low- to high-latitude areas.
  157. [157]
    [PDF] Report on the Impact of Climate Change on Migration
    2 When combined with physical, social, economic, and/or environmental vulnerabilities, climate change can undermine food, water, and economic security.
  158. [158]
    Evidence of climate and economic drivers affecting migration in an ...
    Oct 1, 2025 · Extreme weather events—including hurricanes, floods, droughts, and heatwaves—are becoming more frequent and severe due to climate change, ...
  159. [159]
    [PDF] OECD Global Long-Run Economic Scenarios: 2025 Update
    The reduction in global output associated with climate change, estimated to be approximately 1¾ per cent of global GDP today, rises to nearly. 9% by 2100. • ...
  160. [160]
    Researchers More Accurately Estimate Economic Impacts of Climate ...
    Oct 16, 2025 · The researchers find that empirical models ignoring convergence can calculate average global income losses of up to 19% by the end of the ...Missing: 2023-2025 | Show results with:2023-2025
  161. [161]
    The economic commitment of climate change | Nature
    Apr 17, 2024 · Here we use recent empirical findings from more than 1,600 regions worldwide over the past 40 years to project sub-national damages from ...
  162. [162]
    The Global Economic Impact of Climate Change: An Empirical ...
    Oct 10, 2025 · Here, I describe the landscape of empirical economic research on global impacts, I explain elements of modern analyses, I summarize recent ...Missing: costs | Show results with:costs
  163. [163]
    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 ...
  164. [164]
    CO2 is making Earth greener—for now - NASA Science
    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 ...
  165. [165]
    Crop yields have increased dramatically in recent decades, but ...
    Sep 30, 2024 · The study estimated that “CO2 fertilization over the 1961-2017 period has raised yields of C3 crops (rice and wheat) by 7.1% and of C4 crops ( ...
  166. [166]
    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 ...
  167. [167]
    5 Ways Climate Change Will Affect You: Crop Changes
    Climate change may actually benefit some plants by lengthening growing seasons and increasing carbon dioxide. ... Northern European potato farmers will see longer ...<|separator|>
  168. [168]
    Climate change leads to longer growing seasons and favours ...
    Jul 18, 2023 · Rising temperatures and a longer growing season may cause favourable growing areas to shift upwards.
  169. [169]
    Are there more cold deaths than heat deaths? - ScienceDirect.com
    Excess cold deaths far outnumber heat deaths, with a ratio of roughly 9:1, corresponding to about 4.6 million cold deaths and 489,000 heat deaths.
  170. [170]
    Global, regional, and national burden of mortality associated with ...
    Most excess deaths were linked to cold temperatures (8·52%), whereas fewer were linked to hot temperatures (0·91%). Globally, from 2000–03 to 2016–19, the cold- ...<|separator|>
  171. [171]
    Misinformed: Hyped heat deaths and ignored cold deaths
    Mar 4, 2025 · Rising temperatures have reduced net global death by 0.3 per cent, meaning some 166,000 deaths have been avoided.
  172. [172]
    The Future of the Northern Sea Route - A “Golden Waterway” or a ...
    Canadian and American maritime experts say two percent of global shipping could be diverted to the Arctic by 2030, reaching 5 percent by 2050. Experts cite a ...
  173. [173]
    Central Arctic Shipping Route
    The opening of these new Arctic shipping passages is expected to benefit northern coastal communities,resource-extraction operations, fisheries, and tourism ...
  174. [174]
    The Climate Epochs That Weren't - State of the Planet
    Jul 24, 2019 · That warming has been connected to improved crop yields in parts of Europe, and the temporary Viking occupation of Greenland.
  175. [175]
    Three Decades of Climate Mitigation Policy: What Has It Delivered?
    We find that mitigation policies have had a discernible impact on emissions and multiple emission drivers. Most notably, they have led to reductions in energy ...
  176. [176]
    Levelized Cost of Energy+ (LCOE+) - Lazard
    Lazard's Levelized Cost of Energy+ is a widely cited report that analyzes the cost competitiveness of renewables, energy storage, and system considerations.
  177. [177]
    [PDF] Lazard LCOE+ (June 2024)
    The results of our Levelized Cost of Energy (“LCOE”)analysis reinforce what we observe across the Power, Energy & Infrastructure Industry—sizable.
  178. [178]
    The cost of mitigation revisited | Nature Climate Change
    Nov 11, 2021 · We synthesize key factors influencing mitigation cost estimates to guide interpretation of estimates, for example from the Intergovernmental Panel on Climate ...
  179. [179]
    [PDF] Policy Implications of Net-Zero Emissions: A Multi-Model Analysis of ...
    Energy spending across the economy decreases relative to today for many models under reference and net-zero policies, especially as a share of GDP (Figure 13).
  180. [180]
    Climate adaptation vs mitigation: two sides of the same coin
    According to the Global Commission on Adaptation6, every USD 1 invested in adaptation could result in up to USD 10 in net economic benefits. ... The information ...
  181. [181]
    [PDF] A systematic global stocktake of evidence on human adaptation to ...
    Our synthesis of the resulting. 1,682 articles presents a systematic and comprehensive global stocktake of implemented human adaptation to climate change.
  182. [182]
    12 great examples of how countries are adapting to climate change
    Sep 17, 2019 · 12 great examples of how countries are adapting to climate change · The Thames Barrier · Risk-pooling in Africa · Nature-based solutions · Community ...
  183. [183]
    Climate change adaptation strategies and their predictors amongst ...
    The present study was conducted in Ambassel district of Northern Ethiopia to understand adaptation strategies employed by rural farmers to the adverse effects ...
  184. [184]
    "The Empirical Analysis of Climate Change Impacts and Adaptation ...
    This chapter reviews recent advancements in the analysis of climate change impacts and adaptation in agriculture with an emphasis on econometric and statistical ...Missing: evidence | Show results with:evidence
  185. [185]
    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 ...
  186. [186]
    Scaling Up of Nature-Based Solutions to Guide Climate Adaptation ...
    Apr 11, 2022 · This paper addresses this lack by providing examples of two case studies, one in the Netherlands (Amersfoort) and one in South Africa (Lephalale) ...
  187. [187]
    [PDF] U.S. Environmental Protection Agency Climate Change Adaptation ...
    Based on lessons learned about the most effective climate change adaptation strategies, EPA can make adjustments to the way adaptation is integrated into its ...
  188. [188]
    Assessing the costs and benefits of climate change adaptation
    Mar 3, 2023 · In general, adaptation costs are easier to quantify than the benefits of adaptation or the costs of inaction. Thus, the costs of adaptation ...
  189. [189]
    Adapting to the Weather: Lessons from U.S. History - PMC
    For example, new technologies (including plant varieties) and economic conditions might reduce the negative effects of drought and extreme temperature on crop ...
  190. [190]
    Enabling Climate Change Adaptation in Coastal Systems: A ...
    Aug 16, 2023 · This paper provides insight into 650 peer-reviewed empirical research studies on coastal climate adaptation from the past two decades.
  191. [191]
    United Nations Framework Convention on Climate Change | UNFCCC
    The United Nations Framework Convention on Climate Change (the Convention or UNFCCC) was adopted at the United Nations Headquarters, New York on the 9 May 1992.
  192. [192]
    History of the Convention - UNFCCC
    The UNFCCC was established in 1992, the Kyoto Protocol was adopted in 1997, and the Paris Agreement in 2015, with the Kyoto Protocol entering into force in ...
  193. [193]
    The Kyoto Protocol | UNFCCC
    Overall, these targets add up to an average 5 per cent emission reduction compared to 1990 levels over the five year period 2008–2012 (the first commitment ...
  194. [194]
    Analyzing the effectiveness of international environmental policies
    Results show that the protocol was successful in reducing the emissions of the ratifying countries approximately by 7% below the emissions expected under a “No ...<|separator|>
  195. [195]
    The Paris Agreement | UNFCCC
    It entered into force on 4 November 2016. Its overarching goal is to hold “the increase in the global average temperature to well below 2°C above pre-industrial ...Paris Agreement Work · Nationally Determined · Report of the Conference · Article 6
  196. [196]
    Emissions Gap Report 2023 | UNEP - UN Environment Programme
    Nov 20, 2023 · The report looks at how stronger implementation can increase the chances of the next round of NDCs, due in 2025, bringing down greenhouse gas ...
  197. [197]
    Global Climate Agreements: Successes and Failures
    Through the Kyoto Protocol and Paris Agreement, countries agreed to reduce greenhouse gas emissions, but the amount of carbon dioxide in the atmosphere keeps ...
  198. [198]
    Success or failure? The Kyoto Protocol's troubled legacy - Foresight
    Dec 8, 2022 · Even papers that see an emissions reduction effect due to the Kyoto Protocol see relatively small gains. The reduction goals were simply too ...Missing: compliance | Show results with:compliance
  199. [199]
    Quantifying the consensus on anthropogenic global warming in the ...
    A claim has been that 97% of the scientific literature endorses anthropogenic climate change (Cook et al., 2013. Environ. Res. Lett. 8, 024024).
  200. [200]
    Surface versus satellite; the temperature data set controversy
    Feb 10, 2016 · It's the basis of the dispute over which year was really the warmest on record, whether the Earth has warmed by 0.59°C since 1979, as indicated ...<|separator|>
  201. [201]
    Urban Heat Island Effects in U.S. Summer Surface Temperature ...
    A novel method is described for quantifying average urban heat island (UHI) warming since 1895 in the contiguous United States (CONUS) summer air temperature ...Abstract · Introduction · Data and methods · Results
  202. [202]
    Evidence of Urban Blending in Homogenized Temperature Records ...
    We show that the homogenization process unintentionally introduces new nonclimatic biases into the data as a result of an “urban blending” problem.
  203. [203]
    Causes of Higher Climate Sensitivity in CMIP6 Models - Zelinka - 2020
    Jan 3, 2020 · Climate sensitivity is larger on average in CMIP6 than in CMIP5 due mostly to a stronger positive low cloud feedback · This is due to greater ...
  204. [204]
    Use of 'too hot' climate models exaggerates impacts of global warming
    May 4, 2022 · Researchers should no longer simply use the average of all the climate model projections, which can result in global temperatures by 2100 up to 0.7°C warmer ...
  205. [205]
    Climate model trend errors are evident in seasonal forecasts at short ...
    Nov 20, 2024 · Previous studies have argued that the rapid mean error development in forecasts may be used to help identify and correct deficiencies with model ...
  206. [206]
    3 apocalyptic climate change predictions that failed to come true
    Apr 16, 2025 · Myth 1: The Arctic will soon be ice-free. It "could already be ice-free by the summer of 2030," shrieks a DW News report.
  207. [207]
    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 ...
  208. [208]
    Role of solar activity and Pacific decadal oscillation in the ...
    In addition to its direct interaction with climate, the PDO is closely associated with solar activity and affects the regional climatic response to radiation ( ...<|control11|><|separator|>
  209. [209]
    [PDF] Climate Change Is Not an Apocalyptic Threat—Let's Address It Smartly
    Bjorn Lomborg researches the smartest ways to do good. With his think tank, the Copenhagen Consensus, he has worked with hundreds of the world's top economists ...
  210. [210]
    Increasing development, reducing inequality, the impact of climate ...
    The Paris Agreement, if fully implemented, will cost $819–$1,890 billion per year in 2030, yet will reduce emissions by just 1% of what is needed to limit ...
  211. [211]
    Global climate finance hits $1.9 trillion—bridging the climate ...
    Jun 24, 2025 · Mitigation finance more than doubled from USD 757 billion in 2018 to USD 1.78 trillion in 2023, driven by private investment in clean energy ...
  212. [212]
    Mitigation efforts to reduce carbon dioxide emissions and meet the ...
    Oct 17, 2025 · In 2015–2024, total CO2 emissions actually increased by 5.6%, an annual rate of 0.6%. This is because world GDP increased faster than carbon ...
  213. [213]
    Does Mitigation ODA Reduce Emissions?
    Feb 20, 2024 · The short answer from the analysis is that there is no obvious association between mitigation spending and lower emissions. ... The World Bank and ...
  214. [214]
    Pre-Judging Paris by Bjørn Lomborg - Project Syndicate
    Nov 17, 2015 · Our approach to climate change is broken. The Paris agreement will likely cost the world at least a trillion dollars each year, yet deliver ...
  215. [215]
  216. [216]
    The consequences of non-participation in the Paris Agreement
    We find that non-participation of the US would eliminate more than a third of the world emissions reduction (31.8% direct effect and 6.4% leakage effect).
  217. [217]
    Climate Change & Energy - Copenhagen Consensus Center
    Climate change from 1900 to 2025 has mostly been a net benefit, rising to increase welfare about 1.5% GDP per year.
  218. [218]
  219. [219]
    Is there a climate change reporting bias? A case study of English ...
    Sep 22, 2022 · This study's results highlight a huge reporting bias in favour of storms and wildfires in the news media. This has a material cost, where storms ...
  220. [220]
    How Media Bias Caused the Moral Panic Surrounding Climate ...
    Jul 20, 2021 · the media now routinely attributes unusual weather events—heat waves, fires, floods, tornados, or hurricanes—to humanity's insufficiently ...<|separator|>
  221. [221]
    The Inconvenient Facts the Media Ignore About Climate Change
    Feb 26, 2016 · Perhaps one of the worst examples of one-sided, biased reporting involves global warming. Those who reject the liberal viewpoint that climate ...
  222. [222]
    Environmental Nonprofit Organizations and Public Opinion on ...
    Jun 13, 2024 · The results show that the overall presence of ENPOs moderately, positively correlates with public opinions on global warming, with advocacy- ...
  223. [223]
    Engaging non-governmental organizations in global climate ...
    In recent decades, NGOs have gained significant influence in the area of global climate governance. Although global efforts have been made to tackle climate ...3 Research Results · 3.3 Research Design · 3.5 Ngo StrategiesMissing: opinion | Show results with:opinion
  224. [224]
    Full article: The Impact of INGO Climate Shaming on National Laws
    We argue that INGO climate shaming can incur reputational costs for governments through two main pathways: public opinion and transnational politics.
  225. [225]
    New Report Reveals Media's Climate Alarmism Irks Readers | IWF
    Nov 30, 2021 · A new paper found climate alarmism in the media turns people off and makes them less engaged in environmental advocacy.<|separator|>
  226. [226]
    The growing divide in media coverage of climate change | Brookings
    Jul 24, 2024 · On any given day in 2011, there was a 30% chance that heartland news outlets would cover climate change more than elite newspapers.
  227. [227]
    Hegemonic stories in environmental advocacy testimonials
    Several environmental advocacy organizations have emerged that use narrative persuasion techniques to change climate change opinions and overcome climate ...
  228. [228]
    Advocacy for Climate Change Matters: We Can and We Should
    Aug 3, 2023 · Advocates must skillfully navigate the complex policy landscape and power dynamics surrounding climate change narratives. Engaging with decision ...