1. Introduction
Observational studies suggest that the Hadley circulation in the tropics has widened over the past 30 years (Hu and Fu 2007; Seidel and Randel 2007; Seidel et al. 2008), although there is a wide spread of the magnitude of the widening among datasets (Adam et al. 2014). The widening has been qualitatively reproduced in general circulation model (GCM) simulations of global warming scenarios. GCM simulations agree that the terminus of the Hadley cells shifts poleward under global warming, but they disagree on the magnitude of this shift (e.g., Lu et al. 2007). The shift is accompanied by a poleward shift of the subtropical dry zones (e.g., Lu et al. 2007) and midlatitude storm tracks (e.g., Yin 2005; Barnes and Polvani 2013; Simpson et al. 2014). At the same time, the Hadley circulation has been observed to contract during El Niño and to expand during La Niña (e.g., Seager et al. 2003; Lu et al. 2008; Nguyen et al. 2013; Adam et al. 2014).
There is no convincing account of what causes the widening trend or ENSO variations of the Hadley circulation; a theory describing the width of the Hadley circulation on an “Earth like” planet is lacking (Schneider 2006; Schneider et al. 2010). The now-prevalent hypothesis is that the Hadley circulation terminates where baroclinic eddies in some sense start controlling the dynamics. Several arguments of how this may occur have been put forward. The simplest version of this argument is based on the linear stability criterion for the quasigeostrophic (QG) two-layer model and states that the Hadley circulation extends up to the latitude at which angular momentum–conserving axisymmetric flow becomes baroclinically unstable (Held 2000). This argument, and its quantitative consequences, has been widely used in the recent literature to account for the poleward shift of the Hadley circulation terminus in comprehensive GCM simulations (Lu et al. 2007; Frierson et al. 2007). Yet the resulting Hadley circulation extent compares poorly with idealized GCM simulations over a much wider range of climates than that sampled in the comprehensive GCM simulations (Walker and Schneider 2006; Korty and Schneider 2008). Moreover, its core assumptions—angular momentum–conserving flow in the upper branch of the Hadley circulation and the existence of a critical shear for baroclinic instability—are violated in Earth’s atmosphere (Schneider 2006; Zurita-Gotor and Lindzen 2007).
A modified version of this argument is based on a diffusive eddy mixing model and posits that the Hadley circulation extends to the latitude where baroclinic eddies become deep enough to reach the upper troposphere and where, as a consequence, the eddy flux divergence of angular momentum changes sign (Schneider and Walker 2006; Korty and Schneider 2008). Determining this latitude from a supercriticality criterion accounts broadly for Hadley circulation changes in simulations with a dry idealized GCM (Korty and Schneider 2008). But applying this criterion to an atmosphere with an active hydrologic cycle has remained challenging (Schneider and O’Gorman 2008). More recently, O’Gorman (2011) suggested a way to include moisture effects in the depth scaling of baroclinic eddies. This modified scaling was found to capture changes in the extent of the Hadley circulation in global warming simulations with an idealized GCM with an active hydrologic cycle (O’Gorman 2011). However, these recent findings have raised new questions: for example, whether the supercriticality in dry and moist GCM simulations can be represented by a uniform framework, despite the differences in which diabatic heating occurs. It is unclear to what extent convective heating in dry GCM simulations affects baroclinic eddies and whether an effective static stability similar to that of O’Gorman (2011) captures its effects. Furthermore, it is unclear whether accounting solely for vertical convection can capture the effects of diabatic heating on baroclinic eddies or whether other processes, such as slantwise convection, need to be represented explicitly as well (Korty and Schneider 2007).
Here, we describe a modified supercriticality criterion, which is very similar to that defined by O’Gorman (2011), and we test its relevance over a wide range of climates simulated with two idealized GCMs. We show that this criterion can discriminate between regions dominated by baroclinic eddy or convective activity and that it constrains changes in the Hadley circulation extent with climate when baroclinic wave activity is strong in the extratropics.
2. Heuristic arguments
The simplest model of the Hadley circulation is that of an angular momentum–conserving axisymmetric overturning circulation that is energetically closed (Schneider 1977; Held and Hou 1980). Held and Hou (1980) showed that such a Hadley circulation extends to a finite latitude, beyond which radiative–convective equilibrium prevails. In our simulations, the prediction of the Hadley cell extent obtained from this dynamical model compares poorly with the actual latitude of the Hadley circulation terminus (see appendix C). This is unsurprising because the tropical circulation is neither angular momentum conserving nor energetically closed (Walker and Schneider 2006; Schneider 2006; Trenberth and Stepaniak 2003).
Axisymmetric flows were not originally viewed as accounting for the behavior of Earth’s Hadley circulation, but as providing a basic state for studies of baroclinic instability (e.g., Schneider 1977). The zonal wind speeds consistent with an angular momentum–conserving mean flow would be large enough in the subtropics to be linearly unstable to baroclinic instability (e.g., Phillips 1954). This led to the notion that the axisymmetric Hadley circulation may extend up to the latitude where baroclinic instability “sets in” (Held 2000). The analytical relation for an angular momentum–conserving axisymmetric Hadley circulation terminated by linear baroclinic instability has been used in the recent literature to explain the poleward shift of the Hadley circulation terminus in comprehensive GCM simulations (e.g., Lu et al. 2007; Frierson et al. 2007; Kang and Lu 2012). But a quantitative comparison of the resulting Hadley circulation extent with that in our idealized GCM simulations shows significant discrepancies (see appendix C), consistent with previous findings (Walker and Schneider 2006; Korty and Schneider 2008). This is again not surprising, because core assumptions, such as the tropical upper troposphere being in an angular momentum–conserving regime, are usually violated in Earth’s atmosphere (Walker and Schneider 2006; Schneider 2006).
More generally, the Hadley circulation terminus may be considered as the equatorward boundary of the region where baroclinic eddies become deep enough to reach the upper troposphere (Fig. 1). Where they reach the upper troposphere, wave activity no longer propagates upward but horizontally, implying angular momentum flux convergence into the latitude band where it is generated (Edmon et al. 1980; Held and Hoskins 1985; Vallis 2006; Ait-Chaalal and Schneider 2015). Because the Rossby number in the descending branch of the Hadley circulation and poleward of it is generally small (Walker and Schneider 2006; Levine and Schneider 2011), this implies equatorward mean meridional flow where baroclinic eddies are generated and reach the upper troposphere. In this view, the terminus of the Hadley circulation defines the equatorward margin of wave activity generation and upward propagation into the upper troposphere (Korty and Schneider 2008).
a. Supercriticality in dry atmospheres
In regions where baroclinic wave activity strongly affects the mean flow and thus sets the depth of the troposphere, Schneider and Walker (2006) showed that for Earth-like climates,
Using an idealized dry GCM like the one that we use here, Schneider and Walker (2006) found that
Supercritical climates with
The supercriticality (1) with
These results for dry atmospheres suggest that one may use
b. Effective static stability
Schneider and Walker (2006) derived the supercriticality (1) neglecting moisture effects on baroclinic eddies. However, latent heat release in large-scale condensation and moist convection are known to affect the intensity and structure of baroclinic eddies (e.g., Emanuel et al. 1987; Gutowski et al. 1992; Lapeyre and Held 2004), as well as the tropospheric thermal stratification (Korty and Schneider 2007). Using an idealized moist GCM, Schneider and O’Gorman (2008) found a considerable effect of moisture on
In a moist atmosphere, the effective static stability measure (10) is always smaller than the dry static stability measure (3), consistent with the added buoyancy provided by condensation and latent heat release. In a dry atmosphere, the static stability
There is no significant difference in the moist simulations between our estimates of the effective static stability measure (10) and that defined by O’Gorman (2011). Indeed, relation (10) is a specific case of that in O’Gorman (2011), because saturated equivalent entropy is conserved in saturated updrafts whether or not it is associated with vertical convection. But the scaling devised by O’Gorman (2011) cannot easily be tested in dry GCM simulations, because an equivalent potential temperature cannot easily be defined there. This contrasts with our scaling, which can be tested in dry GCM simulations using the prescribed convective lapse rate. The more restrictive interpretation of our effective static stability measure (10) allows for a wider applicability when compared to that defined in O’Gorman (2011). But both estimates are nearly identical when both versions of effective static stability are applicable, such as in moist GCM simulations.
With the effective static stability measure (10), a heuristic effective supercriticality
3. Idealized GCMs
We examine whether the effective supercriticality and the constraint
a. Idealized dry GCM
The idealized dry GCM is described in Schneider and Walker (2006). It simulates an atmosphere bounded by a spatially uniform and thermally insulating spherical surface. A spectral dynamical core solves for the large-scale motions in the atmosphere, with a resolution of T85 in all experiments. The vertical coordinate is discretized with 30 sigma levels. Momentum and dry entropy are diffused in a boundary layer of fixed height (2.5 km), with a turbulent Prandtl number of 1 (Smagorinsky et al. 1965). Frictional dissipation at the surface is parameterized by a bulk aerodynamic formula. Radiative heating is represented by Newtonian relaxation of potential temperatures toward a zonally symmetric radiative equilibrium profile, which is quadratic in the cosine of latitude and statically unstable to dry convection. The time scale for this Newtonian relaxation is zonally symmetric and varies with latitude and height from 7 days at the equator near the surface to 50 days in the interior atmosphere (Schneider 2004). Convection is parameterized by a quasi-equilibrium convection scheme. The convective lapse rate is varied by rescaling the dry adiabatic lapse rate by a rescaling factor γ ≤ 1. Using a reduced convective lapse rate mimics changes in the stability of a convective atmosphere due to an increase in the near-surface moisture content: decreasing γ from its dry adiabatic value of 1 is similar to increasing the moisture content just above surface, which usually happens in a moist atmosphere as a response to warmer surface temperatures.
We simulated a wide range of climates by varying the equator-to-pole temperature contrast in radiative equilibrium from 15 to 300 K and the convective lapse rate rescaling factor from 0.4 to 1.0. All 96 simulations were run for 2000 days. Statistics were accumulated after a statistically steady state was reached: they are averages over the last 600 days, sampled four times daily.
b. Idealized moist GCM
The idealized moist GCM is described in O’Gorman and Schneider (2008) and is similar to that of Frierson (2007). It simulates an atmosphere with an interactive hydrologic cycle, over a uniform ocean surface with a thermal inertia equivalent to 1 m of water. The shortwave albedo of the planet is set to a uniform value of 0.38. Other constants (insolation, rotation rate, gravity, etc.) are kept the same as on Earth. A spectral dynamical core solves for large-scale motions in the atmosphere with a resolution of T42. This lower horizontal resolution was chosen to reduce computational time. We reran a small number of simulations at T85 resolution; we found no significant differences from the results prescribed here at this higher resolution. As in the dry GCM, the vertical coordinate is discretized with 30 sigma levels. Water vapor is advected by the flow, and it condenses whenever saturation occurs on the grid scale. Momentum, moisture, and dry entropy fluxes at the surface are parameterized by bulk aerodynamic formulas; a k-profile boundary layer scheme similar to Troen and Mahrt (1986) represents vertical transport by subgrid-scale motions. Moist convection is parameterized by a quasi-equilibrium convection scheme, which is like that in Frierson (2007), with the modifications described in O’Gorman and Schneider (2008). Radiative heating rates are computed for a gray atmosphere with no clouds, in which a prescribed longwave optical thickness roughly accounts for longwave absorption by water vapor and well-mixed greenhouse gases. The GCM is forced by an approximation of annual-mean insolation at the top of the atmosphere. This insolation is steady (i.e., there is no diurnal or seasonal cycle). The reference longwave optical thickness profile is chosen to lead to an extratropical climate resembling Earth’s in the annual mean (O’Gorman and Schneider 2008). To obtain climate scenarios with a tropical climate and Hadley circulation resembling Earth’s, it is necessary to account for ocean heat transport in low latitudes. We use a simple representation of ocean heat transport, which couples the heat transport to surface temperature gradients and surface wind stress (Klinger and Marotzke 2000; Levine and Schneider 2011).
We simulate a wide range of climates by varying the longwave optical depth and the insolation profile, in climate scenarios with and without the coupled ocean heat transport. The reference profile for the longwave optical depth is rescaled by a constant factor α to mimic changes in greenhouse gas concentrations; we use seven different rescaling factors, ranging from α = 0.2 to 4.0. The insolation profile is varied by varying the pole-to-equator insolation contrast, while keeping the global mean fixed. This is achieved by varying the nondimensional factor Δ in the top-of-atmosphere insolation
List of all configurations for moist GCM simulations (with or without ocean heat flux). Global-mean surface temperature and equator-to-pole contrast are shown in each configuration for the slab simulations. Columns show variations due to change in the rescaling coefficient of longwave optical depth. Rows show variations with insolation contrast. The reference simulation is defined by
c. Earth’s climate
In addition to these simulations, we analyzed Earth’s annual-mean climate as given by ERA-Interim (Dee et al. 2011). The reanalysis data were averaged annually over a 30-yr period (1979–2008). To ease comparison between theory and observation, all variables were computed for the Southern Hemisphere only. This was done to mitigate added complications arising from the existence of continents and large mountain ranges in the Northern Hemisphere. Selecting the Southern Hemisphere allows for a comparison with Earth’s climate under conditions more compatible with the assumptions used to derive the supercriticality and effective static stability.
4. Results
This ensemble of dry and moist simulations (218 moist and 96 dry simulations) spans a very wide range of climates. In the moist GCM simulations without ocean heat transport, global-mean surface temperatures vary from 253 to 312 K, and equator-to-pole temperature contrasts vary from 11 to 144 K (Table 1); the ranges are essentially the same for the simulations with ocean heat transport. Among the moist simulations, we identify a reference simulation (
Slab and ocean moist simulations have nearly identical global means and equator-to-pole contrasts of the surface temperature. Yet the Hadley circulation was found to be sensitive to the presence of ocean heat transport in the tropics: Levine and Schneider (2011) found that the Hadley circulation can become significantly wider (by 5° or more) and weaker (by 50% or more) when ocean heat transport is introduced. For example, in the reference climate (
As the global-mean surface temperature increases, the tropopause shifts upward (Fig. 2), consistent with the radiative–convective response of an atmosphere to an increase in longwave opacity (Thuburn and Craig 2000; Schneider 2007). Increasing the equator-to-pole temperature contrast also deepens the troposphere through an increase in static stability that, other factors equal, usually accompanies increased meridional temperature contrasts. The increase in tropopause height with static stability can likewise be understood from radiative constraints on the tropopause (Held 1982; Schneider 2007).
a. Supercriticality in midlatitudes
Schneider and Walker (2006) found
Figure 3 shows the traditional static stability (3) versus the meridional potential temperature contrast (2) in all dry and moist GCM simulations. Simulations with
Using the effective static stability (10), with the specified convective lapse rate in the case of the dry simulations, better characterizes dynamical regimes in the extratropics. Figure 4 shows the effective static stability (10) versus the meridional potential temperature contrast (2). Comparing Figs. 3 and 4, we see that now a much larger fraction of the simulations, both moist and dry, lie on the one-to-one lines that signify
Comparing Figs. 4b and 4c, we see that moist simulations with and without parameterized ocean heat flux show similar behavior. Despite its large effect on the dynamics of the tropical atmosphere, ocean heat transport in the tropics has little effect on the extratropical atmosphere (Levine and Schneider 2011).
b. Extent of Hadley circulation
Despite a large body of theoretical and observational studies, there is no unique definition for the terminus of the Hadley circulation (cf. Levine and Schneider 2011). Here, we define the terminus as the subtropical latitude where the mean meridional mass flux, integrated from the 750-hPa pressure level to the top of the atmosphere, attains 10% of its tropical extremum. In most climates, the 750-hPa pressure level is near the upper bound of the lower branch of the Hadley circulation. This definition is similar to others used in the literature (e.g., Lu et al. 2007). Figure 5 shows that the latitude of the terminus in both moist and dry simulations varies widely, from 8° to 36°. Using slightly different definitions for the terminus based on the mass flux streamfunction (for instance, the latitude of its zero crossing) does not change our results significantly. Dry GCM simulations suggest the existence of a maximum width for the Hadley circulation on an Earth-like planet: in simulations with large convective static stabilities and large meridional potential temperature contrasts, the terminus appears to asymptotically approach 36°. This maximum extent is generally reached in simulations with large baroclinicity. One might think this maximum latitude is given by the Held–Hou extent of angular momentum–conserving Hadley circulations [see (C1)]. However, unlike the angular momentum–conserving extent, the terminus latitude in the dry simulations depends on the static stability, and it does not appear to continue to increase as the meridional potential temperature contrast increases.
In the moist simulations, the Hadley circulation widens as the climate warms, in agreement with comprehensive global warming simulations (e.g., Lu et al. 2007; Medeiros et al. 2015). The Hadley circulation terminus moves from 16° to 27° when the longwave opacity in the simulations increases, while the insolation is kept fixed to Earth’s annual-mean profile (dashed lines in Fig. 5). The terminus also shifts poleward as the insolation contrast, and with it meridional temperature contrasts, increase: when the longwave opacity is kept fixed to its reference profile while the insolation contrast is increased from zero to its maximum value, consistent with nonnegative insolation at the poles, the terminus moves from 16° to 26° (i.e., simulations on Fig. 5, denoted by a green color, shifting from the dashed red line to the dashed blue line as insolation contrast increases). In the reference simulation without ocean heat transport, the Hadley circulation extends to 23°. This is a 7° smaller extent than that of Earth’s Hadley circulation in the annual mean. Once ocean heat transport is included, the Hadley circulation in the reference simulation is 26° wide and closer to Earth’s in the annual mean (cf. Figs. 5b,c), consistent with the results of Levine and Schneider (2011).
In the dry simulations, Fig. 5 shows that the terminus shifts poleward as the equator-to-pole contrast in radiative equilibrium increases or the convective lapse rate decreases. These variations are qualitatively consistent with a widening of the circulation that accompanies a deepening of the troposphere or an increase in extratropical baroclinicity, as seen in the moist GCM simulations and in observations (Adam et al. 2014).
c. Supercriticality at the terminus of the Hadley circulation
In the simulations in which the extratropical thermal stratification is baroclinically controlled and
Similar to what we found over the baroclinic zones, using the dry static stability in lieu of the effective static stability significantly degrades the ability of the supercriticality to predict the latitude of the Hadley circulation terminus. That is, accounting for convection and large-scale condensation is as important near the terminus as it is over the storm tracks.
To the extent that
5. Discussion
In our simulations, the Hadley circulation was found to widen with global warming, associated with an increase in static stability and a lifting of the extratropical tropopause height. This widening with global warming accompanies a poleward shift of the storm tracks (Schneider 2004; Lorenz and DeWeaver 2007; Mbengue and Schneider 2013). For example, increasing the longwave optical depth of the atmosphere by a factor of 30 in moist simulations leads to a 10° shift of the terminus, from 15° to 25°; the storm tracks shift as much, from 44° to 54°. Both shifts appear to be caused by the increase in tropospheric latent heating (increasing static stability) and lifting of the tropopause associated with the global warming. Previous studies using comprehensive GCM simulations have also recognized that the widening of the Hadley circulation is associated with an increase in static stability (Lu et al. 2008), a lifting of the extratropical tropopause height (Lu et al. 2007), or a decrease in the wind shear (Tandon et al. 2013). Our study quantifies this effect using the invariance of the effective supercriticality at the latitude of the terminus.
The Hadley circulation was generally found to widen when the equator-to-pole temperature contrast increased in both moist and dry GCM simulations. In this scenario, the subtropical static stability was found to increase faster than the local meridional potential temperature contrast, and this leads to a poleward shift of the terminus. This is the opposite of what happens during an El Niño, when an anomalous warming of the tropical troposphere (and thus an increase in the equator-to-pole temperature gradient) leads to a contraction of the Hadley circulation. In the latter case, anomalous heating in the deep tropics increases the meridional temperature gradient in the subtropics, but without large compensation from a static stability increase, and this leads to an equatorward shift of the terminus. Our results are consistent with the findings of Brayshaw et al. (2008), who showed that the storm tracks (and perhaps also the Hadley circulation terminus) could shift poleward or equatorward when increasing baroclinicity (e.g., by prescribing an anomalous SST gradient), depending on whether this anomalous baroclinicity is applied in the subtropics (in which case storm tracks shift equatorward) or in the midlatitudes (in which case storm tracks shift poleward).
It is also important to note that the terminus of the Hadley circulation and the storm-track latitude are not equally sensitive to climate changes, depending on which forcings are varied (Mbengue and Schneider 2013). For instance, increasing the insolation contrast from
6. Conclusions
Using a dry and a moist idealized GCM, we have simulated a wide range of climates. In the simulations, the terminus of the Hadley circulation varied between about 8° and 36°. A supercriticality criterion was modified to account for the tight coupling between baroclinic eddies and convection. This effective supercriticality criterion successfully discriminated between simulations in which the extratropical thermal stratification was controlled by baroclinic eddies and those in which it is controlled locally by convection. The latitude of the Hadley circulation terminus was found to be well constrained by a constant
We found that the extratropical thermal stratification in idealized moist GCM simulations is controlled by baroclinic eddies when the meridional potential temperature contrast becomes larger than the effective static stability in local convective equilibrium. The Earth-like reference climate was found to be near the transition between baroclinic and convective regimes. Earth’s Southern Hemisphere climate in the annual mean similarly shows an important influence from convection on the thermal stratification. More generally, we found convection to influence the static stability and tropopause height, even when baroclinic eddies were strong.
The finding that the effective supercriticality typically assumes a constant
Acknowledgments
We thank Paul O’Gorman for helpful clarifications on the effective static stability, Timothy M. Merlis for discussions on both linear baroclinic wave theories and ocean–atmosphere interactions, and Tobias Bischoff for his helpful comments on this work and its relation to ENSO. We are grateful for support by the National Science Foundation (Grants AGS-1019211 and AGS-1049201) and a Yale Climate and Energy Institute Fellowship. The simulations were performed on the Division of Geological and Planetary Sciences’ Dell cluster at the California Institute of Technology (the program code for the simulations described in this paper is available at www.clidyn.ethz.ch/gcms/).
APPENDIX A
Determination of Baroclinic Zones
APPENDIX B
Determination of Tropopause Height
A reliable determination of the tropopause height is essential when estimating the supercriticality in our simulations. The World Meteorological Organization (WMO) defines the tropopause as the lowest level of the atmosphere where the lapse rate is equal to or lower than a threshold lapse rate of 2 K km−1. We applied this definition to the bulk lapse rate (averaged over the baroclinic zones) for all climates. Tropopause heights estimated from this definition were then used to define the supercriticality (1). The same definition was used to define the local tropopause height in Fig. 2. We found that simulations with near-neutral convective lapse rate, which corresponds to cold climates in the moist GCM, have a poorly defined tropopause when applying the WMO definition. To circumvent this issue, we use a different definition of the tropopause height based on the meridional circulation structure.
In the coldest set of simulations, the tropopause height was redefined as the level where the mean meridional mass flux in the upper troposphere reaches its maximum value in the Northern Hemisphere. This maximum is always achieved in the deep tropics, and its latitude usually defines the center of the Hadley circulation in the Northern Hemisphere. This level usually corresponds to the levels of both maximum eddy momentum flux divergence in the tropics and convergence in the extratropics, consistent with quasi-horizontal propagation of eddies from the storm tracks to the subtropics (Ait-Chaalal and Schneider 2015). The height obtained from this definition is nearly identical to that provided by the lapse rate definition, except in the climates with near-neutral thermal stratification. In the latter simulations, the lapse rate definition fails to estimate the tropopause height because of the existence of strong inversions between the boundary layer and the free troposphere, while our dynamical definition gives a more robust estimate.
APPENDIX C
Other Theories for Hadley Cell Terminus
REFERENCES
Adam, O., T. Schneider, and N. Harnik, 2014: Role of changes in mean temperatures versus temperature gradients in the recent widening of the Hadley circulation. J. Climate, 27, 7450–7461, doi:10.1175/JCLI-D-14-00140.1.
Ait-Chaalal, F., and T. Schneider, 2015: Why eddy momentum fluxes are concentrated in the upper troposphere. J. Atmos. Sci., 72, 1585–1604, doi:10.1175/JAS-D-14-0243.1.
Barnes, E. A., and L. Polvani, 2013: Response of the midlatitude jets, and of their variability, to increased greenhouse gases in the CMIP5 models. J. Climate, 26, 7117–7135, doi:10.1175/JCLI-D-12-00536.1.
Barry, L., G. C. Craig, and J. Thuburn, 2000: A GCM investigation into the nature of baroclinic adjustment. J. Atmos. Sci., 57, 1141–1155, doi:10.1175/1520-0469(2000)057<1141:AGIITN>2.0.CO;2.
Booth, J. F., L. Polvani, P. A. O’Gorman, and S. Wang, 2015: Effective stability in a moist baroclinic wave. Atmos. Sci. Lett., 16, 56–62, doi:10.1002/asl2.520.
Brayshaw, D. J., B. Hoskins, and M. Blackburn, 2008: The storm-track response to idealized SST perturbations in an aquaplanet GCM. J. Atmos. Sci., 65, 2842–2860, doi:10.1175/2008JAS2657.1.
Butler, A. H., D. W. J. Thompson, and T. Birner, 2011: Isentropic slopes, downgradient eddy fluxes, and the extratropical atmospheric circulation response to tropical tropospheric heating. J. Atmos. Sci., 68, 2292–2305, doi:10.1175/JAS-D-10-05025.1.
Chai, J., and G. K. Vallis, 2014: The role of criticality on the horizontal and vertical scales of extratropical eddies in a dry GCM. J. Atmos. Sci., 71, 2300–2318, doi:10.1175/JAS-D-13-0351.1.
Dee, D. P., and Coauthors, 2011: The ERA-Interim reanalysis: Configuration and performance of the data assimilation system. Quart. J. Roy. Meteor. Soc., 137, 553–597, doi:10.1002/qj.828.
Edmon, H. J., Jr., B. J. Hoskins, and M. E. McIntyre, 1980: Eliassen–Palm cross sections for the troposphere. J. Atmos. Sci., 37, 2600–2616, doi:10.1175/1520-0469(1980)037<2600:EPCSFT>2.0.CO;2.
Emanuel, K. A., M. Fantini, and A. J. Thorpe, 1987: Baroclinic instability in an environment of small stability to slantwise moist convection. Part I: Two-dimensional models. J. Atmos. Sci., 44, 1559–1573, doi:10.1175/1520-0469(1987)044<1559:BIIAEO>2.0.CO;2.
Frierson, D. M. W., 2007: The dynamics of idealized convection schemes and their effect on the zonally averaged tropical circulation. J. Atmos. Sci., 64, 1959–1976, doi:10.1175/JAS3935.1.
Frierson, D. M. W., I. M. Held, and P. Zurita-Gotor, 2006: A gray-radiation aquaplanet moist GCM. Part I: Static stability and eddy scale. J. Atmos. Sci., 63, 2548–2566, doi:10.1175/JAS3753.1.
Frierson, D. M. W., J. Lu, and G. Chen, 2007: The width of the Hadley cell in simple and comprehensive general circulation models. Geophys. Res. Lett.,34, L18804, doi:10.1029/2007GL031115.
Gutowski, W. J., Jr., L. E. Branscome, and D. A. Stewart, 1992: Life cycles of moist baroclinic eddies. J. Atmos. Sci., 49, 306–319, doi:10.1175/1520-0469(1992)049<0306:LCOMBE>2.0.CO;2.
Held, I. M., 1978: The vertical scale of an unstable baroclinic wave and its importance for eddy heat flux parameterizations. J. Atmos. Sci., 35, 572–576, doi:10.1175/1520-0469(1978)035<0572:TVSOAU>2.0.CO;2.
Held, I. M., 1982: On the height of the tropopause and the static stability of the troposphere. J. Atmos. Sci., 39, 412–417, doi:10.1175/1520-0469(1982)039<0412:OTHOTT>2.0.CO;2.
Held, I. M., 2000: The general circulation of the atmosphere. Proc. Program in Geophysical Fluid Dynamics, Woods Hole, MA, Woods Hole Oceanographic Institute, 54 pp. [Available online at http://www.whoi.edu/fileserver.do?id=21464&pt=10&p=17332.]
Held, I. M., and A. Y. Hou, 1980: Nonlinear axially symmetric circulations in a nearly inviscid atmosphere. J. Atmos. Sci., 37, 515–533, doi:10.1175/1520-0469(1980)037<0515:NASCIA>2.0.CO;2.
Held, I. M., and B. J. Hoskins, 1985: Large-scale eddies and the general circulation of the troposphere. Adv. Geophys., 28, 3–31, doi:10.1016/S0065-2687(08)60218-6.
Hu, Y., and Q. Fu, 2007: Observed poleward expansion of the Hadley circulation since 1979. Atmos. Chem. Phys., 7, 5229–5236, doi:10.5194/acp-7-5229-2007.
Jansen, M., and R. Ferrari, 2012: Macroturbulent equilibration in a thermally forced primitive equation system. J. Atmos. Sci., 69, 695–713, doi:10.1175/JAS-D-11-041.1.
Kang, S. M., and J. Lu, 2012: Expansion of the Hadley cell under global warming: Winter versus summer. J. Climate, 25, 8387–8393, doi:10.1175/JCLI-D-12-00323.1.
Klinger, B. A., and J. Marotzke, 2000: Meridional heat transport by the subtropical cell. J. Phys. Oceanogr., 30, 696–705, doi:10.1175/1520-0485(2000)030<0696:MHTBTS>2.0.CO;2.
Knippertz, P., and H. Wernli, 2010: A Lagrangian climatology of tropical moisture exports to the Northern Hemispheric extratropics. J. Climate, 23, 987–1003, doi:10.1175/2009JCLI3333.1.
Korty, R. L., and T. Schneider, 2007: A climatology of the tropospheric thermal stratification using saturation potential vorticity. J. Climate, 20, 5977–5991, doi:10.1175/2007JCLI1788.1.
Korty, R. L., and T. Schneider, 2008: Extent of Hadley circulations in dry atmospheres. Geophys. Res. Lett.,35, L23803, doi:10.1029/2008GL035847.
Lapeyre, G., and I. M. Held, 2004: The role of moisture in the dynamics and energetics of turbulent baroclinic eddies. J. Atmos. Sci., 61, 1693–1710, doi:10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2.
Levine, X. J., and T. Schneider, 2011: Response of the Hadley circulation to climate change in an aquaplanet GCM coupled to a simple representation of ocean heat transport. J. Atmos. Sci., 68, 769–782, doi:10.1175/2010JAS3553.1.
Lorenz, D. J., and E. T. DeWeaver, 2007: Tropopause height and zonal wind response to global warming in the IPCC scenario integrations. J. Geophys. Res.,112, D10119, doi:10.1029/2006JD008087.
Lu, J., G. A. Vecchi, and T. Reichler, 2007: Expansion of the Hadley cell under global warming. Geophys. Res. Lett.,34, L06805, doi:10.1029/2006GL028443.
Lu, J., G. Chen, and D. M. W. Frierson, 2008: Response of the zonal mean atmospheric circulation to El Niño versus global warming. J. Climate, 21, 5835–5851, doi:10.1175/2008JCLI2200.1.
Mbengue, C., and T. Schneider, 2013: Storm track shifts under climate change: What can be learned from large-scale dry dynamics. J. Climate, 26, 9923–9930, doi:10.1175/JCLI-D-13-00404.1.
Medeiros, B., B. Stevens, and S. Bony, 2015: Using aquaplanets to understand the robust responses of comprehensive climate models to forcing. Climate Dyn.,44, 1957–1977, doi:10.1007/s00382-014-2138-0.
Merlis, T. M., and T. Schneider, 2009: Scales of linear baroclinic instability and macroturbulence in dry atmospheres. J. Atmos. Sci., 66, 1821–1833, doi:10.1175/2008JAS2884.1.
Nguyen, H., A. Evans, C. Lucas, I. Smith, and B. Timbal, 2013: The Hadley circulation in reanalyses: Climatology, variability, and change. J. Climate, 26, 3357–3376, doi:10.1175/JCLI-D-12-00224.1.
O’Gorman, P. A., 2011: The effective static stability experienced by eddies in a moist atmosphere. J. Atmos. Sci., 68, 75–90, doi:10.1175/2010JAS3537.1.
O’Gorman, P. A., and T. Schneider, 2008: The hydrological cycle over a wide range of climates simulated with an idealized GCM. J. Climate, 21, 3815–3832, doi:10.1175/2007JCLI2065.1.
Phillips, N. A., 1954: Energy transformations and meridional circulations associated with simple baroclinic waves in a two-level, quasi-geostrophic model. Tellus, 6, 273–286, doi:10.1111/j.2153-3490.1954.tb01123.x.
Ralph, F. M., P. J. Neiman, and G. A. Wick, 2004: Satellite and CALJET aircraft observations of atmospheric rivers over the eastern North Pacific Ocean during the winter of 1997/98. Mon. Wea. Rev., 132, 1721–1745, doi:10.1175/1520-0493(2004)132<1721:SACAOO>2.0.CO;2.
Schneider, E. K., 1977: Axially symmetric steady-state models of the basic state for instability and climate studies. Part II. Nonlinear calculations. J. Atmos. Sci., 34, 280–296, doi:10.1175/1520-0469(1977)034<0280:ASSSMO>2.0.CO;2.
Schneider, T., 2004: The tropopause and the thermal stratification in the extratropics of a dry atmosphere. J. Atmos. Sci., 61, 1317–1340, doi:10.1175/1520-0469(2004)061<1317:TTATTS>2.0.CO;2.
Schneider, T., 2006: The general circulation of the atmosphere. Annu. Rev. Earth Planet. Sci., 34, 655–688, doi:10.1146/annurev.earth.34.031405.125144.
Schneider, T., 2007: The thermal stratification of the extratropical atmosphere. The Global Circulation of the Atmosphere, T. Schneider and A. H. Sobel, Eds., Princeton University Press, 47–77.
Schneider, T., and C. C. Walker, 2006: Self-organization of atmospheric macroturbulence into critical states of weak nonlinear eddy–eddy interactions. J. Atmos. Sci., 63, 1569–1586, doi:10.1175/JAS3699.1.
Schneider, T., and P. A. O’Gorman, 2008: Moist convection and the thermal stratification of the extratropical troposphere. J. Atmos. Sci., 65, 3571–3583, doi:10.1175/2008JAS2652.1.
Schneider, T., P. A. O’Gorman, and X. J. Levine, 2010: Water vapor and the dynamics of climate changes. Rev. Geophys., 48, RG3001, doi:10.1029/2009RG000302.
Seager, R., N. Harnik, Y. Kushnir, W. Robinson, and J. Miller, 2003: Mechanisms of hemispherically symmetric climate variability. J. Climate, 16, 2960–2978, doi:10.1175/1520-0442(2003)016<2960:MOHSCV>2.0.CO;2.
Seidel, D. J., and W. J. Randel, 2007: Recent widening of the tropical belt: Evidence from tropopause observations. J. Geophys. Res.,112, D20113, doi:10.1029/2007JD008861.
Seidel, D. J., Q. Fu, W. J. Randel, and T. J. Reichler, 2008: Widening of the tropical belt in a changing climate. Nat. Geosci., 1, 21–24, doi:10.1038/ngeo.2007.38.
Simmons, A. J., and B. J. Hoskins, 1980: Barotropic influences on the growth and decay of nonlinear baroclinic waves. J. Atmos. Sci., 37, 1679–1684, doi:10.1175/1520-0469(1980)037<1679:BIOTGA>2.0.CO;2.
Simpson, I. R., T. A. Shaw, and R. Seager, 2014: A diagnosis of the seasonally and longitudinally varying midlatitude circulation response to global warming. J. Atmos. Sci., 71, 2489–2515, doi:10.1175/JAS-D-13-0325.1.
Smagorinsky, J., S. Manabe, and J. L. Holloway Jr., 1965: Numerical results from a nine-level general circulation model of the atmosphere. Mon. Wea. Rev., 93, 727–768, doi:10.1175/1520-0493(1965)093<0727:NRFANL>2.3.CO;2.
Stone, P. H., 1978: Baroclinic adjustment. J. Atmos. Sci., 35, 561–571, doi:10.1175/1520-0469(1978)035<0561:BA>2.0.CO;2.
Stone, P. H., and B. Nemet, 1996: Baroclinic adjustment: A comparison between theory, observations, and models. J. Atmos. Sci., 53, 1663–1674, doi:10.1175/1520-0469(1996)053<1663:BAACBT>2.0.CO;2.
Tandon, N. F., E. P. Gerber, A. H. Sobel, and L. M. Polvani, 2013: Understanding Hadley cell expansion versus contraction: Insights from simplified models and implications for recent observations. J. Climate, 26, 4304–4321, doi:10.1175/JCLI-D-12-00598.1.
Thuburn, J., and G. C. Craig, 2000: Stratospheric influence on tropopause height: The radiative constraint. J. Atmos. Sci., 57, 17–28, doi:10.1175/1520-0469(2000)057<0017:SIOTHT>2.0.CO;2.
Trenberth, K. E., and D. P. Stepaniak, 2003: Seamless poleward atmospheric energy transports and implications for the Hadley circulation. J. Climate, 16, 3706–3722, doi:10.1175/1520-0442(2003)016<3706:SPAETA>2.0.CO;2.
Troen, I. B., and L. Mahrt, 1986: A simple model of the atmospheric boundary layer: Sensitivity to surface evaporation. Bound.-Layer Meteor., 37, 129–148, doi:10.1007/BF00122760.
Vallis, G. K., 2006: Wave-mean flow interaction. Atmospheric and Oceanic Fluid Dynamics: Fundamentals and Large-Scale Circulation, Cambridge University Press, 295–336.
Vecchi, G. A., and B. J. Soden, 2007: Global warming and the weakening of the tropical circulation. J. Climate, 20, 4316–4340, doi:10.1175/JCLI4258.1.
Walker, C. C., and T. Schneider, 2006: Eddy influences on Hadley circulations: Simulations with an idealized GCM. J. Atmos. Sci., 63, 3333–3350, doi:10.1175/JAS3821.1.
Yin, J. H., 2005: A consistent poleward shift of the storm tracks in simulations of 21st century climate. Geophys. Res. Lett.,32, L18701, doi:10.1029/2005GL023684.
Zurita-Gotor, P., and R. S. Lindzen, 2007: Theories of baroclinic adjustment and eddy equilibration. The Global Circulation of the Atmosphere, T. Schneider and A. H. Sobel, Eds., Princeton University Press, 22–46.
Seven dry GCM simulations with low meridional potential temperature contrasts and low convective lapse rates were dominated by numerical noise. This prevented a robust estimate of baroclinic zones and of the Hadley circulation terminus. They were excluded from Fig. 3 and all subsequent figures. In them, the tropopause height is controlled by convection.
Figure 4 shows four moist GCM simulations below the one-to-one line (i.e., with
The relevance of supercriticality to ENSO events is difficult to assess quantitatively from observations, as the Northern Hemisphere Hadley circulation contracts by less than 2° during a typical El Niño event, with significant discrepancies among observational datasets (e.g., Nguyen et al. 2013). But prescribing a heat source in the deep tropics in our idealized moist GCM induces a contraction of the Hadley circulation as during an El Niño event, and the effective supercriticality criterion captures these changes (T. Bischoff 2014, personal communication). These results will be described in a forthcoming publication.