1. Introduction
The surface temperature discontinuity δTs refers to the difference between the temperature of the skin layer Ts and of the overlying air Tas, typically at a height of 2 m. The standard convention is that δTs = Ts − Tas and it is generally positive, driving turbulent heat and moisture fluxes from the surface to the atmosphere above. Changes in δTs are intricately coupled to the surface energy balance and, via the latent heat flux, to the hydrological cycle. Climate models simulate regionally varying changes in δTs (ΔδTs, henceforth Δ denotes a change) under global warming. Model–observation comparisons are then complicated, since global-mean surface temperature (GMST) records combine measurements of SST over ocean with Tas over land and sea ice. In situ data typically estimate SST from near-surface water samples rather than the ocean skin layer, but observed SST trends are very similar to those in Ts (Merchant et al. 2014; Hausfather et al. 2017). In models, changes in Ts and near-surface water temperatures are similar enough that they are interchangeable for the purposes of this study (e.g., Fig. S1 in the online supplemental material). The Coupled Model Intercomparison Project (CMIP) Tas diagnosis method is left to modeling groups, for an example see section 5 of Oleson et al. (2010), where it depends on properties including Ts and the temperature of the lowest atmospheric layer.
Historically, many comparisons used model global-mean Tas [global surface air temperature (GSAT)], and in CMIP5 models, 1850–2015 ΔGSAT averaged 24% larger than when model data were handled similarly to the HadCRUT4 GMST record (Richardson et al. 2016). This difference was mostly due to incomplete spatial data coverage, and a further part was due to the treatment of sea ice changes (Cowtan et al. 2015). The effect of using ocean Tas rather than Ts on warming through 2100 averaged 4% of global warming, but even this small bias has consequences for the actions necessary to achieve global climate change targets. For example, the Paris Agreement aims to limit global warming below 2°C, but did not specify whether this is in terms of ΔGMST or ΔGSAT. If ΔGSAT were selected rather than ΔGMST, then cumulative emissions would have to be 37 GtC lower to achieve a 2°C target (Richardson et al. 2018), which is equivalent to over 3 years of human emissions.
GMST datasets have since improved spatial coverage (Morice et al. 2021; Vose et al. 2021) and some changed sea ice treatment (Rohde and Hausfather 2020), but with no standard treatment for Tas − Ts blending, ΔGMST is sometimes scaled by a model-derived factor to estimate ΔGSAT (Clarke and Richardson 2021). Ship and buoy data disagree even about the sign of ocean δTs trends (Junod and Christy 2020; Rubino et al. 2020), but dataset uncertainties are currently too large to allow confident conclusions (Morice et al. 2021).
Referring to δTs, the Intergovernmental Panel on Climate Change Sixth Assessment Report (IPCC AR6) stated that “there is no simple explanation based on physical grounds alone for how this difference responds to climate change” (Gulev et al. 2021). This paper’s primary motivation is to provide a simple physical explanation for differences in model ΔGSAT and ΔGMST, so the analysis targets nonpolar ocean within ±60°S/N which includes most global ocean. Complications from polar and land processes are therefore excluded, but future work there could allow understanding of how, for example, in situ land ΔTas should be compared with satellite-retrieved ΔTs.
A secondary motivation is that the model treatment of δTs and near-surface processes is also intricately linked to the hydrological cycle and its anticipated intensification under global warming (Andrews et al. 2009). Modeled global precipitation increases at just ∼2% °C−1 while specific humidity rises at ∼7% °C−1, and several studies have investigated this difference with reference to δTs. Richter and Xie (2008) stated that shrinking δTs reduces evaporation by ∼1 W m−2, and they argued that this “is consistent with the high heat capacity of the ocean.” Lorenz et al. (2010) also described δTs as limiting evaporation increases, but stated that causality could act the other way and that evaporation “will act to increase Ta relative to Ts.” More recently, Siler et al. (2019) used a Penman–Monteith framework to link changes in the surface energy budget to evaporation. They noted that evaporation transports heat away from the surface so efficiently that δTs must change to maintain energy conservation, as previously described for some extreme paleoclimate states by Pierrehumbert (2002). Here I show that contrary to Richter and Xie (2008)’s stated causality, ocean heat uptake does not cause long-term ΔδTs < 0, rather there is stronger support for the energetic description of Siler et al. (2019) in which evaporation drives ΔδTs < 0. In addition to the temperature-mediated (“feedback”) effect, there is also a direct effect of atmospheric adjustments to CO2 radiative forcing.
This study only considers CO2-related changes, since CO2 is the dominant cause of recent warming and there is a suite of model tools to evaluate CO2 forcing scenarios. The conclusions may not extend to other forcings such as aerosols, which have been important for recent ΔGMST progression. The results show an effective equilibrium attribution of 34% ± 13% to forcing, and 66% ± 13% to feedbacks (ensemble mean ± standard deviation). CO2 forcing directly reduces δTs through rapid adjustments driven primarily by atmospheric longwave heating (e.g., Kamae et al. 2015). Warming then strongly increases evaporation, which cools Ts and further decreases δTs.
I will proceed as follows: section 2 shows model ΔδTs and uses millennium-length simulation output to demonstrate that ocean heat uptake cannot explain the long-term ΔδTs. The long simulation shows a distinction between a short-term forcing-dominated ΔδTs followed by a longer-term feedback-related ΔδTs. This section also summarizes canonical rapid atmospheric changes to CO2 forcing and argues that they explain the short-term ΔδTs. The remainder of the paper is devoted to a feedback analysis, beginning with section 3, which derives how surface feedbacks drive ΔδTs. Feedback calculations begin in section 4 using radiative kernels from the HadGEM2-ES climate model. Kernels are ideal for this purpose since they uniquely isolate surface contributions from those of the atmosphere, but are only available for some energy-budget terms and for few models. The analysis then proceeds to the standard Gregory regression (Gregory et al. 2004) decomposition using abrupt4xCO2 output to evaluate all feedback components in 23 models. The regression methodology and validation for surface feedbacks is described in section 5 and the results are presented in section 6. Section 7 discusses the study and section 8 presents the conclusions.
2. Model changes in surface temperature discontinuity under CO2-forced warming
This section describes the typical phase 5 of the Coupled Model Intercomparison Project (CMIP5) model δTs behavior in CO2-forced simulations. It highlights how ocean differs from other surfaces and uses long-run simulation output to show that there are both forcing and feedback contributions. Figure 1 maps ΔδTs between years 1–10 and 61–80 in eight CMIP5 1% yr−1 CO2 ramping simulations (1pctCO2). These eight models provide all properties necessary for later analysis, and years 61–80 cover atmospheric CO2 doubling. There is widespread ΔδTs < 0°C over nonpolar ocean, intermodel disagreement over land, and ΔδTs > 0°C in polar regions. By year 140, mean nonpolar ocean Tas warms by 5%–10% more than Ts in all models (Fig. S2).
As proposed by Richter and Xie (2008), water may warm less than the overlying air due to the oceans’ large heat capacity, so ΔδTs < 0 may disappear once the ocean warms to equilibrium. This hypothesis is testable using recently available LongRunMIP output (Rugenstein et al. 2019) from an abrupt4xCO2 simulation in which atmospheric CO2 is quadrupled and then held fixed. If ocean heat uptake dominates, then ΔδTs should transiently decrease then increase back to equilibrium. However, Fig. 2a shows that for nonpolar ocean, ΔδTs < 0 even after 1300 years. Figures 2b and 2c changes differ for nonpolar land and polar regions, so this study’s conclusions only apply to nonpolar oceans. Returning to Fig. 2a, the year 1 abrupt4xCO2 δTs is roughly 0.1°C below preindustrial, and this is followed by a longer-term decrease of approximately 0.2°C. I argue that these two features can be understood in terms of a standard forcing–feedback decomposition.
Importantly, the effect of CO2 forcing on the atmosphere is well-established and can explain the rapid modeled δTs decrease. To the author’s knowledge, the link between ΔδTs and CO2 forcing adjustments has not previously been made explicit, so this understanding is now summarized. Additional CO2 causes instantaneous radiative forcing (iRF) that heats the lower atmosphere. Doubled-CO2 longwave heating rates approach 0.1°C day−1 in the low to midtroposphere, with smaller positive values extending to the surface (Collins et al. 2006). Rather than iRF, a more commonly used value is effective radiative forcing (ERF), which includes atmospheric adjustments to the forcing agent that are independent of surface temperature. They often occur within days to weeks and so can dominate short-term atmospheric changes (Cao et al. 2012; Dong et al. 2009). The canonical over-ocean adjustments include warming air temperature (Ta) with drying aloft and PBL moistening, which suppresses evaporation. These adjustments are described in Kamae et al. (2015). Since ERF refers to changes that are independent of Ts, CO2-driven ERF adjustments with ΔTas > 0 therefore mean ΔδTs < 0. Model ERF is commonly approximated with reference to globally fixed Tas (Gregory et al. 2004) or fixed Ts over oceans (Hansen et al. 2005) and Hansen-method CMIP5 simulations indeed show ΔδTs < 0 over oceans (Fig. S3). Not all of Fig. 2’s first year changes are due to forcing since warming begins immediately, but most of the rapid changes will be forcing related.
Henceforth the CO2 ERF adjustments are taken as explained, and the analysis focusses on (i) interpreting how individual feedbacks drive ΔδTs and (ii) separately quantifying the contributions of ERF and feedbacks to ΔδTs. Addressing (i) and (ii) requires establishing the forcing–feedback framework and linking individual feedback terms to changes in δTs.
3. Interpreting feedback-driven changes in the surface temperature discontinuity
I define
4. Radiative kernel analysis
a. Theory, data, and processing
I use the published HadGEM2-ES kernels since they fully decouple Ts and Ta (Smith et al. 2018; Smith 2018). Several other kernel datasets combine changes in Ts and in the lowest atmospheric layer Ta,i=1, so cannot separate
b. Kernel results
Figure 3 shows non–cloud radiative
5. CMIP5 methods
a. Overview of approach
Gregory regression (Gregory et al. 2004) was developed to estimate forcing and feedback strengths for global-mean TOA flux in response to global-mean ΔTas. It was subsequently used to study regional feedbacks (e.g., Andrews et al. 2015). Here I use it to calculate individual surface energy budget terms over nonpolar oceans. To justify interpretation based on the Gregory-derived parameters, I require that the parameters accurately predict output from fully coupled 1pctCO2 simulations.
b. Data sources
I use CMIP5 preindustrial control (piControl) and abrupt4xCO2 scenarios, taking the first run from each model (labeled r1i1p1) if they provide monthly Ts, Tas, and all surface and TOA fluxes (N = 23 models, see Table S1). Those models with matching 1pctCO2 output (N = 8) are used to test the regression performance. Model drift in abrupt4xCO2 is removed following DeAngelis et al. (2015) by subtracting the centered 21-yr mean from the corresponding model piControl simulation in each grid cell. If the abrupt4xCO2 fork year is not provided, then I match piControl years 1–21 to abrupt4xCO2 year 1.
c. Separation of forcing and feedbacks
1) Regional Gregory decomposition
The surface–atmosphere feedback relationship described in Eq. (6) defines whether δTs increases or decreases in response to feedbacks. The term
2) Evaluating Gregory method
The spatial distribution of ΔNs and ΔTas activates regional feedbacks and changes global-mean responses (Dong et al. 2019; Rose et al. 2014; Armour et al. 2013; Gregory et al. 2015; Knutti and Rugenstein 2015). The recently observed spatial ΔTas pattern implies 30-yr mean TOA λ that can be 100% different from that associated with the simulated long-term warming pattern (Zhou et al. 2016). Furthermore, quadrupled-CO2 responses are not precisely twice those experienced under doubled-CO2 (Mitevski et al. 2022).
3) Attributing discontinuity change to feedbacks and forcing
6. Results
a. Validation of Gregory decomposition
Figure 5 shows that the Eq. (11) Gregory-based reconstruction provides accurate predictions of 1pctCO2 output. However, Gregory regression produces some nonphysical results since its ERF is inferred from a statistical fit extended to global ΔTas = 0°C, which implies land warming and ocean cooling. The ocean cooling, seen as the blue lines in Figs. 5a and 5b, is clearly unphysical and means that some local feedback effects will be included as forcing. From Fig. 5a, the cooling is equivalent to approximately −7% of the transient warming, so forcing terms in Figs. 5c–5h will include an artificial component equivalent to approximately −7% of each feedback term. Section 7c will discuss how ERF estimates depend on the calculation method, and concludes that forcing definition has a minor effect on this study’s conclusions.
In terms of reconstructing the coupled model output, the worst performance is for SWnet, which has large cloud-driven interannual variability relative to its long-term change. Across the eight 1pctCO2 runs, errors are typically <10% by years 61–80 (Fig. S5). The results support a Gregory-method decomposition over nonpolar oceans so the derived forcing and feedback parameters for all 23 models can be used to interpret the physical drivers of ΔδTs.
Also from Fig. 5, for properties where feedbacks are relatively more important than ERF, there is more curvature in the response. This is because global ΔTas,global is delayed by heat uptake Ns and slowly accelerates as atmospheric CO2 accumulates. Properties with large feedback components, such as evaporation, therefore, show more notable curvature. This is consistent with the results of Yeh et al. (2021), who noted asymmetric precipitation responses in models when CO2 is ramped up versus ramped down.
b. Ensemble Gregory forcing and feedback responses
1) Changes in ΔTs
The derived forcing, feedback and effective equilibrium responses of δTs and surface fluxes are displayed for nonpolar ocean in Fig. 6a. Feedbacks are divided by oceanic
Section 7 argues that the small-magnitude
2) Changes in surface fluxes
Figures 6b–6d show Gregory-derived flux responses and include cloud radiative effects (CRE), the difference between all-sky radiation and what the radiation would be in the absence of clouds. In these panels, signs are selected so that values greater than zero are of the correct direction to support ΔδTs < 0. For forcing, Fig. 6b changes follow the previously described atmospheric adjustments to CO2, with net longwave surface heating and suppressed L and H establishing a large ΔNs.
For feedbacks in Fig. 6c results must be interpreted with care, as Eq. (12) states that the Gregory feedbacks reported here are a combination of response to changes in Ts, Tas, and δTs so cannot be directly applied to Eq. (6). There is a large range in SW CRE that fits with the changes in oceanic low clouds that cause large intermodel spread in ECSeff (Bony and Dufresne 2005). Atmospheric moistening shifts all-sky SWnet toward less surface absorption relative to SW CRE alone. The LWnet feedback is of the wrong sign to support ΔδTs < 0 and the net surface SW feedback has models with either sign. The LW and SW CRE components anticorrelate, resulting in a residual net CRE feedback of +0.6 W m−2 °C−1, of a sign that acts to decrease δTs. The Gregory-derived radiation feedback parameters are qualitatively consistent with the kernel results, suggesting that
The sensible heat flux H is typically parameterized in models as proportional to δTs, and so its Gregory feedback is likely to be dominated by
Figure 6d’s effective-equilibrium values show the combined forcing and feedback effects at ΔTas = ECSeff, and the range of ECSeff causes much of the spread in results. Crucially, the implied equilibrium ΔδTs (Fig. 6a) is negative in all models, while ΔNs is very small (from −1.2 to +0.2 W m−2) relative to the changes in other terms such as evaporation (4.2–11.7 W m−2). This further supports that heat uptake changes do not explain ΔδTs. The equilibrium ΔNs represents transport between nonpolar oceans and other areas, in particular increased transport to the polar regions (Hwang et al. 2011; Holland and Bitz 2003; Newsom et al. 2021).
Taken together:
-
Figure 6d’s minor ΔNs at equilibrium justifies the approach resulting in Eq. (6).
-
Figure 6c’s radiation terms plus the kernel results rule out radiation feedbacks from explaining ΔδTs on their own.
-
Figure 6c’s H feedback sign rules out turbulent heat, while the L feedback represents a large cooling effect on Ts consistent with evaporation driving decreased δTs.
Assuming that λL,greg is a minimum bound and the Clausius-Clapeyron limit an upper bound, then over nonpolar oceans climate models likely have
c. Attribution of δTs change between forcing and feedbacks for nonpolar ocean
Figure 7 applies Eq. (14) and expresses the fraction of ΔδTs due to feedbacks as a function of global ΔTas divided by ECSeff. ΔδTs is initially dominated by ERF, before warming triggers feedbacks that drive further ΔδTs. According to the Gregory regression parameters, when warming reaches ECSeff the feedbacks explain a mean of 66% of ΔδTs with an intermodel range of 48%–92%. Year 70 in 1pctCO2 simulations is commonly used as a measure of transient response, with warming approximately 50%–60% of ECSeff, and at this point the forcing–feedback split is roughly even. Nonpolar oceanic ΔδTs in the modern day therefore likely contains similar-magnitude contributions both from CO2 forcing adjustments and temperature-mediated feedbacks. Neither the intermodel ΔδTs nor the attribution fraction strongly correlate with any surface energy budget term, so understanding intermodel variability would require a more detailed analysis. Section 7c will discuss how this results in a small overestimate of the feedback contribution.
7. Discussion
a. Summary of results
Changes in the surface temperature discontinuity δTs can bias model–observation comparisons of global mean temperature change and are intricately tied to the surface energy budget. This paper provides a simple, physical mechanism to explain modeled ΔδTs over nonpolar oceans, which IPCC AR6 identified as lacking. The explanation comes in two parts:
-
CO2 heats the lower atmosphere and subsequent rapid adjustments cause Tas to warm more than Ts, therefore ΔδTs < 0,
-
Warming then increases surface evaporation so that Ts warms less than Tas, therefore ΔδTs < 0.
Both processes have been reported elsewhere, but here I confirm this model behavior and quantify their contributions to ΔδTs. Heat uptake ΔNs shrinks to near zero by ECSeff over oceans, so it cannot explain equilibrium ΔδTs. This is consistent with the interpretation of Siler et al. (2019) rather than Richter and Xie (2008), the latter of which proposed that Ns reduces δTs. The presented results are only for large-area means, however, and Ns may be important locally.
This paper’s analysis began when only CMIP5 had sufficient outputs available, but CMIP6 shows similar ERF adjustments (Smith et al. 2020), evaporation feedback (Pendergrass 2020), and decreased ΔδTs (e.g., Liang et al. 2020), so the process conclusions likely apply to CMIP6 as well. CMIP6 features more models with higher ECSeff, which may result in larger equilibrium ΔδTs, but with minor present-day differences.
It is only nonpolar ocean where ΔδTs < 0 in all models so observational datasets, which combine Ts over ocean and Tas elsewhere, will report less long-term warming than either global Ts or global Tas. Other regions may have limited moisture availability that limits surface evaporation (Gregory and Webb 2008) or changes in surface type, such as from sea ice retreat, that would further complicate changes (see Fig. S7 for all surface Gregory results). Note that the effect of ΔδTs changes reported here are proportionally smaller than those reported in Richardson et al. (2016) since sea ice changes are excluded here.
b. Radiative kernel interpretation and model treatment of non–cloud radiation changes
The radiative kernel results use just HadGEM2-ES, which is not an outlier in terms of ΔδTs or flux changes so its findings may extend more broadly. The kernel-based
For ΔδTs, ongoing model mean-state biases (Xu et al. 2022) can influence calculated
The surface is both warmer and more emissive than the atmosphere so for uniform vertical warming the temperature component of
The moisture kernel calculations assumed fixed relative humidity (RH), while models typically show increases over ocean (Byrne and O’Gorman 2016) since specific humidity changes similarly over land and ocean but land warms faster (Simmons et al. 2010). Higher RH would favor ΔδTs > 0, so relaxing the fixed-RH assumption would likely only further rule out non–cloud radiative effects from explaining the coupled-model ΔδTs. Finally, cloud masking will change with warming, but the clear-sky kernel-derived feedbacks differ by just 0.3 W m−2 °C−1, which is far too small to overturn the main conclusions. Taken together, the kernel results suggest a minor role for non–cloud radiation in explaining the feedback-driven ΔδTs.
c. Gregory decomposition limitations and interpretation
As discussed in section 6a, Gregory regression can produce nonphysical forced oceanic cooling since it is fit to global ΔTas = 0. The Hansen method avoids artificial ocean cooling but allows land temperature response (Fig. S3). Either standard method has errors in separating forcing from feedback, and for the four models that also provide fixed-SST outputs the Hansen
Land Tas warming in the Hansen calculations could affect coastal areas, so one may anticipate that Hansen
Gregory regression also assumes linear responses to global Tas, which is violated over multiple centuries (Rugenstein et al. 2020). It seems unlikely that this would counteract short-term forcing adjustments and there is no evidence of later suppression of evaporation (Schwarzwald et al. 2021). Therefore, the effective equilibrium mechanisms should qualitatively apply in the long term. Importantly, when 1pctCO2 outputs of nonpolar ocean properties are reconstructed using Gregory-derived parameters, errors are of order <10% in the transient representation of all terms (Fig. S5), so conclusions relevant to historical warming seem unaffected by nonlinearity. Ultimately, the feedback-related L changes are large and consistent, and I argue that their physical link to a cooling of Ts is simple enough to justify the argument that it drives decreased δTs.
d. Model parameterizations and ΔδTs
1) Model treatment of near-surface properties
Common modeling choices could consistently bias CMIP simulations. For example, IPCC AR6 notes that all CMIP5 and CMIP6 models use Monin–Obukhov similarity theory (Monin and Obukhov 1954; Obukhov 1971) in their PBL schemes to diagnose Tas and turbulent fluxes. There is evidence that BCC-CSM1-1-M has some error in either Tas or Ts derivation, with a preindustrial ocean δTs mean of just 0.06°C (Table S1). This is far smaller than the other 22 models or in situ measurements from, for example TAO (McPhaden et al. 1998) or PIRATA (Servain et al. 1998) buoys. I discuss below the credibility and testability of model representations of forcing- and feedback-driven ΔδTs. For feedbacks in particular, the energetics framework adopted here means that errors in the PBL scheme alone would be insufficient to overturn the conclusions unless they changed surface fluxes in ways that are potentially measurable.
2) Forcing-related changes
Strong evidence validates line-by-line radiation codes (e.g., Tjemkes et al. 2003), CO2’s surface iRF spectral signature has been observed (Feldman et al. 2015) and known iRF uncertainties are a small fraction of the mean (Mlynczak et al. 2016). It is unlikely that all CMIP models have consequential errors in CO2’s direct radiative effects, but the atmospheric adjustments that translate from iRF to ERF and result in Tas warming are not, to my knowledge, directly validated with observations. The real world experiences simultaneous changes in forcings, temperature, and internal variability, so isolating CO2-only ERF adjustments seems exceptionally challenging. Some confidence may be drawn from how CO2-driven rapid adjustments are similar across the model hierarchy, including in a multiscale modeling framework (Xu et al. 2020) and in large-eddy simulations (LES) of marine stratocumulus regions (Bretherton and Blossey 2014; Blossey et al. 2016). LES resolution is up to 1000 times finer than that of CMIP models, ruling out errors in some CMIP subgrid parameterizations. However, there could be common errors across the full model hierarchy.
3) Feedback-related changes
Feedback-driven ΔδTs < 0 can, I argue, be understood from the relative strengths of feedbacks defined with respect to Tas and Ts (for radiation) or Ts and δTs (for turbulent fluxes), with a requirement that
The Gregory-derived λL,greg combines both
It may be possible to indirectly infer whether real-world evaporation changes are consistent with models. First, ΔL will be matched by changes in precipitation, which in turn is controlled by the atmospheric energy budget (Manabe and Wetherald 1975; Takahashi 2009). Model errors would likely have to extend above the PBL to restrict surface evaporation from playing a large role. Some inference could also be obtained from trends in PBL RH or height, since rising PBL RH limits evaporation increase (Lorenz et al. 2010) and PBL RH tends to anticorrelate with PBL height (e.g., Zhang et al. 2013). In situ ocean evaporation datasets show increasing trends, but have sparse spatial sampling and large trend uncertainties (Zhang et al. 2018). Other signatures of evaporation include amplified tropical upper troposphere warming. Satellite microwave sounding unit (MSU) and weather-balloon datasets disagree about the existence of this amplification (Santer et al. 2017; Spencer et al. 2017; Mitchell et al. 2020; Allen and Sherwood 2008). Global Navigation Satellite System-Radio Occultation (GNSS-RO) records (Gleisner et al. 2022) are more stable in time than MSU or radiosonde records, and could potentially support or refute the feedback effect identified here. Lakes have been called “sentinels” of climate change (Adrian et al. 2009) and have similar surface physics to oceans. Detailed lake modeling reports low-latitude lake changes that are similar to my nonpolar ocean results here, with ΔδTs < 0 and increased evaporation (Wang et al. 2018). However, lakes are strongly affected by nearby land and are sensitive to stratification changes that are dissimilar to those of oceans (Kraemer et al. 2015; Anderson et al. 2021; Stetler et al. 2021).
e. Future research priorities
If observations supported multidecadal ΔδTs > 0, that would suggest serious errors in modeled surface energy exchange. It could, for example, require smaller precipitation increase with warming or imply large errors in currently understood atmospheric forcing adjustments. Therefore, improved observational evidence is exceptionally important. This includes indirect information, for example the PREFIRE (L’Ecuyer et al. 2021) and FORUM (Palchetti et al. 2020) missions will improve knowledge of far-infrared ε and therefore reduce errors in calculated surface feedbacks.
Confirmation of similar CMIP6 behavior would be useful, although section 7a argues the conclusions are unlikely to change. Conclusions may differ over land and polar regions, especially where moisture is limited. Results may also be sensitive to the forcing agent, either due to different ERF adjustments or if they express a different spatial pattern of heating or cooling and so activate a feedback pattern effect. The results of the Precipitation Driver Model Intercomparison Project [PDRMIP, Samset et al. (2016)] indicate large disagreements in ERF-related changes depending on the forcing agent while in Andrews et al. (2021), solar ERF has approximately half of CO2’s effect on ΔδTs. Models disagree strongly on the effects of absorbing aerosols which warm the atmosphere (Johnson et al. 2019), driving larger ΔδTs than seen for CO2. Meanwhile, sulfate aerosols are a major source of forcing uncertainty and were linked to decreased pan evaporation over land during the late twentieth century (Roderick and Farquhar 2002).
Finally, the results presented here are relevant for long-term forced changes in CMIP models. Changes in the shorter term, or for periods when feedbacks differ from those simulated in CMIP models may be different. Recent decades may show different ΔδTs due to the “pattern effect” since recent temperature trend patterns, particularly in the Pacific, are not commonly simulated by CMIP models (Fueglistaler and Silvers 2021; Watanabe et al. 2021).
8. Conclusions
This paper explains why models report more warming of global air temperatures than the combined air–water global surface temperature datasets (Cowtan et al. 2015; Richardson et al. 2016). Decreased δTs can be understood from surface–atmosphere energy exchange rather than the oceans’ large transient heat uptake, so models project ΔδTs < 0 even at equilibrium. CO2 forcing adjustments preferentially warm the air, and warming-driven evaporation preferentially cools the surface. Combined, these provide a simple physical explanation for changes in the surface temperature discontinuity.
Future model research could explain changes over land or polar surfaces, or in response to non-CO2 forcing agents such as aerosol while current datasets are too uncertain to allow confident conclusions about trends in δTs. An alternative approach would be to indirectly constrain the relevant processes. For example, modeled extra warming of air relative to water is fundamentally linked to increased evaporation under global warming. Observations could be targeted to determine whether changes in precipitation and atmospheric heating structure are consistent with the simulated surface evaporative cooling, and this would be a key step in evaluating whether model simulation of near-surface changes are realistic.
Acknowledgments.
Government sponsorship acknowledged. This research was carried out at the Jet Propulsion Laboratory, California Institute of Technology under a contract with NASA. The author thanks Dr. Matthew D. Lebsock and Prof. Graeme L. Stephens for helpful discussions; Dr. Ryan Kramer and Dr. Chris Smith for assistance with radiative kernels; and Prof. Maria Rugenstein and Dr. Jonah Bloch-Johnson for help with the LongRunMIP output.
Data availability statement.
The CMIP5 data used here are available from the https://esgf-node.llnl.gov/search/cmip5/ and the HadGEM2 radiative kernels are at https://archive.researchdata.leeds.ac.uk/382/. LongRunMIP data for Fig. 2 are accessible by following the instructions at http://www.longrunmip.org/ (access is free but a password must be requested). The relevant kernel and LongRunMIP citations are in the text.
REFERENCES
Adrian, R., and Coauthors, 2009: Lakes as sentinels of climate change. Limnol. Oceanogr., 54, 2283–2297, https://doi.org/10.4319/lo.2009.54.6_part_2.2283.
Allen, R. J., and S. C. Sherwood, 2008: Warming maximum in the tropical upper troposphere deduced from thermal winds. Nat. Geosci., 1, 399–403, https://doi.org/10.1038/ngeo208.
Anderson, E. J., and Coauthors, 2021: Seasonal overturn and stratification changes drive deep-water warming in one of Earth’s largest lakes. Nat. Commun., 12, 1688, https://doi.org/10.1038/s41467-021-21971-1.
Andrews, T., P. M. Forster, and J. M. Gregory, 2009: A surface energy perspective on climate change. J. Climate, 22, 2557–2570, https://doi.org/10.1175/2008JCLI2759.1.
Andrews, T., J. M. Gregory, and M. J. Webb, 2015: The dependence of radiative forcing and feedback on evolving patterns of surface temperature change in climate models. J. Climate, 28, 1630–1648, https://doi.org/10.1175/JCLI-D-14-00545.1.
Andrews, T., C. J. Smith, G. Myhre, P. M. Forster, R. Chadwick, and D. Ackerley, 2021: Effective radiative forcing in a GCM with fixed surface temperatures. J. Geophys. Res. Atmos., 126, e2020JD033880, https://doi.org/10.1029/2020JD033880.
Armour, K. C., C. M. Bitz, and G. H. Roe, 2013: Time-varying climate sensitivity from regional feedbacks. J. Climate, 26, 4518–4534, https://doi.org/10.1175/JCLI-D-12-00544.1.
Blossey, P. N., C. S. Bretherton, A. Cheng, S. Endo, T. Heus, A. P. Lock, and J. J. van der Dussen, 2016: CGILS phase 2 LES intercomparison of response of subtropical marine low cloud regimes to CO2 quadrupling and a CMIP3 composite forcing change. J. Adv. Model. Earth Syst., 8, 1714–1726, https://doi.org/10.1002/2016MS000765.
Bony, S., and J.-L. Dufresne, 2005: Marine boundary layer clouds at the heart of tropical cloud feedback uncertainties in climate models. Geophys. Res. Lett., 32, L20806, https://doi.org/10.1029/2005GL023851.
Bretherton, C. S., and P. N. Blossey, 2014: Low cloud reduction in a greenhouse-warmed climate: Results from Lagrangian LES of a subtropical marine cloudiness transition. J. Adv. Model. Earth Syst., 6, 91–114, https://doi.org/10.1002/2013MS000250.
Byrne, M. P., and P. A. O’Gorman, 2016: Understanding decreases in land relative humidity with global warming: Conceptual model and GCM simulations. J. Climate, 29, 9045–9061, https://doi.org/10.1175/JCLI-D-16-0351.1.
Cao, L., G. Bala, and K. Caldeira, 2012: Climate response to changes in atmospheric carbon dioxide and solar irradiance on the time scale of days to weeks. Environ. Res. Lett., 7, 034015, https://doi.org/10.1088/1748-9326/7/3/034015.
Clarke, D. C., and M. Richardson, 2021: The benefits of continuous local regression for quantifying global warming. Earth Space Sci., 8, e2020EA001082, https://doi.org/10.1029/2020EA001082.
Collins, W. D., and Coauthors, 2006: Radiative forcing by well-mixed greenhouse gases: Estimates from climate models in the Intergovernmental Panel on Climate Change (IPCC) Fourth Assessment Report (AR4). J. Geophys. Res., 111, D14317, https://doi.org/10.1029/2005JD006713.
Cowtan, K., and Coauthors, 2015: Robust comparison of climate models with observations using blended land air and ocean sea surface temperatures. Geophys. Res. Lett., 42, 6526–6534, https://doi.org/10.1002/2015GL064888.
Dai, N., R. J. Kramer, B. J. Soden, and T. S. L’Ecuyer, 2021: Evaluation of CloudSat radiative kernels using ARM and CERES observations and ERA5 reanalysis. J. Geophys. Res. Atmos., 126, e2020JD034510, https://doi.org/10.1029/2020JD034510.
DeAngelis, A. M., X. Qu, M. D. Zelinka, and A. Hall, 2015: An observational radiative constraint on hydrologic cycle intensification. Nature, 528, 249–253, https://doi.org/10.1038/nature15770.
Dee, D. P., and Coauthors, 2011: The ERA-Interim reanalysis: Configuration and performance of the data assimilation system. Quart. J. Roy. Meteor. Soc., 137, 553–597, https://doi.org/10.1002/qj.828.
Dong, B., J. M. Gregory, and R. T. Sutton, 2009: Understanding land–sea warming contrast in response to increasing greenhouse gases. Part I: Transient adjustment. J. Climate, 22, 3079–3097, https://doi.org/10.1175/2009JCLI2652.1.
Dong, Y., C. Proistosescu, K. C. Armour, and D. S. Battisti, 2019: Attributing historical and future evolution of radiative feedbacks to regional warming patterns using a Green’s function approach: The preeminence of the western Pacific. J. Climate, 32, 5471–5491, https://doi.org/10.1175/JCLI-D-18-0843.1.
Feldman, D. R., W. D. Collins, R. Pincus, X. Huang, and X. Chen, 2014: Far-infrared surface emissivity and climate. Proc. Natl. Acad. Sci. USA, 111, 16 297–16 302, https://doi.org/10.1073/pnas.1413640111.
Feldman, D. R., W. D. Collins, P. J. Gero, M. S. Torn, E. J. Mlawer, and T. R. Shippert, 2015: Observational determination of surface radiative forcing by CO2 from 2000 to 2010. Nature, 519, 339–343, https://doi.org/10.1038/nature14240.
Fueglistaler, S., and L. G. Silvers, 2021: The peculiar trajectory of global warming. J. Geophys. Res. Atmos., 126, e2020JD033629, https://doi.org/10.1029/2020JD033629.
Gleisner, H., M. A. Ringer, and S. B. Healy, 2022: Monitoring global climate change using GNSS radio occultation. npj Climate Atmos. Sci., 5, 6, https://doi.org/10.1038/s41612-022-00229-7.
Graversen, R. G., P. L. Langen, and T. Mauritsen, 2014: Polar amplification in CCSM4: Contributions from the lapse rate and surface albedo feedbacks. J. Climate, 27, 4433–4450, https://doi.org/10.1175/JCLI-D-13-00551.1.
Gregory, J. M., and M. Webb, 2008: Tropospheric adjustment induces a cloud component in CO2 forcing. J. Climate, 21, 58–71, https://doi.org/10.1175/2007JCLI1834.1.
Gregory, J. M., and Coauthors, 2004: A new method for diagnosing radiative forcing and climate sensitivity. Geophys. Res. Lett., 31, L03205, https://doi.org/10.1029/2003GL018747.
Gregory, J. M., T. Andrews, and P. Good, 2015: The inconstancy of the transient climate response parameter under increasing CO2. Philos. Trans. Roy. Soc., A373, 20140417, https://doi.org/10.1098/rsta.2014.0417.
Grose, M. R., J. Gregory, R. Colman, and T. Andrews, 2018: What climate sensitivity index is most useful for projections? Geophys. Res. Lett., 45, 1559–1566, https://doi.org/10.1002/2017GL075742.
Gulev, S. K., and Coauthors, 2021: Changing state of the climate system. Climate Change 2021: The Physical Science Basis, V. Masson-Delmotte et al., Eds., Cambridge University Press, 287–422.
Hansen, J., and Coauthors, 2005: Efficacy of climate forcings. J. Geophys. Res., 110, D18104, https://doi.org/10.1029/2005JD005776.
Hausfather, Z., K. Cowtan, D. C. Clarke, P. Jacobs, M. Richardson, and R. Rohde, 2017: Assessing recent warming using instrumentally homogeneous sea surface temperature records. Sci. Adv., 3, e1601207, https://doi.org/10.1126/sciadv.1601207.
Hofmann, M., and M. A. Morales Maqueda, 2009: Geothermal heat flux and its influence on the oceanic abyssal circulation and radiocarbon distribution. Geophys. Res. Lett., 36, L03603, https://doi.org/10.1029/2008GL036078.
Holland, M. M., and C. M. Bitz, 2003: Polar amplification of climate change in coupled models. Climate Dyn., 21, 221–232, https://doi.org/10.1007/s00382-003-0332-6.
Hwang, Y.-T., D. M. W. Frierson, and J. E. Kay, 2011: Coupling between Arctic feedbacks and changes in poleward energy transport. Geophys. Res. Lett., 38, L17704, https://doi.org/10.1029/2011GL048546.
Johnson, B. T., J. M. Haywood, and M. K. Hawcroft, 2019: Are changes in atmospheric circulation important for black carbon aerosol impacts on clouds, precipitation, and radiation? J. Geophys. Res. Atmos., 124, 7930–7950, https://doi.org/10.1029/2019JD030568.
Junod, R. A., and J. R. Christy, 2020: A new compilation of globally gridded night‐time marine air temperatures: The UAHNMATv1 dataset. Int. J. Climatol., 40, 2609–2623, https://doi.org/10.1002/joc.6354.
Kamae, Y., M. Watanabe, T. Ogura, M. Yoshimori, and H. Shiogama, 2015: Rapid adjustments of cloud and hydrological cycle to increasing CO2: A review. Curr. Climate Change Rep., 1, 103–113, https://doi.org/10.1007/s40641-015-0007-5.
Kim, S., S. Park, and J. Shin, 2020: Impact of subgrid variation of water vapor on longwave radiation in a general circulation model. J. Adv. Model. Earth Syst., 12, e2019MS001926, https://doi.org/10.1029/2019MS001926.
Knutti, R., and M. A. A. Rugenstein, 2015: Feedbacks, climate sensitivity and the limits of linear models. Philos. Trans. Roy. Soc., A373, 20150146, https://doi.org/10.1098/rsta.2015.0146.
Kraemer, B. M., and Coauthors, 2015: Morphometry and average temperature affect lake stratification responses to climate change. Geophys. Res. Lett., 42, 4981–4988, https://doi.org/10.1002/2015GL064097.
Kramer, R. J., A. V. Matus, B. J. Soden, and T. S. L’Ecuyer, 2019: Observation‐based radiative kernels from CloudSat/CALIPSO. J. Geophys. Res. Atmos., 124, 5431–5444, https://doi.org/10.1029/2018JD029021.
L’Ecuyer, T. S., and Coauthors, 2021: The polar radiant energy in the far infrared experiment: A new perspective on polar longwave energy exchanges. Bull. Amer. Meteor. Soc., 102, E1431–E1449, https://doi.org/10.1175/BAMS-D-20-0155.1.
Liang, Y., N. P. Gillett, and A. H. Monahan, 2020: Climate model projections of 21st century global warming constrained using the observed warming trend. Geophys. Res. Lett., 47, e2019GL086757, https://doi.org/10.1029/2019GL086757.
Lorenz, D. J., E. T. DeWeaver, and D. J. Vimont, 2010: Evaporation change and global warming: The role of net radiation and relative humidity. J. Geophys. Res., 115, D20118, https://doi.org/10.1029/2010JD013949.
Manabe, S., and R. T. Wetherald, 1975: The effects of doubling the CO2 concentration on the climate of a general circulation model. J. Atmos. Sci., 32, 3–15, https://doi.org/10.1175/1520-0469(1975)032<0003:TEODTC>2.0.CO;2.
McPhaden, M. J., and Coauthors, 1998: The tropical ocean-global atmosphere observing system: A decade of progress. J. Geophys. Res., 103, 14 169–14 240, https://doi.org/10.1029/97JC02906.
Merchant, C. J., and Coauthors, 2014: Sea surface temperature datasets for climate applications from phase 1 of the European Space Agency climate change initiative (SST CCI). Geosci. Data J., 1, 179–191, https://doi.org/10.1002/gdj3.20.
Mitchell, D. M., Y. T. Eunice Lo, W. J. M. Seviour, L. Haimberger, and L. M. Polvani, 2020: The vertical profile of recent tropical temperature trends: Persistent model biases in the context of internal variability. Environ. Res. Lett, 15, 1040b4, https://doi.org/10.1088/1748-9326/ab9af7.
Mitevski, I., L. M. Polvani, and C. Orbe, 2022: Asymmetric warming/cooling response to CO2 increase/decrease mainly due to non‐logarithmic forcing, not feedbacks. Geophys. Res. Lett., 49, e2021GL097133, https://doi.org/10.1029/2021GL097133.
Mlynczak, M. G., and Coauthors, 2016: The spectroscopic foundation of radiative forcing of climate by carbon dioxide. Geophys. Res. Lett., 43, 5318–5325, https://doi.org/10.1002/2016GL068837.
Monin, A. S., and A. M. Obukhov, 1954: Basic laws of turbulent mixing in the surface layer of the atmosphere. Tr. Geophiz. Inst., Akad. Nauk. SSSR, 24, 163–187.
Morice, C. P., and Coauthors, 2021: An updated assessment of near‐surface temperature change from 1850: The HadCRUT5 data set. J. Geophys. Res. Atmos., 126, e2019JD032361, https://doi.org/10.1029/2019JD032361.
Newsom, E. R., A. F. Thompson, J. F. Adkins, and E. D. Galbraith, 2021: A hemispheric asymmetry in poleward ocean heat transport across climates: Implications for overturning and polar warming. Earth Planet. Sci. Lett., 568, 117033, https://doi.org/10.1016/j.epsl.2021.117033.
Obukhov, A. M., 1971: Turbulence in an atmosphere with a non-uniform temperature. Bound.-Layer Meteor., 2, 7–29, https://doi.org/10.1007/BF00718085.
Oleson, K. W., and Coauthors, 2010: Technical description of version 4.0 of the Community Land Model. NCAR Tech. Note NCAR/TN-478+STR, 257 pp., https://doi.org/10.5065/D6FB50WZ.
Palchetti, L., and Coauthors, 2020: FORUM: Unique far-infrared satellite observations to better understand how Earth radiates energy to space. Bull. Amer. Meteor. Soc., 101, E2030–E2046, https://doi.org/10.1175/BAMS-D-19-0322.1.
Pendergrass, A. G., 2020: The global‐mean precipitation response to CO2‐induced warming in CMIP6 models. Geophys. Res. Lett., 47, e2020GL089964, https://doi.org/10.1029/2020GL089964.
Pendergrass, A. G., A. Conley, and F. M. Vitt, 2018: Surface and top-of-atmosphere radiative feedback kernels for CESM-CAM5. Earth Syst. Sci. Data, 10, 317–324, https://doi.org/10.5194/essd-10-317-2018.
Pierrehumbert, R. T., 2002: The hydrologic cycle in deep-time climate problems. Nature, 419, 191–198, https://doi.org/10.1038/nature01088.
Pithan, F., and T. Mauritsen, 2014: Arctic amplification dominated by temperature feedbacks in contemporary climate models. Nat. Geosci., 7, 181–184, https://doi.org/10.1038/ngeo2071.
Räisänen, P., 1996: The effect of vertical resolution on clear-sky radiation calculations: Tests with two schemes. Tellus, 48A, 403–423, https://doi.org/10.3402/tellusa.v48i3.12068.
Richardson, M., K. Cowtan, E. Hawkins, and M. B. Stolpe, 2016: Reconciled climate response estimates from climate models and the energy budget of Earth. Nat. Climate Change, 6, 931–935, https://doi.org/10.1038/nclimate3066.
Richardson, M., K. Cowtan, and R. J. Millar, 2018: Global temperature definition affects achievement of long-term climate goals. Environ. Res. Lett., 13, 054004, https://doi.org/10.1088/1748-9326/aab305.
Richter, I., and S.-P. Xie, 2008: Muted precipitation increase in global warming simulations: A surface evaporation perspective. J. Geophys. Res., 113, D24118, https://doi.org/10.1029/2008JD010561.
Richter, I., S.-P. Xie, S. K. Behera, T. Doi, and Y. Masumoto, 2014: Equatorial Atlantic variability and its relation to mean state biases in CMIP5. Climate Dyn., 42, 171–188, https://doi.org/10.1007/s00382-012-1624-5.
Roderick, M. L., and G. D. Farquhar, 2002: The cause of decreased pan evaporation over the past 50 years. Science, 298, 1410–1411, https://doi.org/10.1126/science.1075390-a.
Rohde, R. A., and Z. Hausfather, 2020: The Berkeley Earth land/ocean temperature record. Earth Syst. Sci. Data, 12, 3469–3479, https://doi.org/10.5194/essd-12-3469-2020.
Rose, B. E. J., K. C. Armour, D. S. Battisti, N. Feldl, and D. D. B. Koll, 2014: The dependence of transient climate sensitivity and radiative feedbacks on the spatial pattern of ocean heat uptake. Geophys. Res. Lett., 41, 1071–1078, https://doi.org/10.1002/2013GL058955.
Rubino, A., D. Zanchettin, F. De Rovere, and M. J. McPhaden, 2020: On the interchangeability of sea-surface and near-surface air temperature anomalies in climatologies. Sci. Rep., 10, 7433, https://doi.org/10.1038/s41598-020-64167-1.
Rugenstein, M., and Coauthors, 2019: LongRunMIP: Motivation and design for a large collection of millennial-length AOGCM simulations. Bull. Amer. Meteor. Soc., 100, 2551–2570, https://doi.org/10.1175/BAMS-D-19-0068.1.
Rugenstein, M., and Coauthors, 2020: Equilibrium climate sensitivity estimated by equilibrating climate models. Geophys. Res. Lett., 47, e2019GL083898, https://doi.org/10.1029/2019GL083898.
Samset, B. H., and Coauthors, 2016: Fast and slow precipitation responses to individual climate forcers: A PDRMIP multimodel study. Geophys. Res. Lett., 43, 2782–2791, https://doi.org/10.1002/2016GL068064.
Santer, B. D., and Coauthors, 2017: Causes of differences in model and satellite tropospheric warming rates. Nat. Geosci., 10, 478–485, https://doi.org/10.1038/ngeo2973.
Schwarzwald, K., A. Poppick, M. Rugenstein, J. Bloch-Johnson, J. Wang, D. McInerney, and E. J. Moyer, 2021: Changes in future precipitation mean and variability across scales. J. Climate, 34, 2741–2758, https://doi.org/10.1175/JCLI-D-20-0001.1.
Servain, J., A. J. Busalacchi, M. J. McPhaden, A. D. Moura, G. Reverdin, M. Vianna, and S. E. Zebiak, 1998: A Pilot Research Moored Array in the Tropical Atlantic (PIRATA). Bull. Amer. Meteor. Soc., 79, 2019–2031, https://doi.org/10.1175/1520-0477(1998)079<2019:APRMAI>2.0.CO;2.
Sherwood, S. C., and Coauthors, 2020: An assessment of Earth’s climate sensitivity using multiple lines of evidence. Rev. Geophys., 58, e2019RG000678, https://doi.org/10.1029/2019RG000678.
Siler, N., G. H. Roe, K. C. Armour, and N. Feldl, 2019: Revisiting the surface-energy-flux perspective on the sensitivity of global precipitation to climate change. Climate Dyn., 52, 3983–3995, https://doi.org/10.1007/s00382-018-4359-0.
Simmons, A. J., K. M. Willett, P. D. Jones, P. W. Thorne, and D. P. Dee, 2010: Low-frequency variations in surface atmospheric humidity, temperature, and precipitation: Inferences from reanalyses and monthly gridded observational data sets. J. Geophys. Res., 115, D01110, https://doi.org/10.1029/2009JD012442.
Smith, C. J., 2018: HadGEM2 radiative kernels. University of Leeds, accessed 24 February 2022, https://doi.org/10.5518/406.
Smith, C. J., and Coauthors, 2018: Understanding rapid adjustments to diverse forcing agents. Geophys. Res. Lett., 45, 12 023–12 031, https://doi.org/10.1029/2018GL079826.
Smith, C. J., and Coauthors, 2020: Effective radiative forcing and adjustments in CMIP6 models. Atmos. Chem. Phys., 20, 9591–9618, https://doi.org/10.5194/acp-20-9591-2020.
Soden, B. J., I. M. Held, R. Colman, K. M. Shell, J. T. Kiehl, and C. A. Shields, 2008: Quantifying climate feedbacks using radiative kernels. J. Climate, 21, 3504–3520, https://doi.org/10.1175/2007JCLI2110.1.
Spencer, R. W., J. R. Christy, and W. D. Braswell, 2017: UAH version 6 global satellite temperature products: Methodology and results. Asia-Pac. J. Atmos. Sci., 53, 121–130, https://doi.org/10.1007/s13143-017-0010-y.
Stephens, G. L., B. H. Kahn, and M. Richardson, 2016: The super greenhouse effect in a changing climate. J. Climate, 29, 5469–5482, https://doi.org/10.1175/JCLI-D-15-0234.1.
Stetler, J. T., S. Girdner, J. Mack, L. A. Winslow, T. H. Leach, and K. C. Rose, 2021: Atmospheric stilling and warming air temperatures drive long‐term changes in lake stratification in a large oligotrophic lake. Limnol. Oceanogr., 66, 954–964, https://doi.org/10.1002/lno.11654.
Takahashi, K., 2009: Radiative constraints on the hydrological cycle in an idealized radiative–convective equilibrium model. J. Atmos. Sci., 66, 77–91, https://doi.org/10.1175/2008JAS2797.1.
Tjemkes, S. A., and Coauthors, 2003: The ISSWG line-by-line inter-comparison experiment. J. Quant. Spectrosc. Radiat. Transfer, 77, 433–453, https://doi.org/10.1016/S0022-4073(02)00174-7.
Vargas Zeppetello, L. R., A. Donohoe, and D. S. Battisti, 2019: Does surface temperature respond to or determine downwelling longwave radiation? Geophys. Res. Lett., 46, 2781–2789, https://doi.org/10.1029/2019GL082220.
Vose, R. S., and Coauthors, 2021: Implementing full spatial coverage in NOAA’s global temperature analysis. Geophys. Res. Lett., 48, e2020GL090873, https://doi.org/10.1029/2020GL090873.
Wang, C., L. Zhang, S.-K. Lee, L. Wu, and C. R. Mechoso, 2014: A global perspective on CMIP5 climate model biases. Nat. Climate Change, 4, 201–205, https://doi.org/10.1038/nclimate2118.
Wang, W., X. Lee, W. Xiao, S. Liu, N. Schultz, Y. Wang, M. Zhang, and L. Zhao, 2018: Global lake evaporation accelerated by changes in surface energy allocation in a warmer climate. Nat. Geosci., 11, 410–414, https://doi.org/10.1038/s41561-018-0114-8.
Watanabe, M., J.-L. Dufresne, Y. Kosaka, T. Mauritsen, and H. Tatebe, 2021: Enhanced warming constrained by past trends in equatorial Pacific sea surface temperature gradient. Nat. Climate Change, 11, 33–37, https://doi.org/10.1038/s41558-020-00933-3.
Xu, J., and Coauthors, 2022: Assessment of surface downward longwave radiation in CMIP6 with comparison to observations and CMIP5. Atmos. Res., 270, 106056, https://doi.org/10.1016/j.atmosres.2022.106056.
Xu, K.-M., Z. Li, A. Cheng, and Y. Hu, 2020: Changes in clouds and atmospheric circulation associated with rapid adjustment induced by increased atmospheric CO2: A multiscale modeling framework study. Climate Dyn., 55, 277–293, https://doi.org/10.1007/s00382-018-4401-2.
Yeh, S.-W., S.-Y. Song, R. P. Allan, S.-I. An, and J. Shin, 2021: Contrasting response of hydrological cycle over land and ocean to a changing CO2 pathway. npj Climate Atmos. Sci., 4