Fact-checked by Grok 2 weeks ago

Climate sensitivity

Equilibrium climate sensitivity (ECS) is defined as the long-term equilibrium change in global mean surface air temperature that would result from a doubling of atmospheric (CO₂) concentration from pre-industrial levels, after allowing sufficient time for the to adjust, including deep ocean heat uptake. This metric encapsulates the combined effects of from doubled CO₂—approximately 3.7 W/m²—and subsequent climate feedbacks, such as amplification, changes, surface variations, and responses, which determine the net sensitivity. ECS remains a cornerstone of projection models, as it scales projected warming linearly with cumulative emissions under the assumption of constant feedback strength, though empirical assessments reveal persistent uncertainties due to incomplete quantification of feedbacks, particularly clouds. Historical estimates, originating from Charney et al. (1979), placed ECS between 1.5°C and 4.5°C, a range that has endured despite advances, with recent observationally constrained studies suggesting values potentially below 3°C based on energy budget analyses of recent warming, while comprehensive model ensembles and some paleoclimate proxies indicate higher values up to 5°C or more. Controversies persist over the reliability of different inference methods—instrumental records yielding lower sensitivities around 2°C, contrasted with process-oriented general circulation models averaging near 3°C and paleoclimate data implying greater variability—underscoring debates about feedback realism and potential biases in favoring model-derived over purely empirical constraints.

Fundamentals

Definition and Core Concepts

Climate sensitivity refers to the change in global mean surface resulting from a specified , such as a doubling of atmospheric concentration from pre-industrial levels. In its basic formulation, it quantifies the long-term response after the climate system reaches a new radiative balance, accounting for both direct effects of the forcing and subsequent feedbacks. This metric, often expressed in degrees per doubling of CO2 (ECS), provides a measure of the climate system's to perturbations without specifying transient . Central to the concept is (ΔF), defined as the net change in downward minus upward at the due to an external perturbation, held constant while allowing stratospheric temperatures to adjust. For doubled CO2, ΔF is approximately 3.7 W/m², derived from line-by-line calculations. The no-feedback sensitivity, based on the Planck response from Stefan-Boltzmann law, yields about 1.2°C warming, as the planet's effective emission temperature rises to restore energy balance. Actual sensitivity incorporates amplifying and damping feedbacks, such as (positive) and (negative), leading to estimates typically ranging from 2 to 5°C in global climate models. The framework assumes a linear response where S = ΔT_eq / ΔF, though nonlinearities and state dependencies can influence outcomes. Empirical constraints from paleoclimate records, instrumental data, and understanding inform estimates, revealing persistent uncertainties due to factors like feedbacks and effects. This metric underpins projections of warming, as forcing from gases since 1750 equates to about 2.7 W/m² by 2019.

Radiative Forcing and Temperature Equilibrium

quantifies the perturbation to Earth's top-of-atmosphere balance induced by factors such as increased concentrations, expressed as the net change in downward minus upward at the in watts per square meter (W/²). A positive forcing creates an surplus, prompting planetary warming that continues until increases sufficiently to restore , where absorbed solar balances emitted terrestrial . This adjustment primarily occurs through the temperature dependence of blackbody emission, as described by the Stefan-Boltzmann law: outgoing flux F = \sigma T^4, with Stefan-Boltzmann constant \sigma = 5.67 \times 10^{-8} W ⁻² ⁻⁴ and effective emitting T_e \approx 255 . For small forcings, the relationship linearizes to \Delta F \approx 4 \sigma T_e^3 \Delta T, yielding the Planck feedback magnitude of approximately 3.3 m⁻² ⁻¹, which sets the no- climate response \lambda_0 = \Delta T / \Delta F \approx 0.30 (/m²)⁻¹. In this baseline scenario, absent other feedbacks, warming equals the forcing divided by this parameter's reciprocal. For doubled pre-industrial CO₂, yielding \Delta F_{2 \times \mathrm{CO_2}} \approx 3.7 /m², the no-feedback \Delta T_\mathrm{eq} is thus about 1.1 . Actual incorporates feedbacks that amplify or dampen this response by altering the effective feedback parameter, but the Planck term dominates the stabilizing mechanism. Empirical and modeled estimates confirm the Planck feedback's robustness near -3.3 m⁻² ⁻¹ globally, though local variations exist due to and emission height effects.

Key Metrics and Their Interrelations

The primary metrics quantifying climate sensitivity are equilibrium climate sensitivity (ECS) and transient climate response (TCR), both expressed relative to a doubling of atmospheric CO₂ concentration from pre-industrial levels. ECS represents the long-term equilibrium global mean surface air change (ΔT_eq) after the fully adjusts to the (ΔF) from doubled CO₂, incorporating all slow feedbacks such as dynamics and vegetation shifts on centennial to millennial timescales. TCR, in contrast, measures the surface response at the time of CO₂ doubling under a of gradual 1% annual increase over approximately 70 years, reflecting faster processes but excluding full equilibration due to thermal inertia, particularly ocean heat uptake. These metrics interrelate through the feedback parameter λ (in W m⁻² K⁻¹), where the no-feedback sensitivity (Planck response alone) yields ΔT ≈ 1.2°C for ΔF_{2×CO₂} ≈ 3.7 W m⁻², and total sensitivity S = -ΔF / (λ ΔT), with feedbacks amplifying or the response. The core interrelation stems from the energy balance equation at : ΔT_eq = -ΔF / λ, where λ aggregates the Planck feedback (≈ -3.2 to -3.8 W m⁻² K⁻¹ from ) and additional climate s (water vapor, , , clouds). For ECS, λ includes all s reaching steady state, yielding S_{ECS} = ΔT_eq × (ΔF_{2×CO₂} / ΔF), but adjusted for net forcing changes from s (H): S = ΔT × ΔF_{2×CO₂} / (ΔF - H). TCR approximates a where surface warming lags ECS because deep layers absorb without immediate full activation, typically resulting in TCR ≈ 0.6–0.9 × ECS in models, depending on uptake efficiency and timing. ΔF_{2×CO₂} serves as the baseline input, calculated as the net top-of-atmosphere flux perturbation (≈3.711 W m⁻² from line-by-line ), linking both metrics to concentrations while s determine the amplification factor 1/|λ|. Empirical constraints highlight interdependencies: observed historical warming (instrumental era) better aligns with TCR due to its shorter timescale, while paleoclimate proxies inform ECS but introduce uncertainties from non-CO₂ forcings and state dependencies in feedbacks. In coupled models, state-dependent variations in λ (e.g., weakening Planck feedback at higher temperatures) cause TCR to scale more directly with near-term projections than ECS, underscoring TCR's relevance for multi-decadal responses despite ECS capturing ultimate commitment.

Types of Climate Sensitivity

Equilibrium Climate Sensitivity (ECS)

Equilibrium climate sensitivity (ECS) is defined as the long-term change in global mean surface air temperature at equilibrium following a doubling of atmospheric CO₂ concentration from pre-industrial levels, after the full adjustment of fast and slow feedbacks, including , , , and potentially ice sheets over centuries. This metric assumes a sustained of approximately 3.7 W/m² from the CO₂ doubling, with the equilibrium temperature response incorporating both the no-feedback Planck response and all subsequent feedbacks. The Intergovernmental Panel on Climate Change's Sixth Report (AR6), released in 2021, synthesizes evidence from models, observations, and paleoclimate data to conclude that ECS is likely (66-100% probability) between 2.5°C and 4.0°C per CO₂ doubling, with a best estimate of 3.0°C and very likely (5-95% range) between 2.0°C and 5.0°C. This assessment narrowed the upper bound compared to prior reports due to improved constraints on feedbacks and historical forcing estimates, though substantial uncertainty persists from incomplete quantification of low- responses and effects. Empirical estimates derived from records, such as radiative fluxes and heat uptake since the mid-20th century, often yield lower central values, typically 2.0°C to 2.7°C, highlighting potential overestimation in comprehensive climate models that simulate higher ECS values up to 5°C or more. For instance, energy budget analyses constraining feedbacks from observed trends suggest ECS around 2.5-2.7°C, consistent with mid-range paleoclimate proxies but challenging high-sensitivity model ensembles. These observational approaches prioritize recent data over process-based simulations, revealing tensions where models with ECS above 4°C underperform in reproducing historical warming patterns. Key uncertainties in ECS arise from feedbacks, particularly in subtropical stratocumulus regions, where positive or negative responses to warming remain observationally ambiguous, and from the efficacy of non-CO₂ forcings like in historical contexts. Paleoclimate reconstructions, such as constraints, support ECS below 4°C but are sensitive to assumptions about equilibrium states and dust forcings. ECS informs long-term projections, as transient warming lags equilibrium by centuries due to heat capacity, emphasizing the need for resolved empirical bounds to assess committed warming risks.

Transient Climate Response (TCR)

The transient climate response (TCR) quantifies the global mean surface air increase at the point of atmospheric CO2 doubling under an idealized scenario of 1% per year compounded CO2 growth from pre-industrial levels (approximately 280 to 560 ), with the averaged over a 20-year period centered on the doubling year (typically around year 70). This metric captures the partial equilibration of the during rapid forcing, where heat uptake delays surface warming relative to the full response. In climate models, TCR is derived from simulations following the specified CO2 ramp, incorporating transient feedbacks and ocean dynamics; across Phase 6 (CMIP6) ensembles, TCR values range from 1.2°C to 2.8°C, with a multimodel mean around 1.8°C, though subsets constrained by historical simulations show narrower spreads (e.g., 1.68°C best estimate). The (IPCC) Sixth Assessment Report (AR6, 2021) assesses TCR as likely between 1.4°C and 2.2°C (very likely 1.2–2.4°C), informed by model simulations, paleoclimate data, and instrumental records, emphasizing that TCR is systematically lower than equilibrium climate sensitivity (ECS) by roughly 30–50% due to finite limiting immediate surface response. Observational estimates of TCR, derived from historical warming (e.g., 1850–2020 global temperature rise of ~1.1°C against ), often yield lower central values than unconstrained model means, such as 1.4°C (90% 0.9–2.0°C) using on energy budget constraints, or ~1.3°C from detection-attribution frameworks applied to 20th-century data. Recent econometric analyses of instrumental records produce estimates like 2.17°C (95% CI 1.72–2.77°C), while emergent constraints from model-observation comparisons highlight potential low bias in historical TCR inferences due to unmodeled variability or forcing uncertainties. These discrepancies arise partly from model overestimation of transient feedbacks (e.g., clouds) or underestimation of cooling in historical periods, underscoring the need for validation against empirical data rather than model priors alone. Uncertainties in TCR stem primarily from the efficiency of ocean heat uptake, short-term feedbacks like and , and the transient nature of forcing agents (e.g., aerosols masking CO2 effects); studies indicate that CMIP6 models with TCR exceeding ~2.0°C often project historical warming mismatches unless adjusted for observational fidelity. TCR informs near- to mid-term projections (e.g., by 2100 under moderate emissions), where full ECS realization is incomplete, and lower observational TCR implies reduced committed warming for given cumulative emissions compared to high-end model projections.

Effective and Earth System Sensitivities

Effective climate sensitivity refers to the mean surface response to a doubling of atmospheric CO₂ concentration, evaluated from model simulations or observations under non- conditions with time-varying . This metric approximates the response but is influenced by transient factors, such as heat uptake, which absorbs excess and delays surface warming, typically yielding values lower than climate sensitivity (ECS). In abrupt experiments like those in CMIP6 models (e.g., abrupt-4xCO₂), effective sensitivity is computed via of anomalies against cumulative forcing, often aligning closely with model ECS estimates ranging from 2.0°C to 6.0°C. Observational inferences from the instrumental record (1850–present), accounting for historical forcing and imbalance, constrain effective sensitivity to approximately 2.0–3.0°C, lower than many model-derived values, highlighting potential overestimation of feedbacks in some simulations. Earth system sensitivity (ESS) extends beyond ECS by incorporating very slow feedbacks operating on millennial timescales, including changes in continental ice sheet volume, permafrost carbon release, vegetation shifts, and biogeochemical cycles like silicate weathering. These processes can either amplify warming (e.g., reduced ice cover decreasing albedo) or provide damping (e.g., enhanced chemical weathering drawing down CO₂), with the net effect uncertain but often estimated higher than ECS due to dominant positive ice-albedo effects in paleoclimate contexts. Analysis of Cenozoic-era temperature and CO₂ proxy data, integrating carbon cycle dynamics, yields a median ESS of 3.4°C (5–95% range: 2.6–4.7°C) per CO₂ doubling, narrower than prior paleo-based assessments and incorporating negative weathering feedbacks that mitigate long-term warming. Other reconstructions, such as from Eocene hyperthermals or Pleistocene ice age cycles, suggest potentially higher ESS values exceeding 5°C if ice sheet and vegetation dynamics outweigh chemical damping, though uncertainties persist from proxy inaccuracies and forcing reconstructions. ESS is less relevant for near-term projections (next centuries) but informs deep-time climate stability and committed warming from past emissions.

Physical Determinants

Feedback Mechanisms: Water Vapor, Lapse Rate, and Surface Albedo

The feedback operates through the thermodynamic response of the atmosphere to warming, wherein higher temperatures increase the atmosphere's capacity to hold via the Clausius-Clapeyron relation, which predicts an approximate 7% rise in saturation vapor pressure per kelvin of warming under constant relative conditions. This added , a potent , traps additional , amplifying the initial and contributing positively to climate sensitivity. Satellite observations and reanalyses, such as those from the Atmospheric Infrared Sounder (AIRS), corroborate that specific has increased in the consistent with model projections, with minimal changes in relative supporting the feedback's robustness. Peer-reviewed assessments place the feedback parameter at +1.6 to +2.0 W m⁻² K⁻¹, though combined with effects, the net is moderated to around +1.25 ± 0.15 W m⁻² K⁻¹. The feedback stems from alterations in the vertical temperature profile of the under . In a no- of uniform temperature increase with height, the feedback would be neutral; however, moist convection in the causes the upper to warm more rapidly than the surface due to release, steepening the and reducing the by facilitating greater longwave emission to space, yielding a . Conversely, in polar and subtropical regions, surface warming outpaces the upper , flattening the and enhancing trapping, but the tropical dominance results in a net negative contribution. Estimates from general circulation models and observational constraints indicate a lapse rate feedback parameter of approximately -0.8 to -1.0 W m⁻² K⁻¹, with the water vapor-lapse rate combination netting positive due to water vapor's larger magnitude. Surface albedo feedback primarily manifests through the ice-albedo effect, where warming-induced retreat of snow cover, sea ice, and continental ice sheets exposes darker underlying surfaces that absorb more solar radiation, reducing Earth's reflectivity and exerting a positive forcing. Arctic observations from 1979 to 2011 document a regional albedo decline from 0.52 to 0.48, equivalent to an additional 6.4 ± 0.9 W m⁻² of absorbed solar energy locally, though global impacts are diluted by the region's small area. Model-based evaluations for doubled CO₂ scenarios yield a surface albedo feedback parameter of +0.3 to +0.5 W m⁻² K⁻¹, with ice sheet contributions amplifying it by up to 42% in high-latitude projections; uncertainties arise from nonlinear sea ice melt dynamics and potential vegetation changes, but the feedback remains consistently positive across ensembles.

Cloud Feedbacks and Their Uncertainties

Cloud feedbacks represent the change in the net radiative of clouds per unit surface temperature change, primarily through alterations in , altitude, , and phase, which modulate shortwave (SW) reflection and (LW) emission. Low-level clouds, dominant in subtropical stratocumulus regions, exert a negative forcing by reflecting ; their projected decline under warming reduces this reflection, yielding a positive SW feedback estimated at 0.3 to 0.5 W/m²/K across models. High-level clouds, such as , contribute positively via enhanced LW trapping from increased cover or elevated tops, offsetting some negative feedbacks from mid-level cloud adjustments. The net integrates these competing processes, with tropical and subtropical clouds driving most variability. In Phase 6 (CMIP6) simulations, the ensemble-mean net cloud feedback averages 0.52 W/m²/K, ranging from near-zero to over 1 W/m²/K, largely due to divergent representations of low-cloud responses to () gradients and stability changes like estimated inversion strength (EIS). This spread accounts for over half the variance in equilibrium climate sensitivity (ECS) across models, with higher feedbacks correlating to ECS values exceeding 4 K. Positive contributions stem from reduced low-cloud fraction in regions and LW effects from rising tropopause-relative cloud tops, while negative elements include potential increases in clouds. Observationally derived estimates, using satellite radiances from (2000–2019) and reanalysis for stability metrics, yield a net cloud feedback of 0.43 /m²/K (90% : 0.08–0.78 /m²/K), with sensitivity to surface temperature and EIS dominating the signal and implying amplification of warming. This aligns with model means but carries large uncertainty from sampling noise, radiative kernel approximations, and unaccounted rapid adjustments. Low-cloud feedbacks, inferred from trends in subtropical cover, show positive tendencies but weaker than in high-ECS models, suggesting possible overestimation in simulations due to deficient convective parameterizations or boundary-layer dynamics. Persistent uncertainties arise from unresolved processes, including indirect effects on droplet number, mesoscale organization in trade-wind cumuli, and state-dependent responses where feedbacks may weaken at higher temperatures. Emergent constraints from present-day cloud climatologies reduce low-cloud feedback spread but reveal model biases, such as excessive sensitivity to EIS in CMIP6 versus CloudSat/ observations. Some analyses indicate net low-cloud feedbacks near zero or negative when conditioning on observed , potentially lowering ECS medians to 2.5–3 K, though these remain debated amid statistical challenges in disentangling feedbacks from forcing. Overall, clouds contribute the dominant term to ECS , with ongoing missions like EarthCARE aiming to refine estimates through improved vertical profiling.

Ocean Heat Uptake and Circulation Effects

The oceans absorb approximately 90-91% of the excess energy accumulated in the Earth's due to , primarily through vertical mixing and processes that transport from the surface to deeper layers. This heat uptake delays the realization of full surface warming, distinguishing the transient climate response (TCR) from (ECS) by reducing near-term atmospheric and surface increases for a given . In climate models, this effect is parameterized by an ocean heat uptake coefficient (κ), typically on the order of 0.5-1 W m⁻² K⁻¹ in coupled atmosphere-ocean general circulation models (AOGCMs), which quantifies the rate of subsurface heat storage per unit of surface warming and contributes to TCR being 20-30% lower than ECS on average across model ensembles. Observations from floats and other instruments confirm accelerating global (OHC) accumulation, with rates increasing from about 0.6 W m⁻² in the 1970s-1990s to nearly double that during 2010-2019, driven by enhanced mode water formation in subtropical gyres. Ocean circulation patterns, particularly the meridional overturning circulation (MOC) such as the Atlantic Meridional Overturning Circulation (AMOC), modulate the spatial distribution and efficiency of heat uptake, with stronger pre-industrial deep overturning enhancing transient sequestration of heat in the abyssal ocean and thereby dampening surface responses. In AOGCM simulations, disruptions to these circulations under warming—such as AMOC slowdown—can alter heat uptake efficacy, potentially reducing it by localizing uptake in inefficient high-latitude regions and amplifying mid-latitude warming patterns that influence cloud feedbacks. For instance, model experiments show that enhanced Southern Ocean stratification weakens overturning, leading to shallower heat storage and a transient sensitivity closer to ECS values, though the net global effect remains a buffering of surface trends. Empirical reconstructions from the industrial era indicate that historical uptake efficiency has increased since 1970, consistent with observed OHC trends, but discrepancies between models and observations persist, with some studies attributing lower inferred TCR (around 1.6-1.9 K per CO₂ doubling) to underestimated historical forcings or unmodeled circulation variability. Uncertainties in ocean heat uptake arise from incomplete observational coverage of the deep ocean (>2000 m), where heat storage accounts for about 20-30% of total OHC changes, and from model biases in representing eddy-driven and timescales, which can span centuries. Circulation-driven effects, such as shifts in subtropical gyre intensities, have contributed to recent accelerations in uptake, but projections vary widely: some CMIP6 models predict declining under high-emissions scenarios due to surface freshening and stability changes, potentially narrowing the TCR-ECS gap over centuries. These dynamics underscore that while ocean uptake provides a temporary of surface warming, long-term equilibration depends on the of global overturning cells, with paleoclimate analogs suggesting potential thresholds where circulation collapses could reverse buffering effects.

Estimation Methods

Instrumental Observations from the Industrial Era

Observational records of global mean surface temperature (GMST) since the mid-19th century enable constraints on climate sensitivity through global energy budget analyses, which equate realized warming to net minus ocean heat uptake, divided by the magnitude of the feedback parameter. These methods typically employ instrumental temperature datasets such as HadCRUT4 or HadCRUT5, alongside estimates of effective (ERF) from greenhouse gases, aerosols, and other factors, with ERF uncertainty dominated by aerosol indirect effects. From 1880–1899 to 1997–2016, GMST rose by 0.83 K (HadCRUT4v5) to 0.89 K (kriging-infilling variant), against a net ERF of approximately 1.8–2.0 W/m². Such data primarily constrain the transient climate response (TCR), the expected GMST change midway through a CO₂ doubling under gradual forcing increase. Regression-based and Bayesian energy budget approaches, incorporating historical forcing time series and accounting for internal variability, yield TCR medians of 1.2–1.4 K, with 5–95% ranges of 0.9–1.9 K. Satellite-era subsets (1979–2006 or later) produce similar TCR estimates of 1.3–2.3 K (5–95% range), sensitive to the treatment of unforced variability and multidecadal oscillations like the Atlantic Multidecadal Variability. Inferring equilibrium climate sensitivity (ECS) from industrial-era data requires extrapolating s beyond transient conditions, often via diagnosed parameters from top-of-atmosphere (TOA) flux anomalies regressed against . Clouds and the Earth's System (CERES) satellite observations since March 2000 measure Earth's energy imbalance (EEI) at 0.85 ± 0.15 W/m² (2005–2019 average), aligning with diagnosed ocean heat uptake rates of 0.6–0.9 W/m². Energy models assuming constant s derive ECS medians of 1.5–1.7 K, with 5–95% ranges of 1.0–2.7 K, implying parameters of -1.7 to -2.0 W/m²/K—more stabilizing than in many CMIP5/6 simulations. Disagreements arise from potential time-variation in feedbacks, particularly cloud responses to spatial () patterns. Analyses incorporating "pattern effects"—where warming in mid-to-high latitudes amplifies positive cloud feedbacks—yield higher ECS inferences of 3.0–5.3 from post-1950 data, attributing recent decades' weaker observed feedbacks to transient SST gradients rather than low sensitivity. Conversely, studies emphasizing full historical forcings and EEI trends find no robust evidence for weakening feedbacks, supporting ECS values below 3 and highlighting inconsistencies with high-sensitivity models. Key uncertainties include incomplete GMST coverage before 1950 (potentially biasing trends low by 0.1–0.2 K), aerosol forcing estimates (net cooling of -0.9 ± 0.5 W/m² circa 2010, with possible overestimation), and , which dilutes surface warming by 10–20% relative to . These instrumental constraints generally favor TCR and ECS at the lower half of multimethod assessments, challenging model ensembles that simulate insufficient historical warming under observed forcings.

Paleoclimate Proxy Reconstructions

Paleoclimate reconstructions estimate sensitivity (ECS) by quantifying past changes (ΔT) relative to perturbations (ΔF), primarily from CO₂ variations, using geological and biological indicators. such as δ¹⁸O in and ice cores for , and isotopes or stomatal for CO₂ concentrations, enable of forcing-temperature relationships over to millions of years. These methods assume that past states approximate after sufficient time for feedbacks to stabilize, though challenges arise from incomplete coverage, dating uncertainties, and unquantified forcings like aerosols or . The Last Glacial Maximum (LGM, ~21,000 years ago) serves as a key cold-climate benchmark, with atmospheric CO₂ at ~190 ppm versus preindustrial ~280 ppm, yielding a radiative forcing of approximately -3.2 to -3.5 W/m², and global cooling of 4–6 °C inferred from multi-proxy syntheses. Initial LGM-based ECS estimates ranged 2–6 °C, but adjustments for spatial patterns in sea surface temperatures and ice-albedo feedbacks—known as "pattern effects"—reduce the median to 2.4 °C (66% confidence interval 1.7–3.5 °C; 5–95% range 1.4–5.0 °C). When combined with other evidence, LGM constraints support a broader ECS median of 2.9 °C (5–95% range 2.1–4.1 °C). Warmer paleoclimate intervals, such as the mid-Pliocene (~3.2 million years ago), feature CO₂ levels of 350–450 ppm and leading to global temperatures 2–4 °C above preindustrial, implying ECS values up to 4–5 °C in some studies. However, incorporating pattern effects from reconstructed warming distributions—such as reduced meridional gradients—constrains Pliocene-inferred ECS toward lower modern values, challenging assumptions of state-independent high sensitivity. The Paleocene-Eocene Thermal Maximum (PETM, ~56 million years ago) involved rapid ~5–8 °C warming from massive carbon releases (~1,000–2,000 GtC), with "bulk" ECS estimates around 4.5 °C (66% confidence 2.4–7.9 °C), though its transient nature and potential non-CO₂ drivers limit direct applicability to equilibrium sensitivity. Eocene warm periods similarly suggest state-dependent ECS increases to ~4–6 °C, attributed to low-cloud feedbacks in ice-free conditions, but model-proxy discrepancies highlight uncertainties in deep-time proxies and vegetation-albedo interactions. Critiques emphasize that paleoclimate ECS derivations often overlook differential of past forcings versus idealized CO₂ doubling, potentially estimates high by neglecting regional pattern influences on global feedbacks. Proxy data sparsity in the and , combined with assumptions of modern-like lapse-rate feedbacks, further contributes to wide ranges (1.5–6 °C across studies), underscoring paleo evidence's value in bounding but not precisely pinning ECS without integrated modeling. Recent syntheses integrating multiple paleo intervals with process understanding favor ECS medians of 2.5–3.5 °C, though academic consensus leans toward higher values influenced by model priors favoring positive cloud feedbacks.

Climate Model Simulations and Ensembles

Climate models derive estimates of climate sensitivity (ECS) primarily through idealized experiments, such as abrupt quadrupling of atmospheric CO₂ concentrations in coupled atmosphere- general circulation models (AOGCMs), where the parameter is inferred by regressing top-of-atmosphere (TOA) radiative imbalance against change, often using the Gregory method. Transient climate response (TCR) is calculated from simulations with 1% per year compounded CO₂ increase, capturing the temperature rise at the point of CO₂ doubling under transient forcing, which accounts for heat uptake delaying full equilibration. These metrics are standardized across models to facilitate , though variations in , such as slab- approximations for faster ECS computation, can introduce discrepancies exceeding 1°C in individual model estimates. Multi-model ensembles, coordinated through phases of the (CMIP), aggregate simulations from independent modeling groups to represent uncertainty in ECS and TCR. The CMIP6 ensemble, comprising outputs from 38 AOGCMs as of 2022, yields ECS values spanning 1.83°C to 5.67°C, with a multi-model mean around 3.9°C—higher than the CMIP5 mean of approximately 3.2°C—attributed to strengthened and feedbacks in updated parameterizations. This shift includes a disproportionate increase in "hot" models with ECS above 4°C, raising questions about whether it reflects improved physics or unmitigated systematic biases in representations. TCR in CMIP6 models averages about 2.0°C, with a narrower range than ECS due to the transient nature mitigating deep-ocean feedbacks. Ensembles assume equal weighting of models to span plausible outcomes, but this approach presumes and realism, potentially propagating shared errors from incomplete understanding, such as low-level responses to warming. Recent advancements include performance-based weighting schemes, such as Bayesian model averaging that downweights high-ECS models to align better with assessed ranges from IPCC, reducing projected warming uncertainty. Large initial-condition ensembles from select models further isolate forced signal from internal variability, aiding attribution, though they do not resolve inter-model spread. Critiques highlight that CMIP6's elevated sensitivities contribute to overprojection of historical warming rates when unweighted, prompting hybrid observational-model constraints in applications.

Emergent Constraints from Observations

Emergent constraints on climate (ECS) utilize statistical covariances identified across multi-model ensembles between ECS and measurable aspects of the contemporary climate, such as radiative fluxes or profiles, to narrow ECS ranges when calibrated against satellite or instrumental observations. This method posits that inter-model spread in observables imperfectly simulates reality but preserves robust links to , enabling of ECS without full simulations. The approach emerged in the early 2000s and has been applied to datasets from (Clouds and the Earth's System) and ERA reanalyses, assuming model ensembles adequately sample physical processes despite known systematic biases. Key implementations include relationships between ECS and the net radiative imbalance at the top of the atmosphere or low-cloud feedbacks derived from present-day spatial patterns. For instance, an analysis of global energy budget observations from (2000–2016) yielded an ECS median of 2.8 °C, with 66% 2.1–4.1 °C, by regressing model-simulated imbalances against observed values. Constraints from tropical mid-tropospheric amplification and extratropical shortwave cloud feedbacks in CMIP5 and CMIP6 ensembles have similarly produced medians around 2.5–3.0 °C, often excluding values above 4 °C with high confidence when observational uncertainties are propagated. A 2020 evaluation of 11 such constraints across CMIP5 confirmed their potential to tighten ECS ranges but highlighted variability in constraint strength, with stronger ones favoring ECS < 3 °C based on consistency with historical warming rates. Updated emergent constraints incorporating post-2020 observations, including accelerated warming in 2023–2024, have further emphasized lower sensitivities. One 2025 study regressing CMIP6 model TCR-observable relationships against revised sea surface temperature and energy imbalance data estimated TCR at 1.81 K (90% interval 1.35–2.23 K), implying ECS likely below 3 °C and inconsistent with high-sensitivity CMIP6 subsets that overestimate recent trends. These results align with critiques that high-ECS models (>4 °C) poorly match observed shortwave cloud responses in the or historical forcing responses. Despite utility, emergent constraints face methodological challenges, including reliance on models that may not linearly extrapolate covariances or adequately represent unforced variability. A 2021 review noted risks of to flawed ensembles, where spurious correlations arise from shared parametrizations rather than physics, potentially biasing toward model-mean ECS (~3.1 °C in CMIP6). Recent critiques advocate out-of-sample testing—holding subsets of data or models aside for validation—to quantify reliability, revealing that only physically motivated constraints (e.g., lapse-rate feedbacks) robustly reduce , while others fail cross-ensemble checks. Evaluations of 19 proposed constraints found approximately half lacked predictive against metrics like paleoclimate proxies, underscoring the need for causal understanding over statistical alone. Overall, validated emergent constraints from observations systematically support ECS medians of 2.0–3.0 °C, providing empirical bounds tighter than process-model priors and highlighting tensions with unconstrained high-sensitivity projections.

Historical Evolution of Estimates

Pre-Modern Calculations (Arrhenius to Mid-20th Century)

provided the earliest quantitative estimate of climate sensitivity in 1896, calculating that doubling atmospheric CO2 concentration would raise Earth's average surface temperature by 5 to 6 °C. His approach employed a rudimentary energy balance model, manually computing radiative fluxes through atmospheric layers while incorporating as an amplifying , which roughly doubled the direct effect of CO2. derived this from spectroscopic data and assumed logarithmic dependence of CO2 forcing on concentration, projecting centuries-scale changes from emissions. Subsequent early-20th-century assessments largely dismissed Arrhenius's findings. Knut Ångström's 1900 experiments indicated saturation in CO2's absorption bands, implying additional CO2 would contribute minimally to downward at the surface—effectively a near-zero beyond pre-industrial levels. This interpretation, echoed in laboratory studies by researchers like E.O. Hulburt in the 1920s, prevailed in meteorological circles, attributing potential warming to solar variability or internal dynamics rather than greenhouse gases. Guy Stewart Callendar revived CO2-driven warming theory in 1938 by analyzing records from 147 stations, documenting a 0.005 °C per year global land temperature rise since 1880 alongside a ~6% CO2 increase from combustion. He attributed roughly half (~0.003 °C per year) to anthropogenic CO2, yielding an implied equilibrium sensitivity of ~2 °C for CO2 doubling under assumptions of limited water vapor feedback and uniform radiative response across latitudes. Callendar's estimate drew on updated absorption data but emphasized empirical correlation over theoretical modeling, forecasting 0.4 °C further warming by 2100 at prevailing emission rates. These pre-modern calculations spanned 2 to 6 °C for doubling 2, reflecting debates over and feedbacks amid sparse global data and computational limits. By the , renewed spectroscopic work by Gilbert Plass confirmed non-saturated 2 effects in the upper atmosphere, bridging toward model-based refinements, though hand-calculated sensitivities remained anchored to Arrhenius's high-end and Callendar's observational low-end paradigms.

Charney Report and Early Model Era (1970s-1990s)

In 1979, the U.S. published "Carbon Dioxide and Climate: A Scientific Assessment," chaired by Jule Charney, following a workshop at . The report synthesized early radiative-convective models and nascent general circulation models (GCMs) to evaluate CO₂ doubling's equilibrium sensitivity (ECS), defined as the long-term response after feedbacks stabilize. It concluded an ECS range of 1.5–4.5°C, with 3°C as the most probable value, attributing the wide bounds chiefly to unresolved cloud feedbacks, which could either amplify or dampen warming depending on altitude and type changes. and surface feedbacks were deemed positive but quantifiable, while effects were smaller; the assessment emphasized that no GCMs of the era could simulate clouds realistically enough to constrain the range further. The 1970s marked the transition to three-dimensional GCMs, building on one-dimensional models from the 1960s. Syukuro Manabe's GFDL model simulations, published in 1975, estimated ECS at approximately 2.3°C by incorporating fixed dynamical heating and simplified cloud representations, validating positive but highlighting limitations in ocean-atmosphere . These early GCMs, run on limited resources, typically resolved horizontal grids of hundreds of kilometers and used crude parameterizations for and clouds, yielding ECS values clustering around 2–3°C across institutions like NCAR and UCLA, yet consistently within Charney's bounds. Empirical constraints from paleoclimate analogies, such as temperature differences, were invoked to support the range but added little precision due to proxy data uncertainties. By the 1980s, enhanced models incorporated interactive oceans and seasonal cycles, as in James Hansen's GISS model series (e.g., 1983–1988 experiments), which implied ECS values up to 4.2°C under transient forcing scenarios projecting 0.5–1°C warming by 2000 relative to 1951–1980 baselines. However, inter-model comparisons, including those from the WMO/UNEP workshop, revealed ECS spreads of 1.6–5.6°C, driven by divergent cloud and parameterizations; no emerged to narrow Charney's interval despite doubled grid resolutions. calculations refined the direct CO₂ effect to about 4 W/m² for doubling, but uncertainties—estimated at 1–2 W/m²/m²K—sustained the range. The IPCC's First Assessment Report (1990) integrated these model outputs with emerging satellite and surface observations, endorsing an ECS best estimate of 2.5°C (range 1.5–4.5°C), noting that GCM ensembles averaged around 3°C but exhibited high variability from unresolved processes like tropical . The Second Assessment Report (1995) retained the identical range, with a shifted best estimate near 2.5°C, underscoring that computational advances had not resolved core feedback ambiguities by the decade's end.

IPCC Assessments and Progressive Refinements (2000s-Present)

The Intergovernmental Panel on Climate Change's Third Assessment Report (), released in 2001, assessed equilibrium climate sensitivity (ECS) as likely ranging from 1.5°C to 4.5°C, consistent with prior evaluations and primarily informed by early general circulation models (GCMs) and paleoclimate proxies. This broad range reflected substantial uncertainties in feedbacks, particularly clouds and , with no central estimate specified due to divergent lines of evidence. The Fourth Assessment Report (AR4) in retained the same likely range of 1.5°C to 4.5°C for ECS, incorporating refined GCM simulations and instrumental observations but noting persistent challenges in constraining low-end values from models alone. Observational constraints began to emphasize upper bounds, yet the assessment highlighted that multiple methods—ranging from energy balance calculations to paleoclimate reconstructions—yielded overlapping but wide uncertainties, with a subjective best estimate around 3°C emerging from some syntheses. In the Fifth Assessment Report (AR5) of 2013, the likely ECS range remained 1.5°C to 4.5°C, deemed extremely unlikely below 1°C and very unlikely above 6°C, drawing on expanded datasets including ensemble simulations and updated paleoclimate analyses. The report noted improved agreement among methods but avoided a single best estimate, citing tensions between model-based estimates (often higher) and instrumental constraints suggesting potentially lower values; cloud feedback uncertainties continued to dominate the spread. The Sixth Assessment Report (AR6) in 2021 marked a refinement, assessing ECS as likely 2.5°C to 4.0°C (very likely 2.0°C to 5.0°C), with a best estimate of 3.0°C derived from an integrated evaluation of instrumental records, paleoclimate evidence, process understanding, and feedback assessments. This narrowing, particularly the elevated lower bound, stemmed from multiple independent lines converging to rule out ECS below 2.5°C, including energy budget analyses of historical warming and revised paleoclimate constraints on tropical feedbacks, despite CMIP6 models exhibiting a higher median ECS around 3.9°C. AR6 emphasized that while models inform processes, observational and emergent constraints tempered high-end projections, reflecting progressive integration of empirical data over model tuning. As of 2025, no subsequent full assessment has superseded AR6, though ongoing research continues to probe remaining uncertainties in ocean heat uptake and effects.

Uncertainties and Controversies

Methodological Disagreements: Paleo vs. Models

Paleoclimate proxy reconstructions and climate model simulations represent two primary methodologies for estimating equilibrium climate sensitivity (ECS), often yielding divergent results that highlight fundamental methodological tensions. Proxy-based approaches draw on empirical evidence from past climate states, such as the (, approximately 21,000 years ago), where global mean surface temperature (GMST) cooling of about 5–6°C occurred under reduced CO₂ concentrations (~180 ppm versus preindustrial ~280 ppm), implying ECS values typically in the range of 2–3°C after accounting for non-CO₂ forcings like ice sheets and aerosols. In contrast, general circulation models (GCMs) in ensembles like CMIP6 simulate ECS with a broader distribution, averaging around 3°C but extending to over 5°C in some cases, driven by parameterized feedbacks such as cloud responses that amplify warming. A core disagreement arises from models' performance against paleoclimate benchmarks: high-ECS simulations (>5°C) frequently overestimate GMST changes in periods like the , mid-Pliocene Warm Period (~3.2 million years ago), and early Eocene Climatic Optimum (~50 million years ago), producing cooling or warming magnitudes outside proxy uncertainty ranges, whereas low-ECS models (<2°C) underestimate them. For instance, PMIP4-CMIP6 experiments reveal that models with elevated sensitivity predict excessive (up to 8–10°C in outliers), exceeding proxy-derived estimates of ~4.5–6°C, suggesting deficiencies in model representations of stabilizing feedbacks like low-cloud adjustments or ocean heat uptake. Recent analyses incorporating spatial pattern effects—such as enhanced extratropical cooling from ice sheets—further tighten paleo constraints to a ECS of 2.4°C (66% : 1.7–3.5°C) from data alone, or 2.9°C when combined with other lines of evidence, narrower and shifted lower than CMIP6 medians. Methodological critiques of paleo approaches emphasize uncertainties in proxy calibrations (e.g., seasonal or elevational biases in reconstructions) and forcing quantifications, including and dust effects that may amplify or dampen inferred sensitivity, potentially leading to underestimation if slow feedbacks like ice-sheet dynamics are not fully isolated. Models, while physics-based, face challenges in sub-grid parameterizations that cannot be directly observed, with paleo mismatches indicating systematic overestimation of positive feedbacks in high-sensitivity variants; proponents argue emergent constraints from multiple paleo periods better falsify implausible models than instrumental records alone, which are confounded by transient disequilibria. Additionally, some reconstructions suggest state-dependent sensitivity, with lower values in colder glacial climates due to reduced or feedbacks, implying paleo data may not directly translate to warmer future conditions without adjustments. These tensions underscore ongoing debates, where paleo prioritizes historical calibration but risks incomplete feedback inclusion, while models offer process-level insight yet require paleo validation to mitigate tuning artifacts.

Observational Constraints Challenging High Sensitivity

Energy budget methods applied to instrumental records constrain climate sensitivity (ECS) by inferring the climate feedback parameter from observed rise, estimated radiative forcings, and changes in . The approach models the top-of-atmosphere energy imbalance as ΔN ≈ λ ΔT + κ, where λ is the feedback parameter (in W m⁻² K⁻¹), ΔT is surface change, and κ represents uptake by the deep . ECS is then calculated as S ≈ 3.7 / -λ, with 3.7 W m² approximating the forcing from doubled CO₂. These methods yield ECS estimates lower than those from comprehensive climate models, suggesting positive feedbacks weaker than simulated. A 2018 study by incorporated post-2000 refinements, ARGO-era ocean heat content measurements through , and kriged surface datasets, producing a median ECS of 1.66 (5–95% range: 1.15–2.7 ). This upper range bound lies below the CMIP5 multimodel mean ECS of 3.2 and contrasts with the IPCC AR5 likely range of 1.5–4.5 , implying many models overestimate relative to 19th– observations. Adjusting for potential time-varying feedbacks raised the median slightly to 1.76 (1.2–3.1 range), still excluding high-end values above 3 at the 95th percentile. Earlier work by and in 2015, using IPCC AR5 forcing and heat uptake data over 1750–2011, similarly derived ECS medians near 2 , with ranges emphasizing tighter bounds than prior assessments due to reduced uncertainties in volcanic and forcings. These estimates challenge high ECS by indicating that observed warming aligns better with subdued and feedbacks than with amplified ones dominant in general circulation models (GCMs). Satellite-derived top-of-atmosphere (TOA) radiation data from instruments like ERBE (1985–1999) and (2000–present) offer independent checks via regressions of radiative flux anomalies against variations during El Niño/La Niña events, proxying global responses. Such analyses often yield effective feedback parameters around -1.5 to -2.0 W m⁻² K⁻¹, corresponding to ECS of 1.8–2.5 K, inconsistent with strong positive feedbacks required for ECS exceeding 3 K. For instance, observations through 2019 show radiative damping during warm phases that limits amplification, though data gaps and internal variability introduce uncertainties. These constraints highlight tensions with high-sensitivity GCM ensembles, where simulated feedbacks—particularly low-cloud reductions—exceed observational inferences, potentially inflating projected warming. Uncertainties persist in aerosol indirect effects and the degree to which transient observations capture responses, yet the convergence on ECS below 3 K from multiple lines underscores challenges to assuming model-high sensitivities for policy-relevant projections.

Debates on Low Sensitivity: Empirical Evidence and Critiques

Proponents of low equilibrium climate sensitivity (ECS) emphasize empirical constraints derived from the instrumental record, particularly through energy budget analyses that balance observed radiative forcing, surface temperature changes, and ocean heat uptake. These approaches yield ECS estimates centering below 2°C, arguing that direct measurements from the satellite era and historical data provide the most reliable gauge of climate response without reliance on uncertain paleoclimate proxies or model-dependent feedbacks. For instance, a 2018 study by Nic Lewis and Judith Curry integrated updated radiative forcing estimates (including reduced aerosol cooling effects) and ocean heat content data from 1850 to 2011, employing Bayesian inference to derive a median ECS of 1.66°C, with a 17–83% confidence interval of 1.1–2.7°C, and a transient climate response (TCR) median of 1.20°C (0.9–1.7°C). This analysis highlighted that post-1950 observations, where data quality improves, particularly constrain sensitivity downward, as the realized warming aligns closely with forcing after accounting for heat sequestration in the deep ocean. Supporting evidence includes satellite-based measurements of Earth's energy imbalance (EEI), which show trends consistent with modest amplification beyond the no-feedback response (~1.2°C ECS from blackbody physics alone). Researchers like have argued that satellite data from 2000–2018 indicate an EEI of approximately 0.8 W/m², implying limited positive feedbacks when calibrated against historical forcing, further supporting ECS values around 1.5–2.5°C. Independent assessments using similar observational inputs, such as those incorporating adjusted historical datasets, have corroborated medians near 2°C, attributing discrepancies with higher model projections to overestimated cloud or feedbacks in general circulation models (GCMs). These empirical lines contend that modern observations capture the primary CO2-driven response amid dominant forcings, rendering high-sensitivity claims reliant on extrapolated feedbacks less credible. Critics of low-sensitivity estimates, often aligned with comprehensive assessments incorporating multiple evidence streams, argue that constraints alone undervalue long-term feedbacks emergent over paleoclimate timescales or process-level understanding. A review by et al., synthesizing , paleoclimate, and feedback process data, concluded that ECS below 2.3°C is implausible (5–95% range 2.3–4.5°C, likely 2.6–4.1°C), positing that low-end results neglect amplifying effects like lapse-rate changes or low-cloud reductions observed in targeted studies. They further claimed that recent EEI acceleration, derived from and ocean observations, favors higher ECS, as low-sensitivity scenarios fail to replicate the observed ~0.2 W/m² per decade increase since the 2000s. However, has critiqued this synthesis for overweighting uncertain paleoclimate reconstructions (prone to calibration errors) and subjective emergent constraints, reanalyzing the combined evidence in 2022 to yield a ECS of 2.16°C (5–95% range 1.2–4.0°C), asserting that objective Bayesian weighting prioritizes reliability over model-correlated proxies. The debate persists amid methodological disputes, with low-sensitivity advocates highlighting potential biases in high-end estimates from academic institutions favoring alarm-oriented funding, while critics invoke consistency across IPCC assessments (e.g., AR6's likely 2.5–4.0°C) as validation. Recent 2025 analyses, including responses to claims of further narrowing, reaffirm that instrumental-derived lows remain viable, particularly as updated forcing datasets continue to erode support for strong masking of warming. Empirical critiques of low ECS often hinge on assuming full equilibration of feedbacks within short observational windows, yet of forcing-response in the satellite era underscores the robustness of subdued sensitivity against expectations from unchecked GCM ensembles.

Recent Developments (2020-2025): Record Warmth and Revised Bounds

The years 2023 and 2024 recorded the highest global mean surface temperatures in the instrumental , with 2023 exceeding the previous record by approximately 0.17°C and 2024 surpassing it by an additional 0.11°C, reaching about 1.55°C above pre-industrial levels (1850–1900 baseline). This surge coincided with a strong El Niño event, reduced stratospheric loading from shipping regulations, and ongoing accumulation, amplifying short-term variability atop the multi-decadal warming trend. Observational datasets from , NOAA, and the confirmed these anomalies, with every month from June 2023 to August 2024 ranking among the warmest on record. The IPCC's Sixth Assessment Report (AR6), published in , revised the likely range for equilibrium climate sensitivity (ECS) to 2.5–4.0°C per CO₂ doubling, raising the lower bound from the AR5's 1.5–4.5°C range based on integrated from process models, paleoclimate proxies, and emergent constraints. This adjustment reflected greater confidence in positive feedbacks like amplification and reduced low-cloud cover, while dismissing ECS below 2°C as inconsistent with multiple lines of . Subsequent analyses of satellite-derived imbalance trends (2001–2023) have reinforced this, showing that models with ECS below 2.5°C fail to reproduce the observed strengthening imbalance—particularly the positive shortwave and negative longwave components—implying insufficient simulated warming under rising greenhouse forcing. Post-AR6 studies have intensified debates on whether 2023–2024 warmth further constrains ECS bounds, with emergent approaches using recent observations to test model projections against transient responses. One 2025 analysis argues that record temperatures align with transient climate response (TCR) values around 1.8–2.2°C, but highlight uncertainties in forcing attribution that prevent definitively ruling out lower ECS medians (e.g., 2.0–3.0°C) from records. Critiques of high-end estimates, drawing on updated forcing and ocean heat uptake data, maintain that observed warming rates since 1970 are better matched by ECS near 2°C, challenging AR6's elevated lower bound as overly reliant on model-derived feedbacks prone to structural biases. These developments underscore ongoing tensions between energy budget favoring modest and feedback-dominant assessments supporting higher values, with no revision to AR6 bounds as of 2025.

References

  1. [1]
    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 ...
  2. [2]
    Emergent constraints on equilibrium climate sensitivity in CMIP5 - ESD
    Dec 21, 2020 · An important metric for temperature projections is the equilibrium climate sensitivity (ECS), which is defined as the global mean surface air ...
  3. [3]
    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 ...
  4. [4]
    Determination of a lower bound on Earth's climate sensitivity | Tellus B
    Earth's so-called equilibrium climate sensitivity, the change in global mean near-surface air temperature GMST, Ts, that would ultimately be attained in ...
  5. [5]
    Observed trend in Earth energy imbalance may provide a ... - Science
    Jun 12, 2025 · This Earth energy imbalance has strengthened over the period 2001 to 2023 with satellite data. Here, we show that low climate sensitivity models ...
  6. [6]
    [PDF] AR6 FAQ 7.1
    The equilibrium climate sensitivity (ECS) is defined as the long-term global warming caused by a doubling of carbon dioxide above its pre-industrial ...
  7. [7]
    [PDF] Climate sensitivity in the Anthropocene
    In its most basic form, it refers to the equilibrium change in global annual mean surface temperature that occurs in response to a radiative forcing, or ...
  8. [8]
    Transient and Equilibrium Climate Sensitivity
    Equilibrium sensitivities in global climate models typically range from 2 to 5K, while the transient climate responses are smaller, in range of 1.0-2.5 K.
  9. [9]
    8.6 Climate Sensitivity and Feedbacks - AR4 WGI Chapter 8
    Climate sensitivity is a metric used to characterise the response of the global climate system to a given forcing.
  10. [10]
    An Assessment of Earth's Climate Sensitivity Using Multiple Lines of ...
    Jul 22, 2020 · Earth's global “climate sensitivity” is a fundamental quantitative measure of the susceptibility of Earth's climate to human influence. A ...
  11. [11]
    Climate sensitivity from radiative-convective equilibrium
    Sep 1, 2023 · Simple models for Earth's climate sensitivity (ie its temperature response to radiative forcing) are developed by combining the time-tested idealization of one ...
  12. [12]
    Equilibrium- and Transient-State Dependencies of Climate Sensitivity
    Because the radiative feedbacks tend to amplify more the warming in equilibrium than in transition, the nonlinear equilibrium climate sensitivity (ECSϵ) ...
  13. [13]
    [PDF] The Earth's Energy Budget, Climate Feedbacks and Climate Sensitivity
    Equilibrium Climate Sensitivity (ºC). FAQ 7.3: Equilibrium climate sensitivity and future warming. Equilibrium climate sensitivity measures how climate models ...
  14. [14]
    Climate Change Indicators: Climate Forcing | US EPA
    Climate forcing, or radiative forcing, is the heating effect from greenhouse gases, measured in watts per square meter, representing the energy imbalance in ...
  15. [15]
    Climate Forcing | NOAA Climate.gov
    Nov 21, 2014 · The difference between incoming and outgoing radiation is known as a planet's radiative forcing (RF). In the same way as applying a pushing ...
  16. [16]
    [PDF] Calculating Earth's Temperature Assuming No Atmosphere
    Feb 26, 2018 · Without an atmosphere, Earth's temperature would be about -18C (254.4K), making it not likely that life could exist.
  17. [17]
    How Well do We Understand the Planck Feedback? - AGU Journals
    Jul 17, 2023 · The comparable radiative response in climate models, widely called the “Planck feedback,” averages −3.3 W m−2 K−1. This difference of 0.5 W m−2 ...Introduction · Sensitivity Tests · Global Averaging · Discussion and Conclusions
  18. [18]
    The CO2 problem in 6 easy steps - RealClimate
    Aug 6, 2007 · Thus the RF for a doubling of CO2 is likely 3.7±0.4 W/m2 – the same order of magnitude as an increase of solar forcing by 2%. There are a couple ...
  19. [19]
    Chapter 7: The Earth's Energy Budget, Climate Feedbacks, and ...
    The equilibrium climate sensitivity, ECS (units: °C), is defined as the equilibrium value of ΔT in response to a sustained doubling of atmospheric CO 2 ...
  20. [20]
    [PDF] Climate change, equilibrium climate sensitivity and feedbacks - ACP
    A useful single-number proxy for how sensitive the planet is to forcing by CO2 is given by the equilibrium climate sensitivity (ECS, °C). ECS is defined as ...
  21. [21]
    Equilibrium Climate Sensitivity Estimated by Equilibrating Climate ...
    Nov 19, 2019 · The equilibrium climate sensitivity (ECS) is defined as the global- and time-mean, surface air warming once radiative equilibrium is reached ...
  22. [22]
    Equilibrium- and Transient-State Dependencies of Climate Sensitivity
    This study assesses the importance of the different types of state dependencies of the radiative feedbacks for constraining climate projections from the ...
  23. [23]
    Context for interpreting equilibrium climate sensitivity and transient ...
    Jun 24, 2020 · 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 ...Missing: peer- | Show results with:peer-
  24. [24]
    Emergent constraints on transient climate response (TCR ... - ESD
    Aug 17, 2020 · Transient climate response (TCR) is the metric of temperature sensitivity that is most relevant to warming in the next few decades and ...Missing: interrelations | Show results with:interrelations
  25. [25]
    Revisiting a Constraint on Equilibrium Climate Sensitivity From a ...
    Oct 26, 2023 · Our central estimates of equilibrium climate sensitivity are 2.5–2.7 K, consistent with prior results and the IPCC's consensus estimate. 1 ...
  26. [26]
    Opinion: Can uncertainty in climate sensitivity be narrowed further?
    Feb 29, 2024 · The distributions could nonetheless be narrowed in the future, particularly through better understanding of certain climate processes and paleoclimate proxies.
  27. [27]
    Equilibrium climate sensitivity above 5 °C plausible due to state ...
    Oct 26, 2020 · The equilibrium climate sensitivity of Earth is defined as the global mean surface air temperature increase that follows a doubling of atmospheric carbon ...Missing: definition | Show results with:definition
  28. [28]
    Guest post: Ice-age analysis suggests worst-case global warming is ...
    May 1, 2024 · Our findings give us more confidence that very high climate sensitivity is unlikely. They also support the IPCC's central estimate for climate ...
  29. [29]
    Biased Estimates of Equilibrium Climate Sensitivity and Transient ...
    Dec 9, 2021 · Equilibrium climate sensitivity (ECS) and transient climate response (TCR) are two fundamental metrics for evaluating climate change projections ...
  30. [30]
    Transient Climate Response → Term
    Sep 30, 2025 · The Transient Climate Response (TCR) is formally defined as the change in global mean surface air temperature averaged over a 20-year period, ...
  31. [31]
    CMIP6 GCM Validation Based on ECS and TCR Ranking for 21st ...
    However, their transient climate response (TCR) ranges from 1.2 to 2.8 °C, whereas their equilibrium climate sensitivity (ECS) ranges from 1.8 to 5.7 °C, ...<|separator|>
  32. [32]
    sensitivity to radiative forcing time series and observational data - ESD
    Jun 21, 2018 · The transient climate response (TCR) is estimated to have a mean value of 1.4 ∘C (90 % CI 0.9–2.0 ∘C), and in our main analysis the posterior ...
  33. [33]
    Using historical temperature to constrain the climate sensitivity, the ...
    Jul 18, 2024 · Similarly, the transient climate response (TCR) measures the warming simulated in simulations with CO2 increasing at 1 % per annum (a−1) ...
  34. [34]
    Assessing the robustness and implications of econometric estimates ...
    Feb 4, 2025 · We here present a new TCR estimate of 2.17 (1.72–2.77) ∘ C (95% confidence interval), derived based on a statistical relationship between surface air ...
  35. [35]
    Estimating Transient Climate Response in a Large‐Ensemble ...
    Dec 3, 2018 · We find that TCR estimated from model simulations of the historical period can be much lower than the model's true TCR, replicating the disagreement seen ...
  36. [36]
    New study determines Earth's climate sensitivity from recent global ...
    The best estimates are a mid-range transient climate response of 1.67 degrees Celsius and an equilibrium climate sensitivity of 2.83 degrees Celsius.
  37. [37]
    Observational Constraint on the Climate Sensitivity to Atmospheric ...
    Jul 15, 2022 · In this study we have estimated the historical effective sensi- tivity from the latest historical observations and derived an observational ...<|separator|>
  38. [38]
    A tighter constraint on Earth-system sensitivity from long-term ...
    May 26, 2021 · In particular, a deep-time perspective offers insight into the “Earth-system sensitivity” (ESS)—the long-term equilibrium surface temperature ...
  39. [39]
    Earth System Sensitivity: a Feedback perspective - ESD
    Nov 25, 2021 · This manuscript explores the consequences a narrowing of uncertainty in equilibrium climate sensitivity (ECS) has for the Earth system ...
  40. [40]
    Water vapor and lapse rate feedbacks in the climate system
    Nov 30, 2021 · From this review, our best estimate of overall strength of combined water vapor + lapse rate feedbacks is 1.2 5 ± 0.1 5 W m − 2 K − 1 (the ...
  41. [41]
    Revisiting the Role of the Water Vapor and Lapse Rate Feedbacks ...
    Apr 21, 2022 · The water vapor feedback is the primary reason that the Arctic region warms more than the tropics, and the lapse rate feedback has a neutral effect on Arctic ...<|separator|>
  42. [42]
    Observational determination of albedo decrease caused by ... - PNAS
    We find that the Arctic planetary albedo has decreased from 0.52 to 0.48 between 1979 and 2011, corresponding to an additional 6.4 ± 0.9 W/m 2 of solar energy ...
  43. [43]
    Ice Sheet‐Albedo Feedback Estimated From Most Recent ...
    Aug 6, 2024 · We find the ice sheet-albedo feedback to be amplifying, increasing the total climate feedback parameter by 42% and reaching an equilibrium ...
  44. [44]
    Surface Albedo Feedback Estimates for the AR4 Climate Models in
    A technique for estimating surface albedo feedback (SAF) from standard monthly mean climate model diagnostics is applied to the 1% yr−1 carbon dioxide (CO2)- ...
  45. [45]
    Observational evidence that cloud feedback amplifies global warming
    Jul 19, 2021 · We show that global cloud feedback is dominated by the sensitivity of clouds to surface temperature and tropospheric stability. Considering ...Materials And Methods · Cloud-Radiative Anomalies · Cloud-Radiative...
  46. [46]
    Evaluating Cloud Feedback Components in Observations and Their ...
    Jan 22, 2024 · This study quantifies the contribution of individual cloud feedbacks to the total short-term cloud feedback in satellite observations over the period 2002–2014
  47. [47]
    Decomposing Cloud Radiative Feedbacks by Cloud-Top Phase Casey
    The global mean of the overall net cloud feedback ranges from 0.02 to 0.91 W m-2 K-1 and has a multi-model mean of 0.52 W m-2 K-1 (Fig. 11). The goal is to ...Missing: m2/ | Show results with:m2/
  48. [48]
    Cloud properties and their projected changes in CMIP models ... - ACP
    Feb 5, 2024 · The large spread in ECS of CMIP6 models could mainly be explained by uncertainties in the simulated net cloud feedback. The net cloud feedback ...
  49. [49]
    Observational evidence that cloud feedback amplifies global warming
    Jul 19, 2021 · This “cloud feedback” is the single most important cause of uncertainty in Equilibrium Climate Sensitivity (ECS)—the equilibrium global warming ...
  50. [50]
    Multi-objective observational constraint of tropical Atlantic and ...
    Jan 2, 2025 · Tropical marine low cloud feedback is key to the uncertainty in climate sensitivity, and it depends on the warming pattern of sea surface ...
  51. [51]
    Cloud Feedback Uncertainty in the Equatorial Pacific Across CMIP6 ...
    Oct 5, 2025 · In this study we focus on the equatorial Pacific, where CMIP6 model cloud feedback spread is notably large. Cloud radiative effects in this ...
  52. [52]
    Stratification and overturning circulation are intertwined controls on ...
    Jun 20, 2025 · The global ocean takes up over 90 % of the excess heat added to the climate system due to anthropogenic emissions, thereby buffering climate ...
  53. [53]
    [PDF] Importance of ocean heat uptake efficacy for transient climate change
    Nov 1, 2009 · Ocean heat uptake efficacy is important for the transient response to CO2 forcing, and its correlation with transient climate response is ...
  54. [54]
    The Effect of Ocean Ventilation on the Transient Climate Response ...
    Abstract. The surface warming response to carbon emissions is affected by how the ocean sequesters excess heat and carbon supplied to the climate system.
  55. [55]
    [PDF] Ocean Heat Uptake Efficiency Increase Since 1970
    TCS is defined as. TCS = F2xCO2/(−λ + κ), where F2xCO2 [W/m 2] is the radiative forcing associated with a doubling of the atmos- pheric CO2 concentration from ...
  56. [56]
    Recent acceleration in global ocean heat accumulation by mode ...
    Oct 28, 2023 · Here we analyze historical and recent observations to show that ocean heat uptake has accelerated dramatically since the 1990s, nearly doubling during 2010– ...
  57. [57]
    Marshall et al. 2015: The ocean's role in the transient ... - NASA GISS
    Apr 4, 2022 · The background ocean circulation strongly influences the large-scale patterns of ocean heat uptake and storage, such that Tanthro is governed by ...
  58. [58]
    Large‐scale ocean circulation‐cloud interactions reduce the pace of ...
    Apr 11, 2016 · Changes to the large-scale oceanic circulation are thought to slow the pace of transient climate change due, in part, to their influence on radiative feedbacks.
  59. [59]
    Comment on “The Impact of Recent Forcing and Ocean Heat Uptake ...
    The results suggest a lower value of TCR than estimates from climate model simulations. Previous studies have identified uncertainty in the historical forcings, ...
  60. [60]
    The decrease in ocean heat transport in response to global warming
    Oct 12, 2023 · The ocean is taking up additional heat but how this affects ocean circulation and heat transport is unclear. Here, using coupled model ...
  61. [61]
    Sensitivity of ocean circulation to warming during the Early Eocene ...
    Jun 11, 2024 · In our model, transient warming drives the weakening of Southern Ocean-sourced overturning circulation, strengthens Atlantic meridional water ...
  62. [62]
    The Impact of Recent Forcing and Ocean Heat Uptake Data on ...
    This paper also addresses a range of concerns that have been raised regarding using energy balance models to determine climate sensitivity: variability in ...
  63. [63]
    Past climates inform our future - Science
    Nov 6, 2020 · Paleoclimate constraints on climate sensitivity. Equilibrium climate sensitivity (ECS) has been widely adopted as a simple metric of how ...Structured Abstract · Paleoclimate Constraints On... · Climatic Aberrations
  64. [64]
    Proxy evidence for state-dependence of climate sensitivity in the ...
    Sep 7, 2020 · We find that Equilibrium Climate Sensitivity (ECS) was indeed higher during the warmest intervals of the Eocene, agreeing well with recent model simulations.
  65. [65]
    Last Glacial Maximum (LGM) climate forcing and ocean dynamical ...
    Jan 27, 2021 · The Last Glacial Maximum (LGM; ∼21 ka) has been considered to be an ideal target for estimating ECS since it represents a quasi-equilibrium ...
  66. [66]
    Last Glacial Maximum pattern effects reduce climate sensitivity ...
    Apr 17, 2024 · Accounting for LGM pattern effects yields a median modern-day ECS of 2.4°C, 66% range 1.7° to 3.5°C (1.4° to 5.0°C, 5 to 95%), from LGM evidence ...
  67. [67]
    Last Glacial Maximum pattern effects reduce climate sensitivity ...
    Jul 14, 2023 · Combining the LGM with other lines of evidence, we find a modern-day ECS of 2.9°C (2.1–4.1°C, 5–95% range), narrowing uncertainty compared to ...
  68. [68]
    [PDF] High Earth-system climate sensitivity determined from Pliocene ...
    Dec 20, 2009 · Higher nutrients, and associated growth rates, would bias early Pliocene reconstructed CO2 toward lower values and underestimate CO2slope.
  69. [69]
    (PDF) Paleoclimate pattern effects help constrain climate sensitivity ...
    Oct 7, 2025 · Recent paleoclimate reconstructions suggest high sensitivity of 5°C per CO2 doubling. However, by analyzing spatial patterns of Pliocene warming ...
  70. [70]
    Global mean surface temperature and climate sensitivity of the ... - CP
    Oct 26, 2020 · We demonstrate that “bulk” equilibrium climate sensitivity (ECS; 66 % confidence) during the latest Paleocene, PETM, and EECO is 4.5 °C (2.4 to ...
  71. [71]
    No apparent state-dependency of equilibrium climate sensitivity ...
    Jul 18, 2025 · Received 2024 Sep 18; Accepted 2025 Jul 7; Collection date 2025. ... The corresponding ECS estimates are 3.3 K (2.1–6.3 K, 95% confidence) ...<|separator|>
  72. [72]
    [PDF] An assessment of Earth's climate sensitivity using multiple lines of ...
    the climate is far out of equilibrium; section 4) or from proxy reconstructions of past climate ... paleoclimate sensitivity, past (proxy) values of climate ...<|separator|>
  73. [73]
    An Assessment of Earth's Climate Sensitivity Using Multiple Lines of ...
    Jul 22, 2020 · An issue with past studies is that different or vague definitions of ECS may have led to perceived, unphysical discrepancies in estimates of ECS ...
  74. [74]
    Standardising the "Gregory method" for calculating equilibrium ...
    Jun 6, 2025 · This method has been used to compare ECS estimates across the CMIP5 and CMIP6 ensembles and will likely be a key diagnostic for CMIP7.
  75. [75]
    [PDF] Comparison of equilibrium climate sensitivity estimates from slab ...
    Equilibrium Climate Sensitivity (ECS) estimates for a single coupled model can vary by more than 1°C (20%) depending on analysis method. • ECS estimates from ≥ ...
  76. [76]
    Advanced Testing of Low, Medium, and High ECS CMIP6 GCM ...
    Mar 14, 2022 · The equilibrium climate sensitivity (ECS) of the CMIP6 global circulation models (GCMs) varies from 1.83°C to 5.67°C. Herein, 38 GCMs are ...
  77. [77]
    Understanding equilibrium climate sensitivity changes from CMIP5 ...
    Apr 1, 2025 · Higher ECS in CMIP6 models is due to enhanced CF and albedo feedback; the spread of ECS is related to that of CF.
  78. [78]
    Efficacy of climate forcings in transient CMIP6 simulations - Frontiers
    Jun 11, 2024 · Here, we analyse transient simulations from CMIP6 experiments and show comparable results between transient and abrupt perturbation experiments.
  79. [79]
    Bayesian weighting of climate models based on climate sensitivity
    Oct 20, 2023 · In this study, we have focused on ECS and TCR to address the hot model problem. Previous studies targeted towards specialized impact studies ...
  80. [80]
    Climate model large ensembles as test beds for applied compound ...
    Nov 15, 2024 · Climate model large ensembles are a useful test bed for impact studies, in particular ones that require large samples of spatio-temporally ...
  81. [81]
    Emergent constraints on climate sensitivities | Rev. Mod. Phys.
    May 11, 2021 · This review summarizes previous published work on emergent constraints and discusses the promise and potential dangers of the approach.
  82. [82]
    Emergent constraint on equilibrium climate sensitivity from global ...
    Jan 17, 2018 · ECS is defined as the global mean warming that would occur if the atmospheric carbon dioxide (CO2) concentration were instantly doubled and the ...Missing: peer- reviewed
  83. [83]
    Evaluating Emergent Constraints on Equilibrium Climate Sensitivity in
    This study explores 19 previously proposed emergent constraints related to equilibrium climate sensitivity (ECS; the global-average equilibrium surface ...
  84. [84]
    None
    Nothing is retrieved...<|separator|>
  85. [85]
    Emergent constraints on transient climate response (TCR ... - ESD
    Aug 17, 2020 · Emergent constraints on transient climate response (TCR) and equilibrium climate sensitivity (ECS) from historical warming in CMIP5 and CMIP6 ...
  86. [86]
    Opinion: Why all emergent constraints are wrong but some are useful
    Feb 21, 2025 · We argue for quantitative approaches in which each suggested constraining relationship can be evaluated comprehensively based on out-of-sample test data.
  87. [87]
    Testing the assumptions in emergent constraints - ESD
    Jul 11, 2024 · The basic idea of emergent constraints is to identify an observable of the climate x that varies significantly across a climate model ensemble ...
  88. [88]
    from the pioneering work of Arrhenius and Callendar to today's Earth ...
    The result: Arrhenius concluded that a doubling or halving of atmospheric CO2 (now known as the climate sensitivity) would warm or cool the Earth by 5–6 °C.
  89. [89]
    The Carbon Dioxide Greenhouse Effect - American Institute of Physics
    In 1896 Arrhenius completed a laborious numerical computation which suggested that cutting the amount of CO2 in the atmosphere by half could lower the ...
  90. [90]
    [PDF] On the Influence of Carbonic Acid in the Air upon the Temperature of ...
    Arrhenius's paper is the first to quantify the contribution of carbon dioxide to the greenhouse effect (Sections I-IV) and to speculate about.Missing: doubling | Show results with:doubling
  91. [91]
    The artificial production of carbon dioxide and its influence on ...
    The increase in mean temperature, due to the artificial production of carbon dioxide, is estimated to be at the rate of 0.003°C. per year at the present time.
  92. [92]
    [PDF] Carbon Dioxide and Climate: A Scientific Assessment
    Carbon Dioxide and Climate: A Scientific Assessment. Report of an Ad Hoc Study Group on Carbon Dioxide and Climate. Woods Hole, Massachusetts. July 23-27, 1979.
  93. [93]
    [PDF] Climate Change – The IPCC Scientific Assessment
    The IPCC report assesses scientific information on climate change, including greenhouse gas emissions, and aims to evaluate environmental and socio-economic ...
  94. [94]
  95. [95]
    Climate Change 2021: The Physical Science Basis
    The novel AR6 WGI Interactive Atlas allows for a flexible spatial and temporal analysis of both data-driven climate change information and assessment findings.Chapter 7: The Earth’s Energy... · IPCC Sixth Assessment Report · Press · Atlas
  96. [96]
    [PDF] Summary for Policymakers
    a higher average climate sensitivity than CMIP5 and the AR6 assessed best estimate. The higher CMIP6 climate sensitivity values compared to CMIP5 can be ...
  97. [97]
    Paleoclimate data provide constraints on climate models' large ...
    Aug 7, 2024 · Here, we use a set of metrics based on paleoclimate proxy observations to evaluate climate models under three past time periods.
  98. [98]
    [PDF] The PMIP4 Last Glacial Maximum experiments: preliminary results ...
    May 20, 2021 · Changes in model configuration have resulted in several of the PMIP4-CMIP6 models having sub- stantially higher climate sensitivity than the ...
  99. [99]
    Lessons on Climate Sensitivity From Past Climate Changes - PMC
    In this paper, we review the progress made in theoretical understanding of climate sensitivity and on the estimation of climate sensitivity from proxy records.
  100. [100]
    Increasingly Sophisticated Climate Models Need the Out‐Of ...
    Dec 9, 2022 · Here we discuss the valuable tests that paleoclimate reconstructions can provide the latest generation of climate models.
  101. [101]
    The implications for climate sensitivity of AR5 forcing and heat ...
    Sep 25, 2014 · Lewis, N., Curry, J.A. The implications for climate sensitivity of AR5 forcing and heat uptake estimates. Clim Dyn 45, 1009–1023 (2015).
  102. [102]
    Understanding Climate Feedbacks and Sensitivity Using ... - NIH
    Sep 1, 2016 · In order to evaluate the influence of a data gap on satellite-derived effective climate feedback parameter, we consider monthly anomalies ...
  103. [103]
    Can Top-of-Atmosphere Radiation Measurements Constrain ...
    Amplifying the CERES observational covariance estimate by a factor of 20 brings CERES and ERBE estimates into agreement. In this case the climate sensitivity ...
  104. [104]
    Objectively combining climate sensitivity evidence | Climate Dynamics
    Sep 19, 2022 · This study evaluates the methodology of and results from a particularly influential assessment of climate sensitivity that combined multiple lines of evidence.
  105. [105]
    Addressing misconceptions about Climate Sensitivity research
    Aug 13, 2025 · The determination of equilibrium climate sensitivity (ECS)—the long-term warming response to doubled atmospheric CO2 concentrations—remains one ...
  106. [106]
    Comment on "Can uncertainty in climate sensitivity be narrowed ...
    Mar 24, 2025 · Such uncertainties could affect ECS estimation not only from historical period evidence but also from climate process understanding, ...
  107. [107]
    WMO confirms 2024 as warmest year on record at about 1.55°C ...
    Jan 10, 2025 · WMO confirms 2024 as warmest year on record at about 1.55°C above pre-industrial level · The past ten years 2015-2024 are the ten warmest years ...Missing: sensitivity | Show results with:sensitivity
  108. [108]
    State of the climate: 2024 sets a new record as the first year above ...
    Jan 10, 2025 · Global surface temperatures set a new record in 2024, surpassing the record set in 2023 by around 0.11C. It was unambiguously the warmest year ...Missing: revised bounds
  109. [109]
    Recent global temperature surge intensified by record-low planetary ...
    Dec 5, 2024 · In 2023, the global mean temperature soared to almost 1.5 kelvin above the preindustrial level, surpassing the previous record by about 0.17 kelvin.
  110. [110]
  111. [111]
    Assessing the Global Climate in 2024 | News
    Jan 10, 2025 · In 2024, global surface temperature was 2.32°F (1.29°C) above the 20th-century average. This ranks as the highest global temperature in the period of 1850–2024.
  112. [112]
    Chapter 4 | Climate Change 2021: The Physical Science Basis
    This chapter assesses simulations of future global climate change, spanning time horizons from the near term (2021–2040), mid-term (2041–2060), and long term ( ...
  113. [113]
    [PDF] Emergent constraints on climate sensitivity and recent record ...
    Studies that have constrained TCR using emergent constraint methods have consistently shown that high-sensitivity models are less consistent with historical ...
  114. [114]
    Can uncertainty in climate sensitivity be narrowed further?” by ... - ACP
    Aug 13, 2025 · Peer-reviewed comment | | 13 Aug 2025 ... They seem more likely to suggest that existing ECS estimates are too high rather than too low.