Fact-checked by Grok 2 weeks ago

Radiative forcing

Radiative forcing is the change in the net downward in the atmosphere due to a in a driver, such as altered concentrations of gases or aerosols, quantified at the after adjustment for stratospheric equilibrium while fixing tropospheric and surface conditions. This metric, expressed in watts per square meter (W/m²), quantifies the radiative imbalance that drives changes in Earth's surface , with positive forcing indicating a net energy gain leading to warming and negative forcing a net loss leading to cooling. The concept originated from early modeling efforts to isolate the direct radiative effects of atmospheric constituents independent of rapid feedbacks like changes. Since the pre-industrial period (circa 1750), greenhouse gases have exerted a dominant positive effective radiative forcing of about 3.2 W/m² as of 2023, primarily from (≈2.2 W/m²), , and , calculated using logarithmic formulas such as ΔF = 5.35 × ln(C/C₀) for CO₂ where C is the current concentration and C₀ the pre-industrial value. This warming influence is partially offset by negative forcing from s (≈ -1.0 to -1.5 W/m²) and changes, yielding a net effective radiative forcing of roughly 1.0 to 1.5 W/m², though recent reductions in aerosol emissions have increased this net value by enhancing clear-sky radiation. Natural forcings, including solar variability (≈0.05 W/m² over the ) and volcanic eruptions (episodic negative spikes), contribute smaller, shorter-term perturbations compared to sustained trends. Key characteristics include the distinction between instantaneous radiative forcing (before any atmospheric adjustment) and effective radiative forcing (incorporating rapid adjustments like responses), with the latter better predicting changes but introducing greater , particularly for s where estimates vary widely due to incomplete observational constraints. Controversies persist over the magnitude of aerosol forcing, with some peer-reviewed analyses highlighting potential overestimation of cooling effects in models relative to satellite-inferred trends, underscoring the need for empirical validation beyond simulations. Radiative forcing remains a foundational tool in assessment, linking atmospheric composition changes to budget imbalances and projected warming, though its application assumes linear that may not fully capture nonlinear feedbacks or historical discrepancies between modeled and observed responses.

Conceptual Foundations

Definition and Physical Principles

Radiative forcing quantifies the perturbation to Earth's top-of-atmosphere caused by changes in atmospheric constituents or surface properties, expressed as the change in net downward (shortwave plus longwave) at the . This metric assumes fixed sea surface temperatures, tropospheric temperatures, and other variables, while allowing stratospheric temperatures to adjust rapidly to , typically within months. Positive forcing (e.g., from increased greenhouse gases) implies an surplus driving planetary warming until outgoing increases to restore ; negative forcing (e.g., from reflective aerosols) implies a leading to cooling. Values are computed as global and annual means in watts per square meter (W m⁻²), providing a standardized measure for comparing forcing agents independent of slower feedbacks like changes or ice-albedo shifts. The physical basis stems from planetary , where absorbed —approximately 240 W m⁻² after averaging the (about 1361 W m⁻² over Earth's cross-section and subtracting planetary (~0.3))—balances emitted longwave from the surface and atmosphere. A alters this balance by modifying , emission, or scattering of : greenhouse gases trap outgoing longwave , increasing downward flux; aerosols can reflect incoming shortwave or absorb it, reducing net . At the , the forcing isolates the direct radiative effect before tropospheric dynamical responses (e.g., ) amplify or dampen it, ensuring comparability across agents. This boundary avoids conflating surface fluxes with atmospheric adjustments, as tropopause-level flux changes directly link to global temperature response via ΔT ≈ λ ΔF, where λ is the no-feedback sensitivity parameter (~1.2 K per W m⁻² from blackbody physics). Derivations from radiative transfer principles involve solving the Schwarzschild equation for photon transport, integrating absorption and emission coefficients over spectral bands, pressures, and temperatures. For well-mixed gases like CO₂, forcing scales logarithmically with concentration due to saturation in strong absorption bands, shifting effective absorption to weaker wings: ΔF ≈ 5.35 ln(C/C₀) W m⁻² for CO₂ changes from preindustrial C₀ = 278 ppm. Such calculations use line-by-line models validated against observations, confirming forcing independence from surface temperature in the fixed-temperature approximation. Uncertainties arise from spectral data, vertical profiles, and cloud overlaps, but core principles hold across line-shape assumptions (e.g., Lorentzian broadening).

Historical Development

The concept of radiative forcing emerged from early investigations into the radiative balance of Earth's atmosphere, with foundational insights dating to the late . In 1896, published calculations quantifying the temperature response to changes in atmospheric concentration, estimating that doubling CO2 would increase global surface temperatures by 5–6 °C through enhanced absorption of infrared radiation, an effect akin to modern radiative forcing computations. These estimates relied on rudimentary data and assumed equilibrium , predating explicit formulations but establishing a causal link between perturbations and net radiative imbalance at the surface. Mid-20th-century advancements built on Arrhenius's work amid growing of CO2 increases. In 1938, revisited Arrhenius's model, compiling global temperature and CO2 records to argue for a detectable 0.005 °C per decade warming driven by industrial emissions, implicitly incorporating radiative forcing by correlating CO2 rise with altered trapping. Post-World War II developments in theory, enabled by electronic computers, allowed more sophisticated modeling; for instance, in 1967, and Richard Wetherald's radiative-convective equilibrium simulations quantified CO2 doubling's tropospheric warming while holding stratospheric temperatures fixed, refining the perturbation-response framework central to forcing estimates. The term "radiative forcing" was formalized in the late 20th century to standardize climate impact assessments. The 1975 Charney Report on carbon dioxide and climate introduced equilibrium climate sensitivity as the temperature change per unit forcing (typically for doubled CO2), linking it to radiative perturbations at the tropopause. The Intergovernmental Panel on Climate Change (IPCC) adopted and defined the concept in its 1990 First Assessment Report as the change in net downward radiative flux at the tropopause following instantaneous forcing and allowing stratospheric temperature adjustment, but excluding tropospheric and surface responses, to isolate direct agent effects from feedbacks. This definition evolved through subsequent IPCC reports, incorporating agent-specific quantifications (e.g., +2.1 W m⁻² from well-mixed greenhouse gases since 1750 by 2001 estimates) and addressing uncertainties in aerosols and indirect effects.

Distinction from Feedbacks and Metrics

Radiative forcing quantifies the perturbation to the Earth's top-of-atmosphere (TOA) radiative energy balance caused by an external driver, such as increased concentrations, excluding subsequent responses like surface changes. This initial imbalance, typically measured in watts per square meter (W m⁻²), precedes feedbacks, which are amplifying or dampening effects arising from internal adjustments to warming, such as increased (positive feedback) or enhanced low-cloud cover (potential negative feedback). The distinction ensures that forcing isolates the direct causal input, while feedbacks represent the system's endogenous , allowing modular analysis of response where equilibrium surface change approximates ΔT_s ≈ λ ΔF, with λ as the parameter incorporating net strength. Climate feedbacks, by contrast, emerge after the forcing induces tropospheric and surface alterations, including changes, shifts from ice melt, and cloud-radiative effects tied to gradients rather than the forcing agent itself. For instance, feedback amplifies forcing by ~1.8 W m⁻² per of warming due to the Clausius-Clapeyron , but this is excluded from forcing calculations to avoid conflating cause and response. Rapid atmospheric adjustments, like stratospheric cooling or tropospheric circulation shifts occurring on timescales faster than ocean mixing (days to months), are sometimes incorporated into "effective radiative forcing" (ERF) to better predict outcomes, but these are delineated from slower, temperature-mediated feedbacks involving deep ocean heat uptake or changes. Among metrics, instantaneous radiative forcing omits all adjustments, yielding higher values (e.g., ~2.16 W m⁻² for doubled CO₂), while stratosphere-adjusted forcing accounts for in the stratosphere alone, reducing it to ~1.68 W m⁻²; ERF further includes tropospheric rapid responses, approximating ~1.5–2.0 W m⁻² depending on model ensembles. These variants serve distinct purposes: traditional forcing compares agent efficacies linearly, but ERF correlates more strongly with simulated across models, as it embeds non-feedback adjustments without surface temperature feedbacks. in distinguishing adjustments from feedbacks arises from model divergences in and responses, with ERF estimates for forcing ranging 1.0–2.5 W m⁻² since 1750, emphasizing the need for observational constraints like satellite-derived TOA fluxes.

Estimation Methods

Radiative Transfer Modeling

Radiative transfer modeling computes radiative forcing by solving the equations governing the propagation of solar and terrestrial radiation through the atmosphere, accounting for gaseous , , and . These models evaluate the difference in net downward at the between unperturbed and perturbed atmospheric states, typically fixing tropospheric temperatures and profiles while permitting stratospheric temperatures to adjust to a new . This fixed dynamical heating approximation isolates the direct radiative perturbation from feedbacks like tropospheric warming. Line-by-line (LBL) models achieve highest fidelity by explicitly resolving millions of individual molecular lines from spectroscopic databases such as HITRAN or GEISA, integrating monochromatic fluxes over the full spectrum from to . Used as benchmarks in intercomparisons like the Radiative Forcing Model Intercomparison (RFMIP), LBL codes such as LBLRTM demonstrate consistency across independent implementations, with forcing uncertainties for long-lived gases below 5% for CO2 doublings yielding approximately 3.7 W m⁻² in clear-sky conditions. approximations, including correlated-k methods, accelerate computations by sorting and reweighting absorption coefficients within spectral bands, enabling their integration into general circulation models (GCMs) while retaining accuracy within 1-2% of LBL results for well-mixed gases. Prominent models include the Rapid Radiative Transfer Model (RRTM), optimized for GCM time steps with validated longwave and shortwave schemes against LBL benchmarks, and MODTRAN, a moderate-resolution code originally developed for that simulates band-averaged transmittances for forcing estimates. For instance, MODTRAN calculations for a CO2 doubling from pre-industrial levels produce a stratospheric-adjusted forcing of about 3.7 W m⁻², aligning with empirical parameterizations like ΔF = 5.35 ln(C/C₀) derived from similar integrations. These models incorporate vertical profiles of , , and from reanalyses or standard atmospheres (e.g., mid-latitude summer), with clear-sky assumptions for direct forcing or effective radiative forcing including adjustments via double-call methods. Uncertainties arise from spectroscopic (e.g., line intensities ±5-10%), continuum absorption in far-IR windows, and minor cloud-aerosol overlaps, but inter-model spreads for GHG forcing remain below 0.2 W m⁻² in recent benchmarks. Validation against aircraft or observations, such as from sites, confirms model accuracy for clear-sky fluxes within 1-3 W m⁻², though GCM-embedded schemes require tuning to match LBL-derived forcings for historical simulations. Ongoing refinements address spectral gaps in continua and non-LTE effects in the upper atmosphere, ensuring robust quantification of forcing agents like CO2, where in strong bands shifts sensitivity to weak peripheral lines.

Observational Approaches

Satellite-based measurements constitute the primary observational approach for estimating global radiative forcing, capturing top-of-atmosphere (TOA) fluxes of incoming solar shortwave radiation and outgoing longwave radiation to quantify perturbations in Earth's energy balance. The Earth Radiation Budget Experiment (ERBE), operational from 1984 to 1990, provided foundational broadband radiance data, enabling initial assessments of the planetary radiation budget with an accuracy of approximately 1% for monthly global means. Its successor, the Clouds and the Earth's Radiant Energy System (CERES), launched in 1997 aboard NASA's Tropical Rainfall Measuring Mission and subsequent platforms like Terra and Aqua, delivers higher-precision observations with radiometric calibration stability better than 0.3% per decade, facilitating detection of forcing trends on the order of 0.1 W m^{-2}. CERES data products, such as Energy Balanced and Filled (EBAF), adjust angular distribution models and incorporate geostationary imager data to produce gridded TOA fluxes at 1° resolution, supporting estimates of Earth's energy imbalance (EEI) as a direct indicator of net radiative forcing integrated over rapid adjustments. To disentangle radiative forcing from climate feedbacks in these datasets, the radiative is employed, which approximates the change in TOA flux due to external perturbations (e.g., concentrations or ) while holding state variables like and fixed, using precomputed sensitivities from models. Applied to CERES observations from 2001 onward, this method isolates forcing signals; for example, analysis of CERES-EBAF data from 2003 to 2018 indicated a global mean increase in effective radiative forcing of 0.21 ± 0.15 W m^{-2} per decade, consistent with rising influences. More recent integrations combine CERES-derived EEI with to predict radiative responses to observed surface warming, yielding observationally constrained effective forcing estimates that align with model-based assessments within uncertainties of ±0.5 W m^{-2}. These approaches prioritize TOA imbalances over surface measurements due to the former's direct linkage to global forcing definitions, though they require corrections for instrumental drift and cloud contamination, validated against from networks like the World Radiation Monitoring Center. Ground-based and in-situ observations supplement data, particularly for regional aerosol direct forcing, by measuring surface , aerosol , and vertical profiles via sun photometers and radiometers. Networks such as the Aerosol Robotic Network (AERONET) provide column-integrated data used to compute surface shortwave forcing efficiencies, with studies reporting values of -47.4 W m^{-2} τ^{-1} (where τ is ) under clear-sky conditions in polluted regions. Closure experiments, comparing measured fluxes to those simulated from concurrent optical and microphysical observations, refine these estimates; for instance, events have been quantified using radiometers to derive forcing offsets of up to +0.5 W m^{-2} at TOA. However, ground-based methods are inherently local and less suited for global forcing due to sparse coverage, serving mainly for validation of retrievals and parameterization of sub-grid processes like aerosol-cloud interactions. Uncertainties in observational forcing arise from sampling biases (e.g., diurnal cycle undersampling in polar regions) and assumptions, typically ranging 10-20% for components but lower (<5%) for well-mixed greenhouse gases when corroborated across platforms.

Uncertainty Quantification

Quantification of uncertainties in radiative forcing estimates involves assessing ranges from radiative transfer models, multi-model ensembles, and observational constraints, often expressed as 5–95% confidence intervals derived from Monte Carlo simulations or inter-model spreads. The dominant source of uncertainty stems from aerosols, particularly aerosol-cloud interactions (ERFaci), which contribute the largest spread in total effective radiative forcing (ERF), while well-mixed greenhouse gas (GHG) forcings exhibit narrower ranges due to precise concentration measurements and radiative efficiencies. Total anthropogenic ERF from 1750 to 2019 is assessed at 2.72 W m⁻² with a 5–95% range of [1.96–3.48] W m⁻², where aerosol contributions account for much of the variance. For GHGs, uncertainties arise mainly from emission inventories, atmospheric lifetimes, and rapid adjustments like tropospheric temperature changes, but these are relatively low; for example, CO₂ ERF is 2.16 [1.90–2.41] W m⁻², reflecting ±10% uncertainty in radiative efficiency and concentration data. Methane (CH₄) and nitrous oxide (N₂O) forcings carry higher relative uncertainties (±20% and ±16%, respectively) due to indirect effects on ozone and stratospheric water vapor. Halogenated gases add 0.41 [0.33–0.49] W m⁻² with ±19–26% uncertainty from chemical adjustments. In contrast, aerosol ERF is -1.1 [-1.7 to -0.4] W m⁻², dominated by ERFaci at -1.0 [-1.7 to -0.3] W m⁻², stemming from model divergences in cloud microphysics, aerosol activation, and clean-sky condition representations. Aerosol-radiation interactions (ERFari) contribute -0.3 [-0.6 to 0.0] W m⁻², with additional spread from vertical distribution and pre-industrial emission assumptions.
Forcing ComponentBest Estimate ERF (W m⁻², 1750–2019)5–95% Uncertainty Range (W m⁻²)Primary Uncertainty Sources
Total GHGs3.84[3.46–4.22]Emission estimates, radiative efficiencies
CO₂2.16[1.90–2.41]Concentration measurements, spectral line data
Aerosols (total)-1.1[-1.7 to -0.4]Cloud interactions, emissions variability
ERFaci-1.0[-1.7 to -0.3]Aerosol activation, clean conditions
Observational approaches, such as satellite-derived top-of-atmosphere (TOA) fluxes from or aerosol optical depth from , constrain models but introduce uncertainties from instrument calibration, sampling biases, and inability to isolate forcing from feedbacks; for instance, direct aerosol radiative forcing estimates vary by up to 50% across retrieval methods due to assumptions in surface albedo and cloud masking. Radiative transfer modeling uncertainties include spectral resolution limitations and parameterization of microphysical processes, with inter-model spreads in ensembles amplifying aerosol ERF variance by factors of 2–3. Volcanic and solar forcings have smaller uncertainties (±5–10% for stratospheric aerosols per unit optical depth), but natural variability complicates attribution. Efforts to reduce uncertainty, such as fixed-sea-surface-temperature simulations and emergent constraints from historical energy budget closure, have narrowed ranges since prior assessments, yet aerosol-cloud effects remain the principal limiter to precise quantification.

Forcing Agents

Greenhouse Gas Contributions

Greenhouse gases exert the primary positive forcing in the Earth's energy budget imbalance, with well-mixed long-lived greenhouse gases (WMGHGs) contributing an assessed effective radiative forcing (ERF) of approximately 3.3 W m⁻² from 1750 to 2019 relative to pre-industrial conditions. This total arises predominantly from anthropogenic emissions, with carbon dioxide (CO₂) providing the largest share at 2.16 W m⁻² (likely range 1.82–2.50 W m⁻²), equivalent to about 65% of the WMGHG total. Methane (CH₄) follows at 0.54 W m⁻² (0.43–0.65 W m⁻²), nitrous oxide (N₂O) at 0.21 W m⁻² (0.17–0.25 W m⁻²), and halogenated compounds (including chlorofluorocarbons and hydrofluorocarbons) at 0.41 W m⁻² (0.35–0.47 W m⁻²). These estimates derive from radiative transfer models calibrated against spectroscopic data and atmospheric measurements, accounting for overlapping absorption bands and indirect effects like methane's influence on tropospheric ozone and stratospheric water vapor. The logarithmic dependence of CO₂ forcing on concentration—approximated as ΔF = 5.35 × ln(C/C₀) W m⁻², where C is the current concentration and C₀ the pre-industrial value—explains its dominant role, as concentrations have risen from ~280 ppm to over 420 ppm by 2024. and N₂O exhibit near-linear forcing responses over observed ranges, but their shorter atmospheric lifetimes (decades for , over a century for N₂O) result in smaller cumulative effects despite rapid emission growth. Halogenated gases, phased under the , peaked mid-century but continue contributing due to long persistence, with hydrofluorocarbons rising post-CFC restrictions. Observations from networks like NOAA's Global Monitoring Laboratory confirm ongoing increases, with total LLGHG forcing rising 51.5% from 1990 to 2023, 81% attributable to CO₂. Tropospheric ozone, while a greenhouse gas, is treated separately as a short-lived climate forcer with forcing linked to precursor emissions rather than direct concentration changes. Water vapor, the most abundant greenhouse gas, acts primarily as a feedback amplifying initial forcings rather than a direct agent, as its atmospheric abundance responds to temperature perturbations via the Clausius-Clapeyron relation. Stratospheric water vapor adjustments from CH₄ oxidation add a minor direct forcing component (~0.05–0.10 W m⁻²). Uncertainties in GHG ERF stem mainly from spectral line data and vertical profile assumptions, with 5–95% ranges typically ±10–20% for individual gases. Recent updates to 2024 indicate continued growth, with the NOAA Annual Greenhouse Gas Index reaching 1.54 relative to 1990, implying total WMGHG forcing exceeding 3.5 W m⁻² from pre-industrial levels.

Carbon Dioxide Effects

The primary mechanism by which carbon dioxide (CO₂) exerts radiative forcing is through absorption of outgoing longwave radiation in its principal vibrational bands centered around 15 μm and weaker bands near 4.3 μm and 2.7 μm, reducing the flux escaping to space and thereby perturbing Earth's energy balance. This effect is quantified as a positive forcing, with the increase in atmospheric CO₂ concentration from pre-industrial levels of 278 ppm in 1750 to 422.7 ppm in 2024 contributing the dominant share of anthropogenic greenhouse gas forcing. The logarithmic scaling of this forcing with concentration—ΔF = 5.35 × ln(C / C₀) W m⁻², where C is the current concentration and C₀ is the reference (pre-industrial) value—arises from the physics of molecular spectroscopy: strong central absorption lines saturate at lower concentrations, shifting marginal contributions to the unsaturated wings of the bands as levels rise, alongside minor shortwave absorption effects. For 2024 concentrations, this yields a direct forcing of approximately 2.20 W m⁻² relative to 1750. ![{\displaystyle \Delta F=5.35\times \ln {C_{0}+\\Delta C \over C_{0}}~~\\mathrm {W} ~\\mathrm {m} ^{-2}\,}][center] This formula, derived from line-by-line radiative transfer calculations across multiple models including shortwave effects, has an estimated uncertainty of ±10% for well-mixed conditions, primarily from spectroscopic data and vertical profile assumptions. The forcing includes stratospheric temperature adjustment, which for CO₂ slightly reduces the instantaneous value due to enhanced emission from the cooling stratosphere, but the net adjusted forcing remains close to the surface-level perturbation. CO₂'s well-mixed nature and atmospheric lifetime exceeding centuries ensure its forcing persists globally, with minimal regional variability beyond latitude-dependent profiles. Modal simulations, such as those using , confirm the integrated forcing for a doubling of CO₂ (to 560 ppm) at around 3.7 W m⁻², aligning with the formula's prediction of 5.35 × ln(2) ≈ 3.71 W m⁻². Overlap with water vapor absorption partially masks CO₂'s central band effects in the troposphere, but CO₂ dominates in clear-sky conditions over arid regions and in the wings where water vapor is weaker, contributing uniquely to the total longwave opacity. Recent analyses indicate no significant state dependence in the forcing-concentration relationship under current climate conditions, though higher temperatures could modestly enhance it via pressure broadening of lines. Empirical validations from satellite observations of outgoing longwave radiation trends corroborate the model's predicted spectral fingerprint of CO₂ forcing, including reduced radiance in the 12–16 μm window. The forcing's attribution to human activities is supported by isotopic ratios (depleted ¹³C/¹²C) and the correlation with fossil fuel emissions since the Industrial Revolution.

Other Trace Gases

Methane (CH₄), the second most important anthropogenic greenhouse gas after carbon dioxide, has increased from pre-industrial concentrations of approximately 0.73 ppm to 1.92 ppm by 2023, primarily due to emissions from agriculture (enteric fermentation and rice cultivation), fossil fuel extraction and use, and biomass burning. Its effective radiative forcing (ERF) from 1750 to 2019 is assessed at 0.54 [0.43 to 0.65] W m⁻², with updates to 2023 yielding 0.565 W m⁻², reflecting both concentration rises and revised radiative efficiencies that account for enhanced absorption in the near-infrared spectrum. Methane's lifetime of about 9–12 years results in a more rapid forcing response compared to longer-lived gases, though its indirect effects—such as ozone formation and stratospheric water vapor enhancement—amplify its total climate impact by roughly 50%. Nitrous oxide (N₂O), emitted mainly from agricultural soil management, nitrogen fertilizer use, and industrial processes like nitric acid production, has risen from 0.27 ppm pre-industrially to 0.335 ppm in 2023. Its ERF is 0.21 [0.18 to 0.24] W m⁻² for 1750–2019, updated to 0.223 W m⁻² by 2023, with low-confidence tropospheric adjustments adding about 7% to the instantaneous forcing. N₂O's atmospheric lifetime exceeds 100 years, contributing persistently to forcing without significant natural sinks beyond stratospheric photolysis. Fluorinated gases, including chlorofluorocarbons (CFCs), hydrochlorofluorocarbons (HCFCs), hydrofluorocarbons (HFCs), and other synthetic halocarbons used in refrigeration, aerosols, and foam blowing, exhibit high global warming potentials due to strong infrared absorption and long lifetimes (up to centuries for some CFCs). Their combined ERF from 1750 to 2019 is 0.41 [0.33 to 0.49] W m⁻², with 2023 values comprising 0.301 W m⁻² from CFCs, 0.061 W m⁻² from HCFCs, and 0.051 W m⁻² from HFCs, reflecting phase-outs under the that have slowed but not reversed their forcing trend. These gases' radiative efficiencies are updated to include tropospheric adjustments of +12% for major CFCs, though ongoing HFC emissions pose future risks absent further controls. Tropospheric ozone (O₃), a short-lived climate pollutant formed from anthropogenic precursors like nitrogen oxides (NOₓ) and volatile organic compounds (VOCs), contributes an ERF of 0.47 [0.24 to 0.71] W m⁻² since 1750, driven by industrial and transport emissions. This forcing lacks full rapid adjustment assessment due to modeling uncertainties but exceeds prior estimates owing to higher precursor levels. Stratospheric ozone depletion, largely from halocarbons, exerts a small opposing ERF of –0.05 [–0.15 to 0.05] W m⁻², with medium confidence. Overall, other trace gases excluding CO₂ account for approximately 1.2 W m⁻² of total well-mixed GHG forcing in 2023, underscoring their cumulative role despite individual magnitudes smaller than CO₂.

Water Vapor Role

Water vapor constitutes the most abundant greenhouse gas in Earth's atmosphere, accounting for the majority of the natural greenhouse effect by absorbing and re-emitting longwave radiation, with its radiative influence exceeding that of all other greenhouse gases combined. Unlike well-mixed gases such as , whose atmospheric concentrations can persist for centuries due to long lifetimes, tropospheric water vapor has a short residence time of approximately 9 days, maintained in near-equilibrium by evaporation from oceans and land surfaces, condensation, and precipitation processes governed by temperature. In the framework of radiative forcing, which quantifies external perturbations to Earth's energy balance prior to rapid adjustments like changes in water vapor, tropospheric water vapor changes are classified as a feedback rather than a primary forcing mechanism. This distinction arises because water vapor concentrations respond dynamically to initial temperature perturbations—such as those from anthropogenic CO₂ increases—rather than being independently driven by human emissions, which are negligible compared to the natural hydrological cycle's scale of about 5.17 × 10¹⁴ kg/year of evaporation. Direct anthropogenic water vapor additions, for instance from fossil fuel combustion (yielding roughly 2.6 × 10¹² kg/year globally), represent less than 0.001% of this cycle and dissipate rapidly without altering the equilibrium state. The water vapor feedback operates positively: warming expands atmospheric moisture-holding capacity per the Clausius-Clapeyron equation (approximately 7% per Kelvin), enhancing evaporation and thus water vapor amounts, which further traps outgoing longwave radiation and amplifies surface warming by a factor of roughly 1.6 to 2.0, depending on the forcing scenario. Global climate models and observational analyses, including satellite-derived humidity profiles from instruments like AIRS, consistently estimate the water vapor feedback parameter at about +1.8 W/m²/K when combined with lapse rate effects, making it the dominant positive feedback in the climate system and contributing over 50% of the total equilibrium climate sensitivity. This feedback's strength is supported by paleoclimate records, such as those from ice cores showing correlated water vapor and temperature variations over glacial-interglacial cycles, though uncertainties persist in upper-tropospheric relative humidity trends, with some studies indicating potential underestimation in models due to convective processes. Stratospheric water vapor, comprising a smaller fraction of total column water vapor (about 0.001% by mass), exhibits some forcing characteristics, as its trends can be influenced by tropospheric injections like methane oxidation or volcanic eruptions, exerting a radiative forcing of approximately +0.05 W/m² per decade increase in the 20th century; however, this is minor compared to tropospheric feedbacks and often treated separately in assessments due to slower adjustment times. Empirical constraints from reanalyses and chemistry-climate models affirm that excluding water vapor feedbacks would underestimate warming by at least half, underscoring its causal role in amplifying rather than initiating radiative imbalances.

Solar Irradiance Changes

Solar irradiance variations primarily manifest through fluctuations in total solar irradiance (TSI), the total amount of solar electromagnetic radiation incident on Earth per unit area at the top of the atmosphere, measured perpendicular to the incoming rays. TSI averages approximately 1361 W m⁻², but its changes directly alter the planetary energy budget, with radiative forcing calculated as ΔF ≈ ΔTSI / 4 to account for Earth's spherical geometry, yielding an effective incoming flux of about 340 W m⁻² under current conditions. Accounting for planetary albedo (A ≈ 0.29–0.30), the net shortwave forcing is roughly ΔF × (1 - A), though adjustments for rapid atmospheric responses reduce the effective radiative forcing (ERF) by 20–30%. These variations are small compared to anthropogenic forcings but have influenced past climate epochs. Satellite observations since 1978, from instruments like and , reveal an 11-year solar cycle with peak-to-trough TSI amplitude of about 1–1.3 W m⁻² (0.1% relative variation), corresponding to an ERF of approximately 0.2 W m⁻² after stratospheric and tropospheric adjustments. Spectral irradiance shows greater variability in ultraviolet (UV) bands (up to 10–50% in 200–400 nm), which preferentially heat the stratosphere and indirectly influence tropospheric circulation, though total forcing remains dominated by broadband changes. No significant long-term upward trend in TSI is evident over the satellite era; instead, a slight decline of ~0.1 W m⁻² per decade has been observed since the 1980s, uncorrelated with rising global temperatures. Historical TSI reconstructions, derived from proxies such as sunspot numbers, ¹⁴C isotopes, and ¹⁰Be in ice cores, indicate net changes from 1750 to 2019 ranging from –0.5 to +0.6 W m⁻² in TSI, translating to an assessed solar ERF of 0.01 [–0.06 to +0.08] W m⁻² (medium confidence). Some models, like CHRONOS, estimate larger increases of 3.8–6.2 W m⁻² from the Maunder Minimum (1645–1715) to modern maxima, implying ERF up to ~1.5 W m⁻², but these are outliers rejected in consensus assessments due to proxy uncertainties and lack of corroboration from low-variability models like SATIRE. The assessed range reflects debates over quiet-Sun background trends and cycle scaling, with recent analyses suggesting minimal net forcing since pre-industrial times. Over longer timescales, solar evolution drives gradual TSI increases of ~30% since Earth's formation and ~0.07–0.1 W m⁻² per century in the Holocene, but these are irrelevant to 1750–present forcing. Cyclic influences, including grand solar minima like the Maunder, correlate with regional cooling via reduced irradiance and amplified ozone/UV effects, though global impacts are muted by ocean heat capacity. In the industrial era, solar forcing is negligible relative to greenhouse gases (~2.7 W m⁻² anthropogenic ERF), explaining why post-1950 warming persists amid flat or declining TSI. Uncertainties stem from proxy calibration and spectral weighting, with peer-reviewed reconstructions favoring low-end estimates despite alternative models proposing higher variability.

Total Solar Irradiance Variations

Total solar irradiance (TSI) is the spatially and spectrally integrated solar radiative flux at Earth's mean orbital distance, measured at the top of the atmosphere. Satellite observations commencing in 1978, including missions such as ACRIM, SORCE, and TIM, have determined the contemporary mean TSI value at 1361 W/m². These measurements reveal systematic variations tied to solar activity, with the 11-year Schwabe cycle producing peak-to-trough changes of approximately 1 W/m², equivalent to a 0.1% fractional variation. The radiative forcing from these TSI fluctuations is calculated as ΔF ≈ (ΔTSI / 4) × (1 - A), where A ≈ 0.3 is Earth's , yielding an effective forcing amplitude of about 0.17 to 0.2 W/m² over the solar cycle. This forcing modulates stratospheric temperatures and influences tropospheric circulation patterns, though its magnitude is small relative to anthropogenic greenhouse gas forcings. Composite TSI records, harmonized across instruments by researchers like Fröhlich and Lean, ensure continuity and minimize calibration drifts between satellite eras. Over the satellite era (1978–present), TSI has exhibited no statistically significant long-term trend in most reconstructions, though a recent analysis reports a modest decline of -0.15 W/m² per decade from 1980 to 2023, with 95% confidence interval -0.17 to -0.13 W/m² per decade. Proxy-based extensions to earlier centuries, using sunspot numbers and cosmogenic isotopes, suggest TSI variations of up to 0.2–0.4% during grand minima like the Maunder Minimum (1645–1715), but these imply forcings below 0.3 W/m², insufficient to explain modern warming trends without amplification mechanisms of uncertain efficacy. Sunspot records, strongly anticorrelated with TSI, provide a visual proxy for these cyclic and secular modulations.

Cyclic and Spectral Influences

The primary cyclic influence on solar irradiance arises from the 11-year Schwabe cycle, during which total solar irradiance (TSI) varies by approximately 1 W m⁻² from trough to peak. This oscillation, driven by solar magnetic activity and sunspot numbers, corresponds to a radiative forcing amplitude of about 0.18 W m⁻² at the tropopause, accounting for planetary albedo. Observations from satellites since 1978 confirm this periodicity, with cycle amplitudes ranging from 0.7 to 1.3 W m⁻² across recent cycles. Spectral variations amplify the cycle's effects unevenly across wavelengths: ultraviolet (UV) irradiance below 400 nm fluctuates by 5–10% peak-to-peak, while visible (400–700 nm) and near-infrared (>700 nm) components change by less than 0.2%. These disparities arise from facular brightening and sunspot darkening, with UV enhancements linked to chromospheric activity. Consequently, solar maximum conditions deposit more energy in the via absorption and , elevating stratospheric temperatures by up to 2–3 K at 30–50 km altitude. The stratosphere-troposphere coupling induced by these spectral changes can modulate tropospheric dynamics, though the net surface forcing remains dominated by TSI totals. Model simulations indicate that UV-driven stratospheric heating alters planetary wave propagation, potentially shifting tropospheric jets poleward by 1–2° during maxima, with associated responses of ~0.1–0.2°C in extratropical regions. However, global tropospheric responses to the average below 0.1 K, underscoring the 's minor role relative to forcings. Longer-term modulations, such as the 80–90-year Gleissberg , superimpose on the 11-year signal but exhibit even smaller amplitudes, with TSI variations under 0.5 W m⁻².

Long-Term Solar Evolution

Standard stellar evolution models predict that the Sun's luminosity has increased by approximately 30% over the past 4.6 billion years, from about 70% of its current value at the time of solar system formation to its present level of roughly 3.828 × 10²⁶ . This gradual brightening arises from progressive core contraction and rising central temperatures as fuses into , enhancing rates. The rate of increase is nonlinear but averages around 1% per 110 million years, with the Sun being 20–25% fainter than today during the Eon (3.8–2.5 billion years ago). This luminosity evolution imposes a long-term positive radiative forcing on Earth's climate system, quantified as ΔF ≈ (ΔL/L) × (S/4), where S is the (approximately 1366 W/m²) and the division by 4 yields the global mean insolation. Over 4.6 billion years, the cumulative forcing from this increase totals about 100 W/m², equivalent to a top-of-atmosphere imbalance that would drive substantial planetary warming absent countervailing effects. For the period, the reduced luminosity alone generated a forcing deficit of roughly -50 to -85 W/m² relative to modern conditions, depending on the exact luminosity scaling and whether planetary is factored into the net absorbed flux. The faint young Sun paradox highlights the climatic implications: despite this negative forcing in Earth's early history, geological proxies indicate surface temperatures permissive of liquid water, suggesting compensating positive forcings from elevated concentrations (e.g., CO₂ levels potentially 10–100 times higher than today) or lower . Resolution requires atmospheric adjustments that offset the deficit, with models showing that a combination of higher CO₂ partial pressures and could provide the necessary +50 W/m² or more to sustain habitable conditions. Ongoing brightening continues this trend, projecting future forcings that will eventually overwhelm regulatory mechanisms, rendering uninhabitable via runaway greenhouse effects within 1–2 billion years.

Albedo and Surface Modifications

Surface , the fraction of incoming radiation reflected by the Earth's and surfaces, typically ranges from 0.05 for dark forests to 0.8 for fresh , with a global land average around 0.15 to 0.25 depending on , , and seasonal cover. Modifications to surface properties alter this reflectivity, changing the net shortwave radiation absorbed at the top-of-atmosphere and contributing to radiative forcing; a decrease in of Δα exerts a positive forcing by increasing , while an increase yields negative forcing. The global-mean instantaneous radiative forcing from surface changes is approximated as ΔF ≈ -I₀ × (1 - α_p) × Δα or similar effective formulations, where I₀ is the mean incident (~340 m⁻²), and α_p is planetary (~0.3), yielding sensitivities around -100 m⁻² per unit Δα after accounting for atmospheric transmission and masking that reduce the effective input to the surface. Natural fluctuations in surface stem from seasonal and interannual variations in extent, coverage, phenology, and desert expansion. In the , reductions in spring cover extent, which averaged a decline of about 2-3% per decade from 1981 to 2020, lower regional by 0.05-0.1 locally, contributing positive radiative forcing estimates of 0.1-0.3 W m⁻² over affected areas, though global means are attenuated to <0.05 W m⁻² due to hemispheric asymmetry and cloud interactions. Arctic loss, with summer extent decreasing by ~13% per decade since 1979, exposes darker ocean surfaces ( ~0.06 versus ice ~0.5-0.7), generating local forcings up to 1-2 W m⁻² but a global contribution of ~0.2 W m⁻² for 1979-2011 changes when integrated over area. shifts, such as boreal greening replacing high- tundra with darker forests, reduce by ~0.02-0.05, amplifying positive forcing by 0.1-0.5 W m⁻² regionally through enhanced absorption. Anthropogenic surface modifications, primarily through land use and land cover changes since ~1750, have produced a net increase in global albedo of ~0.001-0.002, driven by widespread conversion of dark forests to brighter croplands and pastures, yielding an effective radiative forcing of -0.2 ± 0.1 W m⁻² (cooling effect) as the best estimate for 1750-2011. This masks a heterogeneous pattern: tropical deforestation darkens surfaces (albedo drop ~0.01-0.03, positive forcing ~0.1 W m⁻² regionally), while mid-latitude agriculture brightens them (albedo rise ~0.02-0.05, negative forcing), with the latter dominating globally. Urbanization introduces mixed effects, with dark impervious surfaces reducing albedo by ~0.1 locally (positive forcing up to 1-3 W m⁻² in cities), though reflective materials can reverse this; overall, post-1850 urban expansion contributed negligible net global forcing (<0.01 W m⁻²) due to small areal fraction (~1%). Black carbon deposition from biomass burning further lowers snow albedo by 0.01-0.05 in affected regions, adding positive forcing of ~0.05 W m⁻² globally since pre-industrial times. These estimates derive from satellite observations and models, with uncertainties from cloud-albedo interactions and rapid adjustments reducing magnitudes by 20-50%.

Natural Albedo Fluctuations

Natural fluctuations in Earth's planetary albedo arise primarily from variations in cryospheric extent, cloud cover modulated by climate oscillations, and transient surface changes from events like wildfires or biological activity in oceans. These alterations affect the fraction of incoming solar radiation reflected back to space, inducing radiative forcing that contributes to interannual and decadal climate variability rather than long-term trends. The global mean albedo, typically around 0.29-0.30, can vary by 0.001-0.005 annually due to such processes, corresponding to forcing magnitudes of ±0.1 to 1 W/m², though effects are often regionally confined and short-lived. Seasonal and interannual changes in snow and ice cover represent a dominant natural driver of surface albedo variability, particularly in the Northern Hemisphere. Fresh snow exhibits albedo values of 0.80-0.90, sharply contrasting with underlying vegetation or soil at 0.10-0.30, thereby elevating planetary reflectivity during winter. Fluctuations in snow extent, influenced by natural temperature anomalies, have led to observed global albedo decreases of about 0.001 from 2002 to 2016, partly attributable to reduced snow persistence amid variable weather patterns; this equates to a positive radiative forcing of roughly 0.2-0.5 W/m² over affected regions. Arctic sea ice variability, driven by oscillations like the , similarly modulates albedo, with ice-free ocean surfaces absorbing up to 90% more solar energy than ice-covered ones, amplifying local forcing during melt seasons. Cloud cover variations tied to modes such as the El Niño-Southern Oscillation (ENSO) induce significant albedo perturbations through shifts in low-level marine stratocumulus decks. El Niño phases reduce cloudiness over subtropical oceans, lowering albedo by 0.01-0.02 regionally and yielding a global positive forcing of approximately 0.1-0.2 W/m², as diminished reflection increases absorbed shortwave radiation. Conversely, La Niña enhances cloud reflectivity, producing negative forcing of similar scale. These effects persist for 6-18 months, influencing Earth's energy imbalance and contributing to ENSO-driven temperature swings. Volcanic eruptions and natural biomass burning episodically elevate albedo via surface ash deposition or, more substantially, stratospheric sulfate aerosols that enhance scattering, though the latter overlaps with aerosol forcing mechanisms. Post-eruption surface albedo increases from ash layers can persist months to years locally, with forcing estimates for major events like the 1991 Mount Pinatubo eruption reaching -0.5 to -1 W/m² from surface effects alone amid total aerosol-driven cooling of -2 to -3 W/m². Wildfires deposit light-absorbing black carbon on snow, reducing albedo by 5-15% in affected areas and exerting positive forcing of 0.1-0.3 W/m² regionally, countering any reflective ash benefits. Ocean biological productivity fluctuations, such as phytoplankton blooms, subtly raise marine albedo by 0.001-0.005 through increased surface scattering, but their net forcing remains minor at <0.05 W/m² globally.

Anthropogenic Land Use Impacts

Anthropogenic land use changes, including , agricultural expansion, and urbanization, alter surface by modifying vegetation cover, soil exposure, and impervious surfaces, thereby influencing shortwave at the top of the atmosphere. These modifications typically increase global mean albedo through the replacement of low-albedo forests with higher-albedo croplands and grasslands, enhancing planetary reflectivity and producing a net cooling effect, though regional variations and non-albedo biogeophysical feedbacks introduce uncertainties. Deforestation, particularly in tropical regions since the pre-industrial era, has been a primary driver, converting dense forests with broadband albedo values around 0.12–0.15 to grasslands or pastures with albedos of 0.18–0.25, resulting in local albedo increases of up to 0.05–0.10 during snow-free periods. This change boosts shortwave reflection, with modeled radiative forcing estimates from such transitions ranging from -0.2 to -0.5 W m⁻² regionally in deforested areas. Globally, integrated assessments of land cover change from 1750 to 2014 attribute a mean top-of-atmosphere radiative forcing of -0.15 ± 0.10 W m⁻² to albedo alterations alone, equivalent to a modest offset against greenhouse gas warming. However, climate models exhibit biases in simulating these sensitivities, often underestimating albedo responses by factors of 2–3 in CMIP5 ensembles, which amplifies uncertainties in historical forcing reconstructions. Agricultural practices, such as irrigation and tillage, further modulate albedo through soil moisture and residue management; for instance, irrigated croplands can darken surfaces via wet soils, counteracting some grassland brightening, while residue retention in no-till systems maintains lower albedos akin to natural vegetation. These effects contribute a secondary negative forcing component, estimated at -0.05 to -0.10 W m⁻² in intensively farmed regions like the from 2000–2010. Urbanization, conversely, generally decreases albedo by introducing dark asphalt and concrete (albedo 0.05–0.15 versus rural 0.20–0.30), yielding positive radiative forcing of +0.07 W m⁻² per 1% urban expansion in simulations, with global historical contributions from 1700–2010 around +0.01 to +0.03 W m⁻² due to limited areal coverage. Future projections under shared socioeconomic pathways indicate urbanization could add +0.05 W m⁻² by 2100, though this remains dwarfed by other forcings. The net anthropogenic land use albedo forcing is small and negative, on the order of -0.1 to -0.2 W m⁻² since pre-industrial times, but recent high-resolution analyses suggest prior estimates may overestimate cooling by neglecting dynamic atmospheric responses and spatiotemporally resolved changes, potentially reducing the magnitude to near zero in some datasets. This forcing interacts with carbon cycle effects, where albedo cooling partially offsets CO₂ emissions from land clearing, but empirical satellite observations confirm persistent but modest global albedo trends linked to these activities.

Aerosol Influences

Aerosols, microscopic solid or liquid particles suspended in the atmosphere from both natural sources such as volcanic eruptions, dust storms, and sea spray, and anthropogenic activities including combustion of fossil fuels, biomass burning, and industrial processes, perturb the Earth's radiative balance through direct and indirect mechanisms. Anthropogenic aerosol concentrations have risen significantly since the pre-industrial period (1750), contributing a net negative effective radiative forcing (ERF) that cools the climate by reflecting sunlight and altering cloud reflectivity, thereby masking approximately 20-50% of the warming from greenhouse gases. In the IPCC Sixth Assessment Report (AR6), the total anthropogenic aerosol ERF for 1750-2019 is assessed at -1.3 W m⁻² (90% confidence interval: -2.0 to -0.6 W m⁻²), dominated by sulfate, organic, and nitrate particles from sulfur dioxide and nitrogen oxide emissions. This cooling effect arises despite regional variations, with stronger influences over landmasses like Asia and Europe where emissions peaked mid-20th century before policy-driven declines reduced concentrations post-1980 in North America and Europe. Direct radiative effects occur when aerosols interact with incoming solar shortwave radiation or outgoing longwave terrestrial radiation without intermediary processes. Scattering aerosols, such as sulfates and sea salt, primarily reflect shortwave radiation back to space, reducing net downward flux at the surface and tropopause by an estimated -0.51 W m⁻² (direct ERF from aerosol-radiation interactions, ERFari, in AR6 multimodel assessments). Absorbing aerosols like black carbon and mineral dust, however, warm the atmosphere by capturing radiation, with black carbon exerting a positive forcing of +0.2 to +0.5 W m⁻² globally, though this is often outweighed by scattering counterparts. Observation-based estimates from satellite data and ground measurements place the global mean direct aerosol radiative effect at -2.40 ± 0.6 W m⁻², highlighting uncertainties from aerosol optical depth retrievals and vertical distribution. Volcanic aerosols, such as those from the 1991 Mount Pinatubo eruption, provide transient examples, injecting stratospheric sulfate that cooled global temperatures by ~0.5°C for 2-3 years via a forcing of -3 to -4 W m⁻². Indirect effects amplify aerosol influences by modifying cloud microphysical and macrophysical properties, primarily through aerosol-cloud interactions (ERFaci). Aerosols acting as cloud condensation nuclei (CCN) increase droplet number concentration in low-level clouds, reducing droplet size and enhancing shortwave reflectivity via the Twomey effect, which boosts cloud albedo and yields a cooling forcing of -0.2 to -1.0 W m⁻². Additional semi-direct effects from absorbing aerosols heat the atmosphere, potentially evaporating cloud droplets and decreasing coverage, while lifetime effects prolong precipitating clouds, further scattering radiation. AR6 estimates ERFaci at approximately -0.8 W m⁻², with very high uncertainty due to model discrepancies in simulating cloud responses and observational challenges in isolating aerosol signals from meteorology. Recent analyses indicate that declining anthropogenic emissions, particularly over oceans from shipping regulations since 2020, have reversed aerosol cooling trends, contributing to accelerated surface warming by reducing the masking effect. Natural aerosols, including biogenic organics and dust, exert baseline forcings but with lower variability than anthropogenic ones. Overall, aerosol forcings exhibit high spatial heterogeneity, with cooling maxima over emission source regions and oceans, and substantial interannual variability from events like wildfires or El Niño-driven dust mobilization. Uncertainties stem from incomplete emission inventories, especially for biomass burning and secondary organic aerosols, and from general circulation models' struggles with subgrid processes, leading to ERF ranges spanning a factor of three across ensembles. As global aerosol emissions decline under air quality policies, the unmasking of underlying greenhouse forcing is projected to enhance warming rates, particularly in the coming decades.

Direct Radiative Effects

The direct radiative effects of aerosols arise from their interactions with incoming solar (shortwave) radiation and outgoing terrestrial (longwave) radiation, altering the planetary energy balance without modifying cloud properties. In the shortwave spectrum, aerosols primarily scatter sunlight, reducing the amount reaching the surface and exerting a cooling influence (negative forcing); absorbing aerosols, such as , instead trap energy in the atmosphere, producing a warming effect (positive forcing). Longwave interactions involve absorption and re-emission of infrared radiation, which generally yield small positive forcings for absorbing species but are minor compared to shortwave effects. Major aerosol types exhibit distinct direct effects: sulfate and nitrate particles, largely from anthropogenic sulfur and nitrogen oxide emissions, dominantly scatter shortwave radiation, contributing substantial cooling; organic carbon scatters similarly but with some absorption; black carbon strongly absorbs shortwave radiation, leading to net warming, especially over bright surfaces like snow; mineral dust and sea salt primarily scatter, with dust showing mixed effects depending on composition and location. Globally, anthropogenic direct effects yield a net cooling, estimated at an effective radiative forcing from aerosol-radiation interactions (ERFari) of -0.35 W m⁻² (range -0.65 to -0.05 W m⁻²) from 1750 to 2014, with sulfate (-0.23 W m⁻²), organic carbon (-0.21 W m⁻²), and nitrate (-0.27 W m⁻²) driving negativity, partially offset by black carbon (+0.11 W m⁻²). These effects peak in regions with high emissions, such as eastern and southern Asia, where negative forcings dominate due to sulfate and nitrate burdens, while absorbing aerosols like black carbon induce localized atmospheric heating and surface dimming, potentially stabilizing or destabilizing the atmosphere vertically. Uncertainties stem from aerosol optical properties, vertical profiles, and mixing states, with model spreads exceeding a factor of two for black carbon; observational constraints, including satellite aerosol optical depth trends, indicate stabilization post-2000 after mid-20th-century increases. Recent emission reductions, such as SO₂ cuts in shipping (post-2020 IMO regulations), have diminished cooling by ~3.9 mW m⁻² globally from direct effects alone, highlighting sensitivity to policy-driven changes. While net global cooling offsets ~20-30% of greenhouse gas warming, regional heterogeneity—stronger in the Northern Hemisphere—complicates attribution, with high confidence in the sign but medium confidence in magnitude due to sparse in-situ data in source regions.

Indirect Effects on Clouds

Aerosols influence clouds indirectly by acting as cloud condensation nuclei (CCN), thereby modifying cloud droplet number concentration (Nd), effective radius (re), (τ), and liquid water path (LWP), which alter , lifetime, and coverage, affecting the shortwave reflected to . These interactions produce a net cooling effective radiative forcing (ERF), as increased Nd enhances cloud reflectivity without proportionally increasing . The Twomey effect, first described in 1977, quantifies the initial microphysical response where elevated aerosol levels increase Nd, reducing re and boosting τ for a given LWP, thereby raising cloud albedo and shortwave forcing by approximately 20-30% for susceptible clouds like marine stratocumulus. Observational constraints from satellite data estimate the global Twomey forcing at around -0.75 W m-2, though this varies regionally and is sensitive to baseline cloud conditions. This effect dominates in clean maritime environments where small perturbations in CCN yield large albedo changes, but it diminishes in polluted or precipitating clouds due to saturation of droplet activation. The lifetime (or second indirect) extends this by positing that smaller droplets inhibit warm formation via reduced collision-coalescence efficiency, suppressing , extending persistence, and increasing LWP and coverage, which amplifies shortwave cooling. Studies using satellite observations confirm enhanced fraction and brightness in aerosol-influenced layers, particularly under conditions, contributing an additional negative forcing estimated at 0.2-0.5 W m-2 globally. However, some analyses indicate weaker LWP adjustments than microphysical changes, with rapid responses like convective invigoration potentially offsetting cooling in deeper s. Incorporating rapid adjustments—such as shifts in and circulation—the total ERF from aerosol-cloud interactions (ERFaci) is assessed at -0.84 W m-2 (90% -1.9 to -0.1 W m-2), representing the largest source of uncertainty in forcing due to challenges in disentangling from natural variability and model parametrizations. Uncertainties stem from sparse clean-air baselines for isolating signals, inter-model spread in droplet activation schemes (up to 50% variance), and underrepresentation of organic aerosols or natural emissions like sea spray, which contribute 45% to historical forcing variability. Recent modeling highlights that aerosol-induced circulation responses, such as stabilized boundary layers, can enhance ERFaci by 50-100% beyond fixed-sea-surface-temperature estimates. Despite progress in retrievals (e.g., from MODIS and CloudSat), persistent discrepancies between observed trends and simulations the need for process-level validation.

Comparative Forcing Dynamics

Anthropogenic vs Natural Forcings

radiative forcings, stemming from , aerosol alterations, land-use changes, and ozone precursors, have produced a net positive effective radiative forcing (ERF) of 2.72 W m⁻² (range: 1.96 to 3.48 W m⁻²) from 1750 to 2019, according to assessments integrating models and observations. This net arises from strong positive contributions of well-mixed greenhouse gases at 3.84 W m⁻² (3.46 to 4.22 W m⁻²), including 2.16 W m⁻² from CO₂ alone, offset by negative aerosol effects at -1.1 W m⁻² (-1.7 to -0.4 W m⁻²) and land-use albedo increases at -0.20 W m⁻² (-0.30 to -0.10 W m⁻²). Tropospheric adds +0.47 W m⁻² (0.24 to 0.71 W m⁻²) from anthropogenic precursors. Natural forcings over the same interval remain small by comparison, with variations yielding +0.01 W m⁻² (-0.06 to 0.08 W m⁻²), based on reconstructed total changes of about 0.1% since the . Volcanic forcings are predominantly negative and transient, driven by stratospheric injections from eruptions; no net long-term ERF accumulates, but individual events like Pinatubo (1991) imposed temporary global coolings equivalent to -2 to -3 W m⁻² for 1–3 years. In recent decades (1979–2015), volcanic ERF averaged near zero, with small-magnitude eruptions contributing -0.08 W m⁻² during 2005–2015 relative to quiescent baselines. The divergence in magnitudes and persistence highlights anthropogenic dominance: from 2011 to 2019, GHG ERF rose by 0.59 W m⁻² due to emission-driven concentration increases, while natural forcings exhibited no comparable trend, with solar output declining post-2014 and subdued . Aerosol ERF uncertainties remain substantial (medium confidence), potentially masking 0.5–1.0 W m⁻² in net forcing variability, but empirical constraints affirm forcings as the primary driver of the post-1950 top-of-atmosphere imbalance.
Forcing TypeKey ComponentsERF (W m⁻², 1750–2019, best estimate [5–95% range])
AnthropogenicGHGs, aerosols, , +2.72 [1.96–3.48]
Natural, volcanicSolar: +0.01 [-0.06–0.08]; Volcanic: episodic ~0 net

Paleoclimate and Historical Context

Paleoclimate proxies indicate that variations imposed substantial radiative forcings during glacial- transitions, amplifying smaller orbital perturbations. ice cores, including records from Dome C, document atmospheric CO2 concentrations of about 180–190 ppm at the circa 20,000 years ago, increasing to 260–280 ppm during the subsequent , yielding a CO2 radiative forcing change of approximately 2.4 W/m² calculated via the relation ΔF = 5.35 ln(C/C₀). shifts from ~350 ppb to ~700 ppb added roughly 0.5 W/m², while and changes contributed negative forcings of -0.5 to -1 W/m² during glacials. These forcings, combined with dynamics, explain much of the 4–7 °C at the LGM relative to pre-industrial conditions. Milankovitch cycles—variations in Earth's , obliquity, and precession—provided the primary initial forcing, with global annual mean insolation changes limited to ~0.1 W/m² amplitude due to redistribution effects rather than net input alterations. However, peak summer insolation at northern high latitudes fluctuated by up to 100 W/m² over 21,000–41,000-year cycles, sufficient to destabilize ice sheets and trigger deglaciations, after which releases from oceans and land amplified the response. from ice cores and sediments confirms this sequence, where orbital changes precede but do not fully account for temperature swings, with greenhouse forcings responsible for 40–50% of the variance in some models. In the historical record spanning the to instrumental era, natural forcings predominated until the industrial period. Solar irradiance reconstructions link grand minima like the (1645–1715) to total solar irradiance reductions of ~0.2–0.24%, equating to a global radiative forcing of -0.1 to -0.3 W/m², which contributed to cooling alongside sporadic volcanic forcings exceeding -2 W/m² for major eruptions. Proxy data show these natural forcings induced temperature anomalies of ~0.5–1 °C regionally, but global variability remained within ±0.5 °C of pre-industrial means, driven by solar and volcanic imbalances rather than sustained greenhouse trends. Anthropogenic forcings since 1750 have shifted this balance, with long-lived greenhouse gases accumulating to produce a net positive effective radiative forcing of ~2.7–3.0 W/m² by 2020, offsetting aerosol cooling and exceeding Holocene natural rates by factors of 10–100 in decadal changes for CO2, CH4, and N2O combined. This rapid escalation contrasts with paleoclimate transitions, where full glacial-interglacial forcings unfolded over millennia, highlighting the unprecedented pace of modern alterations attributable to combustion and .

Interannual and Decadal Variability

Interannual variability in Earth's radiative forcing arises primarily from transient volcanic aerosol injections and coupled ocean-atmosphere phenomena like ENSO, which modulate and atmospheric . The exemplifies volcanic impacts, delivering stratospheric sulfate s that peaked at a global-mean effective radiative forcing of approximately -2.0 W/m² in mid-1992 before decaying over 2–3 years due to gravitational and radiative removal. Smaller eruptions, such as those clustered in 2005–2015, contributed a net forcing of -0.08 W/m² relative to quiescent periods. ENSO drives fluctuations in top-of-atmosphere radiative fluxes via shortwave adjustments and longwave changes, with observed standard deviations of 0.24 W/m² in longwave instantaneous forcing and 0.10 W/m² in shortwave -related components from measurements spanning 2003–2018. Decadal-scale variations stem mainly from the 11-year and multiyear ocean circulation modes that alter radiative effects. The induces total changes of ~1 W/ peak-to-trough (0.07–0.1% fractional variation), yielding a global radiative forcing amplitude of 0.1–0.2 W/ after accounting for planetary averaging. This forcing manifests in stratospheric and circulation responses that amplify surface impacts beyond direct . Pacific multidecadal variability, potentially triggered by volcanic sequences, influences decadal forcing through sea surface temperature patterns and teleconnections, contributing to net top-of-atmosphere flux anomalies on the order of 0.1–0.5 W/. CERES-derived energy imbalance records confirm these timescales dominate natural fluctuations, with interannual-to-decadal standard deviations exceeding 0.5 W/, underscoring the role of clouds in masking or amplifying external signals.

Applications and Interpretations

Climate Attribution

Climate attribution refers to the process of identifying and quantifying the contributions of specific radiative forcings to observed changes in Earth's , particularly trends. Detection and attribution studies typically employ statistical methods and climate models to distinguish forced signals from internal variability, comparing simulations with all forcings to those excluding influences. Effective radiative forcing (ERF) serves as a key metric, linking perturbations in the energy balance to climate responses via equilibrium . In assessments of global mean surface temperature (GMST) rise from 1850–1900 to 2011–2020, forcings account for 0.8–1.3°C of the observed 0.95–1.20°C warming, with greenhouse gases (GHGs) providing the dominant positive contribution of approximately 1.0–2.0 W m⁻² ERF, offset partially by cooling effects of -0.5 to -1.0 W m⁻². Models driven solely by natural forcings, including variations and volcanic eruptions, simulate negligible warming or slight cooling over this period, failing to reproduce the observed upward trend. This discrepancy underscores the causal role of anthropogenic RF, as natural factors alone cannot explain the post-1950 acceleration in warming, which aligns closely with cumulative GHG forcing. Attribution extends to regional and extreme events, where GHG forcing enhances heatwaves and heavy probabilities, though uncertainties arise from model deficiencies in simulating natural variability and indirect effects. For instance, forcing has been linked to a 2–3°C cooling in the upper since 1979, consistent with GHG-induced radiative changes rather than alone. However, some studies highlight persistent challenges, such as overestimation of low-frequency variability in models, potentially inflating attribution confidence to drivers. Overall, while empirical fingerprints like tropospheric warming and stratospheric cooling support RF-based attribution, ongoing debates emphasize the need for improved observational constraints on transient response.

Equilibrium Climate Sensitivity

Equilibrium climate sensitivity (ECS) quantifies the long-term global mean surface temperature response to a doubling of atmospheric CO₂ concentration from pre-industrial levels, incorporating all climate feedbacks such as , , clouds, and surface changes. It is formally defined as the equilibrium surface warming ΔT following a sustained increase in CO₂ from 280 to 560 , after the reaches a new radiative balance. ECS relates to radiative forcing via ΔT = λ ΔF, where ΔF is the effective radiative forcing for doubled CO₂ (approximately 3.9 W m⁻² in recent assessments) and λ is the climate feedback parameter, with no-feedback yielding λ ≈ 0.8 K (W m⁻²)⁻¹ or about 3 K per doubling. Estimates of ECS derive from three primary approaches: climate model simulations, paleoclimate proxies, and instrumental observations via energy budget constraints. The Charney Report in 1979 first established a range of 1.5–4.5 °C based on early models. The (AR6) in 2021 assessed ECS as likely between 2.5–4.0 °C, with a best estimate of 3.0 °C and very likely 2–5 °C, narrowing the lower bound from prior reports by integrating multiple lines of evidence while excluding values below 2 °C as inconsistent with observed warming patterns and cloud feedback physics. Observationally constrained estimates, using historical , forcing, and uptake data, often yield lower central values than process-based models, typically 1.5–2.5 °C, highlighting discrepancies attributed to incomplete treatment of spatial patterns in forcing and feedbacks or unaccounted variability. For instance, Phase 6 (CMIP6) models exhibit a broader ECS range up to 5.6 °C, with high-sensitivity models overestimating recent warming rates unless adjusted for pattern effects that temporarily reduce effective sensitivity. Paleoclimate records from ice ages support ECS values around 2–4 °C but carry uncertainties from proxy calibrations and non-CO₂ forcings. Uncertainties persist due to feedbacks, which dominate spread in model ECS, and limitations in observational constraints over short records that capture transient rather than responses. Recent analyses suggest that while high ECS (>4 °C) remains possible, it is less consistent with mid-twentieth-century warming slowdowns and requires stronger positive feedbacks than empirically supported, prompting calls for refined diagnostics to reconcile model and observational divergences. ECS informs projections of committed warming but underscores the need for empirical validation over reliance on ensemble means potentially inflated by structural model biases.

Limitations in Projections

Projections of future radiative forcing face substantial uncertainties stemming from incomplete knowledge of both and natural drivers, as well as challenges in modeling their interactions. Effective radiative forcing (ERF), which accounts for rapid atmospheric adjustments to perturbations, is central to these projections, yet estimates for components like aerosols exhibit very likely ranges exceeding ±1 W m⁻² due to difficulties in simulating microphysical processes and regional distributions. Similarly, uncertainties in land-use change forcings arise from variable assumptions about rates and dynamics, contributing to projected ERF spreads of 0.5–2 W m⁻² by 2100 across scenarios. A primary limitation is the divergence in model responses to identical forcing inputs, with ensembles revealing that structural differences—such as parameterization of and processes—can amplify projection uncertainties by factors of 1.5–2 for global temperature changes linked to forcing. Aerosol-cloud interactions, in particular, introduce indirect effects that are poorly constrained observationally, leading to ERF uncertainties that dominate total forcing variability in mid-century projections. Natural forcings, including cycles and volcanic eruptions, add unpredictable decadal-scale fluctuations; for instance, a major eruption could offset GHG forcing by 0.1–0.5 W m⁻² temporarily, but their timing and magnitude defy long-term forecasting. Feedback mechanisms further complicate projections, as water vapor and lapse-rate feedbacks, while amplifying CO₂ forcing by approximately 1.5–2 times, exhibit model spreads that propagate into equilibrium climate sensitivity (ECS) estimates ranging from 1.5–4.5°C per CO₂ doubling in AR6 assessments. Cloud feedback uncertainties, assessed at 0.0–0.5 W m⁻² K⁻¹ with low confidence in sign for low clouds, can shift projected forcing efficacy by up to 20%, particularly in high-emission scenarios where polar amplification alters stratocumulus regimes. Empirical constraints from paleoclimate data and instrumental records suggest some models overestimate ECS, implying projections may inflate warming risks beyond observed patterns, though institutional assessments like IPCC maintain broad ranges to encompass ensemble means. Scenario dependence exacerbates this, as shared socioeconomic pathways (SSPs) embed optimistic or pessimistic emission trajectories without robust validation against policy realities, rendering near-term (2021–2040) forcing projections unreliable for adaptation planning.

Controversies and Criticisms

Discrepancies Between Models and Observations

Satellite measurements from the Clouds and the Earth's Radiant Energy System () reveal that Earth's energy imbalance (EEI), which reflects the net radiative forcing after rapid adjustments and partial temperature response, has intensified more rapidly than anticipated by Phase 6 (CMIP6) simulations. From 2001 to 2023, CERES observations indicate a stronger positive trend in EEI compared to the CMIP6 multimodel ensemble mean, with EEI values in 2023 exceeding those produced by the models despite similar forcings applied. This divergence suggests potential underestimation in models of recent anthropogenic forcing enhancements, such as diminished aerosol cooling from pollution controls in regions like and , or inadequate representation of cloud adjustments that amplify the imbalance. Observation-based inferences of effective radiative forcing (ERF) further highlight model limitations. Applying to TOA fluxes and surface temperatures yields an ERF increase of 0.71 ± 0.21 W/m² per decade over 2001–2024, a rate consistent with physical expectations but independent of model assumptions prone to biases in aerosol indirect effects and radiative properties. In contrast, CMIP6 models exhibit variability in ERF decomposition, often failing to replicate observed optical depth (AOD) declines over in the late historical period, which contribute to underestimated aerosol forcing trends and thus overstated model-observation alignment in net forcing. Discrepancies also manifest in aerosol-cloud interactions (ACI), where CERES-derived ERFaci totals -1.11 ± 0.43 W/m² (), exceeding some model estimates due to challenges in simulating clean-sky conditions that amplify indirect cooling. CMIP6 simulations frequently underestimate ACI magnitude under low-aerosol regimes, leading to less negative forcing and potential overprediction of net positive ERF. Earlier CMIP5 assessments similarly overestimated accumulation, implying excessive net TOA forcing relative to CERES-constrained EEI trends during 1970–2012. Spatial pattern effects exacerbate these issues, as CERES data from sequential periods (e.g., 2000–2010 vs. 2010–2020) show substantial unforced variability in cloud feedback, altering effective forcing estimates by up to 1 W/m² regionally—dynamics incompletely captured in models reliant on global-mean approximations. Such pattern dependencies underscore how model physics, including convective organization and feedbacks, diverge from observed TOA flux anomalies, complicating forcing attribution. These observational-model gaps persist despite advances, with CERES trends in reflected indicating unmodeled albedo reductions that boost EEI beyond forcing inputs alone.

Debates on Forcing Attribution

Debates on the attribution of radiative forcing center on the relative magnitudes and interactions of versus natural components in driving observed variations, with significant uncertainties arising from incomplete observational records and model discrepancies. forcings, dominated by well-mixed gases (contributing approximately +2.83 W m⁻² from 1750 to 2011) and offset by aerosols (estimated at -0.9 W m⁻² with high uncertainty), are contrasted against natural forcings like variations (peaking at +0.05 W m⁻² during the ) and volcanic aerosols. Critics, including analyses from non-governmental assessments, contend that mainstream models undervalue natural forcings' role in multidecadal warming, arguing that solar cycles and ocean-atmosphere oscillations explain much of the pre-1950 temperature rise without invoking dominant effects. A key contention involves forcing, where direct radiative changes are small but potential amplification via mechanisms like modulation of clouds or stratospheric alterations could amplify impacts. Studies indicate that climate models may underestimate the observed 20th-century response to variability by a of 2–3, suggesting indirect effects contribute up to 0.3–0.5 W m⁻² equivalent forcing during grand maxima like the (peaking around 1950). However, attribution analyses partitioning surface warming find contributions limited to less than 10% of post-1950 trends, with greenhouse gases accounting for over 100% when cooling is in, though skeptics highlight that such partitions rely on assumed where forcing has lower temperature response per W m⁻² than CO₂ due to stratospheric cooling effects. Aerosol forcing attribution remains highly debated due to its dual direct and indirect (cloud-mediated) effects, with effective radiative forcing estimates ranging from -0.1 to -2.0 W m⁻², representing the largest source of uncertainty in total anthropogenic forcing. Observational constraints suggest aerosol-cloud interactions dominate this spread, particularly in clean marine environments where low aerosol burdens amplify sensitivity, yet models diverge on biomass burning and sulfate contributions, potentially overestimating cooling by ignoring rapid adjustments in circulation. This uncertainty complicates isolating greenhouse gas dominance, as aerosol reductions (e.g., post-1970s sulfur controls) may unmask warming, but attribution studies struggle with co-variability, leading some to argue that resolving aerosol efficacy could shift attributed warming fractions by 20–50%. Peer-reviewed critiques emphasize that institutional consensus favors strong negative aerosol forcing to bolster anthropogenic signals, potentially overlooking natural emission variability.

Uncertainties in Aerosol and Solar Components

Aerosols exert a net cooling effective radiative forcing (ERF) estimated at -1.1 W/m² (-1.7 to -0.4 W/m²) from 1750 to 2019, representing the largest source of uncertainty in total anthropogenic forcing due to complex direct scattering, absorption, and indirect cloud modification effects. This range arises primarily from sulfate aerosol processes, biomass burning emissions, particle size distributions, and interactions with natural aerosols, which are difficult to isolate observationally amid regional variability and rapid atmospheric adjustments. Aerosol-cloud interactions (ACI), including droplet activation and precipitation suppression, amplify uncertainty, particularly in clean-sky conditions where low aerosol burdens allow greater cloud susceptibility but challenge detection in satellite data. Observational constraints remain limited by sparse vertical profiling and short-term campaigns, leading to model-observation discrepancies where global chemistry-transport models overestimate direct radiative effects compared to top-of-atmosphere measurements. Recent emission reductions, such as in sulfate from shipping regulations post-2020, have reversed aerosol forcing trends toward less cooling, unmasking underlying greenhouse gas warming but introducing further parametric uncertainties in projections. Solar radiative forcing from total (TSI) variations contributes a modest 0.05 to 0.1 W/m² over recent decades, dwarfed by greenhouse gases, yet historical reconstructions carry substantial uncertainty due to proxy-based scaling of numbers, cosmogenic isotopes, and tree-ring data before era measurements in 1978. The 11-year modulates TSI by about 1 W/m² at the top of the atmosphere, but its climate impact is debated, with empirical analyses suggesting potential amplification via ultraviolet-driven stratospheric changes and ocean-atmosphere coupling, though mainstream assessments like IPCC AR6 attribute minimal net forcing over the . Critiques highlight underestimation in general circulation models, where balanced multi-proxy total activity indices imply stronger correlations with surface temperatures than TSI alone predicts, challenging attributions that downplay roles in multidecadal variability. Uncertainties persist in distinguishing signals from internal climate modes like ENSO or volcanic forcings, compounded by sparse pre-1950 data and debates over cosmic ray-cloud links, which lack robust empirical support despite theoretical plausibility. These gaps fuel controversies, as over-reliance on low forcing in estimates may overlook causal pathways evident in paleoclimate proxies, such as the Maunder Minimum's cooler epochs.

Post-2020 Observations

![ESSD Radiative Forcing 1750 to 2022][float-right] Post-2020 observations indicate a continued acceleration in effective radiative forcing, driven primarily by rising concentrations of long-lived greenhouse gases. The NOAA Annual Greenhouse Gas Index (AGGI) for 2023 reported a value of 1.51, reflecting a 51% increase in radiative forcing from these gases relative to 1990 levels, with a total forcing of approximately 3.49 W m⁻². This represents a 1.6% rise from 2022, consistent with uninterrupted emissions growth post-COVID recovery. Satellite observations from NASA's CERES instrument reveal an intensifying Earth's energy imbalance (EEI), serving as an empirical proxy for net radiative forcing trends. By 2023, EEI reached 1.8 ± 0.5 W m⁻², more than double the 0.8 W m⁻² observed around 2005, exceeding multimodel CMIP6 projections by a factor of two. An independent analysis of and other datasets estimated an effective radiative forcing trend of 0.71 ± 0.21 W m⁻² per decade from 2001 to 2024, with substantial increases since 2021 unoffset by negative feedbacks. Regulatory changes, such as the 2020 sulfur cap on shipping fuels, contributed a detectable positive forcing by reducing aerosol cooling. applied to data quantified this at +0.074 ± 0.005 W m⁻² globally, while multimodel assessments pegged the effective radiative forcing at 0.06 to 0.09 W m⁻². Concurrently, declining planetary —linked to reduced low-cloud cover and loss—amplified shortwave absorption, further elevating net forcing in 2023–2024. These observations underscore discrepancies between measured imbalances and model simulations, highlighting potential underestimation of forcing agents like tropospheric adjustments or aerosol reductions.

2023-2025 Developments

The Annual Greenhouse Gas Index, calculated by NOAA, reached 1.51 in 2023, reflecting a 51% rise in effective radiative forcing from human-emitted well-mixed greenhouse gases compared to levels. The reported that radiative forcing from long-lived greenhouse gases increased by 51.5% over the same period, with CO₂ responsible for about 81% of the total, driven by sustained emissions despite natural variability. Effective radiative forcing estimates, derived from observation-based methods, showed a marked uptick after , amplifying Earth's energy imbalance without a commensurate negative radiative response from rapid adjustments like cloud feedbacks until later years. satellite observations indicated the planetary imbalance peaked at around 1.8 W/ in 2023—more than double the rate anticipated by many models—before declining in the second half of 2023 and into 2024, possibly due to evolving temperature patterns and the transition to La Niña conditions. This observed acceleration in imbalance exceeded model projections, highlighting potential underestimations in forcing-response dynamics or reductions from measures. The 2024 Indicators of Climate Change update, published in mid-2025, incorporated new data on effective radiative forcing components through 2023, confirming forcings as the primary driver of a human-induced warming estimate nearing 1.5°C above pre-industrial levels, with total ERF dominated by greenhouse gases offset partially by s. Atmospheric CO₂ surged by a record 3.5 ppm from 2023 to 2024—the largest annual increment since systematic monitoring began in 1957—intensifying forcing amid combustion and reduced carbon sinks. Preliminary 2025 analyses, including updated effect evaluations in models, suggest ongoing refinements to historical forcing attributions, though uncertainties in short-lived species persist.

References

  1. [1]
    [PDF] Anthropogenic and Natural Radiative Forcing
    Box 8.1 | Definition of Radiative Forcing and Effective Radiative Forcing ... forcing in this chapter are the radiative forcing (RF) and the effective radiative.
  2. [2]
    [PDF] CHAPTERS Radiative Forcing and Temperature Trends
    Radiative forcing is defined as the global-mean change in the net irradiance at the tropopause following a change in the radiative properties of the atmosphere ...
  3. [3]
    [PDF] Chapter 14 Radiative Forcing of Climate
    We describe the historical evolution of the conceptualization, formulation, quantification, application, and utilization of ''radiative forcing'' (RF) of ...<|separator|>
  4. [4]
    Annual Greenhouse Gas Index (AGGI) - Global Monitoring Laboratory
    The perturbation to climate forcing (also termed “radiative forcing”) that has the largest magnitude and the smallest scientific uncertainty is the forcing ...
  5. [5]
    New estimates of radiative forcing due to well mixed greenhouse ...
    Jul 15, 1998 · The radiative forcing due to all the WMGG is calculated to 2.25 Wm−2, which we estimate to be accurate to within about 5%.Missing: formula | Show results with:formula
  6. [6]
    Figure AR6 WG1 | Climate Change 2021: The Physical Science Basis
    Figure 7.6 shows the change in effective radiative forcing (ERF) from 1750 to 2019, broken down by contributing agents like carbon dioxide, methane, and ...
  7. [7]
    Recent reductions in aerosol emissions have increased Earth's ...
    Apr 3, 2024 · We find that the effective radiative forcing due to anthropogenic aerosol emission reductions has led to a 0.2 ± 0.1 W m−2 decade−1 ...
  8. [8]
    Effective radiative forcing and adjustments in CMIP6 models - ACP
    Aug 17, 2020 · We evaluate effective radiative forcing and adjustments in 17 contemporary climate models that are participating in the Coupled Model Intercomparison Project ( ...
  9. [9]
    Bounding Global Aerosol Radiative Forcing of Climate Change
    Nov 1, 2019 · It finds that there are two chances out of three that aerosols from human activities have increased scattering and absorption of solar radiation ...
  10. [10]
    Observation-based estimate of Earth's effective radiative forcing
    Jun 10, 2025 · By construction, our estimate of the forcing trend (0.71 ± 0.21 Wm − 2 decade − 1 for 2001–2024) is consistent with stabilizing radiative ...
  11. [11]
    [PDF] Radiative Forcing of Climate Change
    used the following definition for the radiative forcing of the climate system: “The radiative forcing of the surface-troposphere system due to the ...
  12. [12]
    [PDF] Box 1.3. Radiative forcing and climate sensitivity - IPCC
    The radiative forcing is defined as the change in net radiative flux at the tropopause after allowing for stratospheric tempera- tures to readjust to ...
  13. [13]
    The spectroscopic foundation of radiative forcing of climate by ...
    Apr 21, 2016 · ... radiative forcing; Overall spectroscopic uncertainty in CO2 radiative forcing is < 1% ... RF is the change in net radiative flux at the tropopause ...
  14. [14]
    [PDF] CHAPTER 7 - NOAA Chemical Sciences Laboratory
    A second definition of radiative forcing, emerg- ing from the discussions in WMO (1986) and IPCC relates to the case when the stratospheric tempera- tures ...
  15. [15]
    A new method for diagnosing radiative forcing and climate sensitivity
    Feb 11, 2004 · A radiative forcing applied to the climate ... For a correct estimate of α based on net radiative flux, we must evaluate N at the TOA.
  16. [16]
    [PDF] Why the Forcing from Carbon Dioxide Scales as the Logarithm of Its ...
    Jul 1, 2022 · The Lorentz line shape can be derived from first principles with some approximations (e.g., Van Vleck and Weisskopf 1945) and is supported ...
  17. [17]
    [PDF] Fermi Resonance and the Quantum Mechanical Basis of Global ...
    Here, we show how CO2 radiative forcing can be expressed via a first-principles description of the molecule's key vibrational-rotational transitions.<|control11|><|separator|>
  18. [18]
    Arrhenius and the Intergovernmental Panel on Climate Change
    Oct 19, 2006 · Svante August Arrhenius was the first to calculate global climate sensitivity to changes in atmospheric CO2 concentration.<|control11|><|separator|>
  19. [19]
    Arrhenius 1896: First Calculation of Global Warming
    Sep 6, 2024 · Arrhenius wants to calculate how the infrared radiation from a body at 15 °C (i.e. the Earth) will be affected by absorption of infrared ...
  20. [20]
    Basic Radiation Calculations - American Institute of Physics
    In 1896 Svante Arrhenius went a step farther, grinding out a numerical computation of the radiation transfer for atmospheres with differing amounts of carbon ...
  21. [21]
    from the pioneering work of Arrhenius and Callendar to today's Earth ...
    The calculations involved balancing the radiative heat budget (thereby assuming a state of equilibrium), namely solar radiation arriving at the Earth's ...
  22. [22]
    The Historical Evolution of the Radiative Forcing ... - AMS Journals
    Abstract We describe the historical evolution of the conceptualization, formulation, quantification, application, and utilization of “radiative forcing” ...
  23. [23]
    [PDF] The concept of climate sensitivity: history and development
    The climate sensitivity concept (CSC) has more than a century of history. It is closely related to the concept of “climate forcing” or “radiative forcing,” ...<|separator|>
  24. [24]
    1 Introduction | Radiative Forcing of Climate Change: Expanding the ...
    The marginal increase in radiative forcing can be calculated as the first derivative of the radiative forcing with respect to concentration. For low ...
  25. [25]
    [PDF] The Earth's Energy Budget, Climate Feedbacks and Climate Sensitivity
    inter-model differences in radiative forcing owing to large variations in aerosol forcing across models (Forster et al., 2013). Likewise, the spread in ...
  26. [26]
    Climate radiative feedbacks and adjustments at the Earth's surface
    Mar 24, 2015 · ... climate feedbacks defined by their impact on top of ... and surface radiative forcing, resulting in reduction in evaporation and rainfall.
  27. [27]
    Chapter 7: The Earth's Energy Budget, Climate Feedbacks, and ...
    The effective radiative forcing, ERF(ΔF; units: W m –2) quantifies the change in the net TOA energy flux of the Earth system due to an imposed perturbation ( ...
  28. [28]
    Comparison of Effective Radiative Forcing Calculations Using ...
    Apr 2, 2019 · We compare six methods of estimating effective radiative forcing (ERF) using a set of atmosphere-ocean general circulation models.
  29. [29]
    Benchmark Calculations of Radiative Forcing by Greenhouse Gases
    Nov 17, 2020 · This paper describes an experiment within the Radiative Forcing Model Intercomparision Project that uses benchmark calculations made with line-by-line models.
  30. [30]
    Radiative forcing by long‐lived greenhouse gases: Calculations with ...
    Jul 2, 2008 · The AER broadband models provide radiative forcing results that are in closer agreement with high-resolution calculations than the GCM radiation codes examined ...
  31. [31]
    Radiative transfer for inhomogeneous atmospheres: RRTM, a ...
    Jul 1, 1997 · A rapid and accurate radiative transfer model (RRTM) for climate applications has been developed and the results extensively evaluated.Missing: MODTRAN | Show results with:MODTRAN
  32. [32]
    Radiative Forcings - Geophysical Fluid Dynamics Laboratory - NOAA
    In a general circulation model (GCM), radiative transfer translates changes in atmospheric constituents, such as clouds, aerosols, water vapor, and CO2, into ...
  33. [33]
    MODTRAN®
    MODTRAN is a computer code used for predicting and analyzing optical measurements through the atmosphere, developed by Spectral Sciences and the Air Force ...About MODTRAN · MODTRAN Web App · Order · Features and BenefitsMissing: RRTM | Show results with:RRTM
  34. [34]
    Contributions of the ARM Program to Radiative Transfer Modeling ...
    Apr 1, 2016 · Clear-sky radiative transfer calculations have to account for thousands of absorption lines due to water vapor, carbon dioxide, and other gases, ...<|separator|>
  35. [35]
    Chapter: Earth Radiation Budget - The National Academies Press
    Current CERES has a precision that can detect radiative forcing. It is perhaps unrealistic to expect significant improvements in the accuracy of ERB ...<|control11|><|separator|>
  36. [36]
    CERES – Clouds and the Earth's Radiant Energy System
    The Clouds and the Earth's Radiant Energy System (CERES) project provides satellite-based observations of ERB and clouds.Science · Data · CERES Operations · CERES Acronyms
  37. [37]
    CERES_EBAF_Edition4.1 - NASA ASDC
    NASA Earth Observatory Article: Net Radiation - The measurements were made by the Clouds and the Earth's Radiant Energy System (CERES) sensors on NASA's Terra ...
  38. [38]
    Observational Evidence of Increasing Global Radiative Forcing
    Mar 25, 2021 · We use the radiative kernel technique to isolate radiative forcing from total radiative changes and find it has increased from 2003 to 2018.Plain Language Summary · Methods · Results · Conclusions
  39. [39]
    Continuity in Top-of-Atmosphere Earth Radiation Budget ...
    Clear-sky fluxes for cloud-free areas are inferred from CERES broadband radiances for completely cloud-free CERES footprints and imager radiances within clear ...
  40. [40]
    Ground-based measurements of aerosol optical properties and ...
    The diurnal aerosol direct radiative forcing efficiency was about −47.4 W/m2. Overall, aerosols led to decrease of the surface net shortwave irradiance in this ...
  41. [41]
    Longwave radiative forcing of Saharan dust aerosols estimated from ...
    Dec 3, 2003 · We present a new technique for studying longwave (LW) radiative forcing of dust aerosols over the Saharan desert for cloud-free conditions.
  42. [42]
    Aerosol optical properties and direct radiative forcing based ... - ACP
    Jan 15, 2018 · The annual mean DARF was −93 ± 44 to −79 ± 39 W m−2 at the Earth's surface and ∼ −40 W m−2 at the top of the atmosphere (for the solar zenith ...
  43. [43]
    Satellite-based analysis of top of atmosphere shortwave radiative ...
    Jun 12, 2024 · This study investigates long-term changes in the shortwave direct aerosol radiative effect (DARE) at the top of the atmosphere (TOA) induced by biomass burning ...
  44. [44]
    Uncertainty in aerosol–cloud radiative forcing is driven by clean ...
    Apr 5, 2023 · Here we show that the behaviour of clouds under these clean conditions is of equal importance for understanding the spread in radiative forcing estimates.
  45. [45]
    [PDF] The Earth's Energy Budget, Climate Feedbacks and Climate Sensitivity
    The forcing components with symmetric uncertainty ranges are CO2 (fractional. 5–95% uncertainty 0.12 of the best estimate), CH4 (0.20), N2O (0.16), halogenated ...
  46. [46]
    Clear-Sky Direct Aerosol Radiative Forcing Uncertainty Associated ...
    Apr 21, 2022 · In this study, we investigate the uncertainty of DARF associated with aerosol vertical distribution, using simulation results from 14 global models.
  47. [47]
    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 ...
  48. [48]
    Importance of Uncertainties in the Spatial Distribution of ...
    Feb 6, 2021 · Uncertainty in preindustrial aerosol emissions, including fires, is one of the largest sources of uncertainty in estimating anthropogenic ...
  49. [49]
    Uncertainty in aerosol radiative forcing impacts the simulated global ...
    Dec 3, 2020 · Anthropogenic aerosols are dominant drivers of historical monsoon rainfall change. However, large uncertainties in the radiative forcing associated with ...<|control11|><|separator|>
  50. [50]
    [PDF] Annex III: Tables of Historical and Projected Well-mixed Greenhouse ...
    Tables AIII.Ia–f provide historical abundances (mixing ratios) and effective radiative forcing (ERF) values for greenhouse gases assessed in this report.Missing: breakdown | Show results with:breakdown
  51. [51]
    Greenhouse gas concentrations surge again to new record in 2023
    From 1990 to 2023, radiative forcing – the warming effect on our climate - by long-lived greenhouse gases increased by 51.5%, with CO2 accounting for about 81% ...
  52. [52]
    [PDF] WMO Greenhouse Gas Bulletin. No 1
    Oct 16, 2025 · The AGGI reached 1.54 in 2024, representing a 54% increase in total radiative forcing(4) from 1990 to 2024 and a 1.5% increase from 2023 to 2024 ...
  53. [53]
    Climate change: atmospheric carbon dioxide
    The global average carbon dioxide set a new record high in 2024: 422.7 parts per million ("ppm" for short). The increase over 2023 amounts was 3.75 ppm—the ...
  54. [54]
    [PDF] Why the forcing from carbon dioxide scales as the logarithm of its ...
    The radiative forcing from carbon dioxide is approximately logarithmic in its concentration, pro- ducing about four watts per square meter of global-mean ...
  55. [55]
    State dependence of CO2 forcing and its implications for climate ...
    Nov 30, 2023 · When evaluating the effect of carbon dioxide (CO2) changes on Earth's climate, it is widely assumed that instantaneous radiative forcing ...
  56. [56]
    Water vapor and lapse rate feedbacks in the climate system
    Nov 30, 2021 · Water vapor is the most important greenhouse gas in Earth's atmosphere, absorbing more solar terrestrial radiation than any other ...Abstract · Article Text · Global Radiative Feedbacks... · Observational Evidence for...
  57. [57]
    The water vapor feedback - Yale Climate Connections
    Feb 4, 2008 · Water vapor serves as a feedback to temperature changes catalyzed by anthropogenic greenhouse gas emissions.
  58. [58]
    8.6.3.1 Water Vapour and Lapse Rate - AR4 WGI Chapter 8: Climate ...
    In GCMs, water vapour provides the largest positive radiative feedback (see Section 8.6.2.3): alone, it roughly doubles the warming in response to forcing ( ...
  59. [59]
    Are direct water vapor emissions endangering anyone? - RealClimate
    Jul 31, 2025 · That's why direct emissions of WV are COMPLETELY INSIGNIFICANT to AGW, and ANY effect of WV on AGW – is only (99.996%) as a positive FEEDBACK – ...
  60. [60]
    Revisiting the Role of the Water Vapor and Lapse Rate Feedbacks ...
    Apr 21, 2022 · The water vapor feedback, while being a positive feedback everywhere, is strongest in low latitudes and is found to contribute more to tropical ...
  61. [61]
    [PDF] water vapor feedback and global warming
    Oct 16, 2000 · To the extent that water vapor concentrations increase in a warmer world, the climatic effects of the other greenhouse gases will be amplified.
  62. [62]
    The Surface Warming Attributable to Stratospheric Water Vapor in ...
    Aug 25, 2020 · Stratospheric water vapor (SWV) is recognized as a potentially important positive feedback in global warming. The SWV change induces ...
  63. [63]
    Climate Change: Incoming Sunlight | NOAA Climate.gov
    On average, the Sun delivers 1,361 Watts of power per square meter at a distance of one astronomical unit. This amount is known as the total solar irradiance.
  64. [64]
    About Solar Irradiance | Earth - NASA
    The sun's total energy input reaching Earth is called total solar irradiance, or TSI. It comes in many different color bands or wavelengths. The distribution of ...
  65. [65]
    Negative trend in total solar irradiance over the satellite era - PMC
    Mar 10, 2025 · Accurate reconstruction of total solar irradiance (TSI) variations is important for explaining historical climate trends (1–3), constraining ...
  66. [66]
    Revised historical solar irradiance forcing - Astronomy & Astrophysics
    The estimated magnitude of the total solar irradiance (TSI) difference between the Maunder minimum and the present time ranges from 0.1 to 6 W m−2 making the ...2 Method · 3 Verification Of Total And... · 4 The Long-Term Evolution Of...
  67. [67]
    Placing limits on long-term variations in quiet-Sun irradiance and ...
    Jun 24, 2020 · Our analysis shows that the most likely upward drift in quiet-Sun radiative forcing since 1700 is between +0.07 and -0.13 W m -2.
  68. [68]
    A Solar Irradiance Climate Data Record in - AMS Journals
    We present a new climate data record for total solar irradiance and solar spectral irradiance between 1610 and the present day with associated wavelength and ...
  69. [69]
    Solar Irradiance - Forcings in GISS Climate Model - NASA
    Jun 4, 2025 · Total solar irradiance (TSI) is derived from Wang et al. (2005), with spectral variations from NRLSSI models. More recent calibration is ~1361 ...
  70. [70]
    Total Solar Irradiance (TSI)
    Total Solar Irradiance (TSI) Composite Database is compiled from many satellite TSI data collected from 1978 to present day by Claus Frohlich and Judith Lean.
  71. [71]
    1.4.3 Solar Variability and the Total Solar Irradiance - AR4 WGI ...
    The solar cycle variation in irradiance corresponds to an 11-year cycle in radiative forcing which varies by about 0.2 W m–2. There is increasingly reliable ...
  72. [72]
    Solar cycle as a distinct line of evidence constraining Earth's ...
    Dec 19, 2023 · Stratospheric temperature and radiative forcing response to 11-year solar cycle changes in irradiance and ozone. J. Atmos. Sci. 66, 2402 ...
  73. [73]
    Negative trend in total solar irradiance over the satellite era - PNAS
    We find a negative trend in TSI of −0.15 W/m 2 per decade over the satellite era of 1980 to 2023 with a 95% CI of −0.17 to −0.13 W/m 2 per decade.
  74. [74]
    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 ...
  75. [75]
    Solar Irradiance Variability: Modeling the Measurements - Lean - 2020
    May 18, 2020 · Spectral irradiance during the solar cycle varies ~50% less at wavelengths between 300 and 400 nm and more at visible and near-infrared ...
  76. [76]
    Global Surface Temperature Response to 11-Yr Solar Cycle Forcing ...
    Mar 10, 2021 · The 11-yr solar cycle is associated with a roughly 1 W m−2 trough-to-peak variation in total solar irradiance (TSI) over the past half century, ...
  77. [77]
    Reconciliation of modeled climate responses to spectral solar forcing
    May 20, 2013 · The percentage changes in VIS and NIR are about 0.16% or less and out of phase with UV irradiance variation. Though the relative changes in UV-A ...
  78. [78]
    Stratospheric Temperature and Radiative Forcing Response to 11 ...
    The 11-yr solar cycle temperature response to spectrally resolved solar irradiance changes and associated ozone changes is calculated using a fixed dynamical ...
  79. [79]
    Stratospheric and tropospheric response to enhanced solar UV ...
    Mar 11, 2003 · The stratosphere-troposphere system shows partly significant response to a realistic solar cycle enhancement of UV radiation.
  80. [80]
    The Footprint of the 11‐Year Solar Cycle in Northeastern Pacific ...
    Feb 16, 2021 · In this study, we show that this solar cycle can induce a ∼0.2°C sea surface temperature (SST) variation in the Northeastern Pacific.
  81. [81]
    3.1 Changes in Solar Output and in the Earth's Atmosphere
    The sun's luminosity has increased substantially over 4.57 billion years, and it is now about one-third brighter that it was originally.
  82. [82]
    Evolutionary variations of solar luminosity - NASA Technical Reports ...
    The Theoretical arguments for a 30% increase in the solar luminosity over the past 4.7 billion years are reviewed. A scaling argument shows that this ...Missing: per | Show results with:per
  83. [83]
    Constraints on the early luminosity history of the Sun
    Stellar evolution theory predicts that the Sun was fainter in the past, which can pose difficulties for understanding Earth's climate history. One proposed ...
  84. [84]
    Life Might Thrive on the Surface of Earth for an Extra Billion Years
    Sep 19, 2024 · The Sun's luminosity has already increased by about 30% since its formation, and the brightening will continue. Any increase in the Sun's ...
  85. [85]
    The faint young Sun problem - Feulner - 2012 - AGU Journals - Wiley
    May 25, 2012 · [43] The faint young Sun problem originates from the fact that the standard solar model implies a considerably lower luminosity for the early ...
  86. [86]
    Clouds and the Faint Young Sun Paradox - CP
    Mar 4, 2011 · Low clouds reflect sunlight but have little greenhouse effect. Removing them entirely gives a forcing of +25 W m−2 whilst more modest reduction ...
  87. [87]
    Faint young Sun paradox remains - Nature
    Jun 1, 2011 · Resolution of the 'faint young Sun paradox' requires a positive radiative forcing—from reducing the albedo or increasing the greenhouse ...Abstract · Main · Methods<|control11|><|separator|>
  88. [88]
    The evolution of habitable climates under the brightening Sun - Wolf
    Jun 16, 2015 · As the Sun brightens due to stellar evolution, Earth will become uninhabitable due to rising temperatures.
  89. [89]
    [PDF] Radiative forcing and temperature response to changes in urban ...
    Regionally, changes to radiative forcing from surface albedo changes can be much larger. For an increase in surface albedo of 0.09, due to an expansion of ...
  90. [90]
    equivalence metrics for surface albedo change based on the ... - ACP
    Jul 1, 2021 · Assessing the impact of a surface albedo change involves employing a measure like radiative forcing (RF) which can be challenging to digest for decision-makers.
  91. [91]
    Snow Cover and Vegetation‐Induced Decrease in Global Albedo ...
    Dec 2, 2017 · Variations in the albedo can affect the surface radiation budget and further impact the global climate. In this study, the interannual variation ...
  92. [92]
    Soot climate forcing via snow and ice albedos - PMC
    The climate forcing due to snow/ice albedo change is of the order of 1 W/m2 at middle- and high-latitude land areas in the Northern Hemisphere and over the ...
  93. [93]
    Global albedo change and radiative cooling from anthropogenic ...
    Oct 24, 2014 · This study quantified the direct consequences of anthropogenic land use change solely on albedo and TOA radiative forcing. It is important ...
  94. [94]
    Radiative forcing due to anthropogenic vegetation change based on ...
    Nov 9, 2005 · Without constraining the surface albedo change we arrive at a radiative forcing of −0.14 Wm−2 which is still weaker than most earlier estimates.
  95. [95]
    Earth's albedo variations 1998–2014 as measured from ground ...
    Apr 4, 2016 · The Earth's albedo is a fundamental climate parameter for understanding the radiation budget of the atmosphere. It has been traditionally ...Missing: impact | Show results with:impact
  96. [96]
  97. [97]
    Measuring Earth's Albedo - NASA Earth Observatory
    Oct 20, 2014 · Changes in ice cover, cloudiness, airborne pollution, or land cover (from forest to farmland, for instance) all have subtle effects on global ...
  98. [98]
    What Is Albedo and What Does It Have to Do With Global Warming?
    Nov 7, 2023 · Other natural events like erupting volcanoes or large wildfires can also change Earth's albedo in an extreme way.
  99. [99]
    Climatic Impact of Global-Scale Deforestation: Radiative versus ...
    Deforestation causes a net cooling of -1K globally due to albedo increase, but also has warming effects from decreased evapotranspiration and surface roughness.
  100. [100]
    Biases in the albedo sensitivity to deforestation in CMIP5 models ...
    Dec 16, 2020 · Biases in the albedo sensitivity to deforestation in CMIP5 models and their impacts on the associated historical radiative forcing.
  101. [101]
    Albedo changes caused by future urbanization contribute to global ...
    Jul 1, 2022 · Radiative forcing. Urbanization worldwide produces a warming effect through reducing albedo in both the past and the projected futures. Relative ...Surface Albedo Change · Radiative Forcing · Estimation Of Albedo Change...
  102. [102]
    [PDF] The response of radiative forcing to high spatiotemporally resolved ...
    A new LUC dataset shows a higher radiative forcing from land-use transition and CO2 emissions in China, reducing the cooling effect from 2000-2010.<|separator|>
  103. [103]
    Highly resolved satellite-remote-sensing-based land-use-change ...
    Apr 15, 2025 · The result reveals that the global LUC-induced surface albedo change may not significantly slow down global warming as was previously anticipated.
  104. [104]
    [PDF] Chapter 6: Short-lived Climate Forcers
    The asymmetric aerosol and greenhouse gas forcing on regional‑ scale climate responses have also been assessed to lead to contrasting effects on ...
  105. [105]
    Robust evidence for reversal of the trend in aerosol effective climate ...
    Sep 21, 2022 · Here, we present and discuss the evolution of the aerosol forcing since 2000. There are multiple lines of evidence that allow us to robustly ...<|control11|><|separator|>
  106. [106]
    New Estimates of Aerosol Direct Radiative Effects and Forcing From ...
    Jun 17, 2019 · We estimate that the global mean aerosol direct radiative effect is −2.40 W/m2 with an error of ± 0.6 W/m2 owing to uncertainties in aerosol ...
  107. [107]
    Uncertainties in an Observation-Based Estimate of the Global ...
    Although the magnitude of the aerosol direct radiative effect (DRE) is estimated to be less than that of the indirect effect, uncertainties are large.
  108. [108]
    [PDF] Comparison of methods to estimate aerosol effective radiative ... - ACP
    Aug 9, 2023 · Uncertainty in the effective radiative forcing (ERF) of climate primarily arises from the unknown contribution of aerosols, which impact ...
  109. [109]
    [PDF] Increasing aerosol direct effect despite declining global emissions in ...
    Sep 25, 2024 · Our results reveal an increase in global mean aerosol radiative forcing in recent decades, despite a global reduction in aerosol emissions.
  110. [110]
    Aerosols and Their Importance | Earth - NASA
    Aerosols scatter (reflect) a portion of the Sun's incoming light, and, dependent on type, may also absorb some. These are known as 'direct radiative effects' of ...
  111. [111]
    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).
  112. [112]
    Global observations of aerosol indirect effects from marine liquid ...
    Oct 18, 2023 · Interactions between aerosols and liquid clouds are one of the largest sources of uncertainty in the historical radiative forcing of climate ...
  113. [113]
    Radiative forcing from aerosol–cloud interactions enhanced by large ...
    Nov 20, 2023 · We show that the effective radiative forcing due to aerosol–cloud interactions is strongly enhanced by adjustments of large-scale circulation to aerosol.
  114. [114]
    Constraining the Twomey effect from satellite observations - ACP
    Dec 4, 2020 · The Twomey effect describes the radiative forcing associated with a change in cloud albedo due to an increase in anthropogenic aerosol emissions.
  115. [115]
    Observational evidence of strong forcing from aerosol effect on low ...
    Nov 8, 2023 · The estimated forcing from the Twomey effect is −0.75 W m−2 ... The radiative forcing due to the Cf adjustment can be 212% of the Twomey ...
  116. [116]
    Aerosols enhance cloud lifetime and brightness along the stratus-to ...
    Jul 13, 2020 · We show that aerosols can enhance cloud fraction and extend the lifetime of overcast cloud fields primarily under stable atmospheric conditions.
  117. [117]
    Constraining cloud lifetime effects of aerosols using A-Train satellite ...
    Feb 17, 2021 · Aerosol indirect effects have remained the largest uncertainty in estimates of the radiative forcing of past and future climate change.
  118. [118]
    Large contribution of natural aerosols to uncertainty in indirect forcing
    Nov 6, 2013 · Our results show that 45 per cent of the variance of aerosol forcing since about 1750 arises from uncertainties in natural emissions.
  119. [119]
    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 ...
  120. [120]
    Volcanic Radiative Forcing From 1979 to 2015 - AGU Journals - Wiley
    Oct 24, 2018 · Small-magnitude eruptions caused a global-mean ERF of −0.08 W/m2 during 2005–2015 relative to the volcanically quiescent 1999–2002 period In ...<|control11|><|separator|>
  121. [121]
    Is Anthropogenic Global Warming Accelerating? - AMS Journals
    (2017) along with radiative forcing estimates from Fig. 2 to estimate the level of anthropogenic warm- ing (black) in comparison with natural forcing (blue), ...
  122. [122]
    Atmospheric CO2 concentrations over the last glacial termination
    Abstract. A record of atmospheric carbon dioxide (CO2) concentration during the transition from the Last Glacial Maximum to the Holocene, obtained from the Dome ...Missing: peer- reviewed
  123. [123]
    [PDF] Ice core records of atmospheric carbon dioxide
    Ice core CO2 records shows variability on timescales ranging from glacial-interglacial (60- 90ppm), millennial (10-30ppm), and centennial (5-15ppm). The stable ...
  124. [124]
    Glacial cycles and carbon dioxide: A conceptual model - AGU Journals
    Jan 1, 2008 · In this model, temperature rises lead CO2 increases at the glacial termination, but it is the feedback between these two quantities that drives ...
  125. [125]
    Milankovitch (Orbital) Cycles and Their Role in Earth's Climate
    Feb 27, 2020 · Milankovitch combined the cycles to create a comprehensive mathematical model for calculating differences in solar radiation at various Earth ...Missing: magnitude | Show results with:magnitude
  126. [126]
    Explainer: How the rise and fall of CO2 levels influenced the ice ages
    Jul 2, 2020 · In this case, CO2 is not the immediate cause of ice ages; rather, it serves as a feedback to amplify changes initiated by orbital variations.
  127. [127]
    Modern Grand Solar Minimum will lead to terrestrial cooling - PMC
    Aug 4, 2020 · The most recent grand solar minimum occurred during Maunder Minimum (1645–1710), which led to reduction of solar irradiance by 0.22% from the ...
  128. [128]
    The Maunder minimum and the Little Ice Age: an update from recent ...
    Dec 4, 2017 · The Maunder minimum (MM) was a period of extremely low solar activity from approximately AD 1650 to 1715.
  129. [129]
    Solar Forcing of Regional Climate Change During the Maunder ...
    A minimum in solar irradiance, the Maunder Minimum, is thought to have occurred from the mid-1600s to the early 1700s (1–3). Concurrently, surface temperatures ...
  130. [130]
    Rates of change in natural and anthropogenic radiative forcing over ...
    We compare rates of change of anthropogenic forcing with rates of natural greenhouse gas forcing since the Last Glacial Maximum and of solar and volcanic ...
  131. [131]
    Assessment of pre-industrial to present-day anthropogenic climate ...
    Jan 29, 2021 · We quantify and analyse a wide range of present-day (PD) anthropogenic effective radiative forcings (ERFs) with the UK's Earth System Model (ESM), UKESM1.
  132. [132]
    Climate forcing by the volcanic eruption of Mount Pinatubo
    Mar 11, 2005 · We determine the volcano climate sensitivity λ and response time τ for the Mount Pinatubo eruption, using observational measurements of the ...
  133. [133]
    Solar influence on climate during the past millennium - PNAS
    The potential role of solar variations in modulating recent climate has been debated for many decades and recent papers suggest that solar forcing may be ...
  134. [134]
    A new approach to the long-term reconstruction of the solar ...
    Effects of solar radiative forcing on the climate. Variations on time-scales of up to the 27-day rotational period have an important influence on space weather, ...Missing: luminosity | Show results with:luminosity
  135. [135]
    Simulating the atmospheric response to the 11-year solar cycle ...
    The UM-UKCA model produces a statistically significant response to the 11-year solar cycle in stratospheric temperatures, ozone and zonal winds.
  136. [136]
    Pacific multidecadal (50–70 year) variability instigated by volcanic ...
    Jan 13, 2022 · The Pacific decadal oscillation (PDO) is the leading mode of decadal climate variability over the North Pacific. However, it remains unknown ...
  137. [137]
    Chapter 3: Human Influence on the Climate System
    Figure 3.10 shows that CMIP6 models forced by combined anthropogenic and natural forcings overestimate temperature trends compared to radiosonde data ( ...
  138. [138]
    Detection and Attribution of Recent Climate Change - AMS Journals
    This paper addresses the question of where we now stand with respect to detection and attribution of an anthropogenic climate signal.
  139. [139]
    [PDF] Chapter 3: Human Influence on the Climate System
    when including anthropogenic forcings, and formal attribution studies. Since the beginning of the modern satellite era in 1979, Northern. Hemisphere sea ice ...
  140. [140]
    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.
  141. [141]
    Analysis: Why scientists think 100% of global warming is due to ...
    Dec 13, 2017 · Overall, warming associated with all human forcings agrees quite well with observed warming, showing that about 104% of the total since the ...
  142. [142]
    Detection and Attribution of Human Influence on the Global Diurnal ...
    Jun 29, 2022 · The DTR is an effective proxy to assess changes in radiative forcing (Thorne et al., 2016; Wang and Clow, 2020), a critical index for climate ...
  143. [143]
    [PDF] Understanding and Attributing Climate Change
    radiative forcing for those periods. A substantial fraction of the ... based on physical principles. This understanding can take the form of conceptual ...
  144. [144]
    On the climate sensitivity and historical warming evolution in recent ...
    Jul 6, 2020 · The equilibrium climate sensitivity (ECS) is defined as the long-term globally averaged amount of surface temperature increase in response ...<|separator|>
  145. [145]
    sensitivity to radiative forcing time series and observational data - ESD
    Jun 21, 2018 · Inferred effective climate sensitivity (ECS inf ) is estimated using a method combining radiative forcing (RF) time series and several series of observed ocean ...
  146. [146]
    A perspective on climate sensitivity - ScienceDirect
    The pioneering work of Charney et al. (1979) estimated the range of equilibrium climate sensitivity (ECS) being from 1.5 K to 4.5 K, in which the ECS is defined ...
  147. [147]
    [PDF] Climate Change 2021
    ... range of equilibrium climate sensitivity is between 2°C (high confidence) and 5°C (medium confidence). The AR6 assessed best estimate is 3°C with a likely ...
  148. [148]
    Opinion: Can uncertainty in climate sensitivity be narrowed further?
    Feb 29, 2024 · Equilibrium climate sensitivity (ECS) – the eventual rise in Earth's mean surface temperature following a doubling of atmospheric CO2 – has a ...
  149. [149]
    Biased Estimates of Equilibrium Climate Sensitivity and Transient ...
    Dec 9, 2021 · Estimates from historical energy budget constraints underestimate equilibrium climate sensitivity and transient climate response in models ...1 Introduction · 2 Data · 5 Discussions And...Missing: criticisms | Show results with:criticisms<|control11|><|separator|>
  150. [150]
    Sea-surface temperature pattern effects have slowed global ... - PNAS
    When driven by observed patterns, even high ECS models produce low EffCS values consistent with the observed global warming rate. The inability of CMIP5/6 ...Abstract · Sign Up For Pnas Alerts · Discussion And ConclusionsMissing: criticisms | Show results with:criticisms
  151. [151]
    Context for interpreting equilibrium climate sensitivity and transient ...
    Here we review and synthesize the latest developments in ECS and TCR values in CMIP, compile possible reasons for the current values as supplied by the ...
  152. [152]
    Quantifying the Sources of Intermodel Spread in Equilibrium Climate ...
    Equilibrium climate sensitivity (ECS; the global-average equilibrium surface temperature change due to CO2 doubling) is one of our primary measures of the ...
  153. [153]
    Addressing misconceptions about Climate Sensitivity research
    Aug 13, 2025 · This predicted pattern change underlies the weakening of climate feedbacks over time in GCM simulations, which contributes to their higher ECS ...
  154. [154]
    Guest post: Ice-age analysis suggests worst-case global warming is ...
    May 1, 2024 · However, observed data from recent warming are not as useful for constraining the upper end of ECS estimates. This is because climate feedbacks ...Missing: criticisms | Show results with:criticisms
  155. [155]
    Comparison of Equilibrium Climate Sensitivity Estimates From Slab ...
    Jul 23, 2020 · Equilibrium climate sensitivity (ECS) estimates for a single coupled model can vary by more than 1°C (20%) depending on analysis method · ECS ...
  156. [156]
    [PDF] Future Global Climate: Scenario-based Projections and Near-term ...
    uncertainty in the radiative forcing in projections (Vial et al., 2013) and ... specifically to idealized CO2 forcing and generally GSAT change, AR6.
  157. [157]
    Quantifying the Uncertainty Sources of Future Climate Projections ...
    Oct 19, 2022 · Model uncertainty is the second source of uncertainty. Different models respond differently to climate change with the same radiative forcing ...
  158. [158]
    Significant impact of forcing uncertainty in a large ensemble of ... - NIH
    Here we use a Earth System Model to evaluate the role of external forcing uncertainty in simulations of past and future climate change.
  159. [159]
    Identifying climate model structural inconsistencies allows for tight ...
    Aug 8, 2023 · Aerosol radiative forcing uncertainty affects estimates of climate sensitivity and limits model skill in terms of making climate projections.Missing: criticisms | Show results with:criticisms
  160. [160]
    High radiative forcing climate scenario relevance analyzed with a ...
    Sep 18, 2024 · While most of this paper discusses radiative forcing as it is the key input to climate models, temperature uncertainty is important as ...Missing: explanation | Show results with:explanation
  161. [161]
    Observed trend in Earth energy imbalance may provide a ... - Science
    Jun 12, 2025 · The CERES data show a stronger trend in EEI than the multimodel CMIP6 mean and higher EEI in 2023 than any of the CMIP6 models. However, for ...
  162. [162]
    Recent reductions in aerosol emissions have increased Earth's ...
    Apr 3, 2024 · The trend in instantaneous and effective radiative forcing due to cooling aerosols has reversed, now showing a robust positive trend for both ...
  163. [163]
    Earth's Energy Imbalance More Than Doubled in Recent Decades
    May 10, 2025 · The root cause of the discrepancy between models and observations is currently not well known, but it seems to be dominated by a decrease in ...
  164. [164]
    Assessing effective radiative forcing from aerosol–cloud interactions ...
    The total ERFaci inferred from CERES observations is − 1.11 ± 0.43 W m−2 (95% CI). This case is considered to be the best estimate of total ERFaci because ...
  165. [165]
    First assessment of the earth heat inventory within CMIP5 historical ...
    May 10, 2021 · The energy imbalance at the top of the atmosphere over the last ... However, the CMIP5 ensemble overestimates the ocean heat content by ...
  166. [166]
    Impacts of the Unforced Pattern Effect on the Cloud Feedback in ...
    Jan 18, 2022 · We find a large unforced pattern effect in CERES data, with cloud feedback estimated from two consecutive 125-month periods (March 2000–July 2010 and August ...
  167. [167]
    Diagnosing the Radiation Biases in Global Climate Models Using ...
    Jul 10, 2023 · Radiation energy balance at the top of atmosphere is crucial to the Earth climate system and is routinely examined in climate model validations.
  168. [168]
    [PDF] Observational Assessment of Changes in Earth's Energy Imbalance ...
    - Decrease in AM4 reflected SW is significantly weaker than CERES, even when climate forcing is included.
  169. [169]
    Debates on the Causes of Global Warming - ScienceDirect.com
    Mar 25, 2012 · The controversy between the IPCC and Non-governmental IPCC (NIPCC) on the attribution of global warming are reviewed.
  170. [170]
    The lure of solar forcing - RealClimate
    Jul 15, 2005 · “It is found that current climate models underestimate the observed climate response to solar forcing over the twentieth century as a whole, ...<|separator|>
  171. [171]
    Why must a solar forcing be larger than a CO 2 forcing to cause the ...
    Apr 7, 2016 · In this study, we investigate the physical mechanisms responsible for the lower efficacy of solar forcing compared to an equivalent CO 2 forcing.Missing: controversies | Show results with:controversies
  172. [172]
    Uncertainties in the attribution of greenhouse gas warming and ...
    Jun 13, 2016 · However, the partitioning of the anthropogenic influence to individual factors, such as greenhouse gases and aerosols, is much less robust.Introduction · Data · Climate Response Patterns... · Results of an Optimal...
  173. [173]
    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.
  174. [174]
    Uncertainty in aerosol effective radiative forcing from anthropogenic ...
    Jun 26, 2025 · We find that uncertainty in aerosol ERF is dominated by sulfate-related processes, biomass burning, size, and natural emissions.
  175. [175]
    Critical Review on Radiative Forcing and Climate Models for Global ...
    Thus, the radiative forcing provides a useful metric to access and compare the impacts of various anthropogenic and natural variations on the Earth's climate.Missing: peer- | Show results with:peer-
  176. [176]
    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).
  177. [177]
    Uncertainty in Observational Estimates of the Aerosol Direct ...
    Aerosols continue to be responsible for the largest uncertainty in determining the anthropogenic radiative forcing of the climate. Aerosols influence ...2. Methodology · A. Radiative Kernels · Share Link
  178. [178]
    Empirical assessment of the role of the Sun in climate change using ...
    As a result, by using the estimated historical forcings, the CMIP6 GCMs concluded that the Sun would have made a negligible contribution to the observed global ...<|control11|><|separator|>
  179. [179]
    Download this table as comma separated values (csv).
    Global Radiative Forcing 1979 - 2023 (W m-2),,,,,,,,CO2-eq,AGGI ... 2023,2.286,0.565,0.223,0.301,0.061,0.051,3.485,534,1.515,1.6 ,,,,,,,,,, "* for the list ...
  180. [180]
    [PDF] WMO Greenhouse Gas Bulletin - UN CC:Learn
    Oct 28, 2024 · The AGGI reached 1.51 in 2023, representing a 51.5% increase in total radiative forcing(4) from 1990 to 2023 and a 1.6% increase from 2022 to ...
  181. [181]
    Radiative forcing from the 2020 shipping fuel regulation is large but ...
    Jan 13, 2025 · Here we employ machine learning to capture cloud natural variability and estimate a radiative forcing of +0.074 ±0.005 W m −2 related to IMO2020.
  182. [182]
    Multi-model effective radiative forcing of the 2020 sulfur cap ... - ACP
    Dec 4, 2024 · The 2020 sulfur cap reduced SO2 emissions by ~80%, weakening aerosol radiative forcing. Model results show an ERF of 0.06 to 0.09 W m-2.Missing: post- | Show results with:post-
  183. [183]
    [PDF] Recent global temperature surge intensified by record-low planetary ...
    Dec 6, 2024 · The contribution of the El Niño albedo signature to the 2023 GMST anomaly is about an order of magnitude smaller compared to the general ...
  184. [184]
    Indicators of Global Climate Change 2024: annual update of key ...
    Jun 19, 2025 · The 2024-observed best estimate of global surface temperature (1.52 °C) is well above the best estimate of human-caused warming (1.36 °C).
  185. [185]
    Carbon dioxide levels increase by record amount to new highs in 2024
    Oct 15, 2025 · From 2023 to 2024, the global average concentration of CO2 surged by 3.5 ppm, the largest increase since modern measurements started in 1957. “ ...
  186. [186]