Fact-checked by Grok 2 weeks ago

Climate variability and change

Climate variability and change refer to fluctuations in climate elements such as temperature, precipitation, and wind patterns that deviate from long-term averages, as well as persistent shifts in those averages, occurring over timescales from seasons to millennia due to internal atmospheric-ocean dynamics and external forcings including solar variations, volcanic eruptions, orbital changes, and anthropogenic greenhouse gas emissions. Natural variability dominates shorter-term changes, exemplified by oscillations like the El Niño-Southern Oscillation (ENSO) that redistribute heat globally, while longer cycles such as Milankovitch orbital forcings have orchestrated alternations over 100,000-year periods by altering solar insolation distribution. Instrumental records indicate global average surface temperatures have increased by about 1.1°C since 1850, with roughly two-thirds of this rise post-1975, corroborated across datasets like HadCRUT and Berkeley Earth, though effects and data homogenization methods introduce uncertainties in precise quantification. Anthropogenic CO2 emissions, rising from pre-industrial ~280 ppm to over 420 ppm, enhance the by trapping outgoing infrared radiation, as evidenced by satellite spectral measurements, yet the net to doubled CO2—estimated at 1.5–4.5°C in mainstream models—remains empirically contested, with observed warming rates falling below many projections and natural factors like solar activity and ocean cycles contributing substantially to recent trends. Controversies persist over attribution, as institutional narratives often emphasize human dominance while downplaying natural variability's role and discrepancies between model hindcasts and satellite-era observations, underscoring the need for skepticism toward sources with evident incentives for alarmist framing.

Terminology and Concepts

Definitions of Variability and Change

variability encompasses fluctuations in statistics, such as averages, extremes, and probabilities of , , and other elements, occurring over timescales from seasons to several decades, distinct from individual events. These variations arise primarily from internal atmospheric and oceanic processes, including phenomena like the El Niño-Southern Oscillation (ENSO), which redistribute heat and moisture globally, leading to temporary deviations from long-term norms without implying a permanent alteration in the 's baseline state. For instance, ENSO events can cause widespread or flooding on interannual scales, but they typically revert toward the mean over time. In contrast, climate change denotes a long-term, statistically detectable shift in the mean state of the climate or its variability, persisting for periods of decades or longer, identifiable through methods like or statistical tests. This definition, as articulated by the (IPCC), encompasses changes attributable to both natural external forcings—such as variations or volcanic eruptions—and factors, including alterations to atmospheric composition from human emissions. Unlike variability, which oscillates around a stable reference mean, climate change involves directional trends, such as sustained increases in global mean surface temperature observed since the late , exceeding natural variability thresholds in multiple datasets. The demarcation between variability and change hinges on duration, persistence, and : short-term anomalies (e.g., multi-year ENSO cycles) represent variability if they do not alter the underlying distribution, whereas prolonged shifts, like the warming or recent warming, qualify as change when they exceed historical ranges or demonstrate non-reverting trends. This distinction is crucial for attribution studies, as natural variability can modulate or mask underlying change signals, requiring separation via modeling and paleoclimate proxies to discern causal drivers. Sources emphasizing only human-induced aspects of change, such as certain policy-oriented reports, may understate natural precedents, but empirical records confirm has undergone multiple natural shifts over millennia, independent of human influence.

Key Metrics and Indicators

anomalies serve as a primary indicator of climate variability and change, with datasets from and NOAA showing the 2024 annual average at approximately 1.28°C above the 20th-century baseline, marking it as the warmest year on record. This warming is modulated by natural oscillations like El Niño-Southern Oscillation (ENSO), which contributed to elevated temperatures in 2023-2024, though the long-term trend since 1880 reflects a rise of about 1.1°C. For 2025 through August, global temperatures ranked second highest, trailing only 2024, amid a transition to neutral ENSO conditions. Atmospheric carbon dioxide (CO₂) concentrations, measured at , reached a monthly average of 425.48 in 2025, continuing an upward trajectory from 315 in 1958, with annual increases accelerating due to emissions. The May 2025 peak hit 430.2 , the second-largest year-over-year jump in the 67-year record, underscoring persistent despite natural sinks absorbing roughly half of emissions. Global mean has risen at an average rate of 3.4 mm per year since 1993, as tracked by satellite altimetry, totaling 8-9 inches (21-24 cm) since 1880, driven by and land ice melt. In 2024, the rate accelerated beyond expectations, linked to extreme warming, reaching a record high of 101.4 mm above 1993 levels by year-end. Upper heat content has increased steadily since the 1970s, absorbing over 90% of Earth's excess energy, with the top 2000 meters gaining heat at rates exceeding prior decades, as evidenced by float and ship-based measurements. This accumulation, equivalent to roughly four bombs per second in recent years, reflects both radiative imbalance and ocean circulation changes. Arctic sea ice minimum extent reached 4.60 million square kilometers in September 2025, tying for the 10th lowest in the satellite era (since 1979), with a long-term decline of 12.2% per decade amid reduced summer melt onset and extent. Variability from weather patterns and ENSO influences annual minima, but the trend indicates diminished ice volume and thickness. Other indicators include shifts in precipitation extremes and drought indices, such as the U.S. Climate Extremes Index, which tracks departures in , , and drought, showing increased frequency of extremes in recent decades, though regional patterns vary due to internal variability. These metrics, derived from instrumental records and proxies, quantify changes against natural baselines but require accounting for data adjustments and urban heat effects in surface observations.

Natural Drivers

Internal Climate Variability


Internal climate variability refers to natural fluctuations in the arising from chaotic internal dynamics and interactions among components such as the atmosphere, oceans, land surface, and , without requiring external forcings. These processes generate preferred spatial patterns of variability, known as modes, that operate on timescales from intraseasonal to multidecadal and contribute substantially to climate predictability on subseasonal to decadal horizons. Internal variability is generally larger in the extratropics than and stronger in winter than summer, influencing regional patterns and masking forced trends in short-term observations.
The El Niño-Southern Oscillation (ENSO) represents a dominant interannual mode, with cycles typically lasting 2-7 years, driven by coupled ocean-atmosphere interactions in the tropical Pacific. During El Niño phases, enhanced easterly weaken, leading to warmer sea surface temperatures (SSTs) in the central-eastern Pacific, which shift circulation and generate teleconnections affecting and temperature; for instance, El Niño events have been linked to increased hurricane activity in the Pacific and droughts in . La Niña phases, conversely, feature cooler SSTs and strengthened trades, often resulting in opposite impacts, such as floods in and drier conditions in the southern U.S. ENSO variability accounts for much of the year-to-year temperature fluctuations, with strong events capable of altering annual averages by up to 0.15°C. Decadal to multidecadal modes include the (PDO), characterized by SST anomalies in the North Pacific north of 20°N, with phases shifting roughly every 20-30 years. Positive PDO phases feature cooling in the central North Pacific and warming along the coasts, impacting marine ecosystems, such as reduced salmon catches during cool phases, and modulating n precipitation patterns. The PDO enhances El Niño teleconnections to during positive phases, amplifying winter precipitation in the southwestern U.S. and diminishing La Niña effects. The Atlantic Multidecadal Oscillation (AMO) exhibits variability in the North Atlantic on 60-80 year cycles, with warm phases since the mid-1990s linked to heightened frequency and intensified multidecadal ENSO variability through alterations in the annual cycle. AMO warm periods increase risk in the U.S. Southwest and north-central regions, while also influencing Sahel rainfall; for example, the positive AMO phase correlates with stronger Pacific ITCZ variability. These oscillations arise from internal ocean circulation changes, such as variations in the Atlantic Meridional Overturning Circulation, rather than direct . Internal modes like ENSO, PDO, and AMO introduce substantial uncertainty in attributing regional climate changes, as their amplitudes can rival forcing on decadal scales, particularly over land areas where variability is amplified. For instance, natural decadal variations may equal global warming-induced surface air temperature changes in magnitude for regional predictions. While models simulate these modes, debates persist on their exact internal origins versus subtle external influences, with some analyses questioning robust multidecadal oscillations in unforced simulations. Nonetheless, observational and modeling evidence confirms their role in generating low-frequency variability that overlays long-term trends.

External Natural Forcings

External natural forcings encompass variations in output, volcanic aerosol emissions, and Earth's orbital parameters that alter the planetary energy balance independently of internal climate dynamics. These factors have driven significant climate shifts across geological timescales, with and volcanic influences affecting decadal to centennial variability, while orbital changes operate on millennial scales. Empirical reconstructions indicate their radiative forcings are typically small in the compared to pre-industrial baselines but have modulated past s through direct energetic imbalances. Solar irradiance variations, tracked via sunspot cycles and proxy records, impose a of about ±0.17 W/m² over the 11-year Schwabe cycle, stemming from a 0.1% fluctuation in total . This equates to global temperature responses of roughly 0.1°C, as observed in measurements since 1978 showing no net upward trend in irradiance amid rising temperatures. Longer-term solar minima, such as the (1645–1715), correlated with cooler European winters, but reconstructions attribute only modest of 0.1–0.3°C, underscoring limited efficacy for explaining centennial-scale changes without amplification mechanisms like cosmic ray-induced cloud cover, which remain debated. Orbital forcings, formalized by Milankovitch, arise from (100,000-year cycle modulating perihelion distance), obliquity (41,000-year axial tilt variation affecting seasonal contrast), and (23,000-year wobble shifting solstice timing). These redistribute insolation, with high-latitude summer peaks driving glacial terminations; for instance, obliquity maxima increase insolation by up to 50 W/m² at 65°N, initiating deglaciations as seen in oxygen records from ice cores. In the , declining obliquity has favored a gradual cooling trajectory, projecting multi-millennial declines of 0.5–1°C per 5,000 years absent countervailing influences, incompatible with 20th-century warming patterns. Volcanic forcings occur via stratospheric aerosol veils from sulfur-rich eruptions, scattering incoming shortwave and yielding negative forcings of 1–5 W/m² for 1–3 years. The 1991 eruption exemplifies this, injecting 20 million tons of SO₂ and inducing 0.5°C detectable in surface temperatures. Larger historical events, like the 1815 Tambora blast, amplified aerosol lifetimes through self-lofting, but aggregate volcanic forcing over the nets near-zero due to episodic nature, contributing to short-term dips amid longer trends rather than sustained directional change.

Anthropogenic Factors

Greenhouse Gas Contributions

Anthropogenic arise predominantly from combustion, industrial processes, agriculture, waste management, and land-use changes such as . These activities have elevated atmospheric concentrations of long-lived es, with (CO2), methane (CH4), and nitrous oxide (N2O) comprising the majority. , though minor in volume, possess high potentials. Global anthropogenic GHG emissions reached 52.9 GtCO2e in , reflecting a 62% increase since 1990. CO2 accounts for approximately 74-76% of total GHG emissions in CO2-equivalent terms, primarily from fossil fuel oxidation in production, transportation, and industry, supplemented by cement manufacturing and net land-use emissions. -related CO2 emissions totaled about 37 Gt in 2023, with coal, oil, and as key contributors; land-use changes, including , added roughly 4-5 GtCO2 annually in recent years, though net land sinks partially offset this.
Greenhouse GasApproximate Share of Total Anthropogenic Emissions (CO2e, %)Primary Anthropogenic Sources
CO274.5 (fossil) + land-use contributionsFossil fuel combustion, production,
CH417.9 in livestock, rice cultivation, fossil fuel extraction and leaks, landfills
N2O4Agricultural (fertilizers), ,
Fluorinated gases~2-3
Methane emissions, equivalent to about 18% of total GHGs, originate largely from biological processes in (roughly 40% of CH4) and systems (30-35%), with the from decomposition. , contributing around 4-6%, is chiefly from application and , enhancing emissions. These breakdowns, derived from inventories like those in IPCC AR6 and , underscore and as dominant sectors, though data uncertainties persist in land-use fluxes and emissions.

Other Human Influences

Anthropogenic aerosols, primarily from emissions, biomass burning, and transportation, exert a net cooling effect on the through direct of solar radiation and indirect modification of cloud properties, with effective estimated at -1.1 W/m² (range -1.7 to -0.4 W/m²) from 1750 to 2019. This forcing offsets a substantial portion of warming, particularly in the , and contributes to regional patterns such as dimming in aerosol-heavy areas like . The spatial distribution of absorbing aerosols, such as , can amplify warming in specific locations like the by reducing surface upon deposition on snow and ice. Land use and land cover changes, including , agricultural expansion, and , alter surface , , and roughness, generating a of approximately -0.2 to -0.5 W/m² globally since pre-industrial times, with stronger regional effects in deforested where reduced lowers flux and enhances sensible heating. These modifications have contributed to amplified warming over land relative to oceans, as bare and crops reflect more than forests, though in some areas provides a counteracting cooling via increased . Peer-reviewed assessments indicate that such changes explain part of the divergence in hemispheric temperature trends, with land alterations exacerbating local variability. The urban heat island effect, arising from impervious surfaces, reduced vegetation, and heat emissions in cities, elevates local temperatures by 1-3°C on average compared to rural surroundings, but its influence on global mean surface temperature trends is minimal, contributing less than 0.1°C to century-scale records due to sparse urban coverage and data homogenization procedures. Analyses of metadata and rural comparisons confirm that adjustments for urbanization bias account for potential overestimation, with ocean and remote land data—unaffected by UHI—dominating global averages. Some estimates suggest UHI may explain up to 25% of land-based trends in certain datasets, though this remains contested and does not alter the overall signal. Other influences, such as tropospheric increases from precursors like oxides, add a warming forcing of about 0.4 W/m² since 1750, while stratospheric induces slight cooling, and contrails enhance cover for transient positive forcing. These short-lived climate forcers collectively rival impacts on patterns and regional variability, underscoring the multifaceted nature of human perturbations beyond long-lived gases.

Observational Data

Instrumental Measurements

Instrumental measurements of climate variables, particularly surface air temperature, precipitation, and , provide direct quantitative data on recent climate variability and change, with systematic records emerging in the for localized regions. The longest continuous series, the record, dates to 1659, capturing regional fluctuations driven by natural patterns. Global-scale instrumental data, however, rely on networks of land stations, ship logs, and buoys, achieving sufficient coverage for hemispheric estimates around 1850, though early records were biased toward the Northern Hemisphere's land areas, limiting precision for global averages. These measurements reveal decadal-scale oscillations, such as the early 20th-century warming and mid-century plateau, alongside a net rise of approximately 1.1°C from the 1850–1900 baseline to 2020, as compiled in independent datasets. Major surface temperature datasets include NOAAGlobalTemp, GISTEMP (NASA), and HadCRUT5 (UK Met Office), each integrating thousands of station readings with sea surface temperatures from buckets and engine intakes pre-1940s, transitioning to buoys and satellites later. These records indicate an accelerated warming rate of about 0.06°C per decade since 1850, intensifying to 0.2°C per decade post-1970, though spatial gaps in polar and oceanic regions require infilling via statistical methods, introducing uncertainties estimated at ±0.05°C for annual global means. Data homogenization adjusts for non-climatic biases like station relocations, instrument changes, and time-of-observation shifts, which raw data analyses show would otherwise underestimate recent warming by 10–20% in some periods; however, critics contend that iterative adjustments, often cooling pre-1950 records more than warming post-1950 ones, amplify trends beyond raw observations, particularly in datasets reliant on urban stations. Urban heat island effects, where concrete and asphalt elevate local readings by 1–2°C in cities, contribute up to 22% of observed U.S. warming in raw data, with mainstream adjustments claiming near-complete mitigation via rural baselines, though residual biases persist in global composites due to increasing station urbanization. Satellite-based microwave sounding units (MSU/AMSU), operational since December 1978, measure lower tropospheric temperatures over land and , offering uniform coverage without surface-specific biases. The (UAH) dataset reports a warming trend of 0.14°C per through 2024, while Remote Sensing Systems () shows 0.21°C per , both lower than surface estimates due to adjustments for , sensor drift, and diurnal drift, with UAH emphasizing lower tropospheric homogeneity. These records highlight tropospheric amplification of surface warming in the but divergences, such as slower polar trends, challenging models expecting uniform vertical profiles under forcing. Ocean heat content measurements advanced with the array of over 3,800 profiling floats deployed since 2000, providing and temperature profiles to 2,000 meters with 0.002°C accuracy, revealing upper- warming of 0.1–0.2°C since 2004 and absorbing over 90% of excess heat inferred from surface trends. Pre-ARGO ship-based , prone to warming biases, underestimated uptake, but ARGO's near-real-time confirm accelerating heat storage below 700 meters, though coverage gaps in marginal seas and deep persist. records, from gauges since the , show increased variability with 1–2% per decade global rise since 1950, but inhomogeneities from gauge undercatch in wind and siting changes complicate attribution. Overall, 's strength lies in empirical directness, yet reliance on adjusted composites underscores the need for transparency to resolve ongoing debates over trend magnitudes.

Proxy Records

Proxy records consist of natural archives that indirectly preserve evidence of past environmental conditions, enabling of variability prior to instrumental observations. These proxies include tree rings, s, sediment layers, coral skeletons, and speleothems, each responding to climatic variables such as , , and atmospheric composition through physical, chemical, or biological mechanisms. Reconstructions derive from calibrating proxy signals against modern data where overlapping periods exist, often using statistical models to infer quantitative estimates. Ice cores from and provide high-resolution records extending back hundreds of thousands of years. Oxygen isotope ratios (δ¹⁸O) and deuterium (δD) in ice reflect past temperatures, with the core documenting orbital-scale cycles where interglacial warmth exceeded glacial minima by 8–10°C in , accompanied by CO₂ variations from 180 ppm to 280 ppm trapped in bubbles. cores reveal abrupt shifts, such as Dansgaard-Oeschger events with warming of 5–10°C over decades during the (circa 115,000–11,700 years ago). These records highlight rapid variability superimposed on longer-term trends driven by orbital forcings. Tree rings offer annual resolution for the past millennium or more, primarily in temperate and zones. Ring width and maximum latewood density correlate with summer temperatures and drought stress; networks reconstruct hemispheric averages showing cooler conditions during the (circa 1450–1850 CE) by 0.5–1°C relative to the 20th century in some datasets. However, species-specific responses, such as divergence in recent decades where ring growth fails to track warming, introduce calibration challenges. Marine and lake sediments extend records to millions of years, using microfossils like for sea surface temperatures via Mg/Ca ratios or alkenone indices, and for terrestrial vegetation shifts indicative of or . Ocean sediment cores confirm mid-Holocene warmth (circa 9,000–5,000 years ago) exceeding current levels in subtropical regions by 1–2°C, linked to orbital enhancing summer insolation. Varved lake sediments preserve seasonal layers for decadal variability. Corals and speleothems provide tropical and cave-based proxies; δ¹⁸O records El Niño-Southern Oscillation variability over centuries, while growth rates and isotopes track strength. These subtropical records show asynchronous regional responses, such as enhanced Indian during the early . Uncertainties in proxy records arise from dating imprecision (e.g., radiocarbon errors beyond 50,000 years), proxy-specific sensitivities to non-climatic factors like biological in trees or effects in isotopes, and sparse global coverage leading to regional biases in syntheses. Multi-proxy ensembles mitigate single-record limitations but require assumptions in weighting and error propagation, with pre-1850 global temperature estimates carrying uncertainties of ±0.5°C or more.

Data Quality and Adjustments

Adjustments to instrumental temperature records aim to correct for non-climatic influences such as changes in times, station relocations, upgrades, and local land-use effects, using algorithms like pairwise homogenization to detect and mitigate discontinuities. These processes, applied by agencies including NOAA and the UK Met Office for HadCRUT, seek to produce homogeneous series reflecting true climatic signals, but their validity depends on accurate break detection and correction direction. Raw surface data often exhibits biases from poor station siting; a comprehensive survey of the U.S. Historical Climatology Network (USHCN) found approximately 70% of stations classified as poorly sited (CRN classes 3-5), located near asphalt, air conditioning exhausts, or urban infrastructure, which can inflate readings via localized heating. Such exposure correlates with increased temperature variance and potentially overstated warming trends, with one analysis estimating a 32% upward bias in raw U.S. trends from substandard siting alone. Homogenization adjustments frequently cool pre-1950 more than recent ones, amplifying post-industrial warming trends; for example, NOAA's methods have been shown to increase the apparent U.S. warming rate by correcting for time-of-observation biases (e.g., shifts from afternoon to morning readings, which artificially lower later minima in ). Globally, versus adjusted comparisons in HadCRUT-like datasets indicate adjustments boost the 1950-2016 warming rate by roughly 10%, though proponents argue this aligns with pristine references like the U.S. Climate Reference Network (USCRN). Critiques highlight inconsistencies in adjustment efficacy; a peer-reviewed of NOAA's Global Historical Climatology Network (GHCN) European subset revealed frequent errors in break identification, with over half of corrections applying the wrong sign, resulting in trends up to 50% higher than unadjusted data in some regions. (UHI) effects, inadequately filtered in global land records, may contribute 20-25% to observed trends, as rural-pristine subsets show subdued warming compared to urban-influenced series. Satellite-derived records (e.g., UAH, microwave sounding units) bypass surface siting issues but require their own and adjustments, yielding lower tropospheric trends of 0.13-0.17°C per decade since 1979, versus 0.18°C per decade in adjusted surface datasets, underscoring unresolved discrepancies between measurement domains. Pristine networks like USCRN, operational since with ideal siting, report U.S. trends approximately half those of adjusted USHCN historical data, suggesting potential overestimation in homogenized long-term series from legacy biases.

Historical Climate Variations

Paleoclimate Evidence

Paleoclimate evidence, derived from natural archives such as ice cores, marine sediments, tree rings, and corals, documents substantial climate variability over millions of years, including repeated glacial-interglacial cycles and shifts between icehouse and greenhouse states, primarily driven by orbital forcings, tectonic changes, and solar variations. These proxies preserve chemical, isotopic, and physical indicators of past temperatures, precipitation, and atmospheric composition, enabling reconstructions of global and regional conditions with varying resolutions from decadal to millennial scales. Ice cores from and provide high-resolution records extending back hundreds of thousands of years, revealing synchronous variations in temperature and greenhouse gases during Quaternary glacial-interglacial transitions. The ice core, spanning 420,000 years, records eight such cycles with deuterium isotope temperatures fluctuating by 8–10°C and atmospheric CO₂ concentrations ranging from 180 to 300 ppm, while varied between 320 and 790 ppb. The longer EPICA Dome C core extends this to 800,000 years, confirming CO₂ levels remained below 300 ppm throughout, with no excursions approaching modern values of over 420 ppm until the Industrial era. In these records, temperature proxies consistently lead CO₂ increases by 800–1,300 years at the onset of deglaciations, suggesting initial orbital-driven warming mobilizes carbon from oceans and , with subsequent CO₂ release amplifying the temperature rise through effects. This lag underscores the role of —variations in Earth's (period ~100,000 years), axial tilt (obliquity, ~41,000 years), and (~23,000 years)—as primary pacemakers of cycles, modulating seasonal insolation contrasts that build or erode continental ice sheets. of oxygen ratios in benthic from ocean sediments aligns closely with these orbital periodicities, providing robust evidence for their causal influence on global ice volume changes. Over longer timescales, cores and proxies indicate more extreme variability, such as during the Eocene epoch (56–34 million years ago), when global mean temperatures exceeded modern levels by 10–15°C, polar regions supported temperate forests without permanent ice caps, and deep ocean temperatures reached 12°C, associated with CO₂ concentrations estimated above 1,000 ppm from stomatal indices and boron isotopes. The Paleocene-Eocene Thermal Maximum (PETM) at ~56 million years ago exemplifies rapid natural warming, with a 5–8°C global increase over ~10,000–20,000 years linked to massive carbon releases from volcanic or methane hydrate sources, disrupting ocean circulation and ecosystems as evidenced by benthic foraminiferal extinctions and carbon isotope excursions in sediments. Tree-ring chronologies and speleothems extend high-resolution evidence into the and , capturing regional variability such as the and , with records from the White Mountains showing cooler conditions during the compared to the 20th, independent of anthropogenic influences. Coral growth bands and lake sediment varves further corroborate decadal to centennial fluctuations in tropical sea surface temperatures and monsoon intensity, highlighting internal variability amplified by ocean-atmosphere interactions like El Niño-Southern Oscillation analogs in the past. Collectively, these proxies demonstrate that Earth's exhibits inherent , with natural forcings capable of producing changes comparable to or exceeding 20th-century warming rates in specific intervals, though without the unprecedented atmospheric CO₂ rise observed since 1850.

Holocene and Recent Pre-Industrial Changes

The Holocene epoch commenced approximately 11,700 years before present, succeeding the Younger Dryas cold interval and initiating a period of relative climatic stability following the last glacial maximum. Proxy-based reconstructions from ice cores, tree rings, lake sediments, and speleothems reveal an initial rapid warming in the early Holocene, driven primarily by retreating ice sheets and rising atmospheric CO2 levels from ~260 ppm to ~280 ppm, with global mean surface temperatures (GMST) increasing by about 4-5°C from glacial lows. This warming culminated in the Holocene Climatic Optimum (HCO), spanning roughly 9,500 to 5,500 years BP, during which multi-proxy GMST estimates indicate peak warmth around 7,000 years BP, approximately 0.5°C above late-Holocene levels but comparable to or slightly exceeding pre-industrial averages in some hemispheric reconstructions. Following the HCO, proxy data document a gradual cooling trend across the mid- to late , often termed the Neoglacial period, with GMST declining by 0.5-1°C over millennia, attributed to decreasing summer insolation from Milankovitch orbital cycles and amplified by feedback mechanisms like alpine glacier advances. This long-term cooling featured centennial-scale fluctuations, including the 4.2 ka event—a abrupt and cooling episode linked to weakening—and regional expressions of the 8.2 ka cooling from glacial meltwater pulses. Empirical reconstructions highlight spatial variability, with land areas showing more pronounced seasonality, where summer temperatures peaked earlier than annual means, resolving some discrepancies between proxy data and model simulations that predict orbital-driven warming. Volcanic eruptions and variations contributed to short-term perturbations, but internal ocean-atmosphere dynamics, such as shifts, played key roles in regional anomalies. In the recent pre-industrial era, spanning the last millennium before 1850, climate exhibited notable variability superimposed on the Holocene cooling trajectory, including the Medieval Warm Period (MWP, ~950-1250 CE) and Little Ice Age (LIA, ~1450-1850 CE). Northern Hemisphere reconstructions indicate MWP temperatures ~0.2-0.5°C above the subsequent LIA baseline, with evidence of broader synchrony in proxy networks suggesting a modest global signal, though amplitudes varied regionally due to land-ocean contrasts. The LIA featured cooling of ~0.5-1°C in hemispheric means relative to pre-industrial norms, correlated with clustered volcanic activity—such as sulfate spikes in ice cores—and low solar irradiance during grand minima like the Maunder Minimum (1645-1715 CE), which reduced total solar forcing by ~0.1-0.2 W/m². These forcings induced stratospheric aerosol cooling and altered tropospheric circulation, with volcanic impacts persisting 2-3 years per event and solar changes modulating decadal modes, while excluding significant anthropogenic influence prior to widespread fossil fuel use. Overall pre-industrial variability remained within ~1°C of GMST excursions, underscoring natural drivers' dominance absent modern greenhouse gas rises.

Observed Changes Since 1850

Instrumental records of global surface air temperature, beginning around 1850 from stations and marine observations, show an overall increase of approximately 1.1°C relative to the 1850–1900 baseline through 2024. This warming has not been uniform, with about two-thirds occurring since 1975 and surfaces warming faster than s, the latter absorbing roughly 90% of excess . temperatures have risen more than in the , attributable to greater coverage and ocean circulation patterns. Atmospheric concentrations, reconstructed from cores and direct measurements, have risen from about 280 around 1850 to over 420 by 2024, with the increase accelerating post-1950 via data. Other greenhouse gases, such as , have also increased, from roughly 700 ppb in the mid-19th century to 1900 ppb today. Global mean sea level, measured by tide gauges since the late and augmented by altimetry since 1993, has risen 21–24 cm since 1880, with rates increasing from about 1.5 mm/year historically to 3.7 mm/year recently. glaciers worldwide have retreated since the mid-, with cumulative ice loss accelerating across decades; for instance, many and North glaciers have lost 30–50% of their volume since 1850. and Antarctic ice sheets have experienced net mass loss, contributing to , though Antarctic trends show regional variability with some gains in . Ocean heat content in the upper 2000 meters has increased steadily since the 1950s, confirmed by float arrays deployed from 2000 onward, indicating uptake of excess energy consistent with surface warming trends. sea ice extent has declined, particularly in summer minima, at 12% per decade since satellite records began in 1979, though pre-satellite proxy data suggest declines predating this period. These changes exhibit variability influenced by natural oscillations like El Niño-Southern Oscillation, with decadal pauses or slowdowns in surface warming observed, such as from the 1940s to 1970s.

Natural vs. Human Attribution Debates

Detection and attribution analyses seek to apportion observed climate changes between forcings, such as elevated concentrations, and natural drivers including fluctuations, volcanic aerosols, and internal variability from ocean-atmosphere oscillations. Mainstream assessments assert that human activities account for approximately 1.0–1.2°C of the observed ~1.1°C global surface warming since 1850–1900, with natural forcings contributing negligibly or slightly negative over that period. These conclusions derive from optimal fingerprinting techniques, which match spatial patterns of change—such as stratospheric cooling juxtaposed with tropospheric warming—to model ensembles incorporating while excluding it yields divergence from observations. Critics contend that attribution methodologies systematically undervalue natural variability by relying on models that inadequately simulate multidecadal oscillations and regional patterns, leading to overattribution to human causes. A analysis of global mean surface records identified synchronized transitions in major climate indices around 1910–1920, 1940, 1960, 1976–1980, and 2000, suggesting that internal variability, rather than monotonic external forcing, underpins the non-linear 20th-century warming trajectory. Such critiques highlight methodological flaws in optimal fingerprinting, including assumptions of Gaussian error distributions that fail under non-stationary conditions and neglect of unresolved forcings like land-use changes or indirect effects. Solar activity variations, proxied by sunspot numbers and cosmogenic isotopes, exhibit correlations with hemispheric temperatures over centuries, with recent studies quantifying a potential role in modulating total (TSI), ultraviolet output, and flux influencing . Empirical reconstructions indicate TSI changes of ~0.1–0.3 W/m² over 11-year cycles and longer grand solar minima, sufficient to explain portions of early 20th-century warming (1900–1940) when anthropogenic emissions were lower. However, post-1950 —declining activity amid rising temperatures—prompts mainstream dismissal of direct TSI dominance, though indirect mechanisms like -modulated galactic s enhancing low-cloud could reconcile discrepancies, with sensitivity estimates up to 0.5–1.0 W/m² effective forcing. Internal ocean-atmosphere modes, including the (PDO), (AMO), and El Niño-Southern Oscillation (ENSO), imprint multidecadal temperature fluctuations on global scales, with positive PDO phases (e.g., 1925–1946, 1977–1998) aligning with accelerated warming epochs and contributing ~0.1–0.2°C to mid-20th-century trends via altered heat redistribution. The AMO's warm phase since the 1990s has amplified North Atlantic and Eurasian temperatures, while ENSO episodes account for ~20–30% of interannual variance, masking or amplifying underlying trends. Detractors of dominant natural attribution argue these modes are oscillatory and net zero over centuries, incapable of sustaining directional warming without external forcing, yet empirical deconstructions reveal their role in explaining ~50% of 20th-century variance when phased with signals. Persistent uncertainties stem from equilibrium climate sensitivity (ECS) estimates spanning 1.5–4.5°C per CO₂ doubling, with observational constraints favoring lower values (~1.5–2.5°C) inconsistent with high-end model projections, and discrepancies like subdued tropospheric warming over the challenging fingerprint predictions. These gaps fuel debate, as attribution hinges on imperfect models exhibiting overestimated historical hindcasts and underestimated natural internal variability at decadal-to-centennial scales. While peer-reviewed consensus leans , dissenting analyses grounded in empirical underscore unresolved contributions from natural processes, urging caution in policy presumptions of near-exclusive human causality.

Modeling and Predictions

Climate Model Mechanics

Climate models simulate the Earth's by numerically solving a set of partial differential equations derived from fundamental physical laws, including conservation of momentum, mass, energy, and . These equations, often referred to as the , approximate the behavior of fluids in the atmosphere and oceans using the Navier-Stokes equations adapted for large-scale geophysical flows, incorporating terms for gradients, Coriolis forces, , and diabatic heating from and release. General circulation models (GCMs), the core of most climate simulations, discretize the globe into a three-dimensional grid—typically with horizontal resolutions of 50 to 300 kilometers and 20 to 100 vertical levels—to compute state variables such as , wind velocity, humidity, and at discrete points and time steps. Numerical methods employed include finite difference schemes for spatial derivatives and time-stepping algorithms like leapfrog or Runge-Kutta integrators to advance solutions forward in time, ensuring stability through constraints such as the Courant-Friedrichs-Lewy condition, which limits time steps to prevent information propagation exceeding grid speeds. Spectral methods, transforming variables into wavenumber space via Fourier or spherical harmonics, offer efficiency for global domains by exploiting periodicity but require additional handling for non-linear terms and topography. Radiation schemes solve the radiative transfer equation to compute shortwave and longwave fluxes, accounting for absorption, scattering, and emission by gases like CO2, water vapor, and ozone, often using band models or correlated-k approximations for computational tractability. Sub-grid scale processes, unresolved by the grid, necessitate parameterization schemes that empirically or theoretically represent their average effects, such as convective triggered when grid-scale moisture exceeds thresholds, or turbulent mixing via eddy diffusivity closures. microphysics and macrophysics are particularly challenging, with parameterizations estimating formation, evolution, and radiative impacts based on simplified assumptions about droplet/ice processes and feedbacks, introducing uncertainties due to the multi-scale nature of and aerosols. Land surface models parameterize dynamics, vegetation transpiration, and variations, while ocean components solve similar with added and sea-ice . Fully coupled Earth system models integrate atmospheric, oceanic, land, ice-sheet, and biogeochemical modules, exchanging fluxes like heat, momentum, and freshwater at interfaces to simulate interactions such as ocean heat uptake or feedbacks. Initialization often draws from reanalysis datasets or spin-up runs to equilibrium, with forcing from observed or projected gases, aerosols, and variability; ensemble simulations perturb initial conditions or parameters to quantify internal variability. High computational demands—requiring supercomputers for simulations spanning centuries—stem from the need for fine resolutions and iterative solver convergence, limiting explicit resolution of processes below ~10 km scales.

Validation Against Observations

Climate models undergo validation through hindcasting, wherein simulations of historical climate forcings are compared against observations, records, and reanalysis datasets to assess fidelity in reproducing trends and variability. This process evaluates key variables such as surface air temperatures, upper-air temperatures, sea surface temperatures, and patterns. While models capture the broad 20th-century signal, systematic biases emerge upon detailed scrutiny, particularly in the magnitude and spatial patterns of change. Surface temperature hindcasts in ensembles like CMIP5 and CMIP6 often exceed observed warming rates; for example, CMIP5 simulations warmed approximately 16% faster than global surface observations since 1970, with about 40% of the divergence attributable to internal variability and the remainder to model errors in forced response. Over the past 50 years, nearly all computerized climate models have projected stronger global warming than the empirically measured rate from datasets like HadCRUT and UAH. These overestimations persist even after accounting for volcanic aerosols and solar variability, suggesting inflated equilibrium climate sensitivity in many models. Tropospheric temperatures provide a stringent test, as models predict enhanced warming in the tropical mid-to-upper due to moist amplification—a "hot spot" signature largely absent in satellite (e.g., UAH, ) and data. Coupled model simulations have shown roughly twice the tropical tropospheric warming relative to satellite observations since 1979, with discrepancies amplified by internal variability and forcing mismatches rather than observational artifacts. Recent analyses confirm that CMIP6 models continue to overestimate these trends, undermining confidence in projections of future amplification. Beyond temperatures, validations reveal mismatches in extent and ocean dynamics; CMIP6 models exhibit biases in September sea ice concentration linked to erroneous atmospheric and oceanic circulations, failing to replicate observed decline rates accurately. trends in regions like the central U.S. also diverge, with models simulating patterns inconsistent with gauge observations due to deficiencies in simulating low-level . These persistent discrepancies highlight limitations in representing natural variability and cloud feedbacks, prompting ongoing refinements but also caution in extrapolating unvalidated projections. Despite some successes in broad-scale energy balance, the prevalence of overpredictions underscores that models tuned to high-sensitivity scenarios do not uniformly align with empirical records.

Projection Uncertainties

Climate projections for future global warming depend on coupled atmosphere-ocean general circulation models (GCMs) integrated over specified emission scenarios, yet these projections are beset by multifaceted uncertainties that span structural model deficiencies, parametric choices, forcing estimates, and internal variability. The IPCC's Sixth Assessment Report (AR6) quantifies global surface air temperature increases of 1.0–1.8°C by 2081–2100 relative to 1850–1900 under low-emission scenarios (SSP1-1.9 to SSP1-2.6), escalating to 3.3–5.7°C under high-emission SSP5-8.5, with 90% confidence intervals reflecting irreducible spreads across model ensembles like CMIP6. These ranges arise partly from equilibrium climate sensitivity (ECS), defined as the long-term temperature response to doubled atmospheric CO₂, assessed at a likely range of 2.5–4.0°C in AR6, narrower than the 1.5–4.5°C of prior reports but still encompassing a factor-of-two spread due to incomplete observational constraints on paleoclimate and process-level feedbacks. A dominant source of projection divergence stems from cloud radiative feedbacks, where low-level stratocumulus and mixed-phase may either enhance warming through reduced or mitigate it via increased reflectivity, with CMIP6 models exhibiting a spread of approximately ±0.35 /m²/°C at 90% confidence. Observational constraints, such as satellite-derived responses in clean environments, suggest positive feedbacks amplifying , rendering low-ECS values (below 2°C) extremely unlikely, yet model biases in schemes and microphysics perpetuate structural . radiative forcing introduces further ambiguity, as sulfates and organics have exerted a cooling effect estimated at -0.9 to -0.1 /m² since pre-industrial times, with the largest inter-model spread among all forcings; future declines under air quality regulations could unmask additional warming, but vertical distribution and interactions remain poorly resolved, contributing up to 0.5°C in transient warming projections. Scenario dependencies compound these issues, as socioeconomic pathways (SSPs) embed assumptions about , , and that diverge sharply—e.g., SSP3-7.0 implies sustained high emissions versus SSP1-1.9's aggressive —yielding non-overlapping warming outcomes despite identical physical models. Internal variability, including decadal oscillations like the , adds noise equivalent to 0.1–0.2°C on multi-decadal scales, while heat uptake efficiency varies across models, delaying peak warming by centuries in some cases. Evaluations of historical projections indicate GCMs have captured global mean warming trends with reasonable skill since the , tracking observed rates of ~0.18°C/decade, but overestimate tropical Pacific warming patterns and changes, highlighting limitations in regional fidelity that amplify for impacts like extremes. Despite advances in and process representation, persistent biases—such as excessive ECS in ~50% of CMIP6 models relative to updated observational inferences—underscore the need for emergent constraints from paleodata and satellites to refine ranges, though consensus views in AR6 reflect averaged model outputs rather than fully reconciled physics.

Potential Impacts

Environmental and Ecological Effects

Rising atmospheric CO2 concentrations have driven a measurable of Earth's vegetated lands, with satellite data from 1982 to 2015 indicating that 25 to 50 percent of global vegetated areas experienced significant greening, 70 percent of which is attributable to CO2 fertilization enhancing and growth. This effect has been particularly pronounced in and agricultural regions, countering some expectations of uniform ecological decline and demonstrating CO2's role as a under elevated levels. However, localized events, such as those observed in the and European mountains, have been linked to compounded stressors including and insect outbreaks intensified by warmer conditions, though pathogens often play a primary role rather than alone. Species distributions have shifted in response to climatic variability, with empirical observations documenting poleward migrations averaging 17 kilometers per decade for terrestrial species and upslope movements of approximately 11 meters per decade in mountainous regions since the late . These redistributions reflect adaptations to changing thermal niches but can disrupt ecosystems, as evidenced by increased tropicalization of temperate communities where warmer waters favor herbivorous , reducing extent. Nonetheless, such changes occur amid dominant anthropogenic pressures like , which exceed climate's current influence on in assessments of imperiled species. In marine environments, warming has triggered mass events, such as the 2014-2017 global episode where sea surface temperatures exceeded thresholds by 1°C for extended periods, expelling symbiotic algae and causing mortality in up to 30 percent of surveyed reefs, though has occurred in less severe cases through larval . Concurrent , with surface declining by 0.1 units since pre-industrial times due to CO2 absorption, impairs in organisms like pteropod mollusks and some corals, potentially altering food webs, yet laboratory and field studies reveal variable among , with non-calcifying often unaffected or benefiting from higher CO2. Overall, while variability contributes to ecological shifts, its role remains secondary to land-use changes and direct exploitation in driving contemporary declines.

Societal and Economic Consequences

Climate variability and change influence societal structures through impacts on health, migration, and inequality, while economic consequences span agriculture, infrastructure, and GDP growth. Empirical assessments indicate that while certain regions face heightened risks from extreme events, human adaptation—such as improved infrastructure and early warning systems—has substantially mitigated adverse outcomes. For instance, global temperature-related deaths have declined by approximately 650,000 annually since the 1990s, driven primarily by reductions in cold-related mortality outpacing increases in heat-related deaths, alongside broader socioeconomic improvements like access to air conditioning and healthcare. Adaptation measures, including resilient building codes and flood defenses, have further reduced vulnerability; in the U.S., the temperature-mortality relationship weakened markedly over the 20th century, with hot days becoming less lethal in warmer regions due to acclimatization and technology. In agriculture, elevated atmospheric CO2 concentrations provide a fertilization effect that enhances crop yields, counteracting some warming-induced stresses. Studies estimate that CO2 fertilization boosted yields of C3 crops like and by 7.1% from 1961 to 2017, contributing to overall global production gains amid . However, this benefit varies by crop type and region; C4 crops such as exhibit smaller responses, and combined with variable patterns, net effects on yields remain regionally heterogeneous, with potential reductions in tropical areas offset by expansions in higher latitudes. Economic analyses suggest that without accounting for CO2 benefits, projected yield losses from warming are overstated, though pests, droughts, and heatwaves pose ongoing risks mitigated by and genetically modified varieties. Extreme weather events, including floods and storms, impose significant economic costs, with U.S. billion-dollar disasters totaling over $2.6 trillion (CPI-adjusted) from 1980 to 2024, though and normalized losses have not escalated proportionally due to and in exposed areas. Globally, such events cost an estimated $2 trillion over the past decade, but investments—such as sea walls and —yield high returns, with every $1 spent generating over $10 in avoided damages over a decade. Macroeconomic models project impacts could reduce GDP by 1-4% by 2100 under moderate warming scenarios, disproportionately affecting developing economies through reduced and , yet historical data show via technological progress and market adjustments. Societal disruptions like and arise from variability, with empirical links to reduced increasing in arid regions, though country-specific factors like dominate. Inequality may widen, as extremes regressively impact lower-income groups with limited , exacerbating within-country disparities. Nonetheless, positive historical precedents exist, where past variability prompted societal innovations that enhanced stability, underscoring 's role over deterministic collapse narratives. Overall, integrated assessment models highlight that benefits from milder winters and CO2 effects partially offset costs, with policy emphasizing cost-effective rather than solely to minimize net economic burdens.

Controversies and Critiques

Scientific Debates

Equilibrium climate sensitivity (ECS), the expected long-term global surface air temperature rise from doubled atmospheric CO₂, constitutes a core debate, with estimates varying substantially across methods. Comprehensive general circulation models and paleoclimate reconstructions informing IPCC AR6 yield a likely ECS range of 2.5–4.0°C, emphasizing strong positive feedbacks from and . In contrast, energy budget approaches using satellite-era observations of radiative fluxes, ocean heat uptake, and aerosols derive medians around 2.0°C, with 95% ranges of 1.2–2.9°C, attributing differences to overstated cloud amplification in models. Recent syntheses combining records, paleo constraints, and emergent constraints further support ECS below 3°C, challenging higher model-derived values as inconsistent with observed historical warming. The vertical structure of atmospheric warming, especially the tropical tropospheric "," highlights another discrepancy between theory, models, and data. Moist adiabatic theory and coupled models predict surface warming by factors of 1.5–2.0 in the mid-to-upper tropical (200–300 levels), a of greenhouse forcing via enhanced convection. networks and MSU/AMSU records from 1979–2020, however, show amplification near or below 1.0, with trends statistically inconsistent with multi-model means at >99% confidence in some analyses. While proponents cite reanalysis adjustments or natural variability to reconcile data, critics maintain the mismatch signals model errors in convective mixing and relative humidity, undermining attribution confidence. Model-observation comparisons reveal further tensions in simulating decadal trends and internal variability. CMIP5/6 ensembles broadly match global surface warming since 1850 but diverge on upper-air profiles, rates, and extremes, often simulating excessive tropospheric warming over land and underestimating stratospheric cooling influences. Structural inconsistencies, such as in microphysics and aerosol-cloud interactions, propagate uncertainties, with some ensembles running "" relative to post-2000 observations when accounting for multi-decadal oscillations like the AMO. These gaps persist despite tuning, prompting calls for better constraint via emergent relationships and paleoclimate analogs, though mainstream assessments deem overall skill sufficient for projections.

Policy and Alarmism Critiques

Critics of climate alarmism contend that exaggerated claims of imminent catastrophe have historically failed to align with empirical observations, fostering public hysteria rather than measured responses. For example, predictions around the first in 1970 included assertions of widespread famines and by 2000 due to and environmental collapse, none of which occurred as forecasted. Similarly, claims of an ice-free by 2013, as suggested by some models and popularized in , have not materialized, with summer persisting despite reductions. These discrepancies highlight a pattern where alarmist projections, often amplified by mainstream outlets, overestimate short-term extremes while underemphasizing natural variability and human adaptability. Economist argues in (2020) that such rhetoric drives policies with disproportionate costs relative to benefits, estimating that aggressive mitigation efforts like the would cost $819–$1,890 billion annually through 2100 while averting only about 0.17°C of warming by century's end. 's cost-benefit analyses, drawing on integrated assessment models, indicate that the is often overstated in alarmist narratives, with benefits from CO2 fertilization—such as enhanced global greening and crop yields—offsetting a portion of projected damages. He critiques the focus on emission cuts over innovation, noting that trillions spent on subsidies for intermittent renewables yield minimal emission reductions compared to targeted R&D in or fusion technologies. Net-zero policies, mandating near-elimination of by mid-century, face scrutiny for their economic toll. In the , pursuit of net zero has correlated with energy prices rising 2–3 times faster than in peers since 2010, contributing to industrial closures and household fuel affecting millions. Global estimates suggest achieving net zero could require $27 trillion annually in redirected spending, imposing opportunity costs by diverting funds from alleviation, , or in developing nations, where climate impacts are often less severe than immediate needs. Critics like Lomborg emphasize that measures—such as resilient —deliver higher returns; for instance, investing in sea walls or drought-resistant crops could mitigate and agricultural risks at fractions of costs. These critiques extend to institutional biases, where summaries from bodies like the IPCC are accused of prioritizing worst-case scenarios to justify expansive interventions, despite underlying data showing moderated projections upon scrutiny. Empirical reviews find no robust evidence linking recent warming to surges in like hurricanes or tornadoes, countering alarmist attributions that drive regulatory overreach. Proponents of rational advocate prioritizing high-impact, low-cost strategies like accelerating green through incentives, arguing that alarmism erodes credibility and hinders pragmatic solutions.

References

  1. [1]
    Climate Variability - UCAR Center for Science Education
    Climate variability is the way aspects of climate (such as temperature and precipitation) differ from an average. Climate variability occurs due to natural and ...
  2. [2]
    [PDF] Climate Variability vs. Climate Change - National Weather Service
    Climate change refers to a statistically significant variation in either the mean state of the climate or in its variability, persisting for an extended period ...
  3. [3]
    Climate Variability
    At its most fundamental, climate variability stems from the redistribution and changes in the amount of energy around the globe, which lead to changes in ...
  4. [4]
    Climate variability and vulnerability to climate change: a review - PMC
    Climate change is inevitably resulting in changes in climate variability and in the frequency, intensity, spatial extent, duration, and timing of extreme ...
  5. [5]
    Climate change: evidence and causes | Royal Society
    Greenhouse gases emitted by human activities alter Earth's energy balance and thus its climate. Humans also affect climate by changing the nature of the land ...
  6. [6]
    Climate change: global temperature
    May 29, 2025 · Earth's temperature has risen by an average of 0.11° Fahrenheit (0.06° Celsius) per decade since 1850, or about 2° F in total.
  7. [7]
    Data Overview - Berkeley Earth
    Global Temperature Data · Global Monthly Averages (1850 – Recent). Berkeley Earth combines our land data with a modified version of the HadSST ocean temperature ...
  8. [8]
    Temperature data (HadCRUT, CRUTEM, HadCRUT5, CRUTEM5 ...
    For our best estimate of how global temperature has changed since 1850, we recommend you use the HadCRUT5 Analysis. The latest version is HadCRUT5, formed ...
  9. [9]
    Climate change: atmospheric carbon dioxide
    In the past 60 years, carbon dioxide in the atmosphere has increased 100-200 times faster than it did during the end of the last ice age.
  10. [10]
    How do we know more CO2 is causing warming? - Skeptical Science
    An enhanced greenhouse effect from CO2 has been confirmed by multiple lines of empirical evidence. Satellite measurements of infrared spectra over the past ...
  11. [11]
    Overstating the effects of anthropogenic climate change? A critical ...
    Feb 8, 2023 · Climate scientists have proposed two methods to link extreme weather events and anthropogenic climate forcing: the probabilistic and the ...
  12. [12]
    Evidence - NASA Science
    Oct 23, 2024 · In 1938, Guy Callendar connected carbon dioxide increases in Earth's atmosphere to global warming. In 1941, Milutin Milankovic linked ice ages ...
  13. [13]
    To understand climate change adaptation, we must characterize ...
    Dec 15, 2023 · Climate variability. Fluctuations of a climate variable from the mean state on timescales including weeks to decades and longer. Climate ...
  14. [14]
    Glossary — Global Warming of 1.5 ºC
    Climate change refers to a change in the state of the climate that can be identified (e.g., by using statistical tests) by changes in the mean and/or the ...
  15. [15]
    Frequently Asked Questions About Climate Change | US EPA
    Aug 25, 2025 · In contrast, climate variability includes changes that occur within shorter timeframes, such as a month, season, or year. Climate variability ...<|control11|><|separator|>
  16. [16]
    Climate Change | United Nations
    Climate change refers to long-term shifts in temperatures and weather patterns. Such shifts can be natural, due to changes in the sun's activity or large ...
  17. [17]
    Global Temperature Anomalies from 1880 to 2024 - NASA SVS
    Jan 10, 2025 · Earth's global surface temperatures in 2024 were the warmest on record -- 1.28 degrees Celsius (2.30 degrees Fahrenheit) above the agency's 20th ...
  18. [18]
    Assessing the Global Temperature and Precipitation Analysis in ...
    Sep 11, 2025 · The January–August global surface temperature was the second-highest on record behind 2024. According to NCEI's Global Annual Temperature ...<|separator|>
  19. [19]
    Trends in CO2 - NOAA Global Monitoring Laboratory
    Sep 5, 2025 · Monthly Average Mauna Loa CO2. August 2025: 425.48 ppm. August 2024: 422.99 ppm. Last updated: Sep ...Last 1 Year · Data · Last Month · Global
  20. [20]
    Annual Carbon Dioxide Peak Passes Another Milestone
    Jun 5, 2025 · Scripps Oceanography scientists calculated a May monthly average of 430.2 ppm for 2025, an increase of 3.5 ppm over May 2024's measurement of ...
  21. [21]
    Overview | Global Sea Level
    The rate of GMSL rise from 1993 to present has been measured at 3.4 millimeters per year, and there are indications that the rate of GMSL rise has increased ...Regional Sea Level · Vital Signs · Ice Melt<|separator|>
  22. [22]
    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) ...
  23. [23]
    Global sea level rose faster than expected in 2024, according to ...
    Mar 14, 2025 · Unprecedented heat led to an unexpected level of sea level rise in 2024, according to a new NASA analysis.
  24. [24]
    Climate Change: Ocean Heat Content
    Upper ocean heat content has increased significantly over the past few decades. Upper layers are accumulating heat faster than deeper layers.
  25. [25]
    Ocean Warming - Earth Indicator - NASA Science
    Sep 25, 2025 · About 90% of the excess heat from planetary warming over the past century has been absorbed by the ocean, causing ocean temperatures to rise.
  26. [26]
    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).
  27. [27]
    Arctic Sea Ice Minimum Extent - Earth Indicator - NASA Science
    The measurements have shown that September Arctic sea ice is shrinking at a rate of 12.2% per decade, compared to its average extent during the period from 1981 ...
  28. [28]
    U.S. Climate Extremes Index (CEI)
    The U.S. Climate Extremes Index (CEI) quantifies observed climate variability in temperature, precipitation, and drought within the contiguous United States.<|separator|>
  29. [29]
    GISS Surface Temperature Analysis (GISTEMP v4) - Data.GISS
    September 11, 2025: The August 2025 report from Maliye Karmakuly seemed inconsistent with history and its neighbors, off by about 12 °C. This was confirmed by ...Global Maps · History · FAQs · News, Updates, and Features
  30. [30]
    [PDF] Annex IV: Modes of Variability
    Many modes of variability are driven by internal climate processes and provide a substantial potential source of climate predictability on sub-seasonal to ...
  31. [31]
    “Certain Uncertainty: The Role of Internal Climate Variability in ...
    Nov 17, 2020 · Generally speaking, internal climate variability is larger in the extra-tropics than the tropics, greater in winter than summer, larger for ...
  32. [32]
    Tropical Pacific Forcing versus Internal Variability in - AMS Journals
    Abstract. The Pacific decadal oscillation (PDO) is the leading mode of sea surface temperature (SST) variability over the North Pacific (north of 20°N).Missing: peer- | Show results with:peer-
  33. [33]
    [PDF] Modulation of ENSO teleconnections over North America by the ...
    Oct 18, 2022 · We find that a positive PDO enhances winter and spring El Niño temperature and precipitation teleconnections and diminishes La Niña ...
  34. [34]
    [PDF] The impact of the AMO on multidecadal ENSO variability
    Apr 30, 2017 · The AMO strongly influences multidecadal ENSO variability, impacting it directly and through changes to the annual cycle, modifying ENSO ...
  35. [35]
    [PDF] Influence of ENSO and the Atlantic Multidecadal Oscillation on ...
    ENSO impacts drought, especially in the Southwest. AMO's warm phase increases drought in the Southwest and north-central US, and AMO can modulate ENSO's ...
  36. [36]
    [PDF] AMO Forcing of Multidecadal Pacific ITCZ Variability
    Jul 15, 2018 · The Atlantic multidecadal oscillation (AMO) has been shown to play a major role in the multidecadal variability of the Northern Hemisphere, ...<|separator|>
  37. [37]
    [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.
  38. [38]
    Absence of internal multidecadal and interdecadal oscillations in ...
    Jan 3, 2020 · For several decades the existence of interdecadal and multidecadal internal climate oscillations has been asserted by numerous studies based ...
  39. [39]
    Internal Variability of the Climate System Mirrored in Decadal‐Scale ...
    Jun 11, 2023 · We seek to find a relationship between 12 well-known climate modes of variability (El Niño–Southern Oscillation, Pacific Decadal Oscillation, ...
  40. [40]
    Interpreting Climate Conditions: What is Attribution
    There are numerous mechanisms that may produce climate variations or change. One is external forcing, which contains both natural and anthropogenic sources.
  41. [41]
    9.1 Introduction - AR4 WGI Chapter 9: Understanding and Attributing ...
    Some external influences, such as changes in solar radiation and volcanism, occur naturally and contribute to the total natural variability of the climate ...
  42. [42]
    Surface Temperature Reconstructions for the Last 2000 Years (2006)
    The main external climate forcings experienced over the last 2,000 years are volcanic eruptions, changes in solar radiation reaching the Earth, ...
  43. [43]
    Climate Change: Incoming Sunlight | NOAA Climate.gov
    During strong solar cycles, the Sun's total average brightness varies by up to 1 Watt per square meter; this variation affects global average temperature by 0.1 ...
  44. [44]
    Solar influence on climate during the past millennium - PNAS
    A small role of solar forcing for late 20th century climate change is additionally supported by the absence of a trend in the satellite-based irradiance record ...
  45. [45]
    Changes in the Total Solar Irradiance and climatic effects
    Jul 22, 2021 · As solar irradiance variations have a global effect there has to be a global climatic solar forcing impact. However, by how much global ...
  46. [46]
    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, ...
  47. [47]
    Orbital forcing of climate 1.4 billion years ago - PNAS
    Much of this climate change is driven by variations of Earth's orbit around the Sun with characteristic frequencies known as Milankovitch cycles.
  48. [48]
    Toward generalized Milankovitch theory (GMT) - CP - Copernicus.org
    Jan 18, 2024 · The eccentricity influences glacial cycles solely through its amplitude modulation of the precession component of orbital forcing, while the ...
  49. [49]
    Volcanoes Can Affect Climate | U.S. Geological Survey - USGS.gov
    Sulfate aerosols can cool the climate and deplete Earth's ozone layer. The most significant climate impacts from volcanic injections into the stratosphere come ...
  50. [50]
    The effect of volcanic aerosols on global climate - ScienceDirect
    Volcanic eruptions that inject large quantities of sulfur-rich gases into the stratosphere have the capability of cooling global climate by 0.2–0.3°C for ...
  51. [51]
    Severe Global Cooling After Volcanic Super-Eruptions? The Answer ...
    By evaluating the range of aerosol size scenarios, we demonstrate that eruptions may be incapable of causing more than 1.5°C cooling no matter how much sulfur ...
  52. [52]
    Greenhouse Gases Factsheet | Center for Sustainable Systems
    In 2023, global anthropogenic GHG emissions totaled 52.9 Gt CO₂e, a 62% increase since 1990. · Average annual GHG emissions were 56 Gt CO₂e from 2010–2019—a ...
  53. [53]
    CO₂ and Greenhouse Gas Emissions - Our World in Data
    Human greenhouse gas emissions have increased global average temperatures · Global emissions have increased rapidly over the last 50 years and have not yet ...Breakdown of carbon dioxide... · CO₂ emissions by fuel · Global aviation
  54. [54]
    Global Greenhouse Gas Overview | US EPA
    Aug 19, 2025 · Includes information on global greenhouse gas emissions trends, and by type of gas, by source, and by country.
  55. [55]
    GHG emissions of all world countries - 2025 Report
    In 2024, the majority of GHG emissions consisted of fossil CO2 accounting for 74.5% of total emissions, while CH4 contributed for 17.9% to the total, N2O for 4 ...Introduction · Main findings · Emissions by country · Sources and references
  56. [56]
    Breakdown of carbon dioxide, methane, and nitrous oxide emissions ...
    This chart shows the breakdown of total greenhouse gases (the sum of all greenhouse gases, measured in tonnes of carbon dioxide equivalents) by sector.Where do global greenhouse... · Greenhouse gas emissions
  57. [57]
    [PDF] Climate Change 2022
    Figure SPM.1 | Global net anthropogenic GHG emissions (GtCO2-eq yr–1) 1990–2019. Global net anthropogenic GHG emissions include CO2 from fossil fuel.
  58. [58]
    Sources of Greenhouse Gas Emissions | US EPA
    Mar 31, 2025 · CO2 emissions from natural gas consumption increased by 5% relative to 2021. CO2 emissions from coal consumption decreased by 6% from 2021.
  59. [59]
    Chapter 6: Energy systems
    The global energy system is the largest source of CO2 emissions (Chapter 2). Reducing energy sector emissions is therefore essential to limit warming. The ...
  60. [60]
    Magnitude uncertainty dominates intermodel spread in zonal-mean ...
    Sep 10, 2025 · Anthropogenic aerosols induced an effective global radiative forcing of −1.1 (−1.7 to −0.4) W/m2 over 1750 to 2019 with large uncertainty (1).
  61. [61]
    Decomposing the effective radiative forcing of anthropogenic ... - ACP
    Jul 10, 2024 · The perturbation induced by changes in anthropogenic aerosols on the Earth's energy balance is quantified in terms of the effective radiative forcing (ERF).
  62. [62]
    Strong control of effective radiative forcing by the spatial pattern of ...
    Jul 21, 2022 · Here, using model experiments, we show that the effective radiative forcing from absorbing aerosol varies strongly depending on their location, ...
  63. [63]
    Heterogeneous Impact of Land-Use on Climate Change - Frontiers
    IPCC also mentioned that some studies have shown that changes in land cover or water available for irrigation will affect the climate in areas hundreds of ...
  64. [64]
    Global Land Use Change and Its Impact on Greenhouse Gas ... - NIH
    Nov 29, 2024 · This study explores how global land use changes, including urbanization, agricultural expansion, and deforestation, have influenced greenhouse gas (GHG) ...
  65. [65]
    Special Report on Climate Change and Land — IPCC site
    An IPCC Special Report on climate change, desertification, land degradation, sustainable land management, food security, and greenhouse gas fluxes in ...Cite Report · Land Degradation · Summary for Policymakers · Technical Summary
  66. [66]
    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.
  67. [67]
    Does Urban Heat Island effect exaggerate global warming trends?
    It has been suggested that UHI has significantly influenced temperature records over the 20th century with rapid growth of urban environments. Scientists have ...
  68. [68]
    Exactly How Large Is the Urban Heat Island Effect in Global Warming?
    Mar 18, 2024 · The UHI effect is estimated to be around 25% of measured temperature trend, but some studies suggest it could be as high as 40% or as low as 6%.
  69. [69]
    The Need to Consider Human Forcings Besides Greenhouse Gases
    In addition to greenhouse gas emissions, other first- order human climate forcings are important to understanding the future behav- ior of Earth's climate.Missing: anthropogenic | Show results with:anthropogenic
  70. [70]
    Comparison of Climate Response to Anthropogenic Aerosol versus ...
    While GHG forcing dominates global mean surface temperature change, its effect is on par with and often opposes the aerosol effect on precipitation, making it ...Missing: besides | Show results with:besides
  71. [71]
    Unlocking Pre-1850 Instrumental Meteorological Records: A Global ...
    Dec 1, 2019 · Very long series such as central England temperatures (start date 1659; Manley 1974; Parker et al. 1992) or Paris temperatures (1658; Rousseau ...
  72. [72]
    Why does the temperature record shown on your "Vital Signs" page ...
    Mar 18, 2024 · The oldest continuous temperature record is the Central England Temperature Data Series, which began in 1659, and the Hadley Centre has some ...
  73. [73]
    NOAA Global Surface Temperature (NOAAGlobalTemp)
    It is a spatially gridded (5° - 5°) global surface temperature dataset, with monthly resolution from January 1850 to present.
  74. [74]
    Global land-ocean surface temperature data: HadCRUT5
    Dec 1, 2020 · HadCRUT5 is one of the main datasets used to monitor global and regional surface temperature variability and trends.<|separator|>
  75. [75]
    Explainer: How data adjustments affect global temperature records
    Jul 19, 2017 · The rate of warming between 1950 and 2016 in the adjusted data is just under 10% faster than the raw data, and only 4% faster since the start of ...Missing: instrumental | Show results with:instrumental
  76. [76]
    Major problems identified in data adjustments applied to a widely ...
    Feb 22, 2022 · For example, changes in the location of the weather station, the types of thermometers used or the growth of urban heat islands around the ...
  77. [77]
    Urban Heat Island Effects in U.S. Summer Surface Temperature ...
    Urban heat island (UHI) is a feature of human settlements, caused by replacing vegetation with buildings. UHI warming is 22% of observed warming in the US.
  78. [78]
    Are surface temperature records reliable? - Skeptical Science
    Adjusting for urban heat island effect. When compiling temperature records, NASA's GISS goes to great pains to remove any possible influence from urban heat ...Missing: issues | Show results with:issues<|separator|>
  79. [79]
    Microwave Sounding Unit Temperature Anomalies
    Global Microwave Sounding Unit (MSU) temperature anomaly records begin in January 1979. The University of Alabama in Huntsville (UAH) dataset extends through ...
  80. [80]
    Latest Global Temps « Roy Spencer, PhD
    Latest Global Average Tropospheric Temperatures​​ Since 1979, NOAA satellites have been carrying instruments which measure the natural microwave thermal ...
  81. [81]
    Global Temperature Report :: The University of Alabama in Huntsville
    Map Data Available for 1978 - 2005. QUESTIONS ABOUT THIS REPORT? For web related issues - Jennifer Geary For content - alabama.climate@nsstc.uah.edu. We try ...
  82. [82]
    Tropospheric temperature change since 1979 from tropical ...
    Mar 16, 2007 · At the 58 sites the UAH data indicate a trend of +0.08 K decade−1, the RSS data, +0.15. When the largest discontinuities in the sondes are ...
  83. [83]
    Argo and climate change
    The figure to the right shows temperature anomalies averaged over the global ocean, down to 1900 meters, since 2004 measured by Argo. While temperatures in the ...
  84. [84]
    Argo Ocean Temperature and Salinity Profiles - Climate Data Guide
    Jul 31, 2025 · The temperature sensors on Argo floats are considered stable and routinely perform within the stated accuracy of 0.002 degrees C. The pressure ...
  85. [85]
    Argo Program - Global Ocean Monitoring and Observing
    Argo provides scientists with measurements of the evolving state of the upper ocean by collecting temperature and salinity profiles from the surface to 2,000 ...
  86. [86]
    Global Surface Temperature Dataset Updated | News
    Feb 14, 2024 · It has data from 1850–Present and is presented on a 5X5 grid. NOAAGlobalTemp is a key component of the Global Climate Report which is updated ...
  87. [87]
    Past Climate | NOAA Climate.gov
    The word "proxy" indicates that these records stand in for direct measurements. One of the most easily recognized types of paleoclimate records is tree ring ...
  88. [88]
    Paleoclimatology: How Can We Infer Past Climates?
    Past climate can be reconstructed using a combination of different types of proxy records. These records can then be integrated with observations of Earth's ...
  89. [89]
    Paleoclimate Proxies | U.S. Geological Survey - USGS.gov
    Scientists combine proxy-based paleoclimate reconstructions with instrumental records (such as thermometer and rain gauge readings) to expand our understanding ...
  90. [90]
    Ice core basics - Antarctic Glaciers
    Through analysis of ice cores, scientists learn about glacial-interglacial cycles, changing atmospheric carbon dioxide levels, and climate stability over the ...
  91. [91]
    What do ice cores reveal about the past?
    Mar 24, 2023 · By preserving evidence of ancient temperatures and greenhouse gases, ice cores show scientists how much our planet has changed.
  92. [92]
    What are some of the limitations and strengths of each of climate ...
    Feb 28, 2024 · The prime limitation is that they are all more or less local. The second that they are proxy data, i.e. do not directly tell the temperature nor ...
  93. [93]
    Mapped: How 'proxy' data reveals the climate of the Earth's distant past
    Mar 29, 2021 · These different types of proxies have been captured by an array of palaeoclimate “archives”, such as sediments, ice cores and cave formations.
  94. [94]
    How accurate are climate proxies in giving us a clear picture of ...
    May 6, 2014 · Individual climate proxies are unreliable; combining multiple proxies helps, but they are not reliable for reliable readings, and have ...
  95. [95]
    On the Development and Use of Homogenized Climate Datasets
    At the National Climatic Data Center, two basic approaches to making homogeneity adjustments to climate data have been developed. The first is based on the ...
  96. [96]
    Understanding adjustments to temperature data - Skeptical Science
    Feb 26, 2015 · This will be the first post in a three-part series examining adjustments in temperature data, with a specific focus on the US land temperatures.Missing: controversies | Show results with:controversies
  97. [97]
    [PDF] On the reliability of the U.S. surface temperature record
    Jun 8, 2010 · Watts [2009], in particular, has speculated that U.S. surface temperature records from the USHCN from the last 30 years or so are likely biased ...
  98. [98]
    [PDF] Is the U.S. Surface Temperature Record Reliable?
    This report, by meteorologist Anthony Watts, presents the results of the first-ever compre- hensive review of the quality of data coming from the National ...
  99. [99]
    [PDF] Has poor station quality biased U.S. temperature estimates?
    Two independent surveys have found that about 70% of the thermometer stations in the U.S. Historical. Climatology Network (USHCN) dataset are currently ...<|separator|>
  100. [100]
    Analysis of the impacts of station exposure on the U.S. Historical ...
    Jul 30, 2011 · The unique opportunity offered by this completed survey permits an examination of the relationship between USHCN station siting characteristics ...
  101. [101]
    Yes, NOAA adjusts its historical weather data: Here's why - ABC News
    Feb 25, 2025 · NOAA makes adjustments to account for the urban heat island effect to ensure the temperature records reflect true regional climate trends ...
  102. [102]
    Evaluation of the Homogenization Adjustments Applied to European ...
    Evaluation of the Homogenization Adjustments Applied to European Temperature Records in the Global Historical Climatology Network Dataset. by. Peter O'Neill.Missing: peer | Show results with:peer
  103. [103]
    Assessing U.S. temperature adjustments using the Climate ...
    Feb 9, 2016 · USCRN does have a noticeably higher maximum temperature trend than both raw and adjusted USHCN data, though the cause of this is still unclear ...
  104. [104]
    Major correction to satellite data shows 140% faster warming since ...
    Jun 30, 2017 · RSS v4 shows about 5% more warming than the NASA record since 1979, when satellite observations began.
  105. [105]
    What Are Proxy Data? | News
    Apr 15, 2016 · Paleoclimatologists gather proxy data from natural recorders of climate variability such as corals, pollen, ice cores, tree rings, caves, pack ...
  106. [106]
    Paleoclimate - Climate Data Guide
    Dec 1, 2005 · Paleoclimate data are derived from Earth's natural climate archives, including tree rings, ice cores, corals, speleothems, and ocean sediments.
  107. [107]
    [PDF] Palaeoclimate - Intergovernmental Panel on Climate Change
    Ice core data indicate that CO2 varied within a range of 180 to 300 ppm and CH4 within 320 to 790 ppb over this period. Over the same period, antarctic ...
  108. [108]
    Paleo Data Search | Study
    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.
  109. [109]
    Asynchrony between Antarctic temperature and CO2 associated ...
    Mar 6, 2018 · Over the past 420 kyr, the Vostok ice core shows that the Antarctic δD temperatures lead the CO2 variations by 1.3 ± 1.0 kyr. During the ...
  110. [110]
    Glacial Cycles and Milankovitch Forcing - NASA ADS
    Using a recent conceptual model of the glacial-interglacial cycles we present more evidence of Milankovitch cycles being the trigger for retreat and forming ...
  111. [111]
    Proxy evidence for state-dependence of climate sensitivity in the ...
    Sep 7, 2020 · We find that Equilibrium Climate Sensitivity (ECS) was indeed higher during the warmest intervals of the Eocene, agreeing well with recent model simulations.
  112. [112]
    Spatial patterns of climate change across the Paleocene–Eocene ...
    The Paleocene–Eocene Thermal Maximum (PETM) is a global warming event that occurred 56 Ma in response to an increase in carbon dioxide.
  113. [113]
    Paleoclimatology: Climate Close-up - NASA Earth Observatory
    Dec 23, 2005 · Both tree rings and similar rings in ocean coral can tell scientists about rainfall and temperatures during a single growing season.<|separator|>
  114. [114]
    Reconstructing Holocene temperatures in time and space using ...
    Dec 15, 2022 · We use paleoclimate data assimilation to create a spatially complete reconstruction of temperature over the past 12 000 years using proxy data.
  115. [115]
    Holocene global mean surface temperature, a multi-method ... - Nature
    Jun 30, 2020 · The only previous GMST reconstruction for the Holocene based on multi-proxy data showed maximum warmth around 7000 ± 2000 years ago (7 ± 2 ka BP ...
  116. [116]
    The Holocene temperature conundrum - PNAS
    Aug 11, 2014 · Here, we show that climate models simulate a robust global annual mean warming in the Holocene, mainly in response to rising CO 2 and the retreat of ice sheets.Missing: peer- | Show results with:peer-
  117. [117]
    Reconstructing Holocene centennial cooling events - Frontiers
    Oct 29, 2024 · Proxy-based reconstructions of hemispheric and global surface temperature variations over the past two millennia. Proc. Natl. Acad. Sci. 105 ...
  118. [118]
    Holocene seasonal temperature evolution and spatial variability ...
    Sep 10, 2022 · Our reconstructions show that Holocene temperature evolution over the NH landmass in both summer and winter seasons was characterised by a ...
  119. [119]
    On the Estimation of Internal Climate Variability During the ...
    Jan 4, 2022 · We use climate model simulations to test methods used in past work to estimate internal climate variability, including the so-called “Atlantic Multidecadal ...Missing: peer- | Show results with:peer-<|control11|><|separator|>
  120. [120]
    An extended Arctic proxy temperature database for the past 2,000 ...
    Aug 19, 2014 · Proxy-based reconstructions of hemispheric and global surface temperature variations over the past two millennia. Proc. Natl Acad. Sci. 105 ...
  121. [121]
    2.3.3 Was there a Little Ice Age and a Medieval Warm Period
    However, viewed hemispherically, the Little Ice Age can only be considered as a modest cooling of the Northern Hemisphere during this period of less than 1°C ...
  122. [122]
    Volcanic and Solar Forcing of Climate Change during the ...
    The two primary forcings for climate in the preindustrial period, solar variations and volcanic eruptions, thus have quite different impacts in the model. For ...Abstract · Comparison of volcanic and... · External forcing of historical...
  123. [123]
    High-frequency climate forcing causes prolonged cold periods in the ...
    May 8, 2024 · Eleven long-lasting cold periods occurred in the Northern Hemisphere during the past 8000 years. These periods correlate with enhanced volcanic activity.
  124. [124]
    Shindell et al. 2003: Volcanic and solar forcing of climate change ...
    The climate response to variability in volcanic aerosols and solar irradiance, the primary forcings during the preindustrial era, is examinedMissing: Pre- industrial
  125. [125]
    World of Change: Global Temperatures - NASA Earth Observatory
    The average global temperature has increased by a little more than 1° Celsius (2° Fahrenheit) since 1880. Two-thirds of the warming has occurred since 1975.
  126. [126]
    Atmospheric Concentrations of Greenhouse Gases | US EPA
    Levels have risen since the 1920s, however, reaching a new high of 336 ppb in 2022 (average of four sites in Figure 3). This increase is primarily due to ...
  127. [127]
    How much is sea level rising? - Skeptical Science
    Aug 20, 2023 · The current rate of sea level rise is about 3.7 mm per year, or nearly 1.5 inches per decade, and is accelerating.
  128. [128]
    Glaciers - Copernicus Climate Change
    Apr 15, 2025 · Both globally and across Europe, glaciers have seen a substantial and prolonged loss of ice mass since the mid 19th century.
  129. [129]
    NASA Study: More Greenland Ice Lost Than Previously Estimated
    Jan 17, 2024 · A recent study found that glacial retreat caused the Greenland Ice Sheet to lose one-fifth more mass than previously estimated.Missing: 1850 | Show results with:1850
  130. [130]
    How was Arctic sea ice measured before the satellite era?
    In other words, multiple lines of evidence show that Arctic sea ice was in decline well before the start of the continuous satellite record in 1979. Indeed, sea ...
  131. [131]
    [PDF] Anthropogenic and Natural Radiative Forcing
    Since AR4, there has been new research on aerosol influences on the hydrologic cycle (also Sections 7.4, 7.6.4, 10.3.3.1 and 11.3.2.4.3). Increased aerosol ...Missing: peer | Show results with:peer
  132. [132]
    Long-term natural variability and 20th century climate change - PNAS
    Observations suggest the warming of the 20th century global mean surface temperature has not been monotonic, even when smoothed by a 10–20 year low-pass filter.
  133. [133]
    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.
  134. [134]
    Empirical assessment of the role of the Sun in climate change using ...
    Changes in solar activity could affect the climate system through several correlated mechanisms, including TSI forcing, UV forcing, cosmic ray forcing, solar ...
  135. [135]
    [PDF] Can the Warming of the 20th Century be Explained by Natural ...
    It is very unlikely that the 20th-century warming can be explained by natural causes. The late 20th century has been unusually warm.Missing: 21st | Show results with:21st
  136. [136]
    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.
  137. [137]
    Climate Models | NOAA Climate.gov
    Nov 21, 2014 · Climate models, also known as general circulation models or GCMs, use mathematical equations to characterize how energy and matter interact in ...
  138. [138]
    [PDF] Modeling the General Circulation of the Atmosphere. Topic 1
    Manabe and Bryan (1969):. ○ First coupled ...
  139. [139]
    General Circulation Models | METEO 469 - Dutton Institute
    GCMs describe the 3D atmosphere and Earth's climate, solving physics and chemistry equations. They can study various climate attributes.
  140. [140]
    [PDF] Numerical Methods for Climate Models
    Sep 8, 2025 · This is the CTCS scheme - Centred in Time, Centred in space, because the approximations for ∂φ/∂t and ∂φ/∂x are both centred.
  141. [141]
    2.2 Climate Modeling — UTCDW Guidebook
    A climate model is a computer program solving mathematical equations from physical laws, representing the climate system, and is a representation, not the ...
  142. [142]
    Q&A: How do climate models work? - Carbon Brief
    Jan 15, 2018 · Climate models use equations to represent the processes and interactions that drive the Earth's climate. These cover the atmosphere, oceans, land and ice- ...
  143. [143]
    a tutorial on climate model diagnostics and parameterization - GMD
    Sep 1, 2023 · Earth system models (ESMs) must represent processes below the grid scale of a model using representations (parameterizations) of physical ...
  144. [144]
    Deep learning to represent subgrid processes in climate models
    Sep 6, 2018 · The representation of nonlinear subgrid processes, especially clouds, has been a major source of uncertainty in climate models for decades.
  145. [145]
    66. Clouds are hard - Geophysical Fluid Dynamics Laboratory - NOAA
    Feb 16, 2016 · ... climate models with uncertain “sub-grid closures”. Unfortunately, the treatment of moist convection affects cloud feedbacks. At GFDL we have ...Missing: limitations | Show results with:limitations
  146. [146]
    Confronting Earth System Model trends with observations - Science
    Mar 12, 2025 · This review covers the state of the science on the ability of models to represent historical trends in the climate system.
  147. [147]
    Persistent Discrepancies between Observed and Modeled Trends in ...
    It is shown that the latest generation of models persist in not reproducing the observations-based SST trends as a response to radiative forcing.
  148. [148]
    Global Warming: Observations vs. Climate Models
    Jan 24, 2024 · The observed rate of global warming over the past 50 years has been weaker than that predicted by almost all computerized climate models.
  149. [149]
    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 ...
  150. [150]
    Discrepancies between observations and climate models of large ...
    Nov 14, 2022 · We show that recent changes involving mid-to-upper-tropospheric anticyclonic wind anomalies – linked with tropical forcing – explain half of the observed ...
  151. [151]
    Comparing Tropospheric Warming in Climate Models and Satellite ...
    We assess the validity of two highly publicized claims: that modeled tropospheric warming is a factor of 3–4 larger than in satellite and radiosonde ...
  152. [152]
    Internal variability and forcing influence model–satellite differences ...
    Climate-model simulations exhibit approximately two times more tropical tropospheric warming than satellite observations since 1979.
  153. [153]
    Multi-decadal climate variability and satellite biases have amplified ...
    Jun 21, 2024 · Most coupled model simulations substantially overestimate tropical tropospheric warming trends over the satellite era, undermining the ...
  154. [154]
    Arctic September Sea Ice Concentration Biases in CMIP6 Models ...
    CMIP6 models show large biases in Arctic sea ice, related to atmospheric and oceanic circulations. Two main modes explain 65% of the intermodel variance.
  155. [155]
    [PDF] Differences in CMIP model simulations of central US precipitation
    explain differences between model-simulated precipitation and observations. Particular areas of emphasis for the diagnostic analysis will be the low-level ...<|separator|>
  156. [156]
    Observed and CMIP6 Modeled Internal Variability Feedbacks and ...
    Dec 5, 2022 · In Summary, both CMIP3 and CMIP5 have shown feedback discrepancies with observations, and although Figure S1 in Supporting Information S1 shows ...
  157. [157]
    Chapter 4 | Climate Change 2021: The Physical Science Basis
    This chapter assesses simulations of future global climate change, spanning time horizons from the near term (2021–2040), mid-term (2041–2060), and long term ( ...
  158. [158]
    Opinion: Can uncertainty in climate sensitivity be narrowed further?
    Feb 29, 2024 · While current climate sensitivity distributions are valid, further narrowing is possible through better understanding of climate processes and ...<|separator|>
  159. [159]
    Clouds study finds that low climate sensitivity is 'extremely unlikely'
    Jul 21, 2021 · The study finds a central cloud feedback estimate of 0.43W/m2/C with a 90% confidence range of ±0.35. This result reduces the uncertainty in ...
  160. [160]
    Implications of a Pervasive Climate Model Bias for Low‐Cloud ...
    Oct 19, 2024 · How low clouds respond to warming constitutes a key uncertainty for climate projections. Here we observationally constrain low-cloud ...
  161. [161]
    On the relationship between aerosol model uncertainty and radiative ...
    The largest uncertainty in the historical radiative forcing of climate is caused by the interaction of aerosols with clouds.
  162. [162]
    Reducing Aerosol Forcing Uncertainty by Combining Models With ...
    May 3, 2023 · Aerosol forcing uncertainty represents the largest climate forcing uncertainty overall. Its magnitude has remained virtually undiminished ...
  163. [163]
    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.
  164. [164]
    Higher-Resolution Climate Models Do Not Consistently Reproduce ...
    We demonstrate that even higher-resolution models do not capture the observed warming pattern of the tropical Pacific Ocean that is closer to observations.
  165. [165]
    Halving of the uncertainty in projected warming over the past decade
    Jun 22, 2024 · Here we show that through progress in climate policy and climate science, these uncertainties have decreased dramatically over the past decade.
  166. [166]
    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 ...
  167. [167]
    Climate-induced forest dieback drives compositional changes in ...
    Jan 18, 2022 · We conclude that forest dieback drives changes in species assemblages that mimic natural forest succession, and markedly increases the risk of catastrophic ...
  168. [168]
    Climate, diseases driving death of several tree species in multiple ...
    Sep 15, 2025 · “The paper proves that latent pathogens are really driving tree decline and climate alone is not responsible – yet – for the dieback,” said ...<|separator|>
  169. [169]
    Climate change and the global redistribution of biodiversity
    Apr 11, 2023 · Range shifts have the potential to reshape ecological communities, alter ecosystem functions and the provision of ecosystem services, impact ...
  170. [170]
    Long-term empirical evidence of ocean warming leading to ... - PNAS
    Nov 14, 2016 · This study shows that an increase in the proportion of warmwater species (“tropicalization”) as oceans warm is increasing fish herbivory in kelp forests.
  171. [171]
    Climate change is not the principal driver of biodiversity loss
    Jan 20, 2022 · The current perception that climate change is the principal threat to biodiversity is at best premature. Although highly relevant, it detracts focus and effort.
  172. [172]
    US Imperiled species and the five drivers of biodiversity loss
    Apr 24, 2025 · They did not include climate change in their assessment, but the next most common factors were invasive species, pollution, and overexploitation ...
  173. [173]
    What is coral bleaching? - NOAA's National Ocean Service
    Jun 16, 2024 · WHAT CAUSES BLEACHING? Change in ocean temperature Increased ocean temperature caused by climate change is the leading cause of coral bleaching.
  174. [174]
    [PDF] Coral Bleaching – A Review of the Causes and Consequences
    On severely damaged reefs, recovery is dependent on the arrival of suitable coral larvae that have survived the bleaching event elsewhere, and their successful.
  175. [175]
    Impacts of ocean acidification on marine fauna and ecosystem ...
    The ability of marine animals, most importantly pteropod molluscs, foraminifera, and some benthic invertebrates, to produce calcareous skeletal structures is ...Abstract · Introduction · Effects of elevated pCO2 on... · Increased pCO2 and other...
  176. [176]
    Is Ocean Acidification Really a Threat to Marine Calcifiers? A ...
    Aug 7, 2022 · In concordance with the prevailing paradigm based on seawater carbonate chemistry, ocean acidification was shown to pose negative effects, ...
  177. [177]
    The direct drivers of recent global anthropogenic biodiversity loss
    Nov 9, 2022 · Climate change is probably the most rapidly intensifying threat to biodiversity, and its impacts are increasingly well quantified (26), but ...
  178. [178]
    How many people die from extreme temperatures, and how this ...
    Jul 1, 2024 · In total, annual temperature-related deaths had fallen by around 650,000 per year. This drop can't entirely be attributed to climate change, as ...
  179. [179]
    [PDF] Adapting to Climate Change: The Remarkable Decline in the US
    This paper examines the temperature-mortality relationship over the course of the 20th century. US both for its own interest and to identify potentially ...
  180. [180]
    Adaptation and the Mortality Effects of Temperature Across U.S. ...
    Using 22 years of Medicare data, we find that both cold and hot days increase mortality. However, hot days are less deadly in warm places while cold days are ...
  181. [181]
    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 ( ...
  182. [182]
    Climate change impacts on crop yield and quality with CO2 ...
    In general, climate change itself tends to reduce crop yield but the fertilization effect of CO2 tends to increase yield.
  183. [183]
    [PDF] CO₂ FERTILIZATION OF US FIELD CROPS Charl
    We argue that rising atmospheric CO₂ is responsible for a significant share of these yield gains. We present a novel methodology to estimate the CO₂.
  184. [184]
    Billion-Dollar Weather and Climate Disasters
    The US sustained 403 weather and climate disasters from 1980–2024 where overall damages/costs reached or exceeded $1 billion (including CPI adjustment to 2024).Events · Summary Stats · Time Series · Disaster Mapping
  185. [185]
    extreme weather events cost economy $2 trillion over the last decade
    Nov 11, 2024 · A new report estimates that climate-related extreme weather events have cost the global economy more than $2 trillion over the past decade.
  186. [186]
    RELEASE: WRI Study Finds Climate Adaptation Investments Yield ...
    Jun 3, 2025 · Every $1 invested in adaptation and resilience generates more than $10 in benefits over ten years. This translates to potential returns of over $1.4 trillion, ...
  187. [187]
    The Economic Effects of Climate Change
    I review the literature on the economic impacts of climate change, an externality that is unprecedentedly large, complex, and uncertain.
  188. [188]
    Climate Change and the U.S. Economic Future - EPIC
    Climate change will harm the U.S. economy, potentially losing 1-4% of GDP by the end of the century, with the poorest areas experiencing greater losses.
  189. [189]
    Country-Specific Effects of Climate Variability on Human Migration
    Abstract. Involuntary human migration is among the social outcomes of greatest concern in the current era of global climate change.
  190. [190]
    Global economic impact of weather variability on the rich and the poor
    Sep 13, 2024 · Empirical studies on the relationship between inequality and climate change have identified a regressive impact of heat extremes and ...
  191. [191]
    Positive Societal Response to Past Climate Variability Sets an ...
    Mar 24, 2021 · Scholars have often argued that climatic changes plunge communities into crisis and create conditions that lead societies to collapse, but a ...
  192. [192]
    A meta-analysis of the total economic impact of climate change
    Only the impact of climate change on time use is positive. Overall, market impacts make up 54% of the total, while the remaining 46% directly affect welfare. On ...
  193. [193]
    Objectively combining climate sensitivity evidence | Climate Dynamics
    Sep 19, 2022 · This study evaluates the methodology of and results from a particularly influential assessment of climate sensitivity that combined multiple lines of evidence.
  194. [194]
    The Impact of Recent Forcing and Ocean Heat Uptake Data on ...
    Using a standard global energy budget approach, this paper seeks to clarify the implications for climate sensitivity (both ECS and TCR) of incorporating the ...
  195. [195]
    [PDF] ENERGY & ENVIRONMENT
    Douglass et al. [2007] examined these tropical discrepancies by specifically comparing upper air temperature trends between the “model-average” available at ...Missing: peer | Show results with:peer
  196. [196]
    [PDF] Evaluation of Climate Models
    very high significance levels of model–observation discrepancies in LT and MT trends that were obtained in some studies (e.g., Douglass et al., 2008 ...Missing: peer | Show results with:peer
  197. [197]
    On the climate sensitivity and historical warming evolution in recent ...
    Jul 6, 2020 · Here we compare the CMIP5 and CMIP6 model ensembles and find a systematic shift between CMIP eras to be unexplained as a process of random sampling.
  198. [198]
    Identifying climate model structural inconsistencies allows for tight ...
    Aug 8, 2023 · We show that structural deficiencies in a climate model, revealed as inconsistencies among observationally constrained cloud properties in the model,<|control11|><|separator|>
  199. [199]
    18 Spectacularly Wrong Predictions Were Made Around the Time of ...
    Apr 21, 2022 · Here are 18 examples of the spectacularly wrong predictions made around 1970 when the “green holy day” (aka Earth Day) started.
  200. [200]
    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.
  201. [201]
    Wrong Again: 50 Years of Failed Eco-pocalyptic Predictions
    Sep 18, 2019 · None of the apocalyptic predictions with due dates as of today have come true. What follows is a collection of notably wild predictions from notable people in ...
  202. [202]
    Increasing development, reducing inequality, the impact of climate ...
    Climate policies also have costs that often vastly outweigh their climate benefits. The Paris Agreement, if fully implemented, will cost $819–$1,890 billion per ...
  203. [203]
    False Alarm: How Climate Change Panic Costs Us Trillions, Hurts ...
    30-day returnsChildren panic about their future, and adults wonder if it is even ethical to bring new life into the world. Enough, argues bestselling author Bjorn Lomborg.
  204. [204]
    False Alarm - Bjorn Lomborg
    Lomborg does not lack solutions. In False Alarm, he advocates a range of cost-benefit tested policies to address both climate change and global poverty....
  205. [205]
    Britain's Disastrous Path to Net Zero Is a Warning to the U.S.
    Feb 6, 2024 · Britain was conned into net zero by deceptive and illusory promises of cheap renewable power. The results have been an economic disaster.
  206. [206]
    Expensive climate policy is dead — and it could be an immense ...
    Nov 26, 2024 · The “net zero” green agenda, based on massive subsidies and expensive legislation, will likely cost $27 trillion per year across the century, ...
  207. [207]
    Visiting Fellow Bjorn Lomborg Analyzes The Financial Costs And ...
    Mar 16, 2020 · Lomborg argued that climate-change policy has largely been ineffective over the past 20 years given the hundreds of billions in dollars spent.
  208. [208]
    The new climate discourse: Alarmist or alarming? - ScienceDirect.com
    This article reviews evidence to support claims that climate change can be viewed as 'catastrophic', 'rapid', 'urgent', 'irreversible', 'chaotic', and 'worse ...
  209. [209]
    [PDF] Refuting 12 Claims Made by Climate Alarmists
    FACT: A NOAA scientist rejected global warming link to tornados stating “no specific consensus or connection between global warming and tornadic activity.”.
  210. [210]
    Expensive climate policy is dead. That's a big opportunity
    Nov 28, 2024 · Money saved as countries back away from Green New Deals can fund research into making green fuel cheap, which will make it popular. Read on.