1. Introduction
Studies of the effect of moisture on baroclinic eddies have focused primarily on initial value problems. One body of work isolates the linear growth phase using quasigeostrophic (Bannon 1986; Mak 1982; Wang and Barcilon 1986), semigeostrophic (Emanuel et al. 1987; Joly and Thorpe 1991; Montgomery and Farrell 1991), or primitive (Fantini 1990; Whitaker and Davis 1994) equations. Others study the eddy life cycle starting from linear growth on zonally symmetric flows with the primitive equations (Balasubramanian and Garner 1997; Balasubramanian and Yau 1996; Gall 1976; Gutowski et al. 1992) or with the quasigeostrophic theory (Beare et al. 2003). Yet another set of studies, in a weather prediction context, examines the importance of latent heat release for particular events (Davis 1992; Gyakum 1983; Zhang and Harvey 1995). There is relatively little discussion of how the presence of latent heat release affects the statistically steady state of a baroclinically unstable system. Full general circulation model simulations in which latent heat release is removed or weakened, or in which the mean temperature is greatly altered, are informative in this regard, but, in addition to other complexities, dramatic changes in the Tropics can confound attempts at isolating the in situ effects of latent heat release in midlatitudes (Boer 1995; Hall et al. 1994; Hayashi and Golder 1981; Stephenson and Held 1993; Zhang 1995). More idealized models of moist midlatitude storm tracks are needed to address these issues. An understanding of moist storm tracks is clearly relevant for studying the responses of midlatitudes to global warming (Held 1993).
We choose a quasigeostrophic (QG) two-layer model in this work to examine some of the fundamental properties of the statistically steady moist baroclinic system. The quantitative limitations of QG dynamics for studies of midlatitude dynamics are well known, and can be particularly evident when examining the effects of moisture on linear baroclinic growth (e.g., Emanuel et al. 1987), but the hope is that the simplicity of the QG framework will lead to new qualitative insights. One of the problems in using QG moist models is that one assumes in QG that horizontal structure in the static stability is small. However, the effective static stability is reduced in regions of latent heat release, potentially creating large changes in the horizontal in this effective stability. The reader should keep this limitation in mind as we try to take advantage of the simplicity of QG dynamics to isolate concepts that might have more general validity.
In constructing our two-layer model, we closely follow Zhang (1995) but we have chosen to simplify the problem even further by considering a horizontally homogeneous, doubly periodic geometry. The close relationship between horizontal homogeneous turbulence models and more traditional inhomogeneous models has been demonstrated by Pavan and Held (1996) in the dry two-level QG model. This close correspondence implies that one can use the homogeneous framework to test closure theories for eddy heat fluxes in the cleanest possible setting (Held and Larichev 1996). Eddy vorticity and momentum fluxes vanish identically in these homogeneous simulations, and eddy potential vorticity fluxes are identical to eddy heat, or thickness, fluxes to within a constant of proportionality. Most importantly, one need not, as in inhomogeneous models, simultaneously consider the effects of baroclinic eddies on the larger-scale environment and the effects of this environment on the baroclinic eddies; rather one can focus exclusively on the latter. These simplifications justify the abstraction of this homogeneous setting, in our view.
In section 2, we introduce our QG model of moist baroclinic turbulence and discuss briefly what range of parameters seems most relevant for the atmosphere. In section 3, limiting cases are presented to introduce the basic dynamics of the moist model. In section 4, we examine the statistically steady states when we vary some of the parameters controlling the moist dynamics. Sections 5 and 6 are devoted to a discussion of the underlying processes that control moist turbulence in this model while section 7 contains our conclusions.
2. Description of the model
a. Model equations






















This is effectively the model of Zhang (1995). We now take the additional step of creating a horizontally homogeneous model. As in the studies of Salmon (1978), Haidvogel and Held (1980), and Held and Larichev (1996), we begin by specifying a uniform zonal flow in each layer Ui, or, equivalently, a uniform interface slope ∂
















One can integrate the moist and dry PV equations [Eqs. (5a), (5b), and (6)] forward in time and then obtain the mixing ratio from the difference, so as to avoid computing the divergence explicitly, or one can obtain the divergence diagnostically from the vorticity and thickness equations, and integrate the moisture equation (5c) directly. We have used this second method. The numerical model integrates Eq. (5a)–(5d) in a doubly periodic domain using a pseudospectral code with a leapfrog adaptive time step, adapted from Smith et al. (2002) and some of its specificities are explained in the appendix. The horizontal resolution is 256 × 256 grid points. The model precipitation would likely be better behaved treating water vapor as a grid rather than spectral variable and advecting it with a monotonic scheme, but we leave this possible improvement for future work. A small number of runs were performed with resolution up to 1024 × 1024. The added resolution results in more aesthetically pleasing precipitation distributions, but our tentative conclusion is that the large-scale structure and energetics of the flow are insensitive to these details.
b. Parameter selection
Even if it is the case that the subgrid-scale diffusion, resolution, and the size of the domain are not significant to the results, we still have five parameters (
The equal-depth dry homogeneous two-layer model is characterized by two nondimensional parameters:
The periodic domain is a square of size 18πλ, corresponding to a smallest nonzero wavenumber of 1/(9λ) that allows some resolution of the cascades of PV and moisture to small scales while leaving room between the energy containing eddies and the size of the domain. In the following, we will call the simulation with
Regarding moist parameters, we first note that
There are several alternatives for estimating
If one chooses U ≈ 20 m s−1, λ ≈ 1000 km, and an evaporation time scale m0/E ≈ 7.5 days, one estimates
We cannot expect dry and moist models in this homogeneous framework to both closely mimic the atmosphere, holding other parameters fixed, especially given the large differences in eddy energy that emerge. But this model allows us to focus on particular dynamical processes more difficult to identify in a more complex framework.
3. Limiting cases
In thinking about the effects of moisture, it is useful to keep in mind two limiting cases: the passive limit (
a. Passive moisture transport










This result is exact for a growing normal mode, but it continues to hold with considerable accuracy for the fully turbulent nonlinear model. We generate a statistically steady state of the dry model, in the presence of Ekman damping and small-scale dissipation, for several values of supercriticality, and with


Another important consequence of the conservation equation for ξ is that moisture is strongly correlated with lower-layer vorticity at small scales. Conservation of ξ intuitively implies a cascade of ξ to small scales and the dissipation of its variance. Indeed, in the case without precipitation, and if the dissipation operator is identical for all variables, then ξ simply tends to 0. If the dissipation operators are different, ξ still tends to be small compared with its component fields. As the cascade of temperature ψ1 − ψ2 is stopped near the deformation radius in baroclinic turbulence (Larichev and Held 1995; Salmon 1980), its spectrum is steeper than that of moisture or vorticity, so there there must be a cancellation between m and (1 +
In a simulation with no precipitation and with different values of
One might expect a strong correlation between lower-layer vorticity and moisture to be related to the Ekman damping in the lower layer, with the associated vertical Ekman pumping causing convergence and moistening. However, this derivation suggests that the Ekman damping reduces what would otherwise be an even more precise correlation by creating a difference in the way that vorticity and moisture are dissipated. In our model, vertical velocities are not caused by Ekman pumping predominately but by baroclinic processes. The mechanism we highlight is more complex than would be expected. First, there is some similarity in the “sources” of moisture and vorticity because of the low-level convergence of moisture (m0∇·u2) and the vortex stretching of vorticity columns (f0∇·u2). However, the important point is that moisture and vorticity are advected by the same horizontal flow. These two facts result in the Lagrangian conservation of ξ, and it is the horizontal flow and the resulting cascade of ξ to small scales that can efficiently create a correlation between m and ζ2.
b. Strong evaporation






4. Numerical results
To begin an exploration of the phase space, we chose to vary parameters
The case of moisture at saturation in section 3b suggests that the dynamics might behave similarly as a function of
In an inhomogeneous environment, with a fixed amount of energy supplied to the system, the eddy energies would not grow this dramatically for the energy throughflow could not be maintained; rather, the mean state would have to find a way to adjust to prevent it from falling deep into regime II. The implication for inhomogeneous flows may be that one should expect adjustment back toward the boundary of this regime.
Concerning the dependence on
Baroclinic turbulence on the β plane is characterized by the presence of jets and vortices (Panetta 1993). To better understand the effect of large-scale precipitation on the storm track dynamics, it is important to examine how these coherent structures are affected by the presence of moisture. Figures 3a–c show the total “dry” potential vorticity [〈q1〉 +
Concerning the moisture distribution, the evaporation rate
5. Effective static stability and energetics
a. Global effective static stability
One can interpret our moist simulations by assuming that the basic effect of latent heat release is a reduction in the effective static stability of the atmosphere. This would tend to enhance the effective supercriticality of the flow and the turbulent field would behave more and more like f-plane turbulence. This picture is in agreement with the simulations as we observe a decrease in the number of jets and the appearance of vortices that eventually grow in size.








We attempt to show in the following that by using Eq. (13) to define the parameter α in Eq. (12), one obtains an effective static stability which, when used in a dry model, does simulate many of the eddy statistics of the moist model. But we have no theory for how α varies with model parameters. We simply take it from the model calculations. Despite this limitation, we think the claim that the effective stability is related to the flux of moisture deficit is important, and potentially relevant for more realistic systems.
To illustrate the usefulness of the effective static stability, one can compare the energy spectra obtained in the moist model and in a equivalent dry model. The dry model uses a reduced deformation radius only in the lower layer. Consistent with the discussion above, we find that this results in a better fit to the moist results than if we modified the upper-level radius of deformation as well. Figure 6a shows the baroclinic and barotropic kinetic energy spectra nondimensioned by ε2/3, where ε is the dissipation of barotropic energy. The reason of this choice comes from the inverse cascade that should imply standard spectra of the form ε2/3 k−5/3 (for the barotropic energy). The dry equivalent model seems to be in agreement with the moist model for simulations in regime I. In regime II, we found that the dry model fails to reproduce the modal energy spectra of the moist simulations (not shown). Very energetic coherent vortices in these simulations create an energy peak at their scales (see the “bump” in the bottom-layer energy spectrum at k ≈ 10 in Fig. 6b). An equivalent dry model would not possess such coherent vortices that dominate the flow energetics.
Another test for our estimation of μeff is to see if the statistics of the moist model are a well-defined function of μeff and if they scatter less than with μsat. Figure 7 displays the total kinetic energy as a function of μeff. The data scatter less if compared to Fig. 2b, which uses μsat (Fig. 7b). Also, we do not see a knee when passing from regime I to regime II (Fig. 7a) which could indicate that the properties of the two regimes are not so statistically different when viewed from this perspective.
The effective static stability we introduced should only affect the lower layer since our argument is based on the lower-layer moisture. What happens in the upper layer? We observe that the vortex asymmetry does not develop in the upper layer (see Fig. 11) and the kinetic energy spectrum in this layer stays similar to the dry case (see Fig. 6b). Therefore the upper-layer eddy statistics seem much less affected by latent heat release than in the lower layer, despite the fact that both layers are forced equally by the latent heat release.
b. Eddy fluxes and energetics
c. Available potential energy
If one had a theory for ε as a function of the parameters (












The MAPE budget will serve to illustrate the difficulties that arise when trying to develop a theory for the eddy heat flux in the moist context. As discussed above, one can relate the moist PV diffusivity to the energy dissipation ε. To close the problem, one needs another relation between ε and the eddy heat flux. Figure 9 shows the standard available potential energy budget of the model, as a function of μsat, holding
This process shows that despite a large latent heat energy available for the total energy, the model extracts only a small portion of this energy, while the remaining serves to balance the moist processes (dehumidification plus diffusion). Thus, there is a high inefficiency of the heat engine, which is unable to convert the latent heat energy into kinetic energy, as explained by Pauluis and Held (2002).
6. Cyclone/anticyclone asymmetry
The moist simulations develop a strong asymmetry between cyclones and anticyclones despite the quasigeostrophic formalism. In the passive case, cyclones and anticyclones have similar properties (amplitude, dynamics, …) because of the internal symmetry of the equations. [See Arbic and Flierl (2001) for an example of a dry QG model with mean easterlies that lead to a cyclone/anticyclone asymmetry.] In the presence of latent heat release, the asymmetry toward stronger cyclones is expected since it is well known that latent heating intensifies cyclonic development near the surface (Emanuel et al. 1987; Stoelinga 1996; Whitaker and Davis 1994). However, the characteristics of vortices in regime II are somewhat different from the ones we might intuitively expect. Therefore, it is useful to examine in details their properties in a similar manner to Polvani et al. (1994) and Legg and McWilliams (2001) in other situations.
The temperature ψ1 − ψ2 associated with the cyclones is strongly positive as expected since latent heat release (associated with precipitation) increases temperature. However, despite this baroclinic forcing, strong vortices have the same sign of vorticity in each layer (see Fig. 10): bottom-layer cyclones are generally cyclonic in vorticity in the upper layer, but with a much smaller magnitude. In contrast, weak bottom-layer anticyclones are associated with stronger anticyclones in the upper layer, consistent with the equivalent barotropic character of the dry model. Starting from a developed dry flow and suddenly turning on latent heat release, we observe a development of small-scale filaments of opposite signs in vorticity in each layer. These filaments are dynamically unstable and roll into vortices. Thus, the forcing is clearly baroclinic but, as the flow equilibrates, vortices tend to barotropize through baroclinic processes.
A natural way to quantify the cyclone/anticyclone asymmetry is to compute the vorticity skewness 〈ω3〉/〈ω2〉3/2 for both layers. Figure 11 shows that the asymmetry develops steadily in both layers for regime I (μsat < 2.5). The upper layer has a weak tendency toward anticyclones, whereas the lower layer displays a strong tendency toward cyclones. When μsat is increased further, a sharp transition occurs and the skewness saturates around 6–7 in the lower layer for regime II. In the upper layer, the skewness remains quite small, toward anticyclones. This diagnostic is consistent with the vorticity spatial field (Fig. 10) as the asymmetry between cyclones and anticyclones is discernible only in the lower layer. In the upper layer, the vortices are less dominant as the vorticity kurtosis (〈ω4〉/〈ω2〉2) is of order 10 there, whereas it is around 50 in the lower layer. This situation contrasts with the passive case and even with regime I as the kurtosis is around 3.6 in the upper layer and 3.4 in the lower one for the passive case and only a bit larger for runs in regime I. Also, the kinetic energy spectra (Fig. 6b) are consistent with the vorticity skewness as we see that bottom-layer vortices are associated with a bump in energy between wavenumbers 3 and 20 near the dry deformation radius (kd = 9) and a very steep slope (close to k−4.5) for high wavenumbers. By contrast, the upper-layer kinetic energy has a standard slope (close to k−3) as the vortices do not dominate the flow there.
Vortices are able to develop in the lower layer despite the presence of friction in this layer. Actually the lower-layer kinetic energy is even larger than the upper-layer energy in regime II, whereas it is the opposite in regime I. Danilov and Gurarie (2001) showed that for the one-layer problem on the f plane, linear friction prevents the appearance of coherent vortices, and this is also the case for the bottom layer of our passive simulation (see Fig. 3d). In the moist case, the additional forcing by latent heat release thus seems to add enough energy so that vortices are able to persist, even in presence of bottom friction. We performed some simulations starting with initial conditions in regime II but with smaller values of
The existence of such cyclones can be related to dynamical feedback between moisture and vorticity in the bottom layer. As explained in section 3a, there is a strong correlation between the lower-layer vorticity and moisture through their conservation properties. This correlation is responsible of the relatively moist character of cyclones and the relatively dry character of anticyclones. This has the consequence that cyclones are able to intensify themselves through latent heat release since their moisture content is near saturation. Such an effect is absent for the anticyclones since they are depleted in moisture. This feedback is likely to act only in the bottom layer because the upper layer is not subject to a correlation between vorticity and moisture. The reason is that the upper-layer vorticity is not advected by the same horizontal flow as is the low-level moisture, and this horizontal advection is central to build the correlation. This explains why the asymmetry is very strong in the lower layer and quite weak in the upper.
To verify that this mechanism is at play in our simulations, we computed the joint probability density functions (pdfs) of bottom-layer vorticity and moisture deficit for different cases (Fig. 12). For the three cases (passive, regime I, and regime II), we see a clear correlation between the two fields: anticyclonic regions tend to have a large moisture deficit, whereas cyclonic regions tend to be saturated. Also for the regimes I and II, the regions at saturation have a broad range of vorticity values preferentially toward cyclonic vorticity. This is most striking for regime II, for which the joint pdf exhibits a local maximum for values of nondimensional vorticity around 5. By contrast, moisture and upper-level vorticity (Fig. 12d) do not show a clear relationship, even if some asymmetry can be seen in the figure. Thus, it is not only the convergence motions (through vortex stretching) that are responsible for the correlation and the intensification of cyclones, but the horizontal advection plays an essential role: moisture and upper-layer vorticity are advected by different flows, which prevent their correlation. To test further if the correlation between vorticity and saturation deficit can be predicted from the passive dynamics, we also added a curve with a slope corresponding to the relation m − ms = (
7. Conclusions
In this paper, we have examined moist baroclinic turbulence in the idealized context of the homogeneous quasigeostrophic two-layer model. We have included moist processes through an explicit moisture equation in the lower layer and with a simple precipitation scheme. Two key parameters (the coefficient of linearization of the Clausius–Clapeyron relationship
We find the existence of two dynamical regimes when increasing
We proposed some interpretation of the mechanisms involved in these simulations. Different diagnostics showed that the moderate regime of moist turbulence behaves as if its global static stability is decreased, which in turn increases the supercriticality of the flow. We have suggested a way of determining a posteriori a global effective static stability and show that it is decreasing with increasing
We find in our model that a diffusive theory for the eddy horizontal flux of moisture using only the horizontal moisture gradient underestimates this flux. Because of correlations between horizontal and vertical motions resulting from baroclinic processes acting in the presence of the β effect, one cannot neglect the influence of off-diagonal components of the diffusivity tensor on the moisture deficit flux. This may suggest modified approaches to moisture transport in more complex models with parameterized eddies.
We have shown that, at least, for the regime of weak latent heating, the concepts of dry baroclinic turbulence can still be applied to these moist simulations, as the flow properties are affected mainly by the variation of the effective static stability. The energy budget can be rewritten in a form that generalizes available potential energy to the moist case. It highlights the role of moisture diffusion and dehumidification of the air in reducing the efficiency of the latent heat release. This result is similar to the conclusions of Pauluis and Held (2002) obtained by an entropy budget for cumulus convection.
One can think that regime I is typical of the conditions of the earth's atmosphere since the extratropics are only weakly baroclinically unstable and latent heating does not modify too strongly the dynamics of synoptic motions. The abrupt transition from regime I to regime II as
Acknowledgments
The authors would like to thank Geoff Vallis and Isidoro Orlanski for helpful discussion about this work. We also thank two anonymous referees for their careful reading of the manuscript and helpful comments. All simulations were performed on the supercomputers at NOAA's Geophysical Fluid Dynamics Laboratory, using a QG code originally written by S. Smith and G. Vallis and parallelized by GL and S. Smith. GL was funded under a grant/cooperative agreement with the National Oceanic and Atmospheric Administration (NA07RG0002).
REFERENCES
Arbic, B. K., and G. R. Flierl, 2004: Effects of mean flow direction on energy, isotropy, and coherence of baroclinically unstable beta-plane geostrophic turbulence. J. Phys. Oceanogr, 34 , 77–93.
Balasubramanian, G., and M. K. Yau, 1996: The life cycle of a simulated marine cyclone: Energetics and PV diagnostics. J. Atmos. Sci, 53 , 639–653.
Balasubramanian, G., and S. T. Garner, 1997: The equilibration of short baroclinic waves. J. Atmos. Sci, 54 , 2850–2871.
Bannon, P. R., 1986: Linear development of quasi-geostrophic baroclinic disturbances with condensational heating. J. Atmos. Sci, 43 , 2261–2274.
Barry, L., G. C. Craig, and J. Thuburn, 2002: Poleward heat transport by the atmospheric heat engine. Nature, 415 , 774–777.
Beare, R. J., A. J. Thorpe, and A. A. White, 2003: The predictability of extratropical cyclones: Nonlinear sensitivity to localized potential-vorticity perturbations. Quart. J. Roy. Meteor. Soc, 129 , 219–237.
Boer, G. J., 1995: Some dynamical consequences of greenhouse gas warming. Atmos.–Ocean, 33 , 731–751.
Cao, Z., and H-R. Cho, 1995: Generation of moist potential vorticity in extratropical cyclones. J. Atmos. Sci, 52 , 3263–3281.
Danilov, S., and D. Gurarie, 2001: Forced two-dimensional turbulence in spectral and physical space. Phys. Rev. E, doi:10.1103/ PhysRevE.63.061208.
Davis, C. A., 1992: A potential vorticity diagnosis of the importance of initial structure and condensational heating in observed extratropical cyclogenesis. Mon. Wea. Rev, 120 , 2409–2428.
Emanuel, K. A., 1985: Frontal circulations in the presence of small moist symmetric stability. J. Atmos. Sci, 42 , 1062–1071.
Emanuel, K. A., and M. Bister, 1996: Moist convective velocity and buoyancy scales. J. Atmos. Sci, 53 , 3276–3285.
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.
Fantini, M., 1990: Nongeostrophic corrections to the eigensolutions of a moist baroclinic instability problem. J. Atmos. Sci, 47 , 1277–1287.
Gall, R., 1976: The effects of released latent heat in growing baroclinic waves. J. Atmos. Sci, 33 , 1686–1701.
Gutowski, W. J., L. E. Branscome, and D. A. Stewart, 1992: Life cycles of moist baroclinic eddies. J. Atmos. Sci, 49 , 306–319.
Gyakum, J. R., 1983: On the evolution of the QE II storm. II: Dynamics and thermodynamic structure. Mon. Wea. Rev, 111 , 1156–1173.
Haidvogel, D. B., and I. M. Held, 1980: Homogeneous quasi-geostrophic turbulence driven by a uniform temperature gradient. J. Atmos. Sci, 37 , 2644–2660.
Hall, N. M. J., B. J. Hoskins, P. J. Valdes, and C. A. Senior, 1994: Storm tracks in a high-resolution GCM with doubled carbon dioxide. Quart. J. Roy. Meteor. Soc, 120 , 1209–1230.
Hayashi, Y., and D. G. Golder, 1981: The effects of condensational heating on midlatitude transient waves in their mature stage: Control experiments with a GFDL general circulation model. J. Atmos. Sci, 38 , 2532–2539.
Held, I. M., 1993: Large-scale dynamics and global warming. Bull. Amer. Meteor. Soc, 74 , 228–241.
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.
Holton, J. R., 1972: Introduction to Dynamic Meteorology. Academic Press, 319 pp.
Joly, A., and A. J. Thorpe, 1991: The stability of time-dependent flows: An application to fronts in developing baroclinic waves. J. Atmos. Sci, 48 , 163–182.
Juckes, M. N., 2000: The static stability of the midlatitude troposphere: The relevance of moisture. J. Atmos. Sci, 57 , 3050–3057.
Klein, P., A. Tréguier, and B. L. Hua, 1998: Three-dimensional stirring of thermohaline fronts. J. Mar. Res, 56 , 589–612.
Lapeyre, G., and I. M. Held, 2003: Diffusivity, kinetic energy dissipation, and closure theories for the poleward eddy heat flux. J. Atmos. Sci, 60 , 2907–2916.
Larichev, V. D., and I. M. Held, 1995: Eddy amplitudes and fluxes in a homogeneous model of fully developed baroclinic instability. J. Phys. Oceanogr, 25 , 2285–2297.
Legg, S., and J. C. McWilliams, 2001: Convective modifications of a geostrophic eddy field. J. Phys. Oceanogr, 31 , 874–891.
Mak, M., 1982: On moist quasi-geostrophic baroclinic instability. J. Atmos. Sci, 39 , 2028–2037.
Montgomery, M. T., and B. F. Farrell, 1991: Moist surface frontogenesis associated with interior potential vorticity anomalies in a semigeostrophic model. J. Atmos. Sci, 48 , 343–367.
Panetta, R. L., 1993: Zonal jets in wide baroclinically unstable regions: Persistence and scale selection. J. Atmos. Sci, 50 , 2073–2106.
Pauluis, O., and I. M. Held, 2002: Entropy budget of an atmosphere in radiative–convective equilibrium. Part II: Latent heat transport and moisture processes. J. Atmos. Sci, 59 , 140–149.
Pavan, V., and I. M. Held, 1996: The diffusive approximation for eddy fluxes in baroclinically unstable jets. J. Atmos. Sci, 53 , 1262–1272.
Peixoto, J. P., and A. H. Oort, 1991: Physics of Climate. American Institute of Physics, 512 pp.
Plumb, R. A., and J. D. Mahlman, 1987: The zonally averaged transport characteristics of the GFDL general circulation/transport model. J. Atmos. Sci, 44 , 298–327.
Polvani, L. M., J. C. McWilliams, M. A. Spall, and R. Ford, 1994: The coherent structures of shallow-water turbulence: Deformation-radius effects, cyclone/anticyclone asymmetry and gravity-wave generation. Chaos, 4 , 177–186.
Rennó, N. O., and A. P. Ingersoll, 1996: Natural convection as a heat engine: A theory for CAPE. J. Atmos. Sci, 53 , 572–585.
Salmon, R., 1978: Two-layer quasi-geostrophic turbulence in a simple special case. Geophys. Astrophys. Fluid Dyn, 10 , 25–52.
Salmon, R., 1980: Baroclinic instability and geostrophic turbulence. Geophys. Astrophys. Fluid Dyn, 15 , 167–211.
Smith, K. S., G. Bocaletti, C. C. Henning, I. N. Marinov, C. Y. Tam, I. M. Held, and G. K. Vallis, 2002: Turbulent diffusion in the geostrophic inverse cascade. J. Fluid Mech, 469 , 14–47.
Stephenson, D. B., and I. M. Held, 1993: GCM response of northern winter stationary waves and storm tracks to increasing amounts of carbon dioxide. J. Climate, 6 , 1859–1870.
Stoelinga, M. T., 1996: A potential vorticity-based study of the role of diabatic heating and friction in a numerically simulated baroclinic cyclone. Mon. Wea. Rev, 124 , 849–874.
Wang, B., and A. Barcilon, 1986: Moist stability of a baroclinic zonal flow with conditionally unstable stratification. J. Atmos. Sci, 43 , 705–719.
Whitaker, J. S., and C. A. Davis, 1994: Cyclogenesis in a saturated environment. J. Atmos. Sci, 51 , 889–907.
Zhang, D-L., and R. Harvey, 1995: Enhancement of extratropical cyclogenesis by a mesoscale convective system. J. Atmos. Sci, 52 , 1107–1127.
Zhang, J., 1995: Effects of latent heating on midlatitude atmospheric general circulation. Ph.D. thesis, Princeton University, 194 pp.
APPENDIX
Numerics







(a) Moisture flux plotted as a function of PV diffusivity in turbulent simulations with no precipitation, for different values of
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

(a) Moisture flux plotted as a function of PV diffusivity in turbulent simulations with no precipitation, for different values of
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2
(a) Moisture flux plotted as a function of PV diffusivity in turbulent simulations with no precipitation, for different values of
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

(a) Total kinetic energy as a function of μsat. Three different sets of runs are represented: plus signs with constant
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

(a) Total kinetic energy as a function of μsat. Three different sets of runs are represented: plus signs with constant
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2
(a) Total kinetic energy as a function of μsat. Three different sets of runs are represented: plus signs with constant
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

Total nondimensional “dry” PV of top layer (first row) and “moist” PV of bottom layer (second row) for (a), (d) a “passive” case (
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

Total nondimensional “dry” PV of top layer (first row) and “moist” PV of bottom layer (second row) for (a), (d) a “passive” case (
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2
Total nondimensional “dry” PV of top layer (first row) and “moist” PV of bottom layer (second row) for (a), (d) a “passive” case (
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

Nondimensional moisture deficit for the same three cases as Fig. 3
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

Nondimensional moisture deficit for the same three cases as Fig. 3
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2
Nondimensional moisture deficit for the same three cases as Fig. 3
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

(a) Nondimensional effective static stability μeff as a function of μsat. The continuous curve corresponds to μsat. A zoom is shown in the insert. Symbols are the same as in Fig. 2. (b) Pseudoprecipitation fraction α (continuous curve) and precipitation fraction (dashed curve) as a function of μsat for set of runs with
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

(a) Nondimensional effective static stability μeff as a function of μsat. The continuous curve corresponds to μsat. A zoom is shown in the insert. Symbols are the same as in Fig. 2. (b) Pseudoprecipitation fraction α (continuous curve) and precipitation fraction (dashed curve) as a function of μsat for set of runs with
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2
(a) Nondimensional effective static stability μeff as a function of μsat. The continuous curve corresponds to μsat. A zoom is shown in the insert. Symbols are the same as in Fig. 2. (b) Pseudoprecipitation fraction α (continuous curve) and precipitation fraction (dashed curve) as a function of μsat for set of runs with
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

(a) Nondimensional barotropic and baroclinic energy spectra for moist simulations (continuous curve) and equivalent dry simulations (dashed curve) with μeff = 1.11, 1.22, 1.39, and μsat = 1.18, 1.58, 2.29 (respectively, lower to upper curves). (b) Kinetic energy spectra of top layer (continuous) and of bottom layer (dashed) for a case in regime II (
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

(a) Nondimensional barotropic and baroclinic energy spectra for moist simulations (continuous curve) and equivalent dry simulations (dashed curve) with μeff = 1.11, 1.22, 1.39, and μsat = 1.18, 1.58, 2.29 (respectively, lower to upper curves). (b) Kinetic energy spectra of top layer (continuous) and of bottom layer (dashed) for a case in regime II (
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2
(a) Nondimensional barotropic and baroclinic energy spectra for moist simulations (continuous curve) and equivalent dry simulations (dashed curve) with μeff = 1.11, 1.22, 1.39, and μsat = 1.18, 1.58, 2.29 (respectively, lower to upper curves). (b) Kinetic energy spectra of top layer (continuous) and of bottom layer (dashed) for a case in regime II (
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

(a) Total kinetic energy as a function of μeff. (b) Same as (a) but for only small μeff. Symbols are the same as in Fig. 2
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

(a) Total kinetic energy as a function of μeff. (b) Same as (a) but for only small μeff. Symbols are the same as in Fig. 2
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2
(a) Total kinetic energy as a function of μeff. (b) Same as (a) but for only small μeff. Symbols are the same as in Fig. 2
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

(a) Moist diffusivity as a function of prediction c ε3/5β−4/5 with c = 1.7 and diffusivities nondimensioned by Uλ0. Symbols are the same as in Fig. 2
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

(a) Moist diffusivity as a function of prediction c ε3/5β−4/5 with c = 1.7 and diffusivities nondimensioned by Uλ0. Symbols are the same as in Fig. 2
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2
(a) Moist diffusivity as a function of prediction c ε3/5β−4/5 with c = 1.7 and diffusivities nondimensioned by Uλ0. Symbols are the same as in Fig. 2
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

Total energy production ε/
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

Total energy production ε/
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2
Total energy production ε/
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

Vorticity field for a case in regime II (i.e., μsat ≈ 3.03): (a) top layer; (b) bottom layer. The square root of the amplitude is shown in order to highlight the vorticity sign
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

Vorticity field for a case in regime II (i.e., μsat ≈ 3.03): (a) top layer; (b) bottom layer. The square root of the amplitude is shown in order to highlight the vorticity sign
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2
Vorticity field for a case in regime II (i.e., μsat ≈ 3.03): (a) top layer; (b) bottom layer. The square root of the amplitude is shown in order to highlight the vorticity sign
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

Vorticity skewness 〈ω3〉/(〈ω2〉)3/2 for (a) upper layer and (b) bottom layer as a function of μsat. Symbols are the same as in Fig. 2
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

Vorticity skewness 〈ω3〉/(〈ω2〉)3/2 for (a) upper layer and (b) bottom layer as a function of μsat. Symbols are the same as in Fig. 2
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2
Vorticity skewness 〈ω3〉/(〈ω2〉)3/2 for (a) upper layer and (b) bottom layer as a function of μsat. Symbols are the same as in Fig. 2
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

Joint pdf of bottom-layer vorticity (abscissa) and moisture deficit (ordinate) in three cases: (a) passive case
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2

Joint pdf of bottom-layer vorticity (abscissa) and moisture deficit (ordinate) in three cases: (a) passive case
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2
Joint pdf of bottom-layer vorticity (abscissa) and moisture deficit (ordinate) in three cases: (a) passive case
Citation: Journal of the Atmospheric Sciences 61, 14; 10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2