1. Introduction
Earth’s climate sensitivity is a crucial factor in understanding and predicting climate change. While uncertainty in climate sensitivity is dominated by cloud feedbacks, the magnitude of climate sensitivity is largely set by the clear-sky longwave feedback λLW. Early studies estimated λLW to be −2.2 to −2.3 W m−2 K−1 (Manabe and Wetherald 1967; Budyko 1969). These estimates were impressively close to the current best estimates from climate models and observations, which agree on a fairly narrow range for λLW of about −1.8 to −2.2 W m−2 K−1 (Andrews et al. 2012; Chung et al. 2010; Kluft et al. 2019; Zhang et al. 2020; Zelinka et al. 2020). By contrast, the recent Sherwood et al. (2020) assessment estimated the total cloud feedback to be both smaller in magnitude and less certain at +0.45 ± 0.33 W m−2 K−1.
The robustness of the clear-sky longwave feedback suggests that one should be able to understand and describe its governing physics in fairly simple form. A simple model for λLW would provide definitive support for the value of −2 W m−2 K−1 derived from observations and climate models. It would also allow us to understand the state dependence of λLW: at warm enough temperatures Earth’s atmosphere transitions to a runaway state, in which λLW becomes zero or even changes sign, but it is unclear how λLW varies between today’s value and the runaway limit. Similarly, there is a long-standing interest in using paleoclimate proxies to constrain present-day climate sensitivity (Tierney et al. 2020), but this effort suffers from uncertainty regarding the state dependence of climate feedbacks (Meraner et al. 2013; Bloch-Johnson et al. 2015). Finally, geographic variation in feedbacks and their importance for the so-called pattern effect is an ongoing topic of research (Armour et al. 2013; Andrews et al. 2015, 2018), but if λLW has state dependence then that dependence should also influence the spatial pattern of λLW. For example, if the global-mean λLW was different in past climates due to changes in the global-mean surface temperature, then present-day λLW should show regional variation due to Earth’s surface temperature pattern, suggesting a close link between state dependence and spatial dependence of λLW.
The match between λLW and the surface Planck feedback λsurf in Eq. (1) is not exact, however. Follow-up work found that λsurf only accounts for 50%–90% of λLW in different regions, with about 60% in the global mean (Raghuraman et al. 2019; Feng et al. 2023), implying a gap in the argument of Koll and Cronin (2018). Similarly, Seeley and Jeevanjee (2021) showed that in hot, high-CO2 climates λsurf becomes negligible yet λLW does not go to zero. As the surface warms the atmosphere is still able to increase its emission to space in spectral regions that are dominated by CO2. This emission mostly comes from the upper atmosphere, and gives rise to a spectral CO2 “radiator fin” feedback. The existence of a CO2 feedback means λLW must depend on CO2 concentration, and thus must have CO2 state dependence. Moreover, the CO2 feedback has to depend on the atmospheric lapse rate: if the atmosphere was isothermal with zero lapse rate, CO2’s forcing and feedback would both have to be zero, in line with previous work which tried to quantify the dependence of CO2 forcing on the lapse rate (Huang and Bani Shahabadi 2014; Dufresne et al. 2020), even if the details of the forcing mechanism are still disputed (Seeley 2018; Romps et al. 2022). So while the “surface-only” feedback picture from Koll and Cronin (2018) gives a reasonable first-order approximation to λLW, more terms are needed to describe λLW quantitatively.
In this study, we aim to derive a simple model of Earth’s feedback that can quantitatively capture the magnitude of λLW as well as its state dependence and regional variations. The model decomposes λLW into the surface Planck feedback (λsurf) plus three atmospheric terms: a CO2 band feedback (
Our model of λLW is based on spectroscopic thinking and hence represents a different perspective than the conventional decomposition which breaks the clear-sky longwave feedback into Planck, lapse rate, and water vapor feedbacks (e.g., Soden et al. 2008; Sherwood et al. 2020; Zelinka et al. 2020). The conventional decomposition has been an important tool for understanding λLW and for diagnosing the physics governing outgoing longwave radiation in climate models. However, it also obscures large cancellations between the atmospheric part of the Planck feedback, the lapse rate feedback, and the water vapor feedback (Held and Shell 2012; Koll and Cronin 2018; Jeevanjee et al. 2021a). By obscuring these cancellations, the conventional decomposition can give a false impression of the uncertainty of climate models. The same cancellations also make it difficult to understand the state dependence of λLW—Planck, lapse rate, and water vapor feedbacks all increase in a warmer climate, but it is far from obvious how these changes add up to affect λLW (Meraner et al. 2013). Building on previous discussions of spectral feedbacks (e.g., Huang et al. 2010; Huang et al. 2014; Koll and Cronin 2018; Pan and Huang 2018; Seeley and Jeevanjee 2021; Jeevanjee et al. 2021a; Kluft et al. 2021; Feng et al. 2023), our goal in this paper is to show that the issues that arise in the conventional decomposition can be resolved by viewing λLW in terms of its spectral components instead.
The layout of the rest of this paper is as follows. Section 2 discusses several preliminaries which are necessary for the main derivations: an idealized Clausius–Clapeyron relation, an analytic approximation for moist lapse rates, and idealized band models for H2O and CO2 spectroscopy. Section 3 lays out our spectral framework and introduces the emission-level approximation, our spectral decomposition of λLW, and a description of the numerical line-by-line calculations. Section 4 derives analytic expressions for Earth’s emission temperature in different parts of the spectrum, which are then used in section 5 to derive analytic feedbacks. Our expressions compare favorably against the state dependence of λLW from line-by-line calculations. Next, section 6 uses these results to understand the spatial pattern of Earth’s clear-sky longwave feedback. We generate global maps of Earth’s clear-sky longwave feedback using a radiative kernel and climate model data. We then show that our analytic expressions recover qualitatively similar feedback patterns, which implies that the spatial pattern of λLW can be largely understood using our analytic model. Breaking λLW up into surface versus atmospheric terms, we find that the surface dominates λLW in the global mean as well as in the dry subtropics, with a spatial pattern set by the pattern of atmospheric relative humidity, while atmospheric feedbacks become significant in the inner tropics, with spatial patterns that are set by regional lapse rate changes under warming. The manuscript closes in section 7 with conclusions and broader discussion of the results.
2. Preliminaries
Our goal is to derive the longwave feedback of a cloud-free vertical column. The column’s state can be specified using five parameters: Ts, γlr, RH,
a. Clausius–Clapeyron
Figure 1 compares the approximations in Eqs. (2) and (3) against a fit based on experimental data (Huang 2018). Considering that a typical tropical atmospheric column spans the vertical temperature range 200–300 K, the quasi-exponential approximation is very accurate, whereas our power-law approximation only matches to roughly a factor of 2. Nevertheless, as shown below, this accuracy is good enough to match numerical calculations.
b. Bulk moist lapse rate
Figure 2 (left) compares profiles based on Eq. (10) to moist adiabatic profiles. The moist adiabats are obtained by numerically integrating a generalized form of the moist adiabat which does not approximate water vapor as a dilute substance and thus remains valid at high temperatures (Ding and Pierrehumbert 2016). In all cases, the tropopause temperature is assumed to be fixed and equal to Tstrat = 200 K. The analytic profiles given by Eq. (10) produce a reasonable fit to the moist adiabats, though at surface temperatures below 340 K they produce slightly colder tropospheres. The tropopause pressure is accurately reproduced, as the analytic profiles always reach the tropopause at roughly the same point as the moist adiabats.
Figure 2 (right) compares the Ts dependence of γlr. First, the moist adiabatic T(p) profiles shown in Fig. 2 (left) are used to numerically compute γlr at individual levels of the troposphere. Because our bulk expression for γlr only depends on temperature, and not pressure, the moist adiabatic values of γlr are similarly shown at fixed temperature levels. Additionally, for each adiabatic T(p) profile we compute the average moist lapse rate using a mass-weighted mean,
c. H2O and CO2 spectroscopy
The third ingredient for our derivations is a model of H2O and CO2 spectroscopy. We follow previous studies and model the absorption cross sections of H2O and CO2 as log-linear band shapes. Despite the simplicity of these models, they are able to explain numerous features of Earth’s climate, including the logarithmic nature of CO2 forcing, the temperature dependence of Earth’s surface feedback, and the vertical structure of radiative cooling (Crisp et al. 1986; Pierrehumbert 2010; Wilson and Gea-Banacloche 2012; Koll and Cronin 2018; Jeevanjee and Fueglistaler 2020; Romps et al. 2022). Because we explore feedbacks over a wide range of temperatures, we additionally need to account for the H2O continuum. We do so by approximating the continuum as a gray absorber.
Our model of H2O spectroscopy has eight parameters: κrot, lrot, rot, κv–r, lv–r, νv–r, κcnt, a. We set νrot = 150 cm−1 and νv–r = 1500 cm−1, and fit the remaining parameters using the median-smoothed H2O cross sections shown in Fig. 3 across the wavenumber range 150 cm−1 ≤ ν ≤ 1500 cm−1. The results are sensitive to the smoothing procedure, that is whether one uses a geometric mean or a median. Because the average transmission across a spectral band tends to be dominated by the most optically thin frequencies (Pierrehumbert 2010), we use a median filter. To perform the fits we use the nonlinear least squares algorithm scipy.optimize.curve_fit, with a reference temperature of T0 = 300 K. We first fit the parameters κrot, lrot, κv–r, lv–r to H2O line opacities only, and then use these parameters to fit κcnt and a to H2O cross sections that include both line and continuum opacity. The resulting values are κrot = 165 m2 kg−1, lrot = 55 cm−1, κv–r = 15 m2 kg−1, lv–r = 38 cm−1, κcnt = 3 × 10−3 m2 kg−1, and a = 7, which broadly match the H2O fits previously reported in Jeevanjee and Fueglistaler (2020). Table 1 summarizes the thermodynamic and spectral parameters used in this paper.
List of parameters and, where applicable, assumed values.
Figure 3 compares the idealized band models with line-by-line absorption cross sections. Overall, the shape of the cross sections is captured fairly well. The median CO2 and H2O cross sections scale linearly with total pressure, as expected for pressure broadening. The increasing H2O absorption in response to warming around 1000 cm−1 is also qualitatively captured by our gray continuum model, even though the H2O continuum itself is actually not gray.
Figure 3 (right column) shows that the slopes of the CO2 and H2O bands flatten as temperature increases, with roughly constant opacity in the band centers but increasing opacity in the band wings. This behavior is not captured by our simple models. Physically, absorption band slopes can depend on temperature due to the shifting population of different molecular excitation states. For example, the wings of the 667 cm−1 CO2 band consist of multiple smaller bands that correspond to transitions between excited states of CO2 (so-called hot bands), while the center of the CO2 band is dominated by transitions to/from the ground state of CO2. As temperature rises more CO2 molecules leave the ground state and access excited states, which in turn preferentially increases the opacity in the wings of the CO2 band. To keep our parameterizations simple, however, we do not attempt to model the temperature dependence of the band slopes.
3. Spectral framework
a. The emission-level approximation
b. Spectral feedback decomposition
Our spectral decomposition complements the conventional feedback decomposition which splits λLW into Planck, lapse rate, and water vapor (or relative humidity) feedbacks. The surface feedback λsurf measures the OLR increase due to surface warming while keeping the atmosphere fixed. This term is identical to the surface contribution of the Planck feedback, or “surface kernel,” in the conventional decomposition (Soden et al. 2008). As for the atmospheric feedback, Eq. (16) shows that it depends on the total derivative of Trad, that is, on dTrad/dTs. The conventional decomposition can be interpreted as splitting the total derivative dTrad/dTs up into various partial derivatives (uniform warming versus lapse rate versus water vapor changes), while using a single, spectrally averaged Trad. In contrast, our decomposition splits the atmosphere’s feedback into three different bands, but still retains the total derivative dTrad/dTs in each band. In principle our decomposition could be split further to recover the conventional decomposition. That is, one could further decompose dTrad/dTs in each band into partial derivatives of Trad that correspond to vertically uniform warming, lapse rate warming, and water vapor changes—see Jeevanjee et al. (2021a) for more details. Here, however, we do not pursue this approach because our analytic expressions are general enough to predict Trad and the total derivative dTrad/dTs.
We use relative humidity as the state variable throughout this paper, so the analytic results are compatible with papers that argue for the use of relative humidity in feedback decompositions instead of specific humidity (Held and Shell 2012; Jeevanjee et al. 2021a). In the fixed-RH framework the conventional water vapor feedback is replaced by a relative humidity feedback, which measures the clear-sky feedback due to RH changes. It is worth noting that the RH feedback is small in individual climate models, and its multimodel mean is close to zero (Zelinka et al. 2020). In the derivations below we therefore treat RH as an external parameter whose value is assumed constant under surface warming.
c. Line-by-line calculations
To calculate spectral feedbacks numerically we use a 1D line-by-line model, PyRADS (Koll and Cronin 2018). The model’s radiative transfer includes HITRAN2016 CO2 and H2O absorption data as well as the H2O component of the MTCKD continuum version 3.2 (Mlawer et al. 2012; Gordon et al. 2017). Calculations cover the spectral range 0.1–2500 cm−1 with a resolution of Δν = 0.01 cm−1, while the vertical resolution is 50 points in log pressure. In general the angular distribution of longwave radiation
The 1D calculations assume the atmosphere’s temperature profile follows either a moist adiabat or a power-law temperature–pressure profile that is consistent with our bulk lapse rate approximation. In both cases the troposphere is capped by a tropopause at 200 K, while the overlying stratosphere is isothermal at the same temperature. Relative humidity in the troposphere is vertically uniform while the H2O mass fraction in the stratosphere is set equal to its value at the tropopause. CO2 is treated as uniformly mixed in the vertical and fixed with respect to surface temperature. Because we are considering a wide range of surface temperatures, across which the tropopause pressure varies substantially, we vary the vertical grid spacing in PyRADS: for each surface temperature, the model-top pressure is set to a slightly lower value than the estimated tropopause pressure based on our bulk lapse rate formulation, which ensures the model’s top is always in the stratosphere and the tropopause is well resolved.
Previous work has used various approaches to interpret line-by-line output. Seeley and Jeevanjee (2021) defined CO2 versus H2O bands based on the column-integrated, spectrally smoothed optical thickness of CO2 and H2O. However, the behavior of H2O differs strongly between the H2O bands and the H2O continuum, and it is difficult to distinguish these terms based on column-integrated optical thicknesses. For example, the H2O continuum might have a larger integrated optical thickness at some wavenumber than the H2O bands, but because continuum absorption decays more rapidly with altitude than band absorption [
4. Emission temperatures
a. CO2
b. Non-Simpsonian H2O
c. H2O continuum
d. Emission temperatures
e. Comparison against LBL calculations
Equations (26a)–(26c) predict how Earth’s emission temperature varies in response to changes in Ts,
Figure 4 demonstrates that the analytic results compare favorably against numerical calculations. Even though the analytic Trad shapes are idealized compared to the numerical calculations, the overall response of Trad to perturbations is captured well. First, increasing CO2 concentration lowers Trad around 667 cm−1, which corresponds to the wings of the CO2 band. This is simply a spectrally resolved view of how increasing CO2 acts as a radiative forcing (Jeevanjee et al. 2021b). Second, warming the surface while keeping all other parameters fixed has multiple effects. The main impact is to increase the emission temperature in the window region between ∼800 and 1200 cm−1. In addition, there are secondary impacts: surface warming also shrinks the width of the CO2 band and slightly increases the emission temperature in the H2O bands below 600 cm−1 and above 1300 cm−1 (this latter effect is hard to see in Fig. 4). The increased emission in the H2O bands shows that Simpson’s law in Eq. (29) is not exact, an effect that is captured by our analytic expressions. Third, reducing the lapse rate γlr preserves the width of the CO2 band, but it flattens the steepness of its slopes and increases the emission temperature in the center of the band. In the H2O bands, a smaller γlr while keeping Ts fixed also leads to a non-Simpsonian increase of the emission temperature in the H2O bands. In contrast to the H2O bands, the emission temperature of the H2O continuum around 1000 cm−1 decreases as γlr is reduced. As discussed above, this is because the atmospheric water path increases with a smaller γlr, which reduces Tcnt. The feedback of the H2O continuum therefore has the opposite sign as the H2O bands, in line with the analytic results. Finally, reducing the relative humidity increases Trad in all regions dominated by water vapor, both in the H2O bands below 600 cm−1 and above 1300 cm−1 and in the H2O continuum around 1000 cm−1, while the CO2 band is unaffected.
Overall, Fig. 4 underlines that comparatively simple physics is sufficient to explain the spectrally resolved response of Trad to different climate perturbations. To connect Fig. 4 back to the total clear-sky longwave feedback we only need to consider how these changes in Trad play out once we average them into spectral bands, and how multiple bands add up to determine the net longwave feedback.
5. Analytic feedbacks
Having derived expressions for the emission temperature in different parts of the LW spectrum, and verified these expressions against line-by-line calculations, we can now derive analytic expressions for the four spectral feedbacks: λsurf,
a. Bandwidths
We define an absorption band as the spectral range in which a given absorber has the coldest emission temperature compared to all other absorbers (this is equivalent to the highest-altitude emission level), and thus dominates the column’s emission to space. For example, the CO2 band is defined as all wavenumbers in which
1) CO2 bandwidth
To estimate the width of the CO2 band we consider three situations: 1) the CO2 concentration
2) H2O bandwidth
b. Surface feedback
c. H2O band feedback
d. H2O continuum feedback
e. CO2 band feedback
Next, we consider the CO2 feedback. Unlike the H2O band and continuum, however, the emission temperature of CO2 varies strongly with wavenumber, which makes it difficult to approximate the CO2 feedback integral via simple multiplication. Instead, we introduce an idealized CO2 “ditch” model, illustrated in Fig. 6. Our approach is closely related to the CO2 forcing models of Wilson and Gea-Banacloche (2012) and Jeevanjee et al. (2021b)—in appendix A we show that our ditch model can also be used to rederive the results of those previous studies, underlining the close relationship between forcing and feedbacks.
We approximate the CO2 band as symmetric around the central frequency ν0 = 667 cm−1. The center of the band emits πBν(Tcold) while outside the band the emission is πBν(Thot). Here Tcold and Thot are cold and hot emission temperatures, while νhot and νcold denote the edges of the CO2 ditch. At low and moderate surface temperatures the CO2 band center around 667 cm−1 radiates from the stratosphere, so Tcold is equal to the stratospheric temperature. However, this situation is no longer true at high surface temperatures. Physically, the tropopause rises as the surface warms, so if one warms the surface while holding CO2 concentration fixed (this is implicit in the definition of a climate feedback), parts of the CO2 band that were previously in the stratosphere have to start radiating from the troposphere. Eventually, even the CO2 band center radiates from the troposphere so the rectangular CO2 ditch turns into a triangular trough (see Fig. 6b). Here we leave our expressions general to allow for either situation.
Finally, when does the CO2 band center change from a stratospheric radiator at low Ts to a tropospheric radiator at high Ts, which also determines the transition between
f. Validation against LBL calculations
To test our analytic feedback expressions, we again use 1D calculations with PyRADS. One potential issue is that our derivations use the bulk lapse rate approximation, and so might differ from realistic feedbacks. Figure 7 compares feedbacks calculated with a moist adiabat to feedbacks with bulk lapse rate profiles. Overall, the bulk lapse rate approximation only introduces minor errors in λLW over the temperature range 250–320 K. We therefore consider the bulk lapse rate approximation sufficiently accurate below 320 K, while care should be taken when applying our analytic expressions to extremely hot climates. To better match the derivations, the PyRADS calculations here also use vertical profiles with a bulk lapse rate, so
To compare our analytic expressions against the 1D calculations we need to specify the scaling constants
Figure 8 shows that our analytic expressions successfully capture the basic state dependence of λLW as well as of its spectral constituents. The longwave feedback λLW is sensitive to changes in surface temperature, but it also varies in response to humidity and CO2 changes. Comparing the left and right columns in Fig. 8, λLW becomes larger with decreasing relative humidity (also see McKim et al. 2021). Comparing the top and bottom rows, adding CO2 to an atmosphere without any CO2 evens out the temperature dependence of λLW, by decreasing λLW at cold temperatures and increasing λLW at high temperatures. Importantly, the analytic expressions capture most of the variation in λLW, including its state dependence.
To understand the behavior of λLW we turn to the individual spectral feedbacks. The surface feedback λsurf is generally the dominant term in the spectral decomposition. Without CO2, λsurf makes up at least 90% of λLW below 300 K. The presence of CO2 decreases λsurf but even in this case λsurf makes up at least 60% of λLW below 300 K. Our analytic expressions thus agree with previous studies which showed that Earth’s longwave feedback is dominated by the surface feedback (Koll and Cronin 2018; Raghuraman et al. 2019). This situation changes at high temperatures, however, once the surface window closes, at which point λLW becomes dominated by atmospheric feedbacks.
In line with section 4, the CO2 band feedback acts to stabilize Earth’s climate in warm climates, and its importance increases with surface temperature. Below 300 K,
Finally, again in line with our analytic results, the two water vapor feedbacks
6. The spatial pattern of λLW
In the previous two sections we demonstrated that the analytic expressions summarized in Table 2 accurately capture the behavior of Earth’s emission temperature Trad as well as the state dependence of λLW. These feedback expressions can be interpreted as either a model for the global-mean feedback or as a model for the local feedback of an isolated atmospheric column, so the state dependence of λLW shown in Fig. 8 should also appear as a spatial dependence in Earth’s clear-sky longwave feedback.
Summary of main theoretical results.
In this section we therefore analyze the spatial pattern of λLW for Earth’s present-day climate. First, we generate a map of λLW using the radiative kernel technique (Soden et al. 2008). Next, we generate a map of λLW using our analytic expressions. The radiative kernel technique cannot be used to determine the feedback contributions of individual gases and our analytic expressions only account for the feedback from Earth’s dominant greenhouse gases, H2O and CO2, whereas the radiative kernel includes additional greenhouse gases such as O3 and CH4. We therefore split λLW into only two terms, namely, the surface feedback λsurf and the atmospheric feedback λatm = λLW − λsurf. Despite the idealizations in our analytic approach compared to a full radiative kernel, we find that the resulting feedback maps are in qualitative agreement. This allows us to attribute the spatial pattern of λLW, as deduced from the radiative kernel, to geographic variations in the inputs of our analytic model.
a. Inputs for feedback maps
For the kernel calculation, we use the HadGEM2 radiative kernel. For consistency with the analytic model (which assumes the stratosphere is isothermal and at a fixed temperature), we set the kernel to zero in the stratosphere. The tropopause is defined as in Soden et al. (2008): the tropopause pressure ptp increases linearly with latitude, from 0.1 bar at the equator to 0.3 bar at the poles. The analytic model also assumes RH stays fixed under surface warming, so we do not include RH changes in the kernel calculation. Doing so is justified because the RH feedback only makes a minor contribution to λLW in individual climate models, and it moreover tends to cancel in the multimodel mean (Zelinka et al. 2020). To compute the forced response, we use HadGEM2 climatologies from the CMIP5 archive for a preindustrial control simulation and an abrupt-4xCO2 simulation, where the climatologies are 50-yr averages (for 4xCO2, years 100–150 after increasing CO2). Multiplying the kernel with the forced response gives a map of the change in top-of-atmosphere (TOA) radiation (Soden et al. 2008). To compute a feedback, one additionally needs to normalize the change in TOA radiation by a change in surface temperature. Consistent with our assumption of an isolated atmospheric column we compute local–local feedback maps, that is, we divide the local change in OLR deduced from the kernel by the local change in surface temperature (Feldl and Roe 2013; Armour et al. 2013; Bloch-Johnson et al. 2020). To distinguish between surface and atmospheric feedbacks in the kernel method we compute the clear-sky longwave feedback λLW and the surface feedback λsurf, where the second is equal to the surface kernel; the atmospheric feedback is then computed as the residual λatm = λLW − λsurf.
Figure 9 shows maps of the input data from HadGEM2, which we use below to evaluate the analytic expressions. In the top two rows, large variations are notable in the maps of surface temperature Ts, column relative humidity RH, and bulk lapse rate γlr. In contrast, apart from minor stationary wave patterns in the northern midlatitudes, the stratospheric temperature Tstrat is zonally fairly uniform and varies by only about 20 K between the equator and poles. The bottom row shows the normalized bulk lapse change, dln(γlr)/dTs = 1/(Δγlr) × (Δγlr/ΔTs), computed using the bulk lapse rate difference Δγlr between 4xCO2 and preindustrial simulations. The bulk lapse rate change shows an equator–pole contrast, with a decrease in γlr at low and midlatitudes and an increase in γlr at high latitudes. This contrast is in line with previous studies—for a moist adiabat the atmospheric temperature–pressure profile becomes less steep under warming, so γlr decreases in the tropics, while the opposite occurs at high latitudes (e.g., Payne et al. 2015; Cronin and Jansen 2016; Stuecker et al. 2018). There is also a noticeable tropical land–ocean contrast in the bulk lapse rate change, with tropical land areas showing near-zero lapse rate change. This is likely due to compensation between moist-adiabatic warming aloft, which is uniform across the tropics and tends to decrease γlr, and amplified land surface warming, which increases γlr (Byrne and O’Gorman 2013). Conversely, subtropical eastern ocean basins have the same moist adiabatic warming aloft but suppressed surface warming, both of which contribute to strong decreases in γlr.
b. Feedback maps
Figure 10 shows the feedback maps resulting from kernel and analytic calculations. Overall, we find good qualitative agreement between kernel-derived feedbacks and our analytic approximations. The global pattern of λLW in both maps shows clear contrasts between the high latitudes, subtropics, and inner tropics (Fig. 10, top row). The value of λLW is smallest in the inner tropics, especially in the intertropical convergence zone (ITCZ), while it is largest in the subtropics, especially over eastern ocean basins. The agreement is less good at small scales, with the analytic map of λLW showing less regional structure and deviating from the kernel-derived map in continental interiors and over the Southern Ocean. This is plausible given the idealizations in our derivations, such as representing realistic vertical temperature profiles by a smooth power law. However, small-scale differences tend to cancel when taking a zonal or global mean. The zonal mean of λLW in our analytic estimate agrees with the zonal mean of the kernel λLW to within 11% at each latitude. The global-mean values of λLW are almost identical, with −2.15 W m−2 K−1 for the kernel calculation and −2.16 W m−2 K−1 for the analytic estimate. Note that these global mean averages are weighted by the HadGEM2 pattern of surface warming, which is required to convert a local–local feedback map into a global mean (Feldl and Roe 2013; Armour et al. 2013).
The qualitative agreement between the λLW maps also holds separately for surface and atmospheric feedbacks, though differences are larger here. The kernel-derived map of λsurf is almost uniform at high latitudes, large in magnitude over subtropical desert regions, and small in magnitude over the ITCZ. The analytic map of λsurf qualitatively matches this pattern, though it overpredicts the magnitude of λsurf in the global mean by 0.18 W m−2 K−1, or 13%. Conversely, the analytic estimate underpredicts λatm relative to the kernel-derived map in the global mean by 0.17 W m−2 K−1, or 22%. In addition, the analytic λatm map predicts that the atmospheric feedback goes almost to zero at the poles, whereas the kernel-derived λatm map shows a small but clearly nonzero feedback. The strong differences at the poles again presumably arise because our derivations fail to capture the atmospheric feedback response in areas with inversions and other complex temperature–pressure profiles.
In addition to an overall spatial agreement, both kernel and analytic feedback calculations agree that the surface dominates the net longwave feedback. Figure 11 shows that the surface’s contribution to the total feedback is about 50% at low latitudes and increases toward the poles, reaching about 75% in the kernel maps and over 90% in the analytic maps. One plausible reason why the analytic maps tend to overestimate λsurf/λLW at high latitudes is that our expressions do not include minor greenhouse gases such as ozone or methane. Any additional atmospheric absorption from such gases reduces the window width via
c. What controls the large-scale pattern of λLW?
The match between our analytic model and the kernel calculation implies that one can explain much of the spatial structure of λLW in terms of the analytic model’s input parameters. We do this by calculating correlations between λsurf and λatm from the kernel-derived feedback maps against the analytic model’s five main inputs: surface temperature Ts, column relative humidity RH, stratospheric temperature Tstrat, bulk lapse rate γlr, and the change in bulk lapse rate under warming dγlr/dTs. Spatial CO2 contrasts are small (e.g., Fraser et al. 1983), and so do not need to be considered here.
Figure 12 shows the resulting spatial correlations between the kernel-derived feedback maps (left column of Fig. 10) and the five inputs from HadGEM2 (Fig. 9). Because the feedback maps differ strongly between tropics and extratropics in terms of zonal variation and magnitude, we compute correlations separately in these two regions (data are split based on being equatorward or poleward of 30° latitude). Based on the inherent correlations between the five input maps, we consider a correlation significant if its coefficient exceeds |r| ≥ 0.75 (the largest intrainput correlations are r = −0.71 between Ts and Tstrat in the tropics, and r = −0.76 between Ts and dγlr/dTs in the extratropics; not shown).
In line with our analytic model, we find that the kernel-derived λsurf is strongly correlated with column RH in the tropics (r = 0.83), while it does not show strong correlation with any inputs in the extratropics (|r| < 0.4). This underlines the importance of the subtropical dry radiator fin regions for λsurf, which are clearly visible as the dark blue regions in Fig. 9 (top right) and the yellow regions in Fig. 10 (center left). As expected, the sign of the correlation is positive which means λsurf becomes less negative, or less stabilizing, as column RH increases.
Next, we find λatm is most strongly correlated with γlr and dγlr/dTs in the tropics (r = 0.75 for both), and with dγlr/dTs in the extratropics (r = 0.88). Of the two parameters that show strong correlations with λatm in the tropics, γlr and dγlr/dTs, which one is more important? We performed a test with the analytical model in which we set dγlr/dTs = 0 (not shown). Doing so eliminates most tropical structure in the map of λatm, which indicates that λatm is largely determined by dγlr/dTs, not γlr. The correlation between λatm and dγlr/dTs is positive, which is intuitive: λatm becomes more negative if the upper atmosphere warms more relative to the surface, i.e., if γlr decreases. The spatial variability of λatm is largest in the tropics, and can be can be understood in terms of the map of dγlr/dTs already discussed in section 6a: tropical λatm is large over subtropical eastern ocean basins due to suppressed surface warming, and small over land due to enhanced surface warming, where these warming patterns are relative to the approximately uniform warming of the tropical free troposphere (Byrne and O’Gorman 2013).
The correlations shown in Fig. 12 are between fields derived from two independent methods, and so are nontrivial. Appendix C shows that the same analysis performed with λsurf and λatm from our analytic feedback maps identifies the same dominant relations (e.g., λsurf is most strongly correlated with column RH in the tropics), though most correlation coefficients are unsurprisingly even larger (e.g., r = 0.93 for the analytic λsurf and tropical column RH). Our results thus underline that the spatial pattern of λLW can be understood, at least in rough terms and on large spatial scales, by Earth’s spatial pattern of relative humidity and lapse rate changes. Relative humidity and lapse rate changes dominate the pattern of λLW in the tropics, where they control λsurf and λatm, respectively, while lapse rate changes dominate the pattern of λatm in the extratropics.
7. Discussion and conclusions
In this paper we have presented a novel decomposition of Earth’s clear-sky longwave feedback λLW into four spectral components, namely, a surface Planck feedback (λsurf) and three atmospheric feedbacks: a CO2 band feedback (
The picture of Earth’s clear-sky longwave feedback that emerges from this perspective is relatively simple, consisting of a surface feedback plus atmospheric feedbacks from CO2 and H2O. At present the surface feedback λsurf is the most important contributor in the global mean and at most latitudes, with its spatial pattern determined by the distribution of atmospheric water vapor. λsurf is largest in the dry subtropics, consistent with the view that these are the locus of Earth’s stabilizing longwave feedback (Pierrehumbert 1995; McKim et al. 2021), and smallest in the inner tropics, where the surface’s emission is blocked by the H2O continuum. The atmospheric feedbacks from the CO2 and H2O bands play a supporting role to λsurf at mid- and high latitudes, but they rival the surface feedback in the inner tropics, with the global pattern of λatm largely determined by the pattern of the atmospheric lapse rate change dγlr/dTs. The H2O continuum provides a negligible feedback below ∼310 K (see section 5), but the continuum itself is still important through its influence on λsurf.
This spectral picture is arguably a more intuitive starting point for reasoning about different climates than the conventional decomposition of λLW into Planck, lapse rate, and water vapor feedbacks. As discussed by Cronin and Dutta (2023), it is nontrivial to accurately estimate the supposedly simple Planck feedback from first principles. Similarly, one can qualitatively reason that lapse rate and water vapor feedbacks both increase in magnitude under global warming, but these are large and of opposite sign, so it is difficult to predict their net change and, by extension, the Ts dependence of λLW, in the conventional decomposition without resorting to numerical models. The strong cancellations between Planck, lapse rate, and water vapor feedbacks can be alleviated by considering conventional feedbacks in a fixed relative humidity framework (Ingram 2010; Held and Shell 2012), but this comes at the cost that the state dependence of the Planck feedback is no longer trivial to understand at fixed RH.
In contrast, the state dependence of λLW is fairly straightforward to understand from a spectral perspective, at least in broad brushstrokes. For present-day Earth the Ts dependence of λLW is dominated by the surface in most regions. If relative humidity is fixed, λsurf increases at very cold temperatures, peaks around 260–290 K depending on RH, and then decreases again (see section 5). The decrease is rapid at high RH due to the H2O continuum, but much slower at low RH. Atmospheric feedbacks also have state dependence. All of them increase in magnitude as the atmosphere warms, and are further amplified by a weakening lapse rate. In the tropics the state dependence of λLW is thus set by the interplay between a decreasing surface feedback and increasing atmospheric feedbacks. This can lead to surprising dynamics—at high RH, λsurf decreases in magnitude more rapidly with warming than the atmospheric feedbacks from
The state dependence of λLW at temperatures far above ∼310 K is beyond the scope of this paper, but a spectral perspective points to the importance of stabilizing H2O and CO2 bands versus the destabilizing H2O continuum as Earth approaches the runaway greenhouse. The main caveat here is that Earth’s net feedback does not necessarily stay dominated by λLW at very high surface temperatures, and atmospheric feedbacks are also complicated at high temperatures by effects such as nondilute thermodynamics and surface pressure changes (Goldblatt et al. 2013; Ramirez et al. 2014).
There are several remaining shortcomings in our analysis of λLW that are beyond the scope of this paper. A major one is our assumption that the atmosphere can be described by a single bulk lapse rate, such that temperature has to monotonically decrease with altitude. In the real world inversions are common, particularly in polar regions and over subtropical oceans. Comparable to the long-standing discussion about how to interpret the lapse rate feedback at high latitudes in the conventional decomposition (e.g., Cai and Lu 2009; Payne et al. 2015; Stuecker et al. 2018; Boeke et al. 2021; Henry et al. 2021), we therefore expect that our approach here only provides a first step toward understanding the processes which shape λLW in inversion regions.
Another assumption is that we ignore stratospheric changes, even though stratospheric cooling induced by rising CO2 levels is a major and robust signal of anthropogenic warming (e.g., Vallis et al. 2014). It is notable that the radiative changes due to stratospheric cooling are also hard to intuitively explain using conventional feedbacks. Climate model analyses typically treat the stratosphere’s fast radiative adjustment to CO2 changes as distinct from Planck, lapse rate, and water vapor feedbacks. Our derivations here sidestep this issue and treat Tstrat as a fixed parameter. Similarly, our derivations ignore the potential feedback from relative humidity changes. In reality there is no guarantee that relative humidity will remain constant under global warming, let alone would have been similar in past climates. In principle our analysis starting from the emission level approximation can be extended to estimate the feedbacks associated with changes in either RH or Tstrat; RH changes would lead to a feedback term proportional to ∂Trad/∂RH, while stratospheric changes would lead to a feedback term proportional to ∂Trad/∂Tstrat.
Acknowledgments.
D.D.B.K. thanks Jeevanjee Gardens in Nairobi. N.J.L. was supported by the NOAA Climate Program Office’s Modeling, Analysis, Predictions, and Projections program through Grant NA20OAR4310387.
Data availability statement.
HadGEM2 GCM data are publicly available in CMIP data archives. The HadGEM2 radiative kernel is available at https://archive.researchdata.leeds.ac.uk/382. Scripts to compute analytical feedbacks are available at https://github.com/danielkoll/spectral_feedbacks.
APPENDIX A
CO2 Forcing
APPENDIX B
Transition from Stratospheric to Tropospheric CO2 Radiator Fin
At high surface temperatures the CO2 band center transitions from mainly radiating from the stratosphere to mainly radiating from the troposphere. Figure B1 shows smoothed brightness temperatures Tb computed from the 1D line-by-line calculations described in section 5, with a CO2 volume-mixing ratio of 400 ppm. In the middle of the CO2 band, at about 667 cm−1, CO2 radiates from the troposphere at surface temperatures above ∼310 K. In rough agreement with the line-by-line results, our analytic CO2 brightness temperatures predict this transition happens at a surface temperature of ∼320 K (dashed lines in Fig. B1). In practice we therefore use a transition temperature of Ts,0 = 310 K for 400 ppm of CO2 to determine when CO2 changes from a stratospheric to a tropospheric radiator.
APPENDIX C
Spatial Correlations in Analytic Feedback Maps
Figure C1 repeats the same analysis as in Fig. 12, but using the analytic feedback maps of λsurf and λatm. Given that the analytic model is computed using the input fields from Fig. 9, it is not surprising that most correlations between inputs and feedback maps are even higher than in Fig. 12. With the exception of λatm in the tropics, for which the correlation between the analytic λatm and γlr is slightly lower than between kernel-derived λatm and γlr, Fig. C1 identifies the same strong correlations as Fig. 12.
REFERENCES
Andrews, T., J. M. Gregory, M. J. Webb, and K. E. Taylor, 2012: Forcing, feedbacks and climate sensitivity in CMIP5 coupled atmosphere-ocean climate models. Geophys. Res. Lett., 39, L09712, https://doi.org/10.1029/2012GL051607.
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., and Coauthors, 2018: Accounting for changing temperature patterns increases historical estimates of climate sensitivity. Geophys. Res. Lett., 45, 8490–8499, https://doi.org/10.1029/2018GL078887.
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.
Bloch-Johnson, J., R. T. Pierrehumbert, and D. S. Abbot, 2015: Feedback temperature dependence determines the risk of high warming. Geophys. Res. Lett., 42, 4973–4980, https://doi.org/10.1002/2015GL064240.
Bloch-Johnson, J., M. Rugenstein, and D. S. Abbot, 2020: Spatial radiative feedbacks from internal variability using multiple regression. J. Climate, 33, 4121–4140, https://doi.org/10.1175/JCLI-D-19-0396.1.
Boeke, R. C., P. C. Taylor, and S. A. Sejas, 2021: On the nature of the Arctic’s positive lapse-rate feedback. Geophys. Res. Lett., 48, e2020GL091109, https://doi.org/10.1029/2020GL091109.
Bretherton, C. S., P. N. Blossey, and M. Khairoutdinov, 2005: An energy-balance analysis of deep convective self-aggregation above uniform SST. J. Atmos. Sci., 62, 4273–4292, https://doi.org/10.1175/JAS3614.1.
Budyko, M. I., 1969: The effect of solar radiation variations on the climate of the Earth. Tellus, 21, 611–619, https://doi.org/10.1111/j.2153-3490.1969.tb00466.x.
Byrne, M. P., and P. A. O’Gorman, 2013: Land–ocean warming contrast over a wide range of climates: Convective quasi-equilibrium theory and idealized simulations. J. Climate, 26, 4000–4016, https://doi.org/10.1175/JCLI-D-12-00262.1.
Cai, M., and J. Lu, 2009: A new framework for isolating individual feedback processes in coupled general circulation climate models. Part II: Method demonstrations and comparisons. Climate Dyn., 32, 887–900, https://doi.org/10.1007/s00382-008-0424-4.
Chung, E.-S., D. Yeomans, and B. J. Soden, 2010: An assessment of climate feedback processes using satellite observations of clear-sky OLR. Geophys. Res. Lett., 37, L02702, https://doi.org/10.1029/2009GL041889.
Crisp, D., S. B. Fels, and M. D. Schwarzkopf, 1986: Approximate methods for finding CO2 15-μm band transmission in planetary atmospheres. J. Geophys. Res., 91, 11 851–11 866, https://doi.org/10.1029/JD091iD11p11851.
Cronin, T. W., and M. F. Jansen, 2016: Analytic radiative-advective equilibrium as a model for high-latitude climate. Geophys. Res. Lett., 43, 449–457, https://doi.org/10.1002/2015GL067172.
Cronin, T. W., and I. Dutta, 2023: How well do we understand the Planck feedback? J. Adv. Model. Earth Syst., 15, e2023MS003729, https://doi.org/10.1029/2023MS003729.
Ding, F., and R. T. Pierrehumbert, 2016: Convection in condensible-rich atmospheres. Astrophys. J., 822, 24, https://doi.org/10.3847/0004-637X/822/1/24.
Dufresne, J.-L., V. Eymet, C. Crevoisier, and J.-Y. Grandpeix, 2020: Greenhouse effect: The relative contributions of emission height and total absorption. J. Climate, 33, 3827–3844, https://doi.org/10.1175/JCLI-D-19-0193.1.
Elsasser, W. M., 1942: Heat Transfer by Infrared Radiation in the Atmosphere. Harv. Meteor. Stud. Monogr., No. 6, 106 pp.
Feldl, N., and G. H. Roe, 2013: Four perspectives on climate feedbacks. Geophys. Res. Lett., 40, 4007–4011, https://doi.org/10.1002/grl.50711.
Feng, J., and Y. Huang, 2019: Diffusivity-factor approximation for spectral outgoing longwave radiation. J. Atmos. Sci., 76, 2171–2180, https://doi.org/10.1175/JAS-D-18-0246.1.
Feng, J., D. Paynter, and R. Menzel, 2023: How a stable greenhouse effect on Earth is maintained under global warming. J. Geophys. Res. Atmos., 128, e2022JD038124, https://doi.org/10.1029/2022JD038124.
Fraser, P. J., G. I. Pearman, and P. Hyson, 1983: The global distribution of atmospheric carbon dioxide: 2. A review of provisional background observations, 1978–1980. J. Geophys. Res., 88, 3591–3598, https://doi.org/10.1029/JC088iC06p03591.
Goldblatt, C., T. D. Robinson, K. J. Zahnle, and D. Crisp, 2013: Low simulated radiation limit for runaway greenhouse climates. Nat. Geosci., 6, 661–667, https://doi.org/10.1038/ngeo1892.
Gordon, I. E., and Coauthors, 2017: The HITRAN2016 molecular spectroscopic database. J. Quant. Spectrosc. Radiat. Transfer, 203, 3–69, https://doi.org/10.1016/j.jqsrt.2017.06.038.
Hartmann, D. L., and K. Larson, 2002: An important constraint on tropical cloud–climate feedback. Geophys. Res. Lett., 29, 1951, https://doi.org/10.1029/2002GL015835.
Held, I. M., and K. M. Shell, 2012: Using relative humidity as a state variable in climate feedback analysis. J. Climate, 25, 2578–2582, https://doi.org/10.1175/JCLI-D-11-00721.1.
Henry, M., T. M. Merlis, N. J. Lutsko, and B. E. J. Rose, 2021: Decomposing the drivers of polar amplification with a single-column model. J. Climate, 34, 2355–2365, https://doi.org/10.1175/JCLI-D-20-0178.1.
Huang, J., 2018: A simple accurate formula for calculating saturation vapor pressure of water and ice. J. Appl. Meteor. Climatol., 57, 1265–1272, https://doi.org/10.1175/JAMC-D-17-0334.1.
Huang, X., X. Chen, B. J. Soden, and X. Liu, 2014: The spectral dimension of longwave feedback in the CMIP3 and CMIP5 experiments. Geophys. Res. Lett., 41, 7830–7837, https://doi.org/10.1002/2014GL061938.
Huang, Y., and M. Bani Shahabadi, 2014: Why logarithmic? A note on the dependence of radiative forcing on gas concentration. J. Geophys. Res. Atmos., 119, 13 683–13 689, https://doi.org/10.1002/2014JD022466.
Huang, Y., S. Leroy, P. J. Gero, J. Dykema, and J. Anderson, 2010: Separation of longwave climate feedbacks from spectral observations. J. Geophys. Res., 115, D07104, https://doi.org/10.1029/2009JD012766.
Ingram, W., 2010: A very simple model for the water vapour feedback on climate change. Quart. J. Roy. Meteor. Soc., 136, 30–40, https://doi.org/10.1002/qj.546.
Jeevanjee, N., and S. Fueglistaler, 2020: Simple spectral models for atmospheric radiative cooling. J. Atmos. Sci., 77, 479–497, https://doi.org/10.1175/JAS-D-18-0347.1.
Jeevanjee, N., D. D. B. Koll, and N. Lutsko, 2021a: “Simpson’s law” and the spectral cancellation of climate feedbacks. Geophys. Res. Lett., 48, e2021GL093699, https://doi.org/10.1029/2021GL093699.
Jeevanjee, N., J. T. Seeley, D. Paynter, and S. Fueglistaler, 2021b: An analytical model for spatially varying clear-sky CO2 forcing. J. Climate, 34, 9463–9480, https://doi.org/10.1175/JCLI-D-19-0756.1.
Kluft, L., S. Dacie, S. A. Buehler, H. Schmidt, and B. Stevens, 2019: Re-examining the first climate models: Climate sensitivity of a modern radiative–convective equilibrium model. J. Climate, 32, 8111–8125, https://doi.org/10.1175/JCLI-D-18-0774.1.
Kluft, L., S. Dacie, M. Brath, S. A. Buehler, and B. Stevens, 2021: Temperature-dependence of the clear-sky feedback in radiative-convective equilibrium. Geophys. Res. Lett., 48, e2021GL094649, https://doi.org/10.1029/2021GL094649.
Koll, D. D. B., and T. W. Cronin, 2018: Earth’s outgoing longwave radiation linear due to H2O greenhouse effect. Proc. Natl. Acad. Sci. USA, 115, 10 293–10 298, https://doi.org/10.1073/pnas.1809868115.
Koll, D. D. B., and T. W. Cronin, 2019: Hot hydrogen climates near the inner edge of the habitable zone. Astrophys. J., 881, 120, https://doi.org/10.3847/1538-4357/ab30c4.
Li, J., 2000: Gaussian quadrature and its application to infrared radiation. J. Atmos. Sci., 57, 753–765, https://doi.org/10.1175/1520-0469(2000)057<0753:GQAIAT>2.0.CO;2.
Manabe, S., and R. T. Wetherald, 1967: Thermal equilibrium of the atmosphere with a given distribution of relative humidity. J. Atmos. Sci., 24, 241–259, https://doi.org/10.1175/1520-0469(1967)024<0241:TEOTAW>2.0.CO;2.
McKim, B. A., N. Jeevanjee, and G. K. Vallis, 2021: Joint dependence of longwave feedback on surface temperature and relative humidity. Geophys. Res. Lett., 48, e2021GL094074, https://doi.org/10.1029/2021GL094074.
Meraner, K., T. Mauritsen, and A. Voigt, 2013: Robust increase in equilibrium climate sensitivity under global warming. Geophys. Res. Lett., 40, 5944–5948, https://doi.org/10.1002/2013GL058118.
Mlawer, E. J., V. H. Payne, J.-L. Moncet, J. S. Delamere, M. J. Alvarado, and D. C. Tobin, 2012: Development and recent evaluation of the MT_CKD model of continuum absorption. Philos. Trans. Roy. Soc. A370, 2520–2556, https://doi.org/10.1098/rsta.2011.0295.
Pan, F., and X. Huang, 2018: The spectral dimension of modeled relative humidity feedbacks in the CMIP5 experiments. J. Climate, 31, 10 021–10 038, https://doi.org/10.1175/JCLI-D-17-0491.1.
Payne, A. E., M. F. Jansen, and T. W. Cronin, 2015: Conceptual model analysis of the influence of temperature feedbacks on polar amplification. Geophys. Res. Lett., 42, 9561–9570, https://doi.org/10.1002/2015GL065889.
Pierrehumbert, R. T., 1995: Thermostats, radiator fins, and the local runaway greenhouse. J. Atmos. Sci., 52, 1784–1806, https://doi.org/10.1175/1520-0469(1995)052<1784:TRFATL>2.0.CO;2.
Pierrehumbert, R. T., 2010: Principles of Planetary Climate. Cambridge University Press, 652 pp.
Raghuraman, S. P., D. Paynter, and V. Ramaswamy, 2019: Quantifying the drivers of the clear sky greenhouse effect, 2000–2016. J. Geophys. Res. Atmos., 124, 11 354–11 371, https://doi.org/10.1029/2019JD031017.
Ramirez, R. M., R. K. Kopparapu, V. Lindner, and J. F. Kasting, 2014: Can increased atmospheric CO2 levels trigger a runaway greenhouse? Astrobiology, 14, 714–731, https://doi.org/10.1089/ast.2014.1153.
Romps, D. M., 2016: Clausius–Clapeyron scaling of CAPE from analytical solutions to RCE. J. Atmos. Sci., 73, 3719–3737, https://doi.org/10.1175/JAS-D-15-0327.1.
Romps, D. M., J. T. Seeley, and J. P. Edman, 2022: Why the forcing from carbon dioxide scales as the logarithm of its concentration. J. Climate, 35, 4027–4047, https://doi.org/10.1175/JCLI-D-21-0275.1.