Vapor-Buoyancy Feedback in an Idealized GCM

Seth Seidel aNASA Goddard Space Flight Center, Greenbelt, Maryland
bDepartment of Land, Air, and Water Resources, University of California, Davis, Davis, California

Search for other papers by Seth Seidel in
Current site
Google Scholar
PubMed
Close
https://orcid.org/0000-0001-6002-9575
and
Da Yang cDepartment of the Geophysical Sciences, University of Chicago, Chicago, Illinois

Search for other papers by Da Yang in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

Humid air is lighter than dry air at a given temperature and pressure, as the molecular weight of water vapor is less than that of dry air. This effect is known as vapor buoyancy (VB). Although VB is a straightforward consequence of the ideal gas law, its influence on climate has been understudied. This study investigates VB’s influence on atmospheric temperature, radiation, and clouds. In mechanism-denial experiments, we remove VB from the dynamics and parameterizations of an idealized general circulation model (GCM). These experiments show that VB warms the tropical free troposphere by approximately 1 K in an Earth-like climate, and the magnitude of this effect increases with warming. This additional atmospheric warming causes greater outgoing longwave radiation (OLR) to be emitted as the climate warms. This constitutes a negative climate feedback which attains a magnitude of 0.1–0.2 W m−2 K−1 in the tropics. The simulations further show that, despite warming the atmosphere, VB does not introduce a countervailing positive water vapor feedback. Finally, we find that VB makes the net cloud feedback in the model more negative than it otherwise would be, reinforcing the clear-sky feedback.

© 2024 American Meteorological Society. This published article is licensed under the terms of the default AMS reuse license. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Seth Seidel, sdseidel@ucdavis.edu

Abstract

Humid air is lighter than dry air at a given temperature and pressure, as the molecular weight of water vapor is less than that of dry air. This effect is known as vapor buoyancy (VB). Although VB is a straightforward consequence of the ideal gas law, its influence on climate has been understudied. This study investigates VB’s influence on atmospheric temperature, radiation, and clouds. In mechanism-denial experiments, we remove VB from the dynamics and parameterizations of an idealized general circulation model (GCM). These experiments show that VB warms the tropical free troposphere by approximately 1 K in an Earth-like climate, and the magnitude of this effect increases with warming. This additional atmospheric warming causes greater outgoing longwave radiation (OLR) to be emitted as the climate warms. This constitutes a negative climate feedback which attains a magnitude of 0.1–0.2 W m−2 K−1 in the tropics. The simulations further show that, despite warming the atmosphere, VB does not introduce a countervailing positive water vapor feedback. Finally, we find that VB makes the net cloud feedback in the model more negative than it otherwise would be, reinforcing the clear-sky feedback.

© 2024 American Meteorological Society. This published article is licensed under the terms of the default AMS reuse license. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Seth Seidel, sdseidel@ucdavis.edu

1. Introduction

The molar mass of water vapor (18 g mol−1) is less than that of dry air (29 g mol−1). According to the ideal gas law, humid air is consequently less dense, or more buoyant, than dry air at a given temperature and pressure. This effect is known as vapor buoyancy (VB). VB is represented in the expression for virtual temperature:
Tυ=T(1+νq),
where Tυ is the temperature a parcel of completely dry air must have in order to have the same density as a given parcel of humid air with temperature T and specific humidity q. The virtual parameter ν=(Md/Mυ)10.61 is a function of the molar masses of water vapor and dry air, Mυ and Md, respectively. Although the origin of VB is well understood, its effect on climate has drawn little theoretical attention. Furthermore, several state-of-the-art global climate models appear to neglect VB in their dynamics, calculating their pressure gradient as if the atmosphere is completely dry. This causes systematic errors in their atmospheric temperature, boundary layer clouds, and clear-sky longwave emission (Yang et al. 2022; Yang and Seidel 2023).

Past studies have shown that VB conspires with the dynamics of Earth’s tropics to warm the midtroposphere (2–6 km in altitude) and causes a negative climate feedback in clear skies (Fig. 4 in Yang and Seidel 2020; Fig. 5 in Seidel and Yang 2020). The warming mechanism is depicted in Fig. 1 and shall be described here. Tropical dynamics constrain the free troposphere to a weak buoyancy gradient (WBG)—that is, a weak horizontal gradient in virtual temperature (Charney 1963; Sobel et al. 2001; Yang 2018a,b). As a result, comparatively dry regions of the atmosphere must be warmer in order to maintain equal density to the moist, convective regions (Yang and Seidel 2020; Bao and Stevens 2021). However, if we were to neglect VB, the weak buoyancy gradient would become a weak temperature gradient instead, so that the wet and dry regions of the tropics would have the same temperature. Since the virtual temperature profile of the tropics is generally set by convection in the humid regions, VB causes the atmosphere to be warmer by the amount ΔTυb.

Fig. 1.
Fig. 1.

(a) The dependence of midtroposphere temperature at a given pressure level (vertical axis) upon relative humidity (horizontal axis) under WBG dynamics. The solid line indicates an atmosphere with VB, whereas the dashed line indicates an atmosphere with VB. (b) Vertical temperature profiles for a cold climate and hot climate. VB causes the atmosphere to warm more than it otherwise would as ΔTVB increases with climate warming.

Citation: Journal of Climate 37, 22; 10.1175/JCLI-D-23-0558.1

Although the influence of VB is usually small in the mean climate, it may increase substantially with warming due to the strong temperature dependence of saturation vapor pressure. Yang and Seidel (2020) argued that ΔTυb would increase as the climate warms, as the Clausius–Clapeyron scaling of saturation vapor pressure leads to an increase in the water vapor difference between the moist and dry regions of the tropics. This vertically nonuniform warming of the atmosphere causes a negative climate feedback in clear skies, as the warmer midtroposphere emits greater longwave radiation to space. Yang and Seidel (2020) demonstrated this VB feedback in a simple one-dimensional model of the tropical atmosphere and estimated its magnitude as O(0.2) W m−2 K−1 in the present-day tropical climate. Seidel and Yang (2020, hereafter SY20) corroborated the VB feedback mechanism in an idealized two-dimensional cloud-resolving model. Although the SY20 study demonstrated the VB feedback mechanism when convection is explicitly resolved, their idealized experiment had several limitations. First, their model lacked planetary rotation. This left it unclear whether the VB feedback would occur at latitudes for which the WBG assumption is only approximate, as strict WBG requires negligible planetary rotation. Such a constraint would confine the VB feedback to a narrow equatorial band, rendering it unimportant to the global energy budget.

A second limitation of the SY20 model is that it relied on the self-organizing nature of convection to give rise to a tropical circulation, whereas a realistic Hadley circulation may have a different water vapor budget and thereby a distinct moisture distribution. This left it unclear whether the negative VB feedback would be accompanied by a countervailing positive component of water vapor feedback, as might be expected with increasing atmospheric temperature (Held and Shell 2012; Jeevanjee et al. 2021). It was also unclear whether there would be a compensating (positive) or reinforcing (negative) cloud feedback due to vapor buoyancy.

The present study will address these unanswered questions regarding the VB feedback. We shall use simulations in a general circulation model to ask whether rotation or changes in dry-region water vapor significantly alter the VB feedback. We shall then clarify whether VB’s influence on the clear-sky climate feedback is principally due to large-scale dynamics and whether its effects on boundary layer processes and convection alter the total climate feedback. Finally, we shall investigate how VB influences cloud feedbacks in this model.

2. Simulation design

We rely on aquaplanet simulations in the Community Atmosphere Model, version 6 (CAM6) (Danabasoglu et al. 2020). CAM6 includes the vapor buoyancy feedback in both its dynamical calculations and its moist parameterizations, making it useful for mechanism-denial experiments. To calculate radiative fluxes and tendencies, CAM6 employs the Rapid Radiative Transfer Model for GCMs (RRTMG, Mlawer et al. 1997). The model is run with fixed CO2 concentration of 348 ppmv. Shallow convection and boundary layer processes are parameterized using the Cloud Layers Unified by Binormals (CLUBB) scheme, and deep convection is parameterized using the Zhang–McFarlane scheme (Bogenschutz et al. 2013; Zhang and McFarlane 1995). The model is run at an approximately 2° horizontal resolution. The prescribed surface temperature Ts is a meridionally symmetric function of latitude ϑ which mimics the observed meridional temperature distribution of temperature. The surface temperature distribution is set according to the following function:
Ts={29°C×[1sin2(65ϑ×π180°)]2°C+δTs,75°<ϑ<75°2°C+δTs,elsewhere.
This is identical to the function used in simulations reported by Yang et al. (2022) but for the addition of a surface temperature increment δTs. Although Ts is prescribed in degrees Celsius, we refer to the simulations by their approximate equatorial temperature in kelvins. For example, the simulations with δTs = 0°C are referred to as the “300-K simulations.” A comparison to the observed surface temperature distribution is given in the online supplemental material.

To test the role of VB, we perform three simulations at each surface temperature: 1) control (CNTL), an unmodified version of the model; 2) dynamics mechanism denial (MD-DYN), in which VB is removed from the model’s dynamics but not its moist parameterizations; and 3) all-physics mechanism denial, in which VB is removed from the dynamics as well as the boundary layer and convection schemes. The MD-DYN experiment is intended to isolate only the physics of the VB feedback, as well as to emulate several climate models who exclude VB from their pressure gradient calculation but from not their moist parameterizations.1 The MD-ALL simulation is intended to test whether the VB feedback is active compared to a more physically consistent counterfactual. The MD-DYN and MD-ALL experiments correspond to the MD1 and MD2 experiments in Yang et al. (2022); however, we use a different atmospheric model here.

This study presents several calculations of clear-sky longwave radiative effects and feedbacks. These are obtained using clear-sky radiative kernels, which represent the linear response of top-of-atmosphere radiation to atmospheric temperature and humidity. The kernels—described in appendix B—are calculated via partial radiative perturbations to temperature and humidity using RRTMG.

3. Hotter atmosphere due to VB

a. WBG dynamics

In this section, we ask whether the atmosphere is hotter due to VB in an environment with appreciable planetary rotation. That is, we are concerned with the difference in simulated temperature due to VB. We begin by obtaining a simple analytic prediction for that temperature difference under strict WBG conditions ΔTwbg. Since density is approximately horizontally homogeneous in the tropical free troposphere under WBG (Charney 1963; Sobel et al. 2001), we equate the virtual temperature of a parcel of saturated air (Tυ,sat) and the virtual temperature of air in the simulated or observed atmosphere Tυ:
Tυ=Tυ,sat.
Substituting the definition of virtual temperature, we obtain
T(1+νq)=Tsat[1+νq*(Tsat)].
We define ΔTwbg = TTsat as the temperature difference due to VB. We further linearize q* around T using the Clausius–Clapeyron relation Tq*=Lq*/RυT2, where L is the latent heat of vaporization of water vapor and Rυ is the gas constant of water vapor. Substituting Tsat = T − ΔTwbg into Eq. (5) and expanding around ΔT:
T(1+νq)=T(1+νq*(T))ΔTwbg[1+νq*(T)+νLRυTq*(T)]+ΔTwbg2νLRυT2q*(T).
We note that νq*(T)1 and exclude the fourth term on the right-hand side. We also exclude the higher-order ΔT2 term and then reorganize
ΔTwbg=νT[q*(T)q]1+ν(LRυT)q*(T).
We will now compare ΔTwbg to the simulated differences in temperature between the CNTL and MD simulations:
ΔTMD-DYN=TCNTLTMD-DYN,
ΔTMD-ALL=TCNTLTMD-ALL,
where TCNTL is the temperature of the atmosphere in the control simulations and TMD-DYN and TMD-ALL are the temperatures in the mechanism-denial simulations. Figures 2a and 2b show ΔTMD-DYN and ΔTMD-ALL, respectively, at the control surface temperature (300 K at the equator). The subtropical midtroposphere is up to 1 K warmer in the CNTL simulation than the MD-DYN simulation and up to 2.5 K warmer in the CNTL simulation than in the MD-ALL simulation.
Fig. 2.
Fig. 2.

Temperature difference due to VB (color contours). (a) The ΔTMD-DYN for an equatorial surface temperature of 300 K. (b) The ΔTMD-ALL for an equatorial surface temperature for 300 K. (c) The ΔTwbg for an equatorial surface temperature of 300 K. (d) The ΔTMD-DYN for an equatorial surface temperature of 306 K. (e) The ΔTMD-ALL for an equatorial surface temperature of 306 K. (f) The ΔTwbg for an equatorial surface temperature of 306 K. The black contours represent the temperature of the CNTL simulations.

Citation: Journal of Climate 37, 22; 10.1175/JCLI-D-23-0558.1

Figures 2d and 2e show ΔTMD-DYN and ΔTMD-ALL, respectively, when the surface temperature is 6 K greater. Both ΔTMD-DYN and ΔTMD-ALL increase with climate warming.

Figures 2c and 2f show ΔTwbg calculated from the control simulation. In the midtroposphere, ΔTwbg is a close match to ΔTMD-DYN (Figs. 2a,d), but substantially underestimates ΔTMD-ALL. This is because ΔTwbg captures the warming due to VB’s interaction with the large-scale dynamics, but it does not account for differences due to boundary layer turbulence or convection. For greater clarity, Fig. 3 compares ΔTMD-DYN, ΔTMD-ALL, and ΔTwbg at the 691-hPa model level. The ΔTwbg closely approximates ΔTMD-DYN, particularly in the tropics (±30°). Even in a model with appreciable planetary rotation where strict WBG does not hold, VB nevertheless warms the atmosphere by the amount predicted by ΔTwbg. This suggests that the vapor-buoyancy feedback is active in the real-world subtropics. We shall evaluate the simulated feedback later in section 4.

Fig. 3.
Fig. 3.

Temperature difference due to VB at the 691-hPa model level, for the simulations with the equatorial surface temperature of 300 K.

Citation: Journal of Climate 37, 22; 10.1175/JCLI-D-23-0558.1

b. Moist adiabats and boundary layer dynamics

In this section, we shall address the discrepancy between ΔTwbg and ΔTMD-ALL. Figure 3 shows that ΔTwbg captures the gross features and scale of the meridional pattern of ΔTMD-ALL within the tropics: Both have local maximums around 15° and local minimums at the equator. However, ΔTMD-ALL is considerably greater than either ΔTwbg or ΔTMD-DYN.

Additionally, ΔTMD-ALL shows substantial warming in the upper troposphere, which is absent in ΔTwbg and ΔTMD-DYN. Free-troposphere temperature in the tropics is tightly linked to boundary layer moist static energy, as the virtual temperature profile is set by convection which ascends from there (Emanuel et al. 1994). CAM6’s Zhang–McFarlane convection scheme, like most deep convection schemes, is in fact formulated based on this relationship to boundary layer energy (Zhang and McFarlane 1995). Therefore, we will draw a link between VB’s influence on boundary layer dynamics and the additional warming indicated by ΔTMD-ALL.

To examine the influence of the boundary layer, we calculate the temperatures of parcels undergoing moist ascent from there. Using an analytic expression provided by Romps (2017), we first calculate the lifting condensation level for an equator-average (±5°) parcel lifted dry-adiabatically from the 913-hPa model level. Then, we calculate the moist-adiabatic temperature profile Tma for a parcel which ascends from that lifting condensation level. The horizontal axis of Fig. 4 shows the calculated moist-adiabatic temperature Tma for each model level between 300 and 800 hPa in the same ±5° latitude band. The vertical axis is the actual simulated temperature T at that level. Filled circles denote the CNTL simulation, and open circles denote MD-ALL. Overall, Tma is a strong predictor of T, as the marks run approximately parallel to the one-to-one line. More importantly, a difference in Tma between CNTL and MD-ALL is matched by an approximately equal difference in T. This suggests that differences in boundary layer moist static energy are responsible for the large values of ΔTMD-ALL at the equator. We may assume that those differences in free-troposphere temperature at the equator are then communicated to the subtropics via WBG dynamics, resulting in the large difference between ΔTMD-DYN and ΔTMD-ALL seen in Figs. 2 and 3.

Fig. 4.
Fig. 4.

Equatorial region (±5°) free-troposphere temperature as predicted by moist-adiabatic ascent from the equatorial boundary layer (horizontal axis) plotted against the simulated difference in temperature. Filled circles denote data from the CNTL simulation, and open circles denote MD-ALL.

Citation: Journal of Climate 37, 22; 10.1175/JCLI-D-23-0558.1

How does VB cause the CNTL simulation to have a warmer boundary layer than the MD-ALL simulation? There are likely two factors at play. First, VB weakens the inversion strength within the boundary layer scheme, which allows greater entrainment of free-troposphere air to warm the equatorial boundary layer. Second, VB accounts for about 50% of surface buoyancy fluxes in Earth’s tropical ocean (Yang et al. 2022). The greater surface buoyancy fluxes in CNTL compared to MD-ALL increase turbulent kinetic energy production in the boundary layer (as shown in Fig. S2). This effect can then enhance the turbulent entrainment of free-troposphere warm air into the boundary layer. Our analysis will focus primarily on the first of these two effects, in which VB weakens the boundary layer inversion.

We begin by assessing the virtual potential temperature structure of the simulated lower troposphere as averaged around the equator (±5°). Figure 5a shows the virtual potential temperature θυ profile between 700 hPa and the surface for the three simulations with the equatorial surface temperature of 300 K:
θυ=T(p0p)R/cp(1+νq),
where p0 = 1000 hPa is a reference pressure, p is atmospheric pressure, R is the gas constant for air, and cp is the specific heat capacity of air at constant pressure. To imitate our treatment of the parameterizations, ν is set to zero for MD-ALL.
Fig. 5.
Fig. 5.

Profiles of lower-troposphere, equator-region (±5°) (a) virtual potential temperature, (b) Δθυ,MD-ALL, (c) temperature, (d) relative humidity with respect to temperature, (e) specific humidity, and (f) moist static energy, for the simulations with equatorial surface temperature 300 K.

Citation: Journal of Climate 37, 22; 10.1175/JCLI-D-23-0558.1

The entrainment of free-troposphere air is governed by the strength of the boundary layer inversion, and the sharp decline in specific humidity with height causes VB to weaken the inversion. Figure 5b shows the difference in θυ between CNTL and MD-ALL:
Δθυ,MD-ALL=θυ,CNTLθυ,MD-ALL.

Δθυ,MD-ALL is greater below 900 hPa (i.e., in the boundary layer) than it is above 800 hPa, by as much as 1 K. This indicates that the boundary layer inversion is weaker due to VB. We would expect a weaker boundary layer inversion to cause greater entrainment of free-troposphere air in CNTL than in MD-ALL. Due to vertical gradients in potential temperature and specific humidity, that entrainment should cause greater boundary layer temperature and reduced boundary layer relative humidity. Indeed, that is what we find in Figs. 5c and 5d: a hotter, less saturated boundary layer. However, the greater temperature in the CNTL boundary layer allows it to attain greater specific humidity q despite having lower relative humidity (Fig. 5e). As a result, the boundary layer moist static energy is greater in CNTL than in MD-ALL (Fig. 5f). The greater moist static energy in the CNTL simulation’s equatorial boundary layer is transmitted upward via convection, resulting in the greater free-troposphere temperature compared to MD-ALL, as seen in Fig. 4.

4. Negative climate feedbacks due to VB

In this section, we investigate the role of VB in the radiation budget, with a special focus on the clear-sky longwave feedback local to the tropics. As a starting point, we quantify the difference in the clear-sky longwave feedbacks Δλcslw between the CNTL and MD simulations:
Δλcslw,MD-DYN=ddTs(Rcslw,CNTLRcslw,MD-DYN),
Δλcslw,MD-ALL=ddTs(Rcslw,CNTLRcslw,MD-ALL).
Here, Rcslw denotes the net upwelling clear-sky longwave radiation at top of atmosphere in each simulation. That is, Rcslw is the clear-sky OLR. The red lines in Fig. 6 show the local (latitude-dependent) Δλcslw as estimated using independent linear regressions of the relevant radiative effect against surface temperature at each latitude. We wish to pay special attention to the clear-sky longwave feedback, so we have chosen a sign convention such that upward fluxes are positive. The values of Δλcslw are positive and strongest in the tropics, indicating a substantial negative climate feedback there. In section 4a, we shall investigate the role of VB in setting the tropical clear-sky longwave feedback.
Fig. 6.
Fig. 6.

Difference in climate feedbacks between CNTL and MD simulations. A positive value of Δλ indicates a more negative climate feedback in the CNTL simulation.

Citation: Journal of Climate 37, 22; 10.1175/JCLI-D-23-0558.1

We also quantify the differences in all-sky feedbacks, including the effects of clouds and shortwave radiation:
Δλall,MD-DYN=ddTs(Rlw,CNTLRlw,MD-DYN+Rsw,CNTLRsw,MD-DYN),
Δλall,MD-ALL=ddTs(Rlw,CNTLRlw,MD-ALL+Rsw,CNTLRsw,MD-ALL).
Here, Rlw denotes the net upwelling all-sky longwave radiation at top of atmosphere for a given simulation and Rsw denotes the net upwelling all-sky shortwave radiation. Since we have chosen a sign convention that upward fluxes are positive, a positive value of Δλall indicates a more negative climate feedback due to VB. The blue lines in Fig. 6 show the local Δλall as estimated using linear regressions. The effects of clouds on the global-average longwave feedback are investigated in section 4b.

a. Clear-sky VB feedback

Here, we investigate the differences in simulated clear-sky outgoing longwave radiation ΔRcslw between the CNTL and MD simulations. The green marks in Fig. 7a show tropical-average (±30° latitude) ΔRcslw for MD-DYN. The value of ΔRcslw increases as the climate warms, suggesting that the vapor-buoyancy feedback is active there. To attribute ΔRcslw to VB, we must decompose this radiative effect according to its contributions from temperature and water vapor.

Fig. 7.
Fig. 7.

Difference in net clear-sky OLR at top of atmosphere for (a) the MD-DYN experiment and (b) the MD-ALL experiment. Closed blue circles indicate the radiative effect implied by ΔTυb as simulated by either experiment. Open blue circles indicate the radiative effect implied by ΔTwbg. The red circles indicate the radiative effect implied by differences in specific humidity between the CNTL and MD simulations. The violet circles indicate the sum of the temperature and specific humidity effects. The green crosses indicate the simulated difference in clear-sky OLR. The blue curve indicates a cubic fit to the open circles.

Citation: Journal of Climate 37, 22; 10.1175/JCLI-D-23-0558.1

To separate ΔRcslw into its contributions from tropospheric temperature and specific humidity, we use clear-sky approximate radiative kernels. These kernels are linear response functions of top-of-atmosphere radiation to perturbations in temperature and humidity. The kernels are described in appendix B. To obtain radiative effects, we take the inner products of the temperature and humidity kernels with the thermodynamic perturbations ΔTMD-DYN and ΔqMD-ALL, using the lapse-rate tropopause2 as the upper limit for the integral. The solid blue circles in Fig. 7a show the differences in ΔRcslw due to tropospheric ΔTMD-DYN, whereas the red circles show the differences due to tropospheric ΔqMD-DYN. The violet circles show the sum of these two radiative effect. These closely approximate ΔRcslw, which suggests that stratospheric effects are negligible. Finally, the open blue circles indicate the temperature radiative effect using ΔTwbg.

We can draw two inferences from Fig. 7a. First, VB is responsible for a robust negative feedback as shown by 1) the robustness of the trend in ΔRcslw with warming and 2) that this trend can be explained well by ΔTwbg. Second, there is essentially no countervailing change in the water vapor feedback due to VB: tropospheric water vapor does not contribute to the trend in ΔRcslw. To supplement this result, appendix A discusses the differences in humidity between CNTL and MD-DYN.

Figure 7b is the same as Fig. 7a, but derived from the MD-ALL experiment. In this experiment, there is an additional positive water vapor feedback due to VB, indicated by the negative trend of the red marks with climate warming. However, this is offset by a greater negative feedback due to ΔTMD-ALL. The sum of the temperature and water vapor radiative effects in the MD-ALL experiment indicates a negative VB feedback similar both to MD-DYN and to the vapor-buoyancy feedback implied by ΔTwbg. Notably, the MD-ALL experiment shows a larger radiative effect of ΔTMD-ALL (solid blue circles) than is predicted by WBG theory (open blue circles). This can be attributed to the additional difference in atmospheric temperature for MD-ALL discussed in the previous section. The warmer atmosphere emits more radiation to space. The greater warming is partially offset by additional atmospheric water vapor, as evidenced by the comparatively smaller discrepancy between WBG theory (open blue circles) and the combined radiative effect of ΔTMD-ALL and Δq (solid violet circles).

The MD-DYN and MD-ALL experiments exhibit similar differences in clear-sky feedback compared to CNTL. The total difference in clear-sky feedback in MD-ALL, Δλcslw,MD-ALL, can be estimated by linear regression as 0.20 W m−2 K−1, compared to 0.22 W m−2 K−1 for Δλcslw,MD-DYN. Evidently, any additional trend in temperature due to the boundary layer mechanisms discussed in section 3b is offset by additional atmospheric water vapor. This is because greater temperature at the sites of deep convection causes greater water vapor detrainment (due to Clausius–Clapeyron). That additional water vapor has a strong countervailing effect on OLR (Jeevanjee et al. 2021).

We refer to the climate feedback derived from the trend in ΔTwbg as the VB feedback. The magnitude of the VB feedback is measured by its feedback parameter, which represents a top-of-atmosphere flux sensitivity to a unit increase of surface temperature:
λυb=dRυbdTs,
where Rυb is the top-of-atmosphere radiative effect (additional OLR) due to ΔTwbg. As before, we define upward fluxes as positive; therefore, a positive value of λυb indicates a negative climate feedback. To calculate λυb, we fit a cubic curve to the ΔTwbg marks in Fig. 7 and take its derivative. We chose a cubic model to capture the nonlinear increase in feedback magnitude found in SY20. Figure 8 shows λυb plotted in red. At an Earth-like equatorial temperature of 300 K, the tropical-average feedback due to VB is 0.11 W m−2 K−1. This is about 5% of the total clear-sky longwave feedback of 2.2 W m−2 K−1 in the CNTL simulation. For context, both idealized column models and comprehensive GCMs similarly find a typical clear-sky longwave feedback of about 2 W m−2 K−1 (Zhang et al. 2020; McKim et al. 2021). The VB feedback increases strongly as the climate warms, reaching a value of 0.36 W m−2 K−1 at the greatest surface temperature we tested. Therefore, the VB feedback may play an outsize role in hot paleoclimates and hot planetary climates.
Fig. 8.
Fig. 8.

Tropical-average VB feedback derived from a hierarchy of climate models. The CRM result is reported in Seidel and Yang (2020), and the simple column result is reported in Yang and Seidel (2020).

Citation: Journal of Climate 37, 22; 10.1175/JCLI-D-23-0558.1

Figure 8 also plots data from two other studies of the VB feedback. The blue curve represents the feedback parameter as estimated by a curve fit in the SY20 study, which employed essentially the same feedback analysis methods as this study, albeit for an idealized 2D cloud-resolving model. The violet curve shows an estimate from a simple 1D column model described in Yang and Seidel (2020). That model employed a simple two-band radiation calculation for an all-troposphere idealized column with and without VB-induced warming. Here, we employ the calculations from that study with the that model’s humidity parameter set to β = 0.5. These curves represent a hierarchy of models for the VB feedback: a 3D GCM, a 2D CRM, and 1D column model. The three models broadly agree upon the magnitude of the VB feedback and its trend with warming, thus corroborating one another. The differences among these three models are due to differences in virtual temperature profiles, in their humidity profiles, and in their treatments of atmospheric radiation.

b. All-sky and global-average feedbacks due to VB

Our analysis of feedbacks due to VB has so far been limited to clear-sky radiation in the tropics. However, it is useful to ask whether VB alters the total climate feedback in a more complete picture: in the global average, with cloud radiative effects included. Figures 9a and 9b show the globally averaged differences in top-of-atmosphere all-sky radiation between CNTL and MD-DYN and CNTL and MD-ALL, respectively. As in the previous sections, we adopt a sign convention that upwelling fluxes are positive so that a positive trend indicates a negative climate feedback. For MD-DYN compared to CNTL, the all-sky radiative effect is greater than the clear-sky radiative effect by 4–6 W m−2. The all-sky radiative effects also exhibit a greater trend (i.e., feedback) with warming. We can estimate the feedback magnitude as 0.41 W m−2 K−1 by way of a linear regression of radiative effect against surface temperature. This is greater than the clear-sky longwave difference of 0.17 W m−2 K−1. Calculating feedbacks in the same way for MD-ALL, the total all-sky feedback is 0.31 W m−2 K−1, compared to 0.14 W m−2 K−1 for clear-sky longwave. In either case, cloud feedbacks add to the vapor buoyancy feedback.

Fig. 9.
Fig. 9.

Global-mean difference in net all-sky radiation (red marks) at top of atmosphere for (a) the MD-DYN experiment and (b) the MD-ALL experiment. Global-mean difference in clear-sky OLR (blue marks) are shown for comparison.

Citation: Journal of Climate 37, 22; 10.1175/JCLI-D-23-0558.1

Next, we shall isolate the differences in cloud feedbacks between our simulations. We define the longwave and shortwave cloud radiative effects (CREs), respectively, as
CRElw=RlwRcslw,
CREsw=RswRcssw,
where all these variables are defined as before and Rcssw is the net upwelling flux of shortwave radiation in clear skies. Using these values of CRElw and CREsw, we calculate the difference in total cloud feedbacks between two simulations, Δλlwcld and Δλswcld:
Δλlwcld,MD-DYN=ddTs(CRElw,CNTLCRElw,MD-DYN),
Δλlwcld,MD-ALL=ddTs(CRElw,CNTLCRElw,MD-ALL),
Δλswcld,MD-DYN=ddTs(CREsw,CNTLCREsw,MD-DYN),
Δλswcld,MD-ALL=ddTs(CREsw,CNTLCREsw,MD-ALL).
As elsewhere in this paper, CRE is defined such that a positive value indicates additional net upward radiation at the top of the atmosphere, so a positive value of Δλlwcld or Δλswcld indicates that the CNTL simulation cloud feedback is more negative than the MD simulation.

The solid lines in Fig. 10 show Δλlwcld,MD-DYN and Δλswcld,MD-DYN. These shortwave and longwave feedbacks approximately cancel one another near the equator. That suggests a high-cloud effect, as greater cloudiness in the upper troposphere has similar but opposite effects on longwave emission and shortwave reflection. However, in the subtropics, a negative cloud shortwave feedback dominates the longwave. That may indicate a difference in low-cloud feedback between CNTL and MD-DYN. Subtropical low clouds are likely to become less extensive as the climate warms (Sherwood et al. 2020; Myers et al. 2021), and our aquaplanet simulations replicate that finding (not shown). The present results suggest that VB causes low clouds to recede less than they otherwise would, causing a more-negative climate feedback. This may be a consequence of the greater midtroposphere temperatures due to VB (Fig. 2), which cause VB to enhance the inversion strength as “seen” by the boundary layer scheme. That would promote subtropical boundary layer clouds by reducing the entrainment of dry free-troposphere air (Yang et al. 2022). If this generalizes to other models and configurations, those comprehensive GCMs which neglect VB in their pressure gradient may achieve a somewhat more negative cloud feedback if they were to instead incorporate VB into their dynamics.

Fig. 10.
Fig. 10.

Simulated differences in cloud feedbacks due to VB.

Citation: Journal of Climate 37, 22; 10.1175/JCLI-D-23-0558.1

The dashed lines in Fig. 10 show Δλlwcld,MD-ALL and Δλswcld,MD-ALL. Here, the deep tropical difference in cloud shortwave feedback is greater in magnitude than its counterpart longwave feedback. Since upper-troposphere clouds dominate there, this difference in magnitude may be due to differences in cloud altitude or in cloud optical depth. A more extensive feedback analysis would be necessary to distinguish these mechanisms. The differences in subtropical cloud feedbacks in MD-ALL are qualitatively similar to what we found for the MD-DYN experiment: VB induces a negative shortwave feedback without much change in longwave. As before, low clouds are presumably responsible for this behavior. However, the mechanism at work is likely different from MD-DYN. One possibility is that VB increases cloud development within the boundary layer by increasing the buoyancy of shallow convection (Dagan et al. 2018).

5. Discussion

This study reports how vapor buoyancy (VB) influences the climate of a simulated atmosphere. We showed that VB conspires with weak buoyancy gradient (WBG) dynamics in the tropics to warm the subtropical midtroposphere. We also showed that VB increases the equilibrium boundary layer moist static energy in the deep tropics and consequently warms the free troposphere. The first of these two temperature effects is responsible for a negative climate feedback, which we call the vapor-buoyancy feedback. Finally, we found that VB alters the trend in subtropical cloudiness in the this model such that the net cloud feedback is more negative than it otherwise would be.

In the present study, we found that the clear-sky VB feedback is spatially extensive and robust even when one considers planetary rotation as well as VB’s influence on the water vapor feedback. This result was not clear in our earlier study (SY20), which used cloud-resolving simulations with zero rotation. Our new results show the VB feedback to be physically valid in a warm, wet, rotating atmosphere such as Earth’s. Furthermore, since VB causes a difference in clear-sky feedback, the VB feedback is not merely an artifact of unconventional feedback decomposition as suggested by Colman and Soden (2021). Without VB, the mean tropical climate feedback would in fact be less negative. This intuition is important for the development of climate models. A review of 23 state-of-the-art GCMs found six which appear to exclude VB from their dynamics (Yang et al. 2022). Those no-VB GCMs consequently emit less outgoing longwave radiation in the tropics (Yang and Seidel 2023), and the VB feedback would presumably not be present in their simulations.

The clear-sky VB feedback is presumably present in those comprehensive climate models which do include VB in their pressure gradient calculations. It would also be included in standard methods for diagnosing feedbacks due to temperature and humidity. However, which of those feedbacks captures VB will depend on our choice of method. Traditionally, temperature feedbacks are calculated with a fixed specific humidity (Soden et al. 2008). Under that assumption, the VB feedback would be part of the “lapse-rate” climate feedback, as it results from greater troposphere warming relative to the surface. However, some have suggested holding (temperature-) relative humidity as fixed when calculating the lapse-rate feedback (Held and Shell 2012; Jeevanjee et al. 2021). In an RH-based feedback decomposition, the negative VB feedback quantified here would instead be split between a negative lapse-rate feedback and a negative relative humidity feedback (due to decreasing relative humidity). This difficulty could be rectified by redefining relative humidity with respect to a constant-density saturation rather than a constant-temperature saturation, as discussed in appendix A. However, that assumption may not be appropriate for the boundary layer. There is no perfect choice of moisture variable for feedback decomposition.

The VB feedback represents a considerable improvement in our understanding of hot-climate longwave feedbacks. Recent studies of the clear-sky longwave feedback have emphasized the importance of the water vapor window through which longwave emission from the surface escapes to space (Koll and Cronin 2018; McKim et al. 2021; Seeley and Jeevanjee 2021; Koll et al. 2023). As the atmosphere becomes hotter and wetter, continuum absorption causes the water vapor window to close, reducing the total clear-sky longwave feedback. Our work suggests that the VB feedback, which strongly increases with surface temperature (Fig. 6), may help to compensate the loss of surface emission and stabilize the climate at hot surface temperatures. In the future, it may be fruitful to investigate whether a VB feedback can delay or halt the transition to a runaway greenhouse state in hot planetary climates (Ingersoll 1969).

Although this and previous studies have addressed the VB feedback using a wide range of methods and models, future studies may still rectify several remaining knowledge gaps. Further modeling experiments could include continents in order to more realistically simulate VB’s influence on cloud feedbacks. It would be worthwhile to also consider the effects of interactive surface temperatures, as the VB feedback’s outsize influence in the tropics may reduce the meridional temperature gradient in a warming climate.

Acknowledgments.

The authors thank three anonymous reviewers whose comments greatly improved this manuscript. The study also benefitted from a discussion with Nathan Arnold regarding boundary layer dynamics. This work was supported by a National Science Foundation CAREER award and a Packard Fellowship for Science and Engineering. Computational resources were provided by the Department of Energy’s National Energy Research Scientific Computing Center (NERSC) and by the National Center for Atmospheric Research Computational and Information Systems Lab (NCAR CISL).

Data availability statement.

Model output and model modifications used to achieve these results can be found at Zenodo: https://zenodo.org/records/10869753. Additional data related to this paper may be requested from the authors.

Footnotes

1

Our design of the MD-DYN experiment is simply to zero out the virtual temperature parameter (zvir) which is used in the dynamical core. This affects not only the dynamical core but also the rest of the model except for the convection and PBL schemes. Notably, this includes surface flux calculations. However, our testing showed these made little difference in the results presented here.

2

We use a lapse-rate threshold of 2 K km−1 to define the tropopause, as in Zelinka et al. (2020).

APPENDIX A

Muted Changes in Humidity Due to VB

Figures A1a and A1c show ΔRHT, the difference in relative humidity between CNTL and MD-DYN. The subscript T denotes that this is a conventional relative humidity, calculated with respect to constant-temperature saturation process. The atmosphere is less humid in the CNTL simulations, particularly in the subtropical midtroposphere, where VB causes the greatest warming. This suggests that VB dries the atmosphere. Furthermore, it appears that the magnitude of ΔRHT increases with warming. This is inconvenient for understanding climate feedbacks, as RHT is often used as the state variable (held fixed) for evaluating temperature-based climate feedbacks (Held and Shell 2012; Zelinka et al. 2020).

For the foregoing reason, it is helpful to explore an alternative notion of relative humidity. We define density-relative humidity as the specific humidity of a parcel of air, q, compared to the specific humidity which would be achieved in a constant-density saturation q*(Tυ,p):
RHρ=qq*(Tυ,p).
RHρ is a useful humidity measure in the free troposphere if one assumes the cloud-free atmosphere is principally moistened by horizontal motions along isentropes (surfaces of constant virtual potential temperature). This alternative definition of relative humidity is not completely novel. Romps (2014) developed an analytical model for tropical relative humidity in which RH was defined as the ratio of specific humidity in a subsaturated environment to the specific humidity of a saturated plume at the same density and pressure, which is how we define RHρ. That model suggested that RH is enhanced by isentropic mixing from moist plumes but reduced by subsidence drying. However, it neglected the lightness of water vapor, so the distinction between RHρ and RHT was not apparent.
In Earth’s atmosphere, RHρ is necessarily greater than RHT since adding more water vapor to a parcel under a constant-density process results in a lower temperature and smaller q*. The relationship between RHρ and RHT can be derived from the relationship q*(Tυ,p)=q*(TΔTwbg). Linearizing q* around the Clausius–Clapeyron relation Tq*=(L/RυT2)q*, this gives
RHρ=qq*(T,p)LRυT2q*(T,p)ΔTwbg.
This simplifies to
RHρ=RHT1LRυT2ΔTwbg,
where RHT is the conventional (temperature-) relative humidity. The departure between RHρ and RHT grows considerably with warming, owing to its dependence on ΔTwbg. For a parcel of air with RHT = 50% and a pressure of 800 hPa, RHρ increases from 52% at a temperature of 280 K to 57% at a temperature of 300 K.

Figures A1b and A1d show ΔRHρ for the MD-DYN experiment at two different surface temperatures. Since ΔTwbg is zero in an atmosphere without VB, RHρ = RHT for the mechanism-denial experiments. Comparing to Figs. A1a and A1c, ΔRHρ exhibits weaker minimums than ΔRHT in the subtropical midtroposphere. In Fig. A1e, we compare ΔRHT and ΔRHρ at the 691-hPa model level, which tells a similar story: ΔRHρ exhibits weaker VB-induced drying of the subtropics compared to ΔRHT and a compensating wetting of the deep tropics. This suggests an intensification or narrowing of the model’s intertropical convergence zone due to VB, which is also evidenced in the cloud field (Fig. S3). Furthermore, RHρ is more nearly fixed than RHT is as the climate warms, especially in the subtropics.

Fig. A1.
Fig. A1.

Difference in zonal-average relative humidity due to VB (color contours); zonal-average CNTL simulation humidity (black contours). (a) The ΔRHT for an equatorial surface temperature of 300 K. (b) The ΔRHρ for an equatorial surface temperature of 300 K. (d) The ΔRHρ for an equatorial surface temperature of 306 K. (e) The ΔRHT and ΔRHρ at 700 hPa.

Citation: Journal of Climate 37, 22; 10.1175/JCLI-D-23-0558.1

APPENDIX B

Clear-Sky Radiative Kernels

To decompose simulated differences in clear-sky OLR into contributions due to temperature and water vapor, we construct clear-sky radiative kernels, shown in Fig. B1. Radiative kernels are linear response functions of top-of-atmosphere radiation to atmospheric properties. We use the approximate kernel technique from Cronin and Wing (2017), performing offline radiative transfer calculations in RRTMG to obtain the response of top-of-atmosphere radiation to small perturbations in the zonal-mean thermodynamic profiles. Our kernels are derived from +0.5 K perturbations in temperature and −1% perturbations in specific humidity from the CNTL simulation zonal average. We calculate a separate radiative kernel for each surface temperature.

Fig. B1.
Fig. B1.

Radiative kernels (W m−2 K−1 100 hPa−1) for (a) temperature and (b) humidity at an equatorial surface temperature of 300 K and for (c) temperature and (d) humidity for an equatorial surface temperature of 306 K. The humidity kernels reflect 1 K of warming at fixed RHT.

Citation: Journal of Climate 37, 22; 10.1175/JCLI-D-23-0558.1

REFERENCES

  • Bao, J., and B. Stevens, 2021: The elements of the thermodynamic structure of the tropical atmosphere. J. Meteor. Soc. Japan, 99, 14831499, https://doi.org/10.2151/jmsj.2021-072.

    • Search Google Scholar
    • Export Citation
  • Bogenschutz, P. A., A. Gettelman, H. Morrison, V. E. Larson, C. Craig, and D. P. Schanen, 2013: Higher-order turbulence closure and its impact on climate simulations in the community atmosphere model. J. Climate, 26, 96559676, https://doi.org/10.1175/JCLI-D-13-00075.1.

    • Search Google Scholar
    • Export Citation
  • Charney, J. G., 1963: A note on large-scale motions in the tropics. J. Atmos. Sci., 20, 607609, https://doi.org/10.1175/1520-0469(1963)020<0607:ANOLSM>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Colman, R., and B. J. Soden, 2021: Water vapor and lapse rate feedbacks in the climate system. Rev. Mod. Phys., 93, 045002, https://doi.org/10.1103/RevModPhys.93.045002.

    • Search Google Scholar
    • Export Citation
  • Cronin, T. W., and A. A. Wing, 2017: Clouds, circulation, and climate sensitivity in a radiative-convective equilibrium channel model. J. Adv. Model. Earth Syst., 9, 28832905, https://doi.org/10.1002/2017MS001111.

    • Search Google Scholar
    • Export Citation
  • Dagan, G., I. Koren, O. Altaratz, and G. Feingold, 2018: Feedback mechanisms of shallow convective clouds in a warmer climate as demonstrated by changes in buoyancy. Environ. Res. Lett., 13, 054033, https://doi.org/10.1088/1748-9326/aac178.

    • Search Google Scholar
    • Export Citation
  • Danabasoglu, G., and Coauthors, 2020: The Community Earth System Model version 2 (CESM2). J. Adv. Model. Earth Syst., 12, e2019MS001916, https://doi.org/10.1029/2019MS001916.

    • Search Google Scholar
    • Export Citation
  • Emanuel, K. A., J. D. Neelin, and C. S. Bretherton, 1994: On large-scale circulations in convecting atmospheres. Quart. J. Roy. Meteor. Soc., 120, 11111143, https://doi.org/10.1002/qj.49712051902.

    • Search Google Scholar
    • Export Citation
  • Held, I. M., and K. M. Shell, 2012: Using relative humidity as a state variable in climate feedback analysis. J. Climate, 25, 25782582, https://doi.org/10.1175/JCLI-D-11-00721.1.

    • Search Google Scholar
    • Export Citation
  • Ingersoll, A. P., 1969: The runaway greenhouse: A history of water on Venus. J. Atmos. Sci., 26, 11911198, https://doi.org/10.1175/1520-0469(1969)026<1191:TRGAHO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Jeevanjee, N., D. D. B. Koll, and N. Lutsko, 2021: “Simpson’s Law” and the spectral cancellation of climate feedbacks. Geophys. Res. Lett., 48, e2021GL093699, https://doi.org/10.1029/2021GL093699.

    • Search Google Scholar
    • Export Citation
  • 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 29310 298, https://doi.org/10.1073/pnas.1809868115.

    • Search Google Scholar
    • Export Citation
  • Koll, D. D. B., N. Jeevanjee, and N. J. Lutsko, 2023: An analytic model of the clear-sky longwave feedback. J. Atmos. Sci., 80, 19231951, https://doi.org/10.1175/JAS-D-22-0178.1.

    • Search Google Scholar
    • Export Citation
  • 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.

    • Search Google Scholar
    • Export Citation
  • Mlawer, E. J., S. J. Taubman, P. D. Brown, M. J. Iacono, and S. A. Clough, 1997: Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated-k model for the longwave. J. Geophys. Res., 102, 16 66316 682, https://doi.org/10.1029/97JD00237.

    • Search Google Scholar
    • Export Citation
  • Myers, T. A., R. C. Scott, M. D. Zelinka, S. A. Klein, J. R. Norris, and P. M. Caldwell, 2021: Observational constraints on low cloud feedback reduce uncertainty of climate sensitivity. Nat. Climate Change, 11, 501507, https://doi.org/10.1038/s41558-021-01039-0.

    • Search Google Scholar
    • Export Citation
  • Romps, D. M., 2014: An analytical model for tropical relative humidity. J. Climate, 27, 74327449, https://doi.org/10.1175/JCLI-D-14-00255.1.

    • Search Google Scholar
    • Export Citation
  • Romps, D. M., 2017: Exact expression for the lifting condensation level. J. Atmos. Sci., 74, 38913900, https://doi.org/10.1175/JAS-D-17-0102.1.

    • Search Google Scholar
    • Export Citation
  • Seeley, J. T., and N. Jeevanjee, 2021: H2O windows and CO2 radiator fins: A clear-sky explanation for the peak in equilibrium climate sensitivity. Geophys. Res. Lett., 48, e2020GL089609, https://doi.org/10.1029/2020GL089609.

    • Search Google Scholar
    • Export Citation
  • Seidel, S. D., and D. Yang, 2020: The lightness of water vapor helps to stabilize tropical climate. Sci. Adv., 6, eaba1951, https://doi.org/10.1126/sciadv.aba1951.

    • Search Google Scholar
    • Export Citation
  • 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.

    • Search Google Scholar
    • Export Citation
  • Sobel, A. H., J. Nilsson, and L. M. Polvani, 2001: The weak temperature gradient approximation and balanced tropical moisture waves. J. Atmos. Sci., 58, 36503665, https://doi.org/10.1175/1520-0469(2001)058<3650:TWTGAA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • 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, 35043520, https://doi.org/10.1175/2007JCLI2110.1.

    • Search Google Scholar
    • Export Citation
  • Yang, D., 2018a: Boundary layer diabatic processes, the virtual effect, and convective self-aggregation. J. Adv. Model. Earth Syst., 10, 21632176, https://doi.org/10.1029/2017MS001261.

    • Search Google Scholar
    • Export Citation
  • Yang, D., 2018b: Boundary layer height and buoyancy determine the horizontal scale of convective self-aggregation. J. Atmos. Sci., 75, 469478, https://doi.org/10.1175/JAS-D-17-0150.1.

    • Search Google Scholar
    • Export Citation
  • Yang, D., and S. D. Seidel, 2020: The incredible lightness of water vapor. J. Climate, 33, 28412851, https://doi.org/10.1175/JCLI-D-19-0260.1.

    • Search Google Scholar
    • Export Citation
  • Yang, D., and S. D. Seidel, 2023: Vapor buoyancy increases clear-sky thermal emission. Environ. Res. Climate, 2, 015006, https://doi.org/10.1088/2752-5295/acba39.

    • Search Google Scholar
    • Export Citation
  • Yang, D., W. Zhou, and S. D. Seidel, 2022: Substantial influence of vapour buoyancy on tropospheric air temperature and subtropical cloud. Nat. Geosci., 15, 781788, https://doi.org/10.1038/s41561-022-01033-x.

    • Search Google Scholar
    • Export Citation
  • Zelinka, M. D., T. A. Myers, D. T. McCoy, S. Po-Chedley, P. M. Caldwell, P. Ceppi, S. A. Klein, and K. E. Taylor, 2020: Causes of higher climate sensitivity in CMIP6 models. Geophys. Res. Lett., 47, e2019GL085782, https://doi.org/10.1029/2019GL085782.

    • Search Google Scholar
    • Export Citation
  • Zhang, G. J., and N. A. McFarlane, 1995: Sensitivity of climate simulations to the parameterization of cumulus convection in the Canadian Climate Centre general circulation model. Atmos.–Ocean, 33, 407446, https://doi.org/10.1080/07055900.1995.9649539.

    • Search Google Scholar
    • Export Citation
  • Zhang, Y., N. Jeevanjee, and S. Fueglistaler, 2020: Linearity of outgoing longwave radiation: From an atmospheric column to global climate models. Geophys. Res. Lett., 47, e2020GL089235, https://doi.org/10.1029/2020GL089235.

    • Search Google Scholar
    • Export Citation

Supplementary Materials

Save
  • Bao, J., and B. Stevens, 2021: The elements of the thermodynamic structure of the tropical atmosphere. J. Meteor. Soc. Japan, 99, 14831499, https://doi.org/10.2151/jmsj.2021-072.

    • Search Google Scholar
    • Export Citation
  • Bogenschutz, P. A., A. Gettelman, H. Morrison, V. E. Larson, C. Craig, and D. P. Schanen, 2013: Higher-order turbulence closure and its impact on climate simulations in the community atmosphere model. J. Climate, 26, 96559676, https://doi.org/10.1175/JCLI-D-13-00075.1.

    • Search Google Scholar
    • Export Citation
  • Charney, J. G., 1963: A note on large-scale motions in the tropics. J. Atmos. Sci., 20, 607609, https://doi.org/10.1175/1520-0469(1963)020<0607:ANOLSM>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Colman, R., and B. J. Soden, 2021: Water vapor and lapse rate feedbacks in the climate system. Rev. Mod. Phys., 93, 045002, https://doi.org/10.1103/RevModPhys.93.045002.

    • Search Google Scholar
    • Export Citation
  • Cronin, T. W., and A. A. Wing, 2017: Clouds, circulation, and climate sensitivity in a radiative-convective equilibrium channel model. J. Adv. Model. Earth Syst., 9, 28832905, https://doi.org/10.1002/2017MS001111.

    • Search Google Scholar
    • Export Citation
  • Dagan, G., I. Koren, O. Altaratz, and G. Feingold, 2018: Feedback mechanisms of shallow convective clouds in a warmer climate as demonstrated by changes in buoyancy. Environ. Res. Lett., 13, 054033, https://doi.org/10.1088/1748-9326/aac178.

    • Search Google Scholar
    • Export Citation
  • Danabasoglu, G., and Coauthors, 2020: The Community Earth System Model version 2 (CESM2). J. Adv. Model. Earth Syst., 12, e2019MS001916, https://doi.org/10.1029/2019MS001916.

    • Search Google Scholar
    • Export Citation
  • Emanuel, K. A., J. D. Neelin, and C. S. Bretherton, 1994: On large-scale circulations in convecting atmospheres. Quart. J. Roy. Meteor. Soc., 120, 11111143, https://doi.org/10.1002/qj.49712051902.

    • Search Google Scholar
    • Export Citation
  • Held, I. M., and K. M. Shell, 2012: Using relative humidity as a state variable in climate feedback analysis. J. Climate, 25, 25782582, https://doi.org/10.1175/JCLI-D-11-00721.1.

    • Search Google Scholar
    • Export Citation
  • Ingersoll, A. P., 1969: The runaway greenhouse: A history of water on Venus. J. Atmos. Sci., 26, 11911198, https://doi.org/10.1175/1520-0469(1969)026<1191:TRGAHO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Jeevanjee, N., D. D. B. Koll, and N. Lutsko, 2021: “Simpson’s Law” and the spectral cancellation of climate feedbacks. Geophys. Res. Lett., 48, e2021GL093699, https://doi.org/10.1029/2021GL093699.

    • Search Google Scholar
    • Export Citation
  • 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 29310 298, https://doi.org/10.1073/pnas.1809868115.

    • Search Google Scholar
    • Export Citation
  • Koll, D. D. B., N. Jeevanjee, and N. J. Lutsko, 2023: An analytic model of the clear-sky longwave feedback. J. Atmos. Sci., 80, 19231951, https://doi.org/10.1175/JAS-D-22-0178.1.

    • Search Google Scholar
    • Export Citation
  • 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.

    • Search Google Scholar
    • Export Citation
  • Mlawer, E. J., S. J. Taubman, P. D. Brown, M. J. Iacono, and S. A. Clough, 1997: Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated-k model for the longwave. J. Geophys. Res., 102, 16 66316 682, https://doi.org/10.1029/97JD00237.

    • Search Google Scholar
    • Export Citation
  • Myers, T. A., R. C. Scott, M. D. Zelinka, S. A. Klein, J. R. Norris, and P. M. Caldwell, 2021: Observational constraints on low cloud feedback reduce uncertainty of climate sensitivity. Nat. Climate Change, 11, 501507, https://doi.org/10.1038/s41558-021-01039-0.

    • Search Google Scholar
    • Export Citation
  • Romps, D. M., 2014: An analytical model for tropical relative humidity. J. Climate, 27, 74327449, https://doi.org/10.1175/JCLI-D-14-00255.1.

    • Search Google Scholar
    • Export Citation
  • Romps, D. M., 2017: Exact expression for the lifting condensation level. J. Atmos. Sci., 74, 38913900, https://doi.org/10.1175/JAS-D-17-0102.1.

    • Search Google Scholar
    • Export Citation
  • Seeley, J. T., and N. Jeevanjee, 2021: H2O windows and CO2 radiator fins: A clear-sky explanation for the peak in equilibrium climate sensitivity. Geophys. Res. Lett., 48, e2020GL089609, https://doi.org/10.1029/2020GL089609.

    • Search Google Scholar
    • Export Citation
  • Seidel, S. D., and D. Yang, 2020: The lightness of water vapor helps to stabilize tropical climate. Sci. Adv., 6, eaba1951, https://doi.org/10.1126/sciadv.aba1951.

    • Search Google Scholar
    • Export Citation
  • 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.

    • Search Google Scholar
    • Export Citation
  • Sobel, A. H., J. Nilsson, and L. M. Polvani, 2001: The weak temperature gradient approximation and balanced tropical moisture waves. J. Atmos. Sci., 58, 36503665, https://doi.org/10.1175/1520-0469(2001)058<3650:TWTGAA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • 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, 35043520, https://doi.org/10.1175/2007JCLI2110.1.

    • Search Google Scholar
    • Export Citation
  • Yang, D., 2018a: Boundary layer diabatic processes, the virtual effect, and convective self-aggregation. J. Adv. Model. Earth Syst., 10, 21632176, https://doi.org/10.1029/2017MS001261.

    • Search Google Scholar
    • Export Citation
  • Yang, D., 2018b: Boundary layer height and buoyancy determine the horizontal scale of convective self-aggregation. J. Atmos. Sci., 75, 469478, https://doi.org/10.1175/JAS-D-17-0150.1.

    • Search Google Scholar
    • Export Citation
  • Yang, D., and S. D. Seidel, 2020: The incredible lightness of water vapor. J. Climate, 33, 28412851, https://doi.org/10.1175/JCLI-D-19-0260.1.

    • Search Google Scholar
    • Export Citation
  • Yang, D., and S. D. Seidel, 2023: Vapor buoyancy increases clear-sky thermal emission. Environ. Res. Climate, 2, 015006, https://doi.org/10.1088/2752-5295/acba39.

    • Search Google Scholar
    • Export Citation
  • Yang, D., W. Zhou, and S. D. Seidel, 2022: Substantial influence of vapour buoyancy on tropospheric air temperature and subtropical cloud. Nat. Geosci., 15, 781788, https://doi.org/10.1038/s41561-022-01033-x.

    • Search Google Scholar
    • Export Citation
  • Zelinka, M. D., T. A. Myers, D. T. McCoy, S. Po-Chedley, P. M. Caldwell, P. Ceppi, S. A. Klein, and K. E. Taylor, 2020: Causes of higher climate sensitivity in CMIP6 models. Geophys. Res. Lett., 47, e2019GL085782, https://doi.org/10.1029/2019GL085782.

    • Search Google Scholar
    • Export Citation
  • Zhang, G. J., and N. A. McFarlane, 1995: Sensitivity of climate simulations to the parameterization of cumulus convection in the Canadian Climate Centre general circulation model. Atmos.–Ocean, 33, 407446, https://doi.org/10.1080/07055900.1995.9649539.

    • Search Google Scholar
    • Export Citation
  • Zhang, Y., N. Jeevanjee, and S. Fueglistaler, 2020: Linearity of outgoing longwave radiation: From an atmospheric column to global climate models. Geophys. Res. Lett., 47, e2020GL089235, https://doi.org/10.1029/2020GL089235.

    • Search Google Scholar
    • Export Citation
  • Fig. 1.

    (a) The dependence of midtroposphere temperature at a given pressure level (vertical axis) upon relative humidity (horizontal axis) under WBG dynamics. The solid line indicates an atmosphere with VB, whereas the dashed line indicates an atmosphere with VB. (b) Vertical temperature profiles for a cold climate and hot climate. VB causes the atmosphere to warm more than it otherwise would as ΔTVB increases with climate warming.

  • Fig. 2.

    Temperature difference due to VB (color contours). (a) The ΔTMD-DYN for an equatorial surface temperature of 300 K. (b) The ΔTMD-ALL for an equatorial surface temperature for 300 K. (c) The ΔTwbg for an equatorial surface temperature of 300 K. (d) The ΔTMD-DYN for an equatorial surface temperature of 306 K. (e) The ΔTMD-ALL for an equatorial surface temperature of 306 K. (f) The ΔTwbg for an equatorial surface temperature of 306 K. The black contours represent the temperature of the CNTL simulations.

  • Fig. 3.

    Temperature difference due to VB at the 691-hPa model level, for the simulations with the equatorial surface temperature of 300 K.

  • Fig. 4.

    Equatorial region (±5°) free-troposphere temperature as predicted by moist-adiabatic ascent from the equatorial boundary layer (horizontal axis) plotted against the simulated difference in temperature. Filled circles denote data from the CNTL simulation, and open circles denote MD-ALL.

  • Fig. 5.

    Profiles of lower-troposphere, equator-region (±5°) (a) virtual potential temperature, (b) Δθυ,MD-ALL, (c) temperature, (d) relative humidity with respect to temperature, (e) specific humidity, and (f) moist static energy, for the simulations with equatorial surface temperature 300 K.

  • Fig. 6.

    Difference in climate feedbacks between CNTL and MD simulations. A positive value of Δλ indicates a more negative climate feedback in the CNTL simulation.

  • Fig. 7.

    Difference in net clear-sky OLR at top of atmosphere for (a) the MD-DYN experiment and (b) the MD-ALL experiment. Closed blue circles indicate the radiative effect implied by ΔTυb as simulated by either experiment. Open blue circles indicate the radiative effect implied by ΔTwbg. The red circles indicate the radiative effect implied by differences in specific humidity between the CNTL and MD simulations. The violet circles indicate the sum of the temperature and specific humidity effects. The green crosses indicate the simulated difference in clear-sky OLR. The blue curve indicates a cubic fit to the open circles.

  • Fig. 8.

    Tropical-average VB feedback derived from a hierarchy of climate models. The CRM result is reported in Seidel and Yang (2020), and the simple column result is reported in Yang and Seidel (2020).

  • Fig. 9.

    Global-mean difference in net all-sky radiation (red marks) at top of atmosphere for (a) the MD-DYN experiment and (b) the MD-ALL experiment. Global-mean difference in clear-sky OLR (blue marks) are shown for comparison.

  • Fig. 10.

    Simulated differences in cloud feedbacks due to VB.

  • Fig. A1.

    Difference in zonal-average relative humidity due to VB (color contours); zonal-average CNTL simulation humidity (black contours). (a) The ΔRHT for an equatorial surface temperature of 300 K. (b) The ΔRHρ for an equatorial surface temperature of 300 K. (d) The ΔRHρ for an equatorial surface temperature of 306 K. (e) The ΔRHT and ΔRHρ at 700 hPa.

  • Fig. B1.

    Radiative kernels (W m−2 K−1 100 hPa−1) for (a) temperature and (b) humidity at an equatorial surface temperature of 300 K and for (c) temperature and (d) humidity for an equatorial surface temperature of 306 K. The humidity kernels reflect 1 K of warming at fixed RHT.

All Time Past Year Past 30 Days
Abstract Views 2634 2634 0
Full Text Views 705 705 476
PDF Downloads 200 200 21