1. Introduction
The manner in which temperatures are expected to change regionally with global warming is not only an important scientific question but also a critical issue for climate change impacts and adaptation. Previous studies using climate models have shown that land warms more than the ocean (e.g., Sutton et al. 2007; Joshi et al. 2008, 2013; Byrne and O’Gorman 2013, 2018) and, over tropical land, days at high temperature percentiles warm more than the average, resulting in an elongated upper tail and exacerbated hot extremes (e.g., Duan et al. 2020; Byrne 2021). This “amplified” warming over land seems to be associated with drier conditions: the land is drier than the ocean, and those days at the high percentiles of the temperature distribution are days with dry soils and low relative humidity.
Two main perspectives have been proposed to explain this drier–warmer relationship. One perspective is based on tropical atmospheric dynamics (Joshi et al. 2013; Byrne and O’Gorman 2013, 2018). The inability of the tropical atmosphere to maintain horizontal temperature gradients leads to approximately spatially uniform temperatures in the tropical upper troposphere, i.e., the weak temperature gradients (WTGs; Sobel and Bretherton 2000; Neelin and Zeng 2000). Below, the uniform free-tropospheric temperatures are connected to the surface by a temperature lapse rate closer to a moist adiabat over a moist surface such as the ocean and closer to a dry adiabat over a drier surface such as the land. The larger lapse rate over the drier surface corresponds to higher surface air temperature. Upon warming, the moist adiabatic lapse rate reduces, while the dry adiabatic lapse rate does not change; this, combined with the lack of temperature gradients in the free troposphere, enhances the base-climate temperature gradient across land and ocean and leads to a land–ocean warming contrast (Joshi et al. 2013; the schematics are reproduced in Fig. 1a).
(a),(b) Schematics of cited studies that explain the amplified warming over land with the atmospheric dynamics perspective. In (a) and (b), lines with faint colors show temperature profiles in the base climate and lines with bright colors show temperature profiles in the warm climate. (c) Schematics of the surface flux perspective.
Citation: Journal of Climate 37, 18; 10.1175/JCLI-D-22-0955.1
Building upon this argument, Byrne and O’Gorman (2013) made the arguments of Joshi et al. (2013) more quantitative by using the concept of quasi-equilibrium (QE; Arakawa and Schubert 1974; Emanuel et al. 1994; Neelin and Zeng 2000). QE together with WTG suggests that the temperature profile in the free tropical troposphere approximately follows the moist adiabat with minimal horizontal gradients. Below the cloud base in the boundary layer, the temperature profile follows a dry adiabat (schematics shown in Fig. 1b). The higher the cloud base, the longer the path in the lower troposphere where temperature follows the dry adiabat, and the warmer the surface.
With the theoretical bases of QE and WTG, this framework (the “QE–WTG” framework) can be expressed into mathematical forms by making two assumptions: 1) an atmospheric dynamics constraint that the increase of the boundary layer moist static energy (MSE) across land and ocean is uniform and 2) a moisture constraint that the change of q over land is a fraction γ of that over the ocean. This fraction γ comes from the base-climate ratio of q over land to q over the ocean and is assumed to remain constant with climate change. Because the base-climate humidity ratio γ is on average <1, land moistens less and warms more than the ocean. This QE–WTG framework was initially applied to explain the mean warming contrast over land and ocean (Byrne and O’Gorman 2013, 2018) and has since been adjusted to explain the amplified warming of the hot tail of the temperature distribution over tropical land (Byrne 2021).
In contrast to the atmospheric-dynamics-based perspective is a land surface–based perspective. The surface perspective instead emphasizes changes in the partitioning between sensible and latent heat fluxes in response to moisture availability and/or changes in plants’ physiology (e.g., Seneviratne et al. 2010; Berg et al. 2014; Donat et al. 2017; Vogel et al. 2017; Duan et al. 2020; Dirmeyer et al. 2021). When sufficient water is available for evapotranspiration, a greater portion of the net radiation at the surface (Rn) is released as latent heat (LH), which does not directly increase the surface air temperature (schematic in Fig. 1c). When the soil gets sufficiently dry and/or plants’ stomata close, the ratio of Rn that can be released through LH reduces. Consequently, the surface warms more, and sensible heat (SH) fluxes increase, which warms the near-surface air. In this surface perspective, soil moisture or the Bowen ratio (the ratio of SH to LH) critically controls the warming magnitude.
Although both perspectives involve a partition of the total energy between moistening and warming, there is a gap between the perspectives. The atmospheric perspective is used primarily by the atmospheric dynamics community, and its reasoning relies on constraints that are the end products of atmospheric variables in an equilibrium state without explicit links to the changes in soil moisture or vegetation dynamics that the surface perspective emphasizes. In fact, the moisture constraint in the QE–WTG framework is derived from an “ocean influence model” (Chadwick et al. 2016; Byrne and O’Gorman 2016, 2018). The surface perspective is familiar to the land–atmosphere interaction community and is process-based, but the physical constraints are based on the local surface energy balance, and there is a lack of connection to atmospheric processes and to the global constraints offered by the atmospheric dynamics perspective. Furthermore, the atmospheric perspective has mainly been invoked to explain large-scale (e.g., averaged across zonal bands and/or over land and ocean) and time-mean changes (only recently applied to daily time scales; Byrne 2021); whereas the surface perspective has widely been used for local extreme days (e.g., Donat et al. 2017; Vogel et al. 2017; Duan et al. 2020; Dirmeyer et al. 2021).
In addition, there is a nuanced question regarding whether the climatological dryness or the change in dryness in response to increased CO2 is more important for amplified warming (dryness is measured by specific humidity in the atmospheric perspective and soil moisture in the surface perspective). In the QE–WTG framework, the moisture constraint parameter γ is calculated based on the base-climate specific humidity alone, i.e., the amplified warming can be predicted for regions that are climatologically dry (base-climate γ smaller than one), whereas the surface perspective often emphasizes changes in the soil moisture or surface flux partition with climate change. This difference can cause confusion especially when the decrease of relative humidity (RH) or the increase of climate aridity (measured by P/PET in the cited work) is discussed together with the amplified warming (Sherwood and Fu 2014; Fu and Feng 2014; Byrne and O’Gorman 2018): one may wonder whether more drying leads to more warming or more warming leads to more drying.
To understand the equivalency, or lack thereof, between the two perspectives, we examine the following questions in the context of climate models:
-
Do the key variables in the two perspectives correspond to each other?
-
How do the dryness measures between the two perspectives relate to each other?
-
How do climatological dryness versus changes in dryness contribute to amplified warming in each perspective?
Motivated by these questions, we first compare the full spatiotemporal distribution of key variables and their changes in the two perspectives using a process-based phase space of daily soil moisture and the climatological aridity index, which captures both the temporal variability and the climatological spatial variations in dryness (Duan et al. 2023). In the phase space, we can visualize how each perspective manifests both across climatologically dry and wet regions as measured by the climatological aridity index (AI) and across temporal variability in dryness and wetness within locations that have similar AIs. The data and frameworks we use are introduced in section 2. We discuss the applicability of the QE–WTG framework to explain the spatiotemporal distribution of warming and examine the moisture constraint’s connection with the land surface in section 3. We then use a diagnostic linearized perturbation model (see also Zeppetello et al. 2020; Chan et al. 2022, for applications in land–atmosphere interactions) to discuss the relative roles of the base-climate dryness versus the change in dryness for the magnitude of warming in section 4. This linearized perturbation model also offers an alternative way to view the connection between the two perspectives. We briefly comment on the mechanistic connection of the two perspectives via the association of the lower-tropospheric temperature lapse rate to surface flux partition in section 5 and provide a final summary in section 6.
2. Data and methods
a. Data
We analyze simulations submitted to phase 6 of the Coupled Model Intercomparison Project (CMIP6; Eyring et al. 2016). We take the first 30 years (1850–79) in the historical experiment as the control (base) climate state and years 121–150 in the abrupt 4 × CO2 (4 × CO2 for short) experiment as the perturbed climate state. Climate changes in variables are characterized by subtracting the base state from the perturbed climate state. We use the outputs from nine models that have reported daily values of near-surface (2 m) air temperature (T), specific humidity (q), RH, surface soil moisture (SM, moisture in the top 10-cm soil layer), surface LH flux, surface SH flux, and upwelling and downwelling short- and longwave radiative fluxes from which we calculate the net radiation at the surface (Rn, downward positive), for both the historical and 4 × CO2 experiments. The nine models are CanESM5, CMCC-ESM2, GFDL-CM4, MIROC6, IPSL-CM6A-LR, INM-CM5.0, MPI-ESM1-2-HAM, MPI-ESM1-2-HR, and MRI-ESM2-0. The limitation of these nine specific models mainly comes from the availability of surface soil moisture and surface flux outputs at daily frequency in the 4 × CO2 experiment (e.g., daily SM is not available for CESM2 and daily surface and radiative fluxes are not available for the ACCESS model and EC-EARTH). Besides the limitation due to the data availability, we only include one model from each modeling institution even if outputs from the different versions of that model are available. In this case, we consider the more recent version or the higher-resolution version that has the more complete set of variables available. NorESM2-MM, FGOALS-g3, and Taiwan ESM (TaiESM1) have also reported daily surface soil moisture, but a large amount of the soil moisture values are invalid, and therefore, they are not included. In addition to the variables listed above, precipitation P in the historical experiment is used in order to calculate the climatological AI. AI is defined as AI = 0.8Rn/(LυP). Arid regions have larger AI under this definition, and the constant 0.8 is set empirically to account for the fact that not all available energy goes into evapotranspiration even if soil moisture is abundant (Milly and Dunne 2016; Koster and Mahanama 2012). To calculate the near-surface air MSE (MSE = cpT + Lυq + gz, where cp is the specific heat of dry air, T is the surface air temperature, Lυ is the latent heat of vaporization, q is the surface air specific humidity, g is the gravitational acceleration, and z is the height above sea level), we also use the orographic height variable. Since the orographic height will not change with climate change, we only need this gz term when calculating MSE in one climate state. Temperature and geopotential height on pressure levels (variable ta and zg) during 1850–79 in the historical experiment are used to calculate the temperature lapse rate between 700 hPa and the surface. Since MRI-ESM2 only reports daily 3D variable after 1950 and daily 3D zg from CMCC-ESM2 is not available, we use only the remaining seven models for temperature lapse rates and profiles. We use the gridcell land area fraction variable to select land grid cells as those with a land area fraction ≥99%. Since we are interested in situations where we expect a strong land surface control on temperature, we focus on the warm season, which we define as the 150 days centered on 15 and 16 July for the Northern Hemisphere and the 150 days centered on 15 and 16 January for the Southern Hemisphere.
b. The daily SM/climatological AI phase space
The land surface is highly heterogeneous: it ranges from very arid regions such as the Sahara Desert to very moist regions such as the Amazon rainforest. For a given location, temporal variability of dryness/wetness can substantially affect the surface latent heat flux: LH that is more than 150 W m−2 during average days can reduce to only a few tens of watts per square meter when the soil is dry (see Fig. A1g). Furthermore, when considering changes in these hydroclimatic variables under climate change, more complexity comes into play due to the nonlinear relationship between LH and soil moisture. In particular, for a fixed amount of net radiation at the surface, LH increases with SM in the transitional regime but is not sensitive to SM in the dry and wet regimes (e.g., Seneviratne et al. 2010) and decreases with SM in the active-rain regime (Duan et al. 2023). As a result, when soil dries under climate change, LH decreases in the transitional regime, while it increases in the wet regime due to increases in net radiation (Duan et al. 2023). These different changes in LH can lead to different responses in surface air temperature: warming will be magnified in the transitional regime due to reductions in latent cooling, in contrast to the behavior in the wet regime. Since places with different climatological dryness have different percentages of days in each of the LH–SM regimes, averaging over time that contains both local wet and dry days or averaging over locations with different local soil moisture distributions can lead to substantial canceling of signals and result in an averaged behavior that is hard to interpret. Moreover, models represent the spatiotemporal distribution of soil moisture and the evapotranspiration–soil moisture functions differently, resulting in further uncertainty.
Duan et al. (2023) presented a process-based phase space constructed using daily soil moisture and the climatological aridity index and showed that when organizing the highly heterogeneous land hydroclimatic variables in this phase space, one can acquire coherent patterns of changes across models and across variables. The coherent patterns in this SM/AI phase space are in contrast to the highly uncertain results in the geographical map view [Fig. S3of Duan et al. (2023)]. Here, we adopt this method and process the model fields that have dimensions of calendar days and geographical locations into phase-space grids consisting of 50 temporal bins and 50 spatial bins. The horizontal axis of the phase space is formed by the temporal bins and calculated by sorting the original data according to the daily SM and placing 2% of the 4500 days in each bin. These temporal bins represent local dry days to wet days. The vertical axis of the phase space is formed by the spatial bins and calculated by sorting the original grid cells (for both the base and the 4 × CO2 climate states) according to the base-climate warm-season mean AI and placing 2% of the tropical land area (30°S–30°N) in each bin. These spatial bins represent the climatologically moist regions to climatologically arid regions. We then take the area-weighted average of data in each bin and plot the bin-averaged values as shadings. Figures A1 and 3 show a number of variables and their changes displayed in this SM/AI phase space. We refer to the average in each of the 50 × 50 bins as data “under a certain spatial–temporal condition.” We normalize changes in variables in the phase space by the tropical mean ocean warming
c. The diagnostic linearized model for the magnitude of warming
Both perspectives consider a partitioning of total energy (MSE) or energy flux (Rn) between a temperature component (cpT or SH) and a humidity component (Lυq or LH). Here, we reduce the variables in the balance relations and rearrange them to relate the temperature component to the total energy/energy flux by a partition factor. When considering climate change, we decompose changes in the total energy/energy flux into a component with an unchanged base-state partition factor and a component representing the contribution from changes in this partition factor. This method of decomposition helps us compare the partitioning in the two perspectives and diagnose the relative importance of the base-climate dryness versus the change in dryness under climate change in contributing to the amplified warming over tropical land. Distinguishing between the role of the base climate state and changes to that state is important because information about the base climate state is also available from observations and can be used to make predictions and constrain climate change projections. We derive the relations below.
A summary of the equations and terms considered for the base-climate physics and climate change perturbations in the atmospheric dynamics and the surface flux perspectives.
The terms (cpψ)−1 in Eq. (6) and (κΨ)−1 in Eq. (7) can be understood as the sensitivity of the temperature response to the forcing analog (the total energy/energy flux). The specific heat of dry air, cp, is a physical constant. For simplicity, we also consider κ to be a constant as discussed above. Therefore, in the following sections, we take ψ−1 and Ψ−1 as the sensitivities in each framework, which are solely determined by the base-climate state.
The terms −cpTΔψ and −SHΔΨ are the contributions of the change in the partition factor (ψ and Ψ, the inverse of which is the sensitivity discussed above) to the magnitude of warming, and we refer to them as the “repartition” terms. Specifically, the terms summarize the changes in the partitioning of MSE into T and q for the atmospheric framework or the partitioning of Rn into SH and LH for the surface energy balance framework. The sign of the repartition terms is consistent with their effect on the temperature response: positive values indicate that the balance is repartitioning in a way that will increase the temperature response and decrease the moisture response.
A summary of the balance relations, their linearized perturbation equations, and relevant terms for the two perspectives are listed in Table 1.
3. Comparing the two perspectives across the spatiotemporal distribution
Previous results based on the QE–WTG framework (Byrne and O’Gorman 2013, 2018; Byrne 2021) are often displayed in terms of zonal averages, and results from the surface perspectives (e.g., Donat et al. 2017; Vogel et al. 2017) are often reported as temporal maxima (or the highest percentiles). Here, we compare the full spatiotemporal distribution of key variables in the two perspectives. To do so, we employ the SM/AI phase space as described in section 2. Recall that in this phase space, the horizontal axis captures temporal variability as measured by local daily SM percentiles and the vertical axis captures spatial variability as measured by the climatological AI percentiles. The isoline of a surface soil moisture value of 20 kg m−2 is shown as a black contour (solid for the base climate and dashed for the 4 × CO2 climate). This line roughly indicates the critical soil moisture, SMcrit, that marks the transitional regime in the Budyko curve where soil moisture limits evapotranspiration and the soil moisture–temperature feedback is the strongest. While its specific value can vary across models, time, and region, this value can be used as a rough indicator in CMIP6 models (Duan et al. 2023). To the lower right of this line are wetter conditions, both spatially and temporally, and to the upper left are drier conditions.
a. The atmospheric perspective
In the atmospheric perspective, the change in the total energy (ΔMSE) is split, by definition, between warming (cpΔT) and moistening (LυΔq). Each term in the phase space in CMIP6 models is shown in Figs. 2a–c. The ΔMSE (Fig. 2a) shows an increase across all 50 × 50 temporal and spatial bins with a value ranging from 3.3 to more than 4.2 kJ kg−1 per degree of mean tropical ocean warming. The warming magnitude (ΔT; Fig. 2b) is smaller in moist conditions (lower right), while it is larger in dry conditions (upper left), maximizing over the desert regions. Warming around the critical soil moisture contours in moist regions is also strong, and this amplified warming center around SMcrit is more evident in changes in the daily maximum temperature (not shown). Specific humidity q increases under all conditions, and the magnitude is generally contrary to that of ΔT, i.e., wetter conditions moisten more and warm less, while drier conditions moisten less and warm more.
Changes in surface air (a) MSE; (b) temperature converted to energy units (cpT); (c) specific humidity converted to energy units (Lυq), between the 4 × CO2 climate state and the base climate state displayed in the phase space of daily SM percentiles (horizontal axis) and climatological AI percentiles (vertical axis). (d) Deviation of the MSE over land from the corresponding (corresponding day and latitude) zonal-mean values over the ocean ([MSE]o) in the base climate; (e) deviation of the MSE over land from the corresponding precipitation-weighted zonal-mean values over the ocean (
Citation: Journal of Climate 37, 18; 10.1175/JCLI-D-22-0955.1
In the QE–WTG framework, ΔMSE is generally assumed to be uniform across various divisions of the tropics. For example, Byrne and O’Gorman (2013, 2018) assumed uniform annual-mean changes across land and ocean for a given latitude, and Byrne (2021) showed uniform changes in CMIP6 models for the upper 50 MSE percentiles across time and locations over land and ocean. With this assumption, amplified warming is then predicted in conditions with less moistening. The general “less moistening–more warming” correspondence we see in Figs. 2b and 2c seems to be consistent with the theory of this QE–WTG framework. However, it is interesting to note that we also see the less moistening–more warming relationship over the desert despite the fact that there is almost no moist convection over the desert, so the region is not expected to be in convective quasi-equilibrium as required by the theory. Indeed, as we can see in Fig. 2a, the change in MSE is noticeably smaller in deserts than in other regions, since there is a lack of coupling between the desert boundary layer and other regions that are dominated by moist convection. Thus, the assumption of equal change in MSE that has been applied previously to other groupings of regions in the tropics appears to break down in the phase space we consider for CMIP6 models, particularly in the drier days and regions that warm the most.
To further examine the QE–WTG framework in CMIP6 models regarding the assumption about the uniform MSE in the AI/SM phase space, we compare the MSE at a certain spatiotemporal condition over land to the corresponding day and latitude zonal-mean values over the ocean in the base climate in Fig. 2d. If all conditions satisfy QE and WTG (especially QE, since WTG holds relatively well across locations in the tropics; see Fig. 6c), the values in Fig. 2d should be approximately 0, indicating that MSE over land roughly equals the corresponding value over the ocean. However, Fig. 2d shows >0 values for wet conditions and <0 values for dry conditions. Zhang and Fueglistaler (2020) showed that this equal MSE assumption holds better when conditioned on precipitating situations (we refer to this as the revised equal MSE assumption below). Following their approach, Fig. 2e shows the MSE at each location over land with the corresponding precipitation-weighted zonal-mean values over the ocean (
QE–WTG suggests equal MSE over land and ocean in each of the base and the warm climate states, from which the equal change of MSE over land and ocean is derived. Figure 2f shows changes in the MSE deviation from the precipitation-weighted ocean zonal-mean value over land between the warm and the base climate states, i.e.,
b. The surface perspective
For the surface flux perspective, changes in the total energy flux, i.e., the net radiation at the surface (Rn), are partitioned between changes in the surface sensible heat flux (ΔSH) and changes in the surface latent heat flux (ΔLH), with a negligible change in ground heat fluxes (not shown). Each term in the phase space in CMIP6 models is shown in Figs. 3a–c. The Rn increases under all spatiotemporal conditions of dryness and wetness, with the maximum increase in wet conditions (around 3 W m−2 per degree of ocean warming) and the smallest increase over the desert (around 1 W m−2 per degree of ocean warming). The pattern of ΔRn across the spatiotemporal states is highly similar to the pattern of ΔMSE (Fig. 2a); their correlation across the 2500 bins in the phase space is 0.85. This close correspondence between the changes in the total energy in the surface air and changes in the total energy flux at the surface may be explained by the following two possibilities: one is that increases in Rn drive the increases in MSE; the other is that increases in q (Fig. 2c), which is the dominant component in ΔMSE (Fig. 2a), result in increases in Rn through the increased radiative emissivity of moister air.
Changes in the surface (a) net radiation (Rn), (b) SH flux, and (c) LH flux, and (d) SM, (e) SH plus upwelling longwave radiation (SH + LWup), and (f) RH between the 4 × CO2 climate state and the base climate state, displayed in the phase space of daily SM percentiles (horizontal axis) and climatological AI percentiles (vertical axis). Changes in these variables are normalized by the mean tropical ocean warming
Citation: Journal of Climate 37, 18; 10.1175/JCLI-D-22-0955.1
Except for the very dry conditions where changes in LH are near 0, and the transitional conditions around SMcrit where LH decreases and SH increases, changes in Rn are predominantly balanced by increases in LH. In particular, in the wet conditions (those to the right of SMcrit), LH increases by about 2–3 W m−2 per degree of mean ocean warming, while SH changes much less (or even decreases), by ±1 W m−2 per degree of mean ocean warming. Recall that in these same moist locations and times, increases in specific humidity also dominate the increases in MSE and the warming magnitude is similar to that over the ocean (Fig. 2b).
For the transitional conditions around SMcrit, decreases in LH in CMIP6 models are substantial (Fig. 3c). The decrease in LH is driven by the drying of SM (Fig. 3d; also indicated by shift of the SMcrit isoline to higher percentiles) and is compensated by strong increases in SH. The increase in SH and decrease in LH in the transitional conditions correspond to amplified warming in these conditions (Fig. 2b), reflecting the role of surface flux repartitioning in exacerbating the warming. It is interesting to note, however, that even though LH decreases in these transitional conditions, specific humidity increases (Fig. 2c). This reflects the role of other sources and sinks of surface air specific humidity besides the local land surface evapotranspiration. Meanwhile, RH decreases (Fig. 3f), suggesting that specific humidity has not increased enough to match the increase in the saturation specific humidity. RH is often taken as a reflection of atmospheric dryness; the decrease of RH along SMcrit despite the increase of the specific humidity indicates that soil moisture–temperature feedback is an important contributor to the decrease of RH over land.
In arid regions (upper quarter of the phase space), in the base climate state, LH is small (deserts are dry), and Rn is predominantly balanced by SH. With climate change, changes in LH will be small (deserts remain deserts), and we might expect large increases in SH, corresponding to the pronounced warming (Fig. 2a). However, we see here that changes in SH in these arid regions are also small. This is because both the surface and near-surface air warm substantially, and the increase of longwave radiation (which scales with T4) is faster than the increase of SH (which scales with T); i.e., upwelling longwave radiation (LWup) is partly playing the role in the energy balance that is played by SH in other regions. Adding changes in LWup, Fig. 3e reproduces the pronounced increases in the arid regions. Nevertheless, increases in Rn in these arid regions are small, and the surface energy balance (Rn ≈ SH + LH) is still obeyed. The comparisons in Figs. 3a–c and 3e are consistent with and supplement the discussion of surface flux changes under different regimes of the temperature distribution change in Duan et al. (2020).
c. A discussion on the moisture constraint
As mentioned in sections 1 and 3a, the stronger warming predicted over land in the QE–WTG framework emerges from the moisture constraint that ΔqL = γΔqO, which is derived based on the transport of atmospheric moisture from the ocean (Byrne and O’Gorman 2016; Chadwick et al. 2016). While the derivation of this moisture constraint emphasizes ocean control—specific humidity over land is assumed to follow specific humidity over the ocean, which increases with warming approximately at the rate of the Clausius–Claperon scaling—land specific humidity is also affected by land–atmosphere exchanges (e.g., van der Ent et al. 2010). A drier land surface will tend to result in lower evapotranspiration and, for a given amount of moisture convergence (although the two do not operate independently), a smaller γ. When the moisture constraint is applied to climate model simulations, the value of γ is calculated as the local climatological qL at each land grid cell divided by the corresponding zonal mean qO (Byrne and O’Gorman 2016, 2018) and is spatially variable. Therefore, the moisture constraint (atmospheric perspective) contains information on the spatially variable surface dryness (surface perspective). Note that the results in Byrne and O’Gorman (2016, 2018) were presented as averages over latitudinal bands or the entire 40°S–40°N domain, which is an average across the underlying variability of γ.
We use the same method as in Byrne and O’Gorman (2016, 2018) to calculate γ on each day and present its spatiotemporal distribution in the SM/AI phase space in Fig. 4a. Examining across the spatiotemporal distribution, γ ranges from <0.2 in the very arid regions to about 1.2 in the wettest days (high local SM percentiles) in semiarid and semimoist regions. The lowest row (the lowest AI percentile bin) contains many grid cells in the upwind slope of the Tibetan Plateau where it rains heavily from the orographic lifting; therefore, AI is small (AI ∝ Rn/P), but the specific humidity (and γ) is not necessarily high (since temperature is low at high orographic altitude).
(a) The land–ocean specific humidity ratio [the γ in Byrne and O’Gorman (2013, 2018), calculated as the specific humidity q at each land grid cell divided by the corresponding ocean zonal-mean specific humidity [q]o]. (b) Fractional changes in γ normalized by the mean tropical ocean warming
Citation: Journal of Climate 37, 18; 10.1175/JCLI-D-22-0955.1
Under climate change, the moisture constraint assumes that the ratio γ remains constant, but this may not hold if evapotranspiration changes are substantial and not compensated by changes in moisture convergence. The change in γ is shown in Fig. 4b. We first draw the readers’ eye to the decrease of γ aligning along the critical soil moisture lines where changes in surface conditions are the strongest as emphasized by the surface perspective (see section 3b). In Fig. 4b, for each degree of warming, γ decreases by about 1% in the water-limited transitional regime for the CMIP6 multimodel mean. This suggests that when qO increases by 7% K−1, qL in these conditions, due to the decrease of γ, will increase by 6% K−1 (taking the log and then taking the climate change difference on both sides of qL = γqO yields ΔqL/qL = ΔqO/qO + Δγ/γ). For qL ≈ 16 g kg−1 (LυqL ≈ 40 kJ kg−1; see Fig. A1c), the 1% change in γ will produce a deviation of about 0.4 kJ kg−1 in LυΔqL (which, assuming an accurate constraint of ΔMSE by QE and WTG, leads to a 0.4-K deviation in the prediction of ΔTL), for each degree of ocean mean warming. In applying the QE–WTG framework to predict the magnitude of land warming, this deviation brought by the moisture constraint assumption will partly compensate for the deviation brought by the equal MSE assumption (the MSE increase in these conditions is smaller; see Figs. 2a,f). Note that in these transitional conditions where γ decreases, LH also decreases (Fig. 3c). While we do not perform a moisture budget analysis here, this correspondence suggests the role of changing land-sourced moisture in controlling changes in γ. In Byrne (2021), where the QE–WTG framework is applied to discuss warming of the upper temperature percentiles, both the drier base-climate q and a decrease of RH on those high temperature percentile days contribute to the amplified warming, and both of these have connections to the land surface.
Another domain in the phase space in Fig. 4b that shows perceptible changes in γ is over the arid land regions: γ increases by a much larger fractional rate. The driving mechanism for this increase is not yet clear: at the top soil moisture percentiles in these arid regions, both rainfall and LH (see Fig. 3c) increase; the overall increase in γ and the moistening at the top soil moisture percentiles in these arid regions can result from both changes in the transport of moisture and local evapotranspiration.
Boundary layer RH in the base climate can be effectively regarded as a key indicator for the warming/moistening partition in the atmospheric perspective, similar to γ: while γ is the key parameter in the mathematical form of the QE–WTG framework, the physical intuition for the importance of base climate moisture in controlling warming magnitude draws upon the cloud-base height above which the temperature lapse rate shifts to follow a moist adiabat from a dry adiabat (see Fig. 1b). The cloud-base height is tightly related to the boundary layer RH (Betts 2009), and therefore, the base climate state of both can be regarded as the key control for the warming magnitude under the atmospheric perspective. In fact, when Byrne and O’Gorman (2013) first formed the QE–WTG framework based on equal equivalent potential temperature θe over land and ocean, the decomposition for the land/ocean warming ratio was based on RH. Altogether, the SFE relation in Eq. (9) that directly associates RH with the Bowen ratio is directly relating the key control in the atmospheric perspective to the key control in the surface perspective.
Figure 4c shows the relationship between RH and the Bowen ratio B for each box in the SM/AI phase space, with the color indicating the soil moisture. RH shows an inverse relationship with B, as expected. This inverse relationship suggests a strong land surface origin for the boundary layer RH: the surface influences RH by both controlling the source of water vapor through evapotranspiration and affecting temperature. The red and blue dots in Fig. 4c show the results when assuming SFE [Eq. (9); red dots use RH to calculate B and blue dots use B to calculate RH]. SFE qualitatively captures the inverse relationship between RH and B, but there are notable deviations from what is simulated by the climate models. For example, in wet conditions (green dots in Fig. 4c), SFE-based RH is higher than the simulated RH, while in dry conditions (brown dots in Fig. 4c), SFE-based RH is lower than the simulated RH. These deviations from the climate model may be expected with the assumptions being made by SFE, where atmospheric heat and moisture convergence are not making a net contribution. While these differences between SFE and CMIP6 make sense given the assumptions made in SFE, we have not compared with observations here, and it is possible that the real world may behave differently. For example, Chen et al. (2021) showed that the SFE-based annual-mean evapotranspiration calculated with CMIP6 model surface air temperature, specific humidity, and net radiation compares better with catchment observations than the evapotranspiration simulated by CMIP6 models. While some of these correspondences may seem intuitive, we make them explicit because soil moisture and its changes do not appear directly in the QE–WTG framework, while studies based on the surface flux perspective do not typically rely on the behavior of the large-scale transport of moisture.
4. Base-climate sensitivity and changes in partition in the two perspectives
To what extent can the two perspectives explain the spatiotemporal warming pattern over land in climate models? Recall that in section 2, we derived a relationship between the magnitude of warming in each perspective and three components: changes in the total energy/energy flux, the base-climate partition between the temperature and the moisture component (the inverse of which is termed the base-climate sensitivity), and the changes in this base-climate partition [Eqs. (6) and (7)]. The comparison of various relations for the two perspectives is listed in Table 1. In the following, we examine the spatiotemporal distribution of the base-climate partition and its changes under warming and the contribution of the base-state partition versus its changes to the warming magnitude. This alternative way of decomposition provides us with a complementary view of the two perspectives.
Figures 5a and 5d show the spatiotemporal distribution of the base-climate sensitivity. For both the surface (Fig. 5a) and the atmospheric (Fig. 5d) perspectives, the base-climate sensitivity is larger for drier conditions (upper left) and smaller for wetter conditions (lower right). Recall that in the surface perspective, base-climate sensitivity Ψ−1 = SH/R ≈ SH/Rn is the ratio of the surface sensible heat flux to the total net radiation; and in the atmospheric perspective, base-climate sensitivity ψ−1 = cpT/ME is the ratio of the temperature component of energy (dry specific enthalpy) to the total moist enthalpy. Both reflect the dryness in the base climate because they indicate how much of the available energy/energy flux is partitioned into temperature/heating rather than moisture/evapotranspiration. The correlation between the atmospheric and surface sensitivity is high (0.99; see also Fig. 6a), reflecting the tight connection between the two perspectives in the base climate.
(a),(d) The base-climate sensitivity (1/Ψ and 1/ψ); (b),(e) the repartition term (−SHΔΨ and −cpTΔψ; a positive sign corresponds to a repartition toward enhancing warming); and (c),(f) the relationship between the magnitude of warming ΔT (normalized by mean tropical ocean warming [ΔT]o) and the base climate sensitivity in the (c) surface and the (f) atmospheric perspectives. Small dots in (c) and (f) represent the 50 × 50 spatiotemporal bins and are color-coded by the repartition terms as shown in (b) and (e); large dots in (c) and (f) are the temporally averaged values of warming and base-climate sensitivity for the 50 spatial bins, color-coded by percentiles of the AI.
Citation: Journal of Climate 37, 18; 10.1175/JCLI-D-22-0955.1
The relationship between the base-climate (a) atmospheric sensitivity 1/ψ and surface sensitivity 1/Ψ, (b) temperature lapse rate between the 700-hPa level and the near-surface level (−dT/dz) and surface sensitivity 1/Ψ (the fraction of SH in the total enthalpy flux R), and (c) warm season mean temperature profiles during 1850–79 from CMIP6 models (see section 2 for details), averaged over tropical (30°S–30°N) ocean (blue), tropical land (black), Amazon rainforest (15°S–0°, 70°–55°W; green), and Sahara Desert (20°–30°N, 10°–30°E, orange). The two pink dashed lines in (c) illustrate the temperature profile below 700 hPa if following the (right) dry adiabat (starting with the temperature at 700 hPa over the Sahara Desert and integrating downward) and the temperature profile below 500 hPa if following the (left) moist adiabat (starting with the temperature at 500 hPa averaged over the tropical ocean and integrating downward).
Citation: Journal of Climate 37, 18; 10.1175/JCLI-D-22-0955.1
The surface sensitivity 1/Ψ (Fig. 5a) ranges from 0 to 1: values over the desert are close to 1, indicating that Rn at the surface is almost entirely balanced by SH, and values in very high SM percentiles over moist and semimoist regions are close to 0, indicating that Rn is almost entirely balanced by LH. The atmospheric sensitivity 1/ψ (Fig. 5d), however, has a smaller range and is generally higher than the surface sensitivity (ranges from 0.88 for moist regions to 0.96 for arid regions). This indicates that in one climate state, for a given location, the near-surface dry specific enthalpy (cpT around 300 kJ k−1g; see Fig. A1b in appendix A) dominates the total moist static energy (around 325 in arid regions to 345 kJ kg−1 in moist regions; see Fig. A1a). This is in contrast to the variability across the spatiotemporal distribution and the climate change values, for which the latent energy (Lυq) values dominate those in surface air MSE.
Figures 5b and 5e show the repartition term in Eqs. (6) and (7). The repartition term captures the contribution of changes in the partition factor (ψ and Ψ) to warming. The patterns of the repartition terms −cpTΔψ and −SHΔΨ are dominated by the patterns of changes in the partition factors −ΔΨ and −Δψ (cf. Figs. 5b,e with Figs. B1d,c in appendix B). In the surface perspective, the surface flux partitions toward higher latent heat flux under wet conditions (blue shadings in Fig. 5b) and toward higher sensible heat flux under intermediate (transitional and some wet) conditions (red shadings in Fig. 5b). Under dry conditions, changes in the partition factor are small, as are changes in SH and LH, except at the high soil moisture percentiles associated with rainy conditions.
The behavior of the repartition term in the atmospheric perspective contrasts with its behavior in the surface perspective: the atmospheric partition factor changes toward moistening under all conditions, although the magnitude of the repartition toward moisture is smallest in dry conditions in correspondence with the surface flux perspective (Fig. 5e). The inconsistency of the pattern in Fig. 5e compared to that in Fig. 5b reflects that the atmosphere is responding to changes in land surface ET but is generally getting moister under the influence of ocean as the globe warms. There is also a possible contribution from changes in the evaporation of precipitation since it is a source of boundary layer specific humidity.
Having established the spatiotemporal structure of the key terms in our diagnostic equations, we test their relationship to the warming magnitude. Figures 5c and 5f show the warming magnitude in each of the 50 × 50 bins as a function of the base-climate sensitivity, with the small dots color-coded by the repartition term. This clearly demonstrates the relationship between the warming magnitude and the base-climate sensitivity in both perspectives: higher base-climate sensitivity (drier base-climate conditions) leads to a larger magnitude of warming. The larger dots show the temporally averaged values of warming and base-climate sensitivity across the 50 AI bins, color-coded by the percentile of AI in each spatial bin. Here, we also see that climatologically arid regions warm the most and climatologically moist regions warm the least.
Equations (6) and (7) suggest that if ΔRn and ΔMSE are relatively uniform and the repartition term is close to 0, the warming magnitude will scale linearly with the base-climate sensitivity. In Figs. 5c and 5f, we indeed see the general linear increase of ΔT with 1/Ψ and 1/ψ. However, during the intermediate conditions between the very wet and dry (those with a surface sensitivity of 0.2 ≤ 1/Ψ ≤ 0.8 and an atmospheric sensitivity of 0.88 ≤ 1/ψ ≤ 0.93), there is a spread in ΔT at a given base climate sensitivity. This spread is consistent with the repartition terms (color shading of the dots). In these intermediate conditions, when the surface flux repartitions toward SH (red dots in Fig. 5c), the warming magnitude is amplified. Conversely, when the surface flux repartitions toward the LH (blue dots in Fig. 5c), the warming magnitude is dampened.
The similar and quasi-linear relationship (especially for the time-mean; the large dots in Figs. 5c,f) between the warming magnitude and the base-climate sensitivity in both perspectives indicates that in climate models, the climatological dryness, both for the surface and the atmosphere, largely explains the spatial pattern of the warming magnitude. This is to say, we do not need the drying of soil to explain the approximate spatial distribution of the mean warming. However, the drying of soil and changes in the surface flux partition are important to explain the variability of warming at shorter time scales, especially during intermediate conditions between wet and dry.
5. Discussion: The lower-tropospheric temperature lapse rate and the surface fluxes
The atmospheric perspective is top-down: starting from the same temperature in the free troposphere constrained by WTG and moving downward, a larger mean lapse rate has to be associated with a higher surface air temperature, as well as a larger magnitude of surface air warming for a given change in free-tropospheric temperatures. Figure 6c shows the temperature profiles over a few regions of different surface aridity in CMIP6 models for the base climate. They are from moist to arid: the ocean averaged over the tropics (blue), the Amazon rainforest (green), the land averaged over the tropics (black), and the Sahara Desert (orange). In the free troposphere (above roughly 3 km), the temperature over the different regions is relatively uniform and approximately follows the moist adiabat (the dashed pink line), as assumed in both Joshi et al. (2013) and the QE–WTG framework. In the lower troposphere, the temperature lapse rate is indeed closer to a dry adiabat for more arid regions and closer to a moist adiabat for more moist regions, although in all cases, the simulated lapse rate falls between these two edge cases. Note that the QE–WTG schematic (Fig. 1b; Byrne and O’Gorman 2013) is comparable to the lapse rate schematic (Fig. 1a; Joshi et al. 2013) and the CMIP6-based lapse rates (Fig. 6c) because the higher the cloud base, the longer the path in the lower troposphere where temperature follows the dry adiabat and the larger the mean lapse rate in the lower troposphere. In this context, the lower-tropospheric lapse rate, the cloud-base height, and the boundary layer RH are closely related and variously act as the key control for the warming magnitude in the atmospheric perspective.
The surface perspective is bottom-up: temperature increases more when SH is higher for a given change in Rn. Figure 6b shows a strong positive relationship between the mean temperature lapse rate in the lower troposphere and the ratio of SH to total turbulent flux R (SH/R, our base-climate sensitivity), where R = Rn − G = SH + LH ≈ Rn. The outliers in the figure are the 50 temporal bins in the first AI bin, which are located at the edge of the Tibetan Plateau around Nepal. As mentioned in the context of Fig. 4a, this region has a very small AI due to a large amount of orographic precipitation, but the high orographic altitude makes the temperature lapse rate between 700 hPa and the surface small since the level of 700 hPa is within the boundary layer. We can qualitatively understand the connection between the lapse rate and SH/R as follows: in drier regions, SH is a higher fraction of R, so the energy received by the surface is turbulently diffused into the boundary layer rather than (potentially) transported out as water vapor. In addition, drier regions also have less horizontal moisture convergence in the lower troposphere, which is associated with less rain. Higher SH and lack of rain at the surface result in warmer surface air relative to aloft; combined with WTG in the free troposphere, the warmer surface air over drier regions yields a larger mean temperature lapse rate in the lower troposphere.
6. Summary and outlook
Motivated to better understand the amplified warming over tropical land and its association with dry conditions, we compare the two existing perspectives in the field that aim to explain the “drier–warmer” phenomenon. One is based on the atmospheric dynamics of the tropical troposphere (the atmospheric dynamics perspective), and the other focuses on land surface processes (the surface flux perspective). The atmospheric dynamics perspective uses moist static energy as a constrained quantity and derives the amplified warming from lower (base-climate) specific humidity. The surface flux perspective uses the net radiation at the surface as a constrained total and centers its argument on reduced water availability due to soil drying and/or plant-moderated reductions in transpiration, which results in a change in the partitioning between the sensible and the latent heat fluxes. Based on climate model outputs, we have sought to understand to what extent equivalency can be drawn from the general correspondence of the two perspectives. In addition, we also sought to understand the role of “getting drier” versus “being drier” in association with “getting warmer.”
Returning to the three questions we posed at the beginning: to answer the first question regarding how key variables in the two perspectives correspond to each other, we first presented a comparison of the full spatiotemporal distribution of relevant variables in the two perspectives (Figs. 2 and 3). We have shown that the spatiotemporal distributions of changes in the total (MSE vs Rn) are highly correlated, the spatiotemporal distributions of changes in the warming component (T vs SH) correlate well besides the very dry regions [here, correlation is regained when considering (SH + LWup)], and the spatiotemporal distributions of changes in the moistening component (q vs LH) also show a general correspondence, although LH decreases in the transitional regime, while q increases as a result of additional moisture advected from the ocean. To answer the second question, how the dryness measures between the two perspectives relate to each other, we have shown that the spatiotemporal variability in the atmospheric dryness (q, or γ and RH for variant forms) relates to but cannot be fully explained by the surface dryness (soil moisture, or similarly the Bowen ratio). However, using a ψ (Ψ) parameter, the inverse of which reflects the partition of the total energy (energy flux) into the warming component, dryness in the two perspectives can be well linked (Figs. 5 and 6a). To answer the third question, how climatological dryness versus changes in dryness contribute to amplified warming in each perspective, we used a diagnostic linearized perturbation framework to relate the temperature response to a base-climate sensitivity, changes in the MSE or Rn, and changes in the warming versus moistening partition factor. We showed that the base-climate sensitivity largely explains the warming magnitude across spatiotemporal conditions: drier base-climate conditions have a larger base-climate sensitivity and experience a larger magnitude of warming. This relationship between the warming magnitude and the base-climate sensitivity holds particularly well for the time mean, and the surface and the atmospheric perspectives correspond well in terms of this role of the base-climate sensitivity. On top of the quasi-linear relationship between the warming magnitude and the base-climate sensitivity, changes in the partition factor between warming and moistening for intermediate (between wet and dry) conditions further enhance or dampen the warming magnitude (Fig. 5).
Last, by linking the lower-tropospheric lapse rate with the surface fluxes (Fig. 6), we provide mechanistic insights on how the top-down atmospheric perspective connects with the bottom-up surface perspective. The comparison of the two perspectives in our study reveals how key arguments of the two perspectives hold across the spatiotemporal conditions and advances our understanding of the drier–warmer relationship. The results we show are based on CMIP6 model output, which have known biases, particularly in their land–atmosphere interactions that can lead them to differ from observations in important ways (e.g., Phillips et al. 2017; Chen et al. 2021; Hsu and Dirmeyer 2022; Simpson et al. 2024). Therefore, the specific terms and their relationships presented here may differ between CMIP6 models and observations. In addition, internal variability (e.g., McKinnon and Deser 2018; Santer et al. 2019; Deser 2020) and transient anthropogenic forcings (e.g., Bonfils et al. 2008; Marvel et al. 2019; Persad et al. 2023) can confound the signals in observations and historical simulations compared to the signals from the strong idealized 4 × CO2 forcing that our results are based on. The linearized perturbation model we derive is diagnostic, and our analyses aim to compare and connect how the two perspectives explain the general spatiotemporal distribution of the warming magnitude over tropical land, instead of making predictions. Making predictions for the warming magnitude in specific regions can involve a more accurate evaluation of the intricate coupling between temperature and radiation (Vargas Zeppetello et al. 2019; Terray 2023), as well as further quantifying the contribution from atmospheric dynamics (Röthlisberger and Papritz 2023).
Acknowledgments.
The authors acknowledge the support from the NSF Grant 1939988. They thank three anonymous reviewers for their helpful reviews and the editor for handling the manuscript. S. Q. D. appreciates helpful conversations with Dr. Zhihong Tan and Dr. Mengxi Wu. The authors also acknowledge the World Climate Research Programme, which, through its Working Group on Coupled Modelling, coordinated and promoted CMIP6. They thank the climate modeling groups for producing and making available their model output, the Earth System Grid Federation (ESGF), for archiving the data and providing access, and the multiple funding agencies that support CMIP6 and ESGF. I. R. S. acknowledges funding from the National Center for Atmospheric Research, which is a major facility sponsored by the National Science Foundation under the Cooperative Agreement 1852977.
Data availability statement.
CMIP6 model outputs used in this study can be downloaded from the CMIP6 data archive (https://esgf-node.llnl.gov/search/cmip6/). ERA5 reanalysis data can be downloaded from the ECMWF Climate Data Store (https://cds.climate.copernicus.eu/). Jupyter Notebooks for analyzing the data will be shared through GitHub upon the acceptance of this paper.
APPENDIX A
Spatiotemporal Distribution of Relevant Variables in the Base Climate State
Figure A1 shows the base-climate spatiotemporal distribution of relevant variables in the atmospheric dynamics and the surface flux perspectives, presented in the phase space of daily soil moisture percentiles and climatological aridity index percentiles.
The spatiotemporal distribution of the base-climate surface air (a) MSE, (b) temperature converted to energy units (cpT), (c) specific humidity converted to energy units (Lυq), and (d) RH and the surface (e) net radiation (Rn), (f) SH flux, (g) LH flux, and (h) SM in the phase space of daily SM percentile and climatological AI percentile. The black contour shows the critical SM of 20 kgm−2.
Citation: Journal of Climate 37, 18; 10.1175/JCLI-D-22-0955.1
APPENDIX B
Additional Information for Terms in the Diagnostic Perturbation Equations
Figure B1 shows the sanity tests for the linearization in Eqs. (3) and (4) and the sensitivity tests for the repartition terms. The fact that the patterns in Figs. B1c and B1d are similar to patterns in Figs. 5b and 5e indicates that the spatiotemporal distribution of changes in the partition factor (−Δψ and −ΔΨ) dominates the spatiotemporal distribution of the repartition terms (−cpTΔψ and −SHΔΨ).
(a) The residuals from the linearization in Eq. (3), (b) the residuals from the linearization in Eq. (4), (c) changes in the atmospheric partition factor −Δψ, and (d) changes in the surface partition factor −ΔΨ.
Citation: Journal of Climate 37, 18; 10.1175/JCLI-D-22-0955.1
APPENDIX C
The Relationship between Perturbations in the Surface Sensible Heat Flux and Perturbations in Temperature
In the perturbation relation we use for the surface energy balance [Eq. (5)], we have linearized the perturbation in the surface sensible heat (SH) flux as a coefficient κ times the temperature perturbation. Figure C1 examines to what extent this linearization is justifiable.
The relationship between (a) the base-climate surface SH flux and surface air temperature (T); (b) the base-climate temporal anomalies in SH and T, with the local temporal-mean value removed; (c) the local temporal-mean SH and T; (d) the climate changes in SH and T; (e) the temporal anomalies of changes in SH and T with the local temporal-mean change removed; and (f) the local temporal-mean changes in SH and T. All scatters are color-coded by the percentile of the AI. In (a), (b), (d), and (e), each scatter represents one of the 50 × 50 spatiotemporal bins, and in (c) and (f), each scatter represents one of the 50 spatial bins, with the temporal dimension averaged out.
Citation: Journal of Climate 37, 18; 10.1175/JCLI-D-22-0955.1
The first row shows the relationships between SH and T in the base climate. Panel (b) informs us that temporal perturbations in SH and T relate to each other linearly by a similar coefficient (approximately 15–20 W m−2 K−1) across locations, except in very dry locations. Panel (c) informs us that spatial anomalies of the mean SH and T also relate to each other linearly except in the very moist and very dry locations.
The second row shows the relationships between changes in SH and T under climate change. For our particular application of the linearized perturbation relation Eq. (5), panel (d) is the most relevant. It shows us that for moist and semimoist conditions, ΔSH and ΔT have a relatively good linear relationship, while the relationship for dry regions is weaker. If we remove the mean at each location and only focus on the temporal anomalies, then all locations show a linear relationship [panel (e)], with the regression coefficient, i.e., κ, being approximately 5 W m−2 K−1 for dry locations and 10 W m−2 K−1 for moist locations. There is no linear relationship between the time-mean change in SH and T across locations [panel (f)].
Since, in our analysis, the data we use retain the temporal variability and we emphasize the qualitative understanding, we accept simply linearizing ΔSH as κΔT. With this linearization, we can derive Eq. (7), and we have the surface sensitivity as 1/(κΨ). We assume κ is the same constant across spatiotemporal conditions and focus on the sensitivity factor 1/Ψ. The nonuniformity of κ will make quantitative predictions of the warming magnitude more complicated although here we see that it may be approached by separating dry and moist conditions.
REFERENCES
Arakawa, A., and W. H. Schubert, 1974: Interaction of a cumulus cloud ensemble with the large-scale environment, Part I. J. Atmos. Sci., 31, 674–701, https://doi.org/10.1175/1520-0469(1974)031<0674:IOACCE>2.0.CO;2.
Berg, A., B. R. Lintner, K. L. Findell, S. Malyshev, P. C. Loikith, and P. Gentine, 2014: Impact of soil moisture–atmosphere interactions on surface temperature distribution. J. Climate, 27, 7976–7993, https://doi.org/10.1175/JCLI-D-13-00591.1.
Betts, A. K., 2009: Land-surface-atmosphere coupling in observations and models. J. Adv. Model. Earth Syst., 1 (3), https://doi.org/10.3894/JAMES.2009.1.4.
Bonfils, C., P. B. Duffy, B. D. Santer, T. M. L. Wigley, D. B. Lobell, T. J. Phillips, and C. Doutriaux, 2008: Identification of external influences on temperatures in California. Climatic Change, 87, 43–55, https://doi.org/10.1007/s10584-007-9374-9.
Byrne, M. P., 2021: Amplified warming of extreme temperatures over tropical land. Nat. Geosci., 14, 837–841, https://doi.org/10.1038/s41561-021-00828-8.
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.
Byrne, M. P., and P. A. O’Gorman, 2016: Understanding decreases in land relative humidity with global warming: Conceptual model and GCM simulations. J. Climate, 29, 9045–9061, https://doi.org/10.1175/JCLI-D-16-0351.1.
Byrne, M. P., and P. A. O’Gorman, 2018: Trends in continental temperature and humidity directly linked to ocean warming. Proc. Natl. Acad. Sci. USA, 115, 4863–4868, https://doi.org/10.1073/pnas.1722312115.
Chadwick, R., P. Good, and K. Willett, 2016: A simple moisture advection model of specific humidity change over land in response to SST warming. J. Climate, 29, 7613–7632, https://doi.org/10.1175/JCLI-D-16-0241.1.
Chan, D., A. Rigden, J. Proctor, P. W. Chan, and P. Huybers, 2022: Differences in radiative forcing, not sensitivity, explain differences in summertime land temperature variance change between CMIP5 and CMIP6. Earth’s Future, 10, e2021EF002402, https://doi.org/10.1029/2021EF002402.
Chen, S., K. A. McColl, A. Berg, and Y. Huang, 2021: Surface flux equilibrium estimates of evapotranspiration at large spatial scales. J. Hydrometeor., 22, 765–779, https://doi.org/10.1175/JHM-D-20-0204.1.
Deser, C., 2020: Certain uncertainty: The role of internal climate variability in projections of regional climate change and risk management. Earth’s Future, 8, e2020EF001854, https://doi.org/10.1029/2020EF001854.
Dirmeyer, P. A., G. Balsamo, E. M. Blyth, R. Morrison, and H. M. Cooper, 2021: Land-atmosphere interactions exacerbated the drought and heatwave over northern Europe during summer 2018. AGU Adv., 2, e2020AV000283, https://doi.org/10.1029/2020AV000283.
Donat, M. G., A. J. Pitman, and S. I. Seneviratne, 2017: Regional warming of hot extremes accelerated by surface energy fluxes. Geophys. Res. Lett., 44, 7011–7019, https://doi.org/10.1002/2017GL073733.
Duan, S. Q., K. L. Findell, and J. S. Wright, 2020: Three regimes of temperature distribution change over dry land, moist land, and oceanic surfaces. Geophys. Res. Lett., 47, e2020GL090997, https://doi.org/10.1029/2020GL090997.
Duan, S. Q., K. L. Findell, and S. A. Fueglistaler, 2023: Coherent mechanistic patterns of tropical land hydroclimate changes. Geophys. Res. Lett., 50, e2022GL102285, https://doi.org/10.1029/2022GL102285.
Emanuel, K. A., J. David Neelin, and C. S. Bretherton, 1994: On large-scale circulations in convecting atmospheres. Quart. J. Roy. Meteor. Soc., 120, 1111–1143, https://doi.org/10.1002/qj.49712051902.
Eyring, V., S. Bony, G. A. Meehl, C. A. Senior, B. Stevens, R. J. Stouffer, and K. E. Taylor, 2016: Overview of the Coupled Model Intercomparison Project Phase 6 (CMIP6) experimental design and organization. Geosci. Model Dev., 9, 1937–1958, https://doi.org/10.5194/gmd-9-1937-2016.
Fu, Q., and S. Feng, 2014: Responses of terrestrial aridity to global warming. J. Geophys. Res. Atmos., 119, 7863–7875, https://doi.org/10.1002/2014JD021608.
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.
Hsu, H., and P. A. Dirmeyer, 2022: Deconstructing the soil moisture–latent heat flux relationship: The range of coupling regimes experienced and the presence of nonlinearity within the sensitive regime. J. Hydrometeor., 23, 1041–1057, https://doi.org/10.1175/JHM-D-21-0224.1.
Joshi, M. M., J. M. Gregory, M. J. Webb, D. M. Sexton, and T. C. Johns, 2008: Mechanisms for the land/sea warming contrast exhibited by simulations of climate change. Climate Dyn., 30, 455–465, https://doi.org/10.1007/s00382-007-0306-1.
Joshi, M. M., F. H. Lambert, and M. J. Webb, 2013: An explanation for the difference between twentieth and twenty-first century land–sea warming ratio in climate models. Climate Dyn., 41, 1853–1869, https://doi.org/10.1007/s00382-013-1664-5.
Kong, W., K. A. McKinnon, I. R. Simpson, and M. M. Laguë, 2023: Understanding responses of summer continental daily temperature variance to perturbations in the land surface evaporative resistance. J. Climate, 36, 1653–1678, https://doi.org/10.1175/JCLI-D-21-1011.1.
Koster, R. D., and S. P. P. Mahanama, 2012: Land surface controls on hydroclimatic means and variability. J. Hydrometeor., 13, 1604–1620, https://doi.org/10.1175/JHM-D-12-050.1.
Marvel, K., B. I. Cook, C. J. Bonfils, P. J. Durack, J. E. Smerdon, and A. P. Williams, 2019: Twentieth-century hydroclimate changes consistent with human influence. Nature, 569, 59–65, https://doi.org/10.1038/s41586-019-1149-8.
McColl, K. A., and A. J. Rigden, 2020: Emergent simplicity of continental evapotranspiration. Geophys. Res. Lett., 47, e2020GL087101, https://doi.org/10.1029/2020GL087101.
McColl, K. A., G. D. Salvucci, and P. Gentine, 2019: Surface flux equilibrium theory explains an empirical estimate of water-limited daily evapotranspiration. J. Adv. Model. Earth Syst., 11, 2036–2049, https://doi.org/10.1029/2019MS001685.
McKinnon, K. A., and C. Deser, 2018: Internal variability and regional climate trends in an observational large ensemble. J. Climate, 31, 6783–6802, https://doi.org/10.1175/JCLI-D-17-0901.1.
Milly, P. C. D., and K. A. Dunne, 2016: Potential evapotranspiration and continental drying. Nat. Climate Change, 6, 946–949, https://doi.org/10.1038/nclimate3046.
Neelin, J. D., and N. Zeng, 2000: A quasi-equilibrium tropical circulation model—Formulation. J. Atmos. Sci., 57, 1741–1766, https://doi.org/10.1175/1520-0469(2000)057<1741:AQETCM>2.0.CO;2.
Persad, G., and Coauthors, 2023: Rapidly evolving aerosol emissions are a dangerous omission from near-term climate risk assessments. Environ. Res. Climate, 2, 032001, https://doi.org/10.1088/2752-5295/acd6af.
Phillips, T. J., and Coauthors, 2017: Using arm observations to evaluate climate model simulations of land-atmosphere coupling on the U.S. Southern Great Plains. J. Geophys. Res. Atmos., 122, 11 524–11 548, https://doi.org/10.1002/2017JD027141.
Roe, G., 2009: Feedbacks, timescales, and seeing red. Annu. Rev. Earth Planet. Sci., 37, 93–115, https://doi.org/10.1146/annurev.earth.061008.134734.
Röthlisberger, M., and L. Papritz, 2023: Quantifying the physical processes leading to atmospheric hot extremes at a global scale. Nat. Geosci., 16, 210–216, https://doi.org/10.1038/s41561-023-01126-1.
Santer, B. D., J. C. Fyfe, S. Solomon, J. F. Painter, C. Bonfils, G. Pallotta, and M. D. Zelinka, 2019: Quantifying stochastic uncertainty in detection time of human-caused climate signals. Proc. Natl. Acad. Sci. USA, 116, 19 821–19 827, https://doi.org/10.1073/pnas.1904586116.
Seneviratne, S. I., T. Corti, E. L. Davin, M. Hirschi, E. B. Jaeger, I. Lehner, B. Orlowsky, and A. J. Teuling, 2010: Investigating soil moisture–climate interactions in a changing climate: A review. Earth-Sci. Rev., 99, 125–161, https://doi.org/10.1016/j.earscirev.2010.02.004.
Sherwood, S., and Q. Fu, 2014: A drier future? Science, 343, 737–739, https://doi.org/10.1126/science.1247620.
Simpson, I. R., K. A. McKinnon, D. Kennedy, D. M. Lawrence, F. Lehner, and R. Seager, 2024: Observed humidity trends in dry regions contradict climate models. Proc. Natl. Acad. Sci. USA, 121, e2302480120, https://doi.org/10.1073/pnas.2302480120.
Sobel, A. H., and C. S. Bretherton, 2000: Modeling tropical precipitation in a single column. J. Climate, 13, 4378–4392, https://doi.org/10.1175/1520-0442(2000)013<4378:MTPIAS>2.0.CO;2.
Sutton, R. T., B. Dong, and J. M. Gregory, 2007: Land/sea warming ratio in response to climate change: IPCC AR4 model results and comparison with observations. Geophys. Res. Lett., 34, L02701, https://doi.org/10.1029/2006GL028164.
Terray, L., 2023: A storyline approach to the June 2021 northwestern North American heatwave. Geophys. Res. Lett., 50, e2022GL101640, https://doi.org/10.1029/2022GL101640.
van der Ent, R. J., H. H. G. Savenije, B. Schaefli, and S. C. Steele-Dunne, 2010: Origin and fate of atmospheric moisture over continents. Water Resour. Res., 46, W09525, https://doi.org/10.1029/2010WR009127.
Vargas Zeppetello, L. R., and D. S. Battisti, 2020: Projected increases in monthly midlatitude summertime temperature variance over land are driven by local thermodynamics. Geophys. Res. Lett., 47, e2020GL090197, https://doi.org/10.1029/2020GL090197.
Vargas Zeppetello, L. R., A. Donohoe, and D. S. Battisti, 2019: Does surface temperature respond to or determine downwelling longwave radiation? Geophys. Res. Lett., 46, 2781–2789, https://doi.org/10.1029/2019GL082220.
Vogel, M. M., R. Orth, F. Cheruy, S. Hagemann, R. Lorenz, B. J. J. M. van den Hurk, and S. I. Seneviratne, 2017: Regional amplification of projected changes in extreme temperatures strongly controlled by soil moisture-temperature feedbacks. Geophys. Res. Lett., 44, 1511–1519, https://doi.org/10.1002/2016GL071235.
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.
Zeppetello, L. R. V., É. Tétreault-Pinard, D. S. Battisti, and M. B. Baker, 2020: Identifying the sources of continental summertime temperature variance using a diagnostic model of land–atmosphere interactions. J. Climate, 33, 3547–3564, https://doi.org/10.1175/JCLI-D-19-0276.1.
Zhang, Y., and S. Fueglistaler, 2020: How tropical convection couples high moist static energy over land and ocean. Geophys. Res. Lett., 47, e2019GL086387, https://doi.org/10.1029/2019GL086387.