1. Introduction
A major challenge to our understanding of midlatitude storm systems lies in the interplay between the atmospheric circulation and the hydrological cycle. On a global scale, higher temperature and humidity in the tropics relative to the poles drives poleward transport of both sensible and latent heat. On the local scale, ascending parcels undergo adiabatic expansion, condensing excess moisture to release latent heat. This additional energy can induce local hydrodynamical instabilities in conditions that would otherwise be stable. The effect of moisture is not isolated to the scales on which condensation occurs, but rather impacts dynamics across a broad range of scales, including the aggregate behavior of storm tracks (Shaw et al. 2016), the extratropical stratification (Frierson et al. 2006; Schneider and O’Gorman 2008; Wu and Pauluis 2014), and the global atmospheric circulation (Pauluis et al. 2010). Understanding the impacts of moist processes across the full range of geophysical scales is necessary to understand how midlatitude storm dynamics will change in a world becoming more humid as a result of climate change.
Many previous studies have focused on scale changes associated with moisture. Stronger moist effects lead to smaller-scale motions and narrower regions of saturation (Emanuel et al. 1987; Fantini 1990; Lapeyre and Held 2004). This correlation obfuscates the effect of different mechanisms by which moisture induces smaller-scale motion. For instance, does the shift arise as a result of highly localized precipitation associated with the cascade of moisture to small scales? Would a similar result persist even if the precipitation characteristically occurred on larger scales? And how do nonlinearities in precipitation and Clausius–Clapeyron change the dynamics? Many studies also reach opposite conclusions regarding the impact of moisture. For instance, moisture’s impact on eddy kinetic energy has been found to be positive (Emanuel et al. 1987; Lapeyre and Held 2004; Lambaerts et al. 2011), negative (Zurita-Gotor 2005; Bembenek et al. 2020; Lutsko and Hell 2021), or about neutral (Lambaerts et al. 2012).
Questions remain about how to synthesize results from different implementations of moisture in idealized systems. The construction and interpretation of a moist energy (ME) and moist potential vorticity (MPV) are key pieces that can help bridge this gap. The changes moisture introduces to the energetics result in changes to the scale at which energy is injected into the flow, its ability to cascade to different scales, and the scale at which it is dissipated. Furthermore, because moisture introduces new processes, moist systems feature new mechanisms of growth and propagation, the impact of which must be understood both individually and in combination. To this end, the study of moist turbulence benefits from a hierarchy of models with implementations of moisture mechanisms in different combinations and at different levels of complexity, including both linear (Emanuel et al. 1987; Adames and Ming 2018; Adames 2021) and nonlinear (Fantini 1990; Lapeyre and Held 2004) frameworks.
The two-layer quasigeostrophic (QG) model is one of the simplest mathematical models to exhibit the basic features of the turbulent midlatitude atmosphere, from planetary-scale barotropic jets to synoptic-scale baroclinic eddies that organize into storm tracks. Its relative simplicity, coupled with its ability to capture key dynamical features, has made it a good choice for studying the broader statistical and scaling properties of a dry atmosphere (e.g., Vallis 2006). While its utility in assessing the moist case is limited due to significant ageostrophy in precipitation regions (Fantini 1990, 1995; Lambaerts et al. 2011, 2012), moist QG (MQG) models can still provide insight into the dynamics without confounding influences from the tropics. Consequently, MQG models have been used in studies of the fundamental dynamics of baroclinic systems, such as mechanisms of growth (Parker and Thorpe 1995; Moore and Montgomery 2004; de Vries et al. 2010; Adames and Ming 2018) and analysis of turbulent spectra (Edwards et al. 2019). MQG systems are also ideal for developing theories of wave–mean flow interaction in moist systems, which is the portion of theory we seek to advance in this paper.
We use the MQG model of Lapeyre and Held (2004) to bring new intuition to the impacts of moisture on geostrophic turbulence. In section 2, we review the model and derive a conservation law for a moist potential vorticity. We argue that in the limit of high evaporation rate, the model approaches a saturated limit with precipitation active everywhere. We show in appendix A that this saturated limit is mathematically equivalent to the classic two-layer problem after replacing the baroclinic potential vorticity by the MPV, and rescaling both the horizontal and temporal dimensions. In particular, existing theory for dry QG turbulence can be readily tested in the MQG model in the saturated limit.
In section 3, we discuss the numerical implementation of the MQG model and analyze the results of numerical simulations. We show that increasing the amount of moisture leads to three main effects: a systematic intensification of turbulence, a shift of energy injection to smaller scales, and an extension of the inverse cascade to larger scales. In section 4, we analyze the energetics of the MQG model and derive an expression for ME that is converted into available potential energy (APE) through precipitation. We show that the intensification of turbulence with increased moisture is directly tied to the increased generation of ME by the barotropic flow acting on the mean temperature and humidity gradient. In section 5, we argue that the shift of the most unstable baroclinic mode in the linear instability analysis is reflected by the shift in the precipitation injection scale. We derive an expression for the Rhines scale in the saturated limit by accounting for the additional generation of ME and show that this captures the impact of moisture on the energy containing scale in our simulations. The study concludes in section 6.
2. Model description
We use the two-layer MQG model of Lapeyre and Held (2004), depicted schematically in Fig. 1. This model consists of two well-stratified layers of equal mean depth H in a doubly periodic domain. Rotational dynamics are captured by a β plane in which the Coriolis parameter is expressed linearly in the meridional coordinate as f = f0 + βy. For guidance, a list of key variables and their definitions can be found in Table 1.
Structure of the two-layer model. Thick flat lines correspond to surfaces that remain fixed and the wavy curve to the interface η, which varies. Each layer has a streamfunction relating to the barotropic and baroclinic modes as described in the text, an associated potential temperature, and a typical height scale H. The interface η captures variations from this typical thickness, which are corrected by vertical motion W. The moisture m is confined to the lower layer and precipitation conditionally triggers mass transport
Citation: Journal of the Atmospheric Sciences 80, 6; 10.1175/JAS-D-22-0215.1
Variables used in the model description.
a. The dry system
b. Incorporating moisture
c. Moist potential vorticity
A schematic for the condensation level ηc as a metric for saturation. When the interface η between the top (white) and bottom (blue) layers is below the condensation level ηc, the system is subsaturated. The interface η evolves by dynamical processes and radiative cooling R, while the condensation level ηc evolves with competing effects from evaporation E and radiative cooling R. When η rises above ηc, precipitation P quickly acts to bring the two to parity by removing water vapor and lowering the interface.
Citation: Journal of the Atmospheric Sciences 80, 6; 10.1175/JAS-D-22-0215.1
d. The saturated limit
The nonlinearity of the precipitation trigger (9) introduces a major complication in the study of moist turbulent dynamics. This nonlinearity, however, is absent in two limiting scenarios: a dry atmosphere with no precipitation, and a fully saturated atmosphere with precipitation everywhere. While the first scenario has been well documented, we argue here that the second scenario, which we refer here to as the saturated limit, can offer additional insights on the impacts of moisture on geostrophic turbulence.
The saturated limit can be achieved if one makes the assumption that precipitation acts quickly enough to maintain the system near saturation. Within the MQG system described above, the limit of complete saturation can be nearly achieved by increasing the evaporation parameter E and decreasing the precipitation relaxation scale τ. The former increases the amount of water vapor added to the system at every time step, ensuring at sufficiently high values that the system is never subsaturated. The latter decreases the amount of time that the system takes to relax to the saturated value, decreasing the value of the moisture surplus m = ms in a supersaturated system. Applying both of these limits corresponds to the strict quasi-equilibrium approximation of Emanuel et al. (1994).
3. Numerical simulations
We perform experiments on a doubly periodic domain in spectral space with a 256 × 256 grid, The domain size L is chosen such that 2πλ = L/9. Simulations were run for a time T = 400λ/U. Time averages are computed over the last half of the run with sampling at intervals of δt = 0.1λ/U. Time stepping uses a third-order Adams–Bashforth method with an integrating factor to remove the stiff portion of the equation and the Jacobian handled pseudospectrally with antialiasing. Time stepping is done for the upper (qBT + qbc), lower (qBT − qbc), and moist lower (qBT − qm) PV, thereby eliminating the need to compute the vertical motion W. The upper and lower streamfunctions are computed diagnostically in Fourier space. Precipitation is computed diagnostically in real space from the moisture surplus,
The simulations used for data in this paper span the parameter space listed in the right column of Table 2. Realistic values are listed in column 3. The estimate for the precipitation relaxation time scale
Tunable parameter space (nondimensionalized), realistic values, and the values used in the simulations here. To enforce the saturated limit, the integrations were done with a very large value of the evaporation
Figure 3 displays snapshots of the barotropic vorticity (first column), baroclinic PV (second column), and MPV (third column) in three of these simulations. The first row shows a dry simulation (μs = 1) at supercriticality ξ = 1.25. As the configuration is only slightly supercritical, the flow is only weakly unstable and is organized in six fairly narrow zonal jets. The second row shows a moist simulation with μs = 4.0 at the same dry criticality ξ = 1.25. An intensification of the flow is evidenced by the increase in the magnitude of the vorticity anomalies. The range of motions is substantially enhanced both at small scales, with the emergence of closed vortices, and at large scale with the organization of the flow around two zonal jets instead of six. Finally, the third row shows a dry simulation (μs = 1.0), but at criticality ξ = 5. This supercriticality is chosen as to match the value of μsξ in the simulation shown in the second row. Qualitatively, the simulations in the second and third rows exhibit similar levels of turbulence, albeit with systematically larger scale of motions in the dry simulation.
Snapshots of the barotropic, baroclinic, and moist baroclinic potential vorticity perturbation for three cases. (top) A dry case with mild supercriticality (ξ = 1.25, μs = 1.0). (middle) A saturated case with the same dry criticality as the first case and moisture (ξ = 1.25, μs = 4.0). (bottom) Another dry case with higher criticality (ξ = 5.0, μs = 1.0), chosen so that the second and third cases have same total saturated criticality. Both exhibit more energetic flows than the top row; note the change in color scale. The middle row, as the only moist system, is the only row to exhibit a smaller magnitude of (center) dry baroclinic PV compared to (right) the moist. This is associated with the inclusion of moisture in the “reservoir” for conversion into baroclinic vorticity. Additionally, the middle row is dominated by small-scale vorticity, consistent with a shift to smaller scales, in contrast with the bottom row with the same saturated criticality.
Citation: Journal of the Atmospheric Sciences 80, 6; 10.1175/JAS-D-22-0215.1
Spectra of the barotropic and baroclinic eddy kinetic energy for a few values of μs with (top) ξ = 1.25 and (bottom) ξ = 0.8. The horizontal scale is the wavelength rescaled by the largest wavelength 2π/L, where L is the domain size. The y axis is plotted in symlog scale, such that the figures are linear for values smaller than 10−4 and log scale for larger values. The vertical lines in the barotropic mode are the Rhines scale, associated with the termination of the inverse cascade. The vertical lines in baroclinic mode mark the centroid of the baroclinic eddy kinetic energy.
Citation: Journal of the Atmospheric Sciences 80, 6; 10.1175/JAS-D-22-0215.1
4. Energetics
The MQG system exhibits features across a broad range of scales. Held and Larichev (1996) argue that the energetics of the (dry) QG system can be understood as an inverse energy cascade associated with barotropic motions, and a direct cascade of APE. The two cascades are coupled in the sense that the APE is mixed to smaller scales by the barotropic flow before being converted into kinetic energy, which in turn sustains the cascade of barotropic kinetic energy to large scales. We revisit how the inclusion of moist processes modifies this picture by quantifying an additional source of energy associated with the poleward transport of moisture.
Energy transfers in the MQG model, with the estimates of the scaling at saturation. At large scales, the background moisture and temperature gradients are redistributed by the barotropic flow, acting as a source for the APE and ME, εAPE and εME, respectively (purple arrows). The energy is mixed to smaller scales by the barotropic flow until near the Rossby radius. The precipitation
Citation: Journal of the Atmospheric Sciences 80, 6; 10.1175/JAS-D-22-0215.1
The balance of (a) the total generation of energy vs the Ekman dissipation, (b) the generation of MAPE vs the injection into barotropic energy, (c) the generation of ME and conversion to precipitation, and (d) the ratio of precipitation injection to sensible heat flux, the two contributions to the generation of APE.
Citation: Journal of the Atmospheric Sciences 80, 6; 10.1175/JAS-D-22-0215.1
5. Scalings
The previous section argues that geostrophic turbulence is characterized by the generation, conversion and dissipation of three different components of the energy budget. Here, we focus on the scales at which these occur. Scaling arguments for the termination of the inverse cascade (e.g., Held and Larichev 1996) often assume that the system has an inertial range: a sufficient scale separation between the injection scale and dissipation scale. In practice, such scaling arguments still offer useful insights, even in the absence of a clear inertial range.
a. Linear stability analysis
Figure 3 shows that the scalings of moist systems cannot be determined by ξ or ξs alone. The top and middle rows depict the potential vorticities of a dry and moist simulation with the same value of ξ, demonstrating the increase in energy at both small and large scales associated with the inclusion of moisture. The middle and bottom rows depict the potential vorticities of a moist and dry simulation with the same value of ξs, demonstrating that the moist system exhibits smaller-scale vortices than the dry with equivalent moist criticality.
Growth rate as a function of scale Kλ = |k|λ and the gross moisture stratification μs, with dry criticality ξ fixed as the value indicated in each title. (right) The dashed horizontal line corresponds to the value of μs necessary to achieve marginal saturated criticality. The two lines enveloping the contour correspond to asymptotic bounds on the unstable region in the limit ξs → ∞.
Citation: Journal of the Atmospheric Sciences 80, 6; 10.1175/JAS-D-22-0215.1
We return to Fig. 3 to see how well these predictions play out. The simulations depicted in the middle and bottom rows have equal values of ξs. However, the middle row has μs = 4.0, while the bottom row is dry. The results of the linear stability analysis predict instability on a length scale that is a factor
b. Baroclinic energy
Baroclinic energy is generated through the downgradient transport of sensible heat εAPE and through precipitation
The spectra of the source terms of the baroclinic energy Ebc. These include a downgradient flux of sensible heat εAPE and a precipitation injection term
Citation: Journal of the Atmospheric Sciences 80, 6; 10.1175/JAS-D-22-0215.1
For large μs, there is a small but noteworthy removal of APE by precipitation at large scale. As in Bembenek et al. (2020) and Lutsko and Hell (2021), this occurs due to regions where precipitation is anticorrelated with temperature. Surface dissipation of barotropic energy at large scales induces regions of mechanically forced ascent and subsidence through Ekman pumping. Descending motions in regions of large-scale subsidence induce warm and dry anomalies. Conversely, ascending regions are associated with colder but moister conditions.
The advective flux of (a) the APE, (b) the baroclinic kinetic energy, and (c) the total baroclinic energy. The y axis is on a symlog scale, such that it is linear for values between ±0.1. The advection term transports energy from the scales where the slope is positive to those where the slope is negative. Cascade behavior corresponds to the regions where the slope is near zero, as energy is added and removed at similar rates.
Citation: Journal of the Atmospheric Sciences 80, 6; 10.1175/JAS-D-22-0215.1
As shown in Fig. 9a, the APE cascade
c. Barotropic energy
The injection into the barotropic energy, decomposed into linear and nonlinear components. The y axis is on a symlog scale, such that it is linear for values less than 0.01.
Citation: Journal of the Atmospheric Sciences 80, 6; 10.1175/JAS-D-22-0215.1
The advective flux of the barotropic energy. The y axis is on a symlog scale, such that it is linear for values between ±0.1.
Citation: Journal of the Atmospheric Sciences 80, 6; 10.1175/JAS-D-22-0215.1
d. Rhines scale and the inverse cascade
The first Eq. (45) indicates that the ratio between the Rhines scale and the moist deformation radius
These scalings are tested in Fig. 12. As in Held and Larichev (1996), the Rhines scale, RMS barotropic velocity, and energy generation increase faster with criticality than predicted. These arguments should apply best in the asymptotic limit ξs → ∞. More data at larger effective criticality would be needed to see if this is the case. A slight shallowing of the slope for values above ξs ≈ 4 indicates that convergence might be possible. At subcriticality, ξs ≈ 1, the barotropic and baroclinic EKE are of similar orders of magnitude. As such, the assumption that the baroclinic PV can be treated as a passive tracer no longer holds.
The scaling of (a) the Rhines scale k0 computed from the RMS barotropic velocity as a function of saturated criticality, (b) the RMS barotropic velocity V as a function of saturated criticality μsξ, and (c) the total energy generated by the MQG system. Each marker corresponds to a fixed value of dry criticality ξ, indicated in the legend.
Citation: Journal of the Atmospheric Sciences 80, 6; 10.1175/JAS-D-22-0215.1
While the proposed scalings indicate that geostrophic turbulence has a very high sensitivity to moisture content through the parameter μs, it should be noted that these scalings only hold in the saturated limit, i.e., in an atmosphere that is raining everywhere. For partial saturation, Lapeyre and Held (2004) show that moist geostrophic turbulence behaves somewhere between the dry and saturated limit. Further investigations of the impacts of moisture on geostrophic turbulence in a partially saturated atmosphere are left to a future study.
6. Conclusions
We have investigated geostrophic turbulence in an idealized moist model analogous to that of Lapeyre and Held (2004). We introduced a framework for the energy centered on the idea that a “condensation level” characterizes a moist energy (ME) that can be transformed into available potential energy (APE) through precipitation. The large-scale gradient of the condensation level provides a reservoir of ME. Eddies extract ME by transporting moisture poleward, which is then converted to APE when precipitation occurs in the warm sector of the eddies. This provides an additional source of baroclinic energy, which can significantly energize moist geostrophic turbulence compared to its dry counterpart under the same temperature gradient. Associated with this new framework is a moist baroclinic potential vorticity which is conserved under latent heat release. This modified PV emphasizes a gradient that takes into account the background meridional configuration of a moist static energy.
By enforcing a high rate of evaporation and a short precipitation relaxation time, we achieve a saturated limit. Under this limit, the MQG system is mathematically equivalent to the dry two-layer QG equations after rescaling both the time and spatial scales. In particular, this saturated limit makes it possible to extend results from geostrophic turbulence to include the effect of moisture on the dynamics of baroclinic eddies.
We analyzed the conditions for baroclinic instability in the saturated limit using a linear stability analysis. We demonstrate that a saturated criticality, the dry criticality rescaled by a parameter μs, better predicts instability, including where dry models would predict stability. We confirm numerically that the fully nonlinear MQG system exhibits instability at smaller scales and an increase in the total energy injection. This conclusion is consistent with the results of Emanuel et al. (1987), Fantini (1995), and Joly and Thorpe (1989), among others.
We examined the impacts of moisture on the injection scale and energy cascade by considering the tendency terms in the full nonlinear equations. The inclusion of moisture results in the energy injection into the barotropic mode broadening to both larger and smaller scales than in the dry case. In particular, as the strength of moist parameters increased, precipitation shifts to smaller scales and becomes the dominant contribution to instability. The increase in eddy kinetic energy generation in from latent heat release enhances the inverse cascade of barotropic kinetic energy, similar to the arguments presented in Held and Larichev (1996). In particular, the Rhines scale associated with the end of the inverse cascade shifts to larger scale in the presence of a moisture gradient. We confirmed this numerically by calculating the Rhines scale from the root-mean-square barotropic velocity.
The moist available potential energy (MAPE) defined here is equivalent to the framework of Lapeyre and Held (2004) with different partitioning. In their framework, a moist static energy and a moisture deficit/surplus term form the basis for the quadratics. Ours keeps the dry APE intact and defines ME as a quadratic of the condensation level, which is conserved in the absence of diabatic forcing. Our ME more closely resembles that of tropical models such as Pauluis et al. (2008) and Frierson et al. (2004). Likewise, the ME of Smith and Stechmann (2017) can be reworked into a quadratic of a quantity proportional to the dry potential temperature of a system brought adiabatically to saturation. This partitioning can be of particular use away from the assumption of strict quasi equilibrium and uniform evaporation rate employed here. These assumptions result in a system that acts as a highly efficient moist heat engine. More realistic moist systems are not fully saturated, and as a result, tend to have reduced mechanical efficiency stemming from diabatic processes that primarily act on the ME. This partitioning emphasizes that moisture contributes to the EKE through the precipitation term, and would allow for a direct study of the factors within the ME tendency that reduce the efficiency in a partially saturated system. More work is needed to determine the analog of the ME described here in more realistic models of moist atmospheric dynamics. Based on our results, it is worth exploring whether the lifting condensation level could be used to define a generalized ME.
The moist criticality parameter suggests that the classic baroclinic adjustment arguments, such as in Stone (1978), might be improved upon by a framework which includes the poleward transport of latent heat and its impact of static stability. Baroclinic adjustment predicts that the dry criticality of Earth’s atmosphere will remain around ξ ≈ 1 across different climate forcings. However, if the threshold of instability is instead determined by a moist criticality, this reduces the temperature gradient necessary to trigger efficient heat transport.
For insight, consider a planet warming uniformly at the surface. We can estimate an increase in the moisture availability by 7% K−1, translating to a comparable increase in the gradient. Naively, one could start by simply increasing parameter μs, leaving the temperature gradient fixed (i.e., fixed dry criticality ξ). This will lead to greater instability and more energy transport poleward. The increase in meridional heat transport, however, would demand a change in the energy balance at the top of the atmosphere. If we instead assume that the top of the atmosphere balance remains about the same, then it is the total meridional transport of energy that is fixed. If the energy transport scales with the saturated criticality ξs = μsξ, then the dry criticality decreases proportionately to compensate for the increase in the moisture gradient.
Further research could help clarify the results of this study and its connections to other work. Our spectral analysis partially supports the findings of studies like Bembenek et al. (2020) and Lutsko and Hell (2021), but a more direct comparison can be achieved by applying our energetic framework to partially saturated systems with nonhomogeneous background states. Additionally, warming predicts an increase in both the dry static stability and the moisture content (Frierson et al. 2006). As such, the absolute scale changes may differ from the relative changes described here. Indeed, studies that consider the case where moisture compensates for changes in dry static stability (Zurita-Gotor 2005; Juckes 2000; Moore and Montgomery 2004) often reach opposite conclusions to ours.
Another natural follow-up is the case of partial saturation, which introduces additional complications. The decorrelation of moisture from temperature adds an additional degree of freedom, requiring a full consideration of the corresponding terms in the energy and moist PV that were neglected here. This opens new possibilities for the cascade behavior of ME, which could result in energy generation at scales smaller than those predicted in either the dry or saturated case. The energetic framework provided here can be used to analyze these additional terms and determine the changes to scalings associated with them.
Acknowledgments.
We thank Shafer Smith for guidance on turbulence and implementing QG, and Ángel F. Adames-Corraliza and two anonymous reviewers for helpful comments on a previous version of this manuscript. MLB and EPG acknowledge support from the U.S. National Science Foundation through Award AGS-1852727. MLB and OP acknowledge support from the National Science Foundation under Grant HDR-1940145 and from the New York University in Abu Dhabi Research Institute under Grant G1102.
Data availability statement.
The code used to generate the data in this study is stored in the repository at https://github.com/margueriti/Moist_QG_public.
APPENDIX A
Equivalence of Dry and Saturated Limits
APPENDIX B
Omega Equation
APPENDIX C
Precipitation Closure
REFERENCES
Adames, Á. F., 2021: Interactions between water vapor, potential vorticity, and vertical wind shear in quasi-geostrophic motions: Implications for rotational tropical motion systems. J. Atmos. Sci., 78, 903–923, https://doi.org/10.1175/JAS-D-20-0205.1.
Adames, Á. F., and Y. Ming, 2018: Interactions between water vapor and potential vorticity in synoptic-scale monsoonal disturbances: Moisture vortex instability. J. Atmos. Sci., 75, 2083–2106, https://doi.org/10.1175/JAS-D-17-0310.1.
Bembenek, E., D. N. Straub, and T. M. Merlis, 2020: Effects of moisture in a two-layer model of the midlatitude jet stream. J. Atmos. Sci., 77, 131–147, https://doi.org/10.1175/JAS-D-19-0021.1.
Betts, A. K., and M. J. Miller, 1986: A new convective adjustment scheme. Part II: Single column tests using GATE wave, BOMEX, ACTEX and arctic air-mass data sets. Quart. J. Roy. Meteor. Soc., 112, 693–709, https://doi.org/10.1002/qj.49711247308.
Bretherton, C. S., M. E. Peters, and L. E. Back, 2004: Relationships between water vapor path and precipitation over the tropical oceans. J. Climate, 17, 1517–1528, https://doi.org/10.1175/1520-0442(2004)017<1517:RBWVPA>2.0.CO;2.
de Vries, H., J. Methven, T. H. A. Frame, and B. J. Hoskins, 2010: Baroclinic waves with parameterized effects of moisture interpreted using Rossby wave components. J. Atmos. Sci., 67, 2766–2784, https://doi.org/10.1175/2010JAS3410.1.
Edwards, T. K., L. M. Smith, and S. N. Stechmann, 2019: Spectra of atmospheric water in precipitating quasi-geostrophic turbulence. Geophys. Astrophys. Fluid Dyn., 114, 715–741, https://doi.org/10.1080/03091929.2019.1692205.
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, https://doi.org/10.1175/1520-0469(1987)044<1559:BIIAEO>2.0.CO;2.
Emanuel, K. A., J. D. 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.
Fantini, M., 1990: Nongeostrophic corrections to the eigensolutions of a moist baroclinic instability problem. J. Atmos. Sci., 47, 1277–1287, https://doi.org/10.1175/1520-0469(1990)047<1277:NCTTEO>2.0.CO;2.
Fantini, M., 1995: Moist Eady waves in a quasigeostrophic three-dimensional model. J. Atmos. Sci., 52, 2473–2485, https://doi.org/10.1175/1520-0469(1995)052<2473:MEWIAQ>2.0.CO;2.
Frierson, D. M. W., A. J. Majda, and O. M. Pauluis, 2004: Large scale dynamics of precipitation fronts in the tropical atmosphere: A novel relaxation limit. Commun. Math. Sci., 2, 591–626, https://doi.org/10.4310/CMS.2004.v2.n4.a3.
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, https://doi.org/10.1175/JAS3753.1.
Held, I. M., and V. D. Larichev, 1996: A scaling theory for horizontally homogeneous, baroclinically unstable flow on a beta plane. J. Atmos. Sci., 53, 946–952, https://doi.org/10.1175/1520-0469(1996)053<0946:ASTFHH>2.0.CO;2.
Inoue, K., and L. E. Back, 2015: Gross moist stability assessment during TOGA COARE: Various interpretations of gross moist stability. J. Atmos. Sci., 72, 4148–4166, https://doi.org/10.1175/JAS-D-15-0092.1.
Inoue, K., and L. E. Back, 2017: Gross moist stability analysis: Assessment of satellite-based products in the GMS plane. J. Atmos. Sci., 74, 1819–1837, https://doi.org/10.1175/JAS-D-16-0218.1.
Joly, A., and A. J. Thorpe, 1989: Warm and occluded fronts in two-dimensional moist baroclinic instability. Quart. J. Roy. Meteor. Soc., 115, 513–534, https://doi.org/10.1002/qj.49711548705.
Juckes, M. N., 2000: The static stability of the midlatitude troposphere: The relevance of moisture. J. Atmos. Sci., 57, 3050–3057, https://doi.org/10.1175/1520-0469(2000)057<3050:TSSOTM>2.0.CO;2.
Lambaerts, J., G. Lapeyre, and V. Zeitlin, 2011: Moist versus dry barotropic instability in a shallow-water model of the atmosphere with moist convection. J. Atmos. Sci., 68, 1234–1252, https://doi.org/10.1175/2011JAS3540.1.
Lambaerts, J., G. Lapeyre, and V. Zeitlin, 2012: Moist versus dry baroclinic instability in a simplified two-layer atmospheric model with condensation and latent heat release. J. Atmos. Sci., 69, 1405–1426, https://doi.org/10.1175/JAS-D-11-0205.1.
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, https://doi.org/10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2.
Lutsko, N. J., and M. C. Hell, 2021: Moisture and the persistence of annular modes. J. Atmos. Sci., 78, 3951–3964, https://doi.org/10.1175/JAS-D-21-0055.1.
Moore, R. W., and M. T. Montgomery, 2004: Reexamining the dynamics of short-scale, diabatic Rossby waves and their role in midlatitude moist cyclogenesis. J. Atmos. Sci., 61, 754–768, https://doi.org/10.1175/1520-0469(2004)061<0754:RTDOSD>2.0.CO;2.
Neelin, J. D., and I. M. Held, 1987: Modeling tropical convergence based on the moist static energy budget. Mon. Wea. Rev., 115, 3–12, https://doi.org/10.1175/1520-0493(1987)115<0003:MTCBOT>2.0.CO;2.
Parker, D. J., and A. J. Thorpe, 1995: Conditional convective heating in a baroclinic atmosphere: A model of convective frontogenesis. J. Atmos. Sci., 52, 1699–1711, https://doi.org/10.1175/1520-0469(1995)052<1699:CCHIAB>2.0.CO;2.
Pauluis, O., D. M. W. Frierson, and A. J. Majda, 2008: Precipitation fronts and the reflection and transmission of tropical disturbances. Quart. J. Roy. Meteor. Soc., 134, 913–930, https://doi.org/10.1002/qj.250.
Pauluis, O., A. Czaja, and R. Korty, 2010: The global atmospheric circulation in moist isentropic coordinates. J. Climate, 23, 3077–3093, https://doi.org/10.1175/2009JCLI2789.1.
Schneider, T., and P. A. O’Gorman, 2008: Moist convection and the thermal stratification of the extratropical troposphere. J. Atmos. Sci., 65, 3571–3583, https://doi.org/10.1175/2008JAS2652.1.
Shaw, T. A., and Coauthors, 2016: Storm track processes and the opposing influences of climate change. Nat. Geosci., 9, 656–664, https://doi.org/10.1038/ngeo2783.
Smith, L. M., and S. N. Stechmann, 2017: Precipitating quasigeostrophic equations and potential vorticity inversion with phase changes. J. Atmos. Sci., 74, 3285–3303, https://doi.org/10.1175/JAS-D-17-0023.1.
Stone, P. H., 1978: Baroclinic adjustment. J. Atmos. Sci., 35, 561–571, https://doi.org/10.1175/1520-0469(1978)035<0561:BA>2.0.CO;2.
Vallis, G. K., 2006: Atmospheric and Oceanic Fluid Dynamics: Fundamentals and Large-Scale Circulation. 1st ed. Cambridge University Press, 769 pp.
Wu, Y., and O. Pauluis, 2014: Midlatitude tropopause and low-level moisture. J. Atmos. Sci., 71, 1187–1200, https://doi.org/10.1175/JAS-D-13-0154.1.
Zurita-Gotor, P., 2005: Updraft/downdraft constraints for moist baroclinic modes and their implications for the short-wave cutoff and maximum growth rate. J. Atmos. Sci., 62, 4450–4458, https://doi.org/10.1175/JAS3630.1.