1. Introduction
The equations governing fluid dynamics are fundamentally conservative, representing conservation of momentum, energy, and mass (Batchelor 1967). The equations of climate modeling, which are based on dynamical approximations to the governing equations of fluid dynamics, preserve the conservation properties to avoid introducing spurious sources or sinks of the conserved quantities (Lorenz 1967). The numerical solution to the equations of climate modeling requires the parameterization of processes that occur on scales smaller than can be represented by the discrete model grid. Physical processes occurring on these subgrid scales are important to the resolved energy and momentum budgets and necessarily respect the same conservation principles.
The tools of systematic multiple-scale perturbation theory from applied mathematics (Kevorkian and Cole 1981) are naturally suited for developing a framework to study the interaction of physical phenomena occurring on multiple spatial and temporal scales (i.e., the resolved and subgrid scales). Klein (2000) and Majda and Klein (2003) have shown how this theory can be used to systematically derive balanced models in the midlatitudes and tropics. Here, the goal is not to derive an asymptotic framework from first principles. Rather, the goal is to find an asymptotic framework that leads to the equations of interest and to use that framework to define a self-consistent treatment of energy and momentum in the context in which there is an imposed separation of horizontal length and time scales but vertical coupling within each column. This is the case of relevance to climate models. Note that the assumed time scale separation imposes statistical stationarity on the subgrid-scale processes.
Section 2 introduces the equations and relevant nondimensionalization. We introduce the multiple-scale framework in section 3, including the derivation of momentum, thermodynamic, and continuity equations for the resolved and subgrid-scale flow including the resolved-scale total energy budget. The horizontal space and time scale separation between the resolved and subgrid scales is used to define wave-activity conservation laws on the subgrid scale in section 4, which are used to close the interaction terms. In section 5 the subgrid-scale dynamics are reduced to a subset satisfying the anelastic constraint, which is an important regime for applications (most subgrid-scale parameterizations are formulated using the anelastic equations), and implications for the wave-activity conservation law closures are discussed. We show how the dissipation of subgrid-scale kinetic energy can be assured to lead to an increase in thermodynamic energy and to entropy production. The paper concludes with a summary and discussion in section 6.
2. Preliminaries








3. Equations for the mesoscale planetary-scale interaction



It is clear from (26) and (32) that mesoscale fluxes of momentum, potential temperature, and pressure drive the planetary-scale flow, which obeys hydrostatic dynamics (15). The quasi-linear mesoscale [(17), (18), and (22)] is coupled nonlinearly through these eddy fluxes to the planetary scale. This interaction has implications for the planetary-scale energy and momentum budgets.

4. Understanding the interaction across scales: Wave-activity conservation laws
Wave-activity conservation laws play a central role in the study of fluid dynamical disturbances to a specified background state. In the case of the large-scale circulation of the atmosphere, the Eliassen–Palm wave activity has been crucial to theoretical analysis (Andrews et al. 1987; Shepherd 2003; Vallis 2006). Wave-activity conservation laws can be generally derived using the Hamiltonian structure of geophysical fluid dynamics (Shepherd 1990). This framework should, in principle, allow one to consider general disturbances to a background flow, without making any Wentzel–Kramers–Brillouin (WKB)-type assumptions, and be extendable to finite amplitude. In the case of the dynamics derived in the previous section, these conservation laws can be derived in the usual way because the planetary scale acts as a horizontally and temporally homogeneous background flow for the mesoscale dynamics according to the ansatz (11). (The scales are, however, coupled in the vertical, as is assumed in the parameterization of subgrid scales in climate models.) In particular, the xm, ym, and tm symmetries in the background flow lead to pseudomomentum and pseudoenergy conservation laws (Shepherd 1990).
Shaw and Shepherd (2008, hereafter SS08) derived wave-activity conservation laws for three-dimensional disturbances to a horizontally homogeneous background flow with disturbances governed by the anelastic or Boussinesq equations. In the current nomenclature, the mesoscale is considered the disturbance and the planetary scale is the background flow. The mesoscale dynamics derived in section 3 are not explicitly anelastic; however, the results of SS08 can be extended to the situation considered here. A detailed derivation can be found in appendix A. The anelastic form of the mesoscale fluxes and multiscale interactions, and their connection to the wave-activity conservation laws, is presented in the next section.






The vertical wave-activity fluxes in (41) are important in driving the planetary scale. In particular, the pseudomomentum (41b) forces the planetary-scale momentum via (33), whereas the pseudoenergy (41a) contributes to forcing the total energy on the planetary scale via (38). The wave-activity conservation laws provide a means of relating the mesoscale fluxes in (33) and (38) to source–sink terms on the mesoscale. The other mesoscale terms on the right-hand side of (38) have already been related to source–sink terms on the mesoscale. In the absence of wave-activity sources–sinks, the subgrid-scales do not contribute to the planetary-scale budgets and thus (41a) and (41b) can be thought of as “non-acceleration” theorems (Charney and Drazin 1961; Eliassen and Palm 1961) for the effects of subgrid-scale disturbances.





Relation (45) represents the generalization of the second Eliassen–Palm theorem to three dimensions with a veering background flow and to dissipative dynamics. Lindzen (1973) extended the second Eliassen–Palm theorem to include diabatic forcing but assumed WKB conditions in the vertical. The plane-parallel version of (45) with Svm = 0, in particular the first line and last equality, corresponds exactly to (8) in Lindzen (1973) and to (8.24) in Lindzen (1990). Here we have shown that the second Eliassen–Palm theorem can be derived systematically using the Hamiltonian structure of geophysical fluid dynamics in the general context of the interaction between a mesoscale (subgrid-scale) flow and a planetary-scale (resolved) flow. Furthermore, we have generalized the relation to three dimensions with a veering background flow, and the wave-activity source–sink terms take full account of the background vertical shear (there is no WKB-type requirement on the background flow).
5. Anelastic dynamics

The neglect of these terms to obtain the reduced planetary-scale dynamics can be justified by considering the small Mach number limit of the dynamics presented in section 3. [Note that section 3 considered an O(1) Mach number.] It is clear from (18) that in the limit of small Mach number, the anelastic constraint ∇m · (ρpvm) = 0 is recovered. In this limit, the integrity of the mesoscale thermodynamic equation (17) can be preserved if the stratification on the planetary scale is assumed to be O(M2). A weak stratification is a well-known assumption of the anelastic equations (Lipps and Hemler 1982; Klein 2000). It is clear that these assumptions result in the neglect of terms proportional to
Zeytounian (1990) derived the small-Mach number limit of the hydrostatic primitive equations, which he called the reduced hydrostatic primitive equations, by first applying the shallow atmosphere approximation (equivalent to ϵ ≪ 1 in section 3) and then applying the small Mach number limit (see his section 7.4). The small Mach number limit was applied to the hydrostatic primitive equations in pressure coordinates, and the leading-order pressure and enthalpy contributions were O(M2). Our reduced planetary-scale dynamics are in agreement with (7.57) of Zeytounian (1990).
Interestingly, the anelastic constraint only affects the horizontal wave-activity fluxes; the vertical wave-activity fluxes are unchanged. A plausible reason for this is that the hydrostatic primitive equations admit only horizontally propagating sound waves, so any coupling with the mesoscale through compressible dynamics can only occur through the horizontal fluxes. Thus, the anelastic limit, which filters out sound waves, has no effect on the vertical fluxes, which are responsible for forcing the planetary-scale momentum and energy.
As in section 4, the wave-activity conservation laws provide a means of relating the mesoscale flux terms in (50a) and (50b) to mesoscale source–sink terms. [The other terms on the right-hand side of (50b) have already been related to source–sink terms on the mesoscale.] As discussed in the previous section, (53) is the generalization of the second Eliassen–Palm theorem for the case of anelastic dynamics.


According to the first Eliassen–Palm theorem (43), upward propagation implies that momentum flux deposition must drag the background flow toward the phase velocity of the waves, such that the nonlocal transfer term in (55) is positive in the wave dissipation region. We obtain the same result from the requirement of entropy production, provided the local transfers are not positive. Our framework, however, shows that a complete statement requires consideration of the local transfers as well as of the wave source region (which is not generally taken into account in gravity wave parameterizations).
It is apparent from (55) that the sign of the various contributions to the entropy budget depends crucially on the structure of the background temperature Tp. If we consider the case of stable stratification on the planetary scale and assume that mesoscale mixing is downgradient, then
However, the second law of thermodynamics is a global constraint and cannot be ensured from local considerations as in (57). For complete consistency, it is necessary to consider the entropy constraint (55), as discussed above. In this case the local transfers do not integrate out in the vertical column, as they do in the enthalpy budget.
6. Summary and discussion
By combining the theories of multiple-scale asymptotics and Hamiltonian geophysical fluid dynamics, we have derived a self-consistent (in terms of energy and momentum) theoretical framework for physical parameterization in climate models. We have derived energy (38) and momentum (33) equations for a nonlinear, hydrostatic resolved (planetary) scale, which include interactions with a quasi-linear, compressible or anelastic subgrid-scale (mesoscale) described by (17), (18), and (22). The temporal and horizontal spatial symmetries in the planetary-scale background flow for the mesoscale allow the construction of wave-activity conservation laws (A10), (A21), (B3), and (B6) on the mesoscale. These conservation laws are used to understand the fluxing of energy and momentum between scales and the ultimate dissipation of mesoscale kinetic energy, conversion to planetary-scale internal energy, and increase in planetary-scale entropy. In particular, they are used to relate mesoscale flux terms in the planetary-scale momentum and energy budgets to mesoscale source–sink terms, thereby providing generalized nonacceleration theorems.
The relationships between the subgrid-scale fluxes (45) and (53) derived here, which result from self-consistency in terms of energy and momentum conservation, are generalizations of the second Eliassen–Palm theorem and place strong constraints on the contributions of subgrid-scale fluxes to the resolved energy and momentum budgets. Any parameterization of subgrid-scale momentum fluxes must satisfy them to ensure the conservation of both energy and momentum and conform to the second law of thermodynamics. This includes wave drag, cumulus, and boundary layer parameterizations. In particular, entropy production is only guaranteed if (49) and (55) are satisfied. These relations also place strong constraints on parameterized fluxes and may be especially useful in cases where observational constraints are lacking. [Note that (45), (53), (49), and (55) are valid whether or not a phase velocity can be usefully defined.] For example, mixing by gravity wave breaking is constrained not only by energy considerations but also by the requirement of entropy production. Current treatments of kinetic energy dissipation that assume local conservation of energy by balancing a kinetic energy tendency locally by a thermodynamic tendency, as in (1), are incorrect according to (57c) and may lead to spurious sources/sinks of energy and a violation of the second law. [The local conservation formulation (1) can also be seen to be in error according to (46) because it neglects the second term on the right-hand side of the second equality, which involves a nonlocal transfer by the turbulent microscale flow.] Ensuring that particular parameterizations satisfy the entropy production constraint in the vertical column, (49) and (55), is the subject of future investigation.
While the motivation of the framework was the parameterization of subgrid-scale processes in climate models, numerical weather prediction models, which are generally of higher spatial resolution, must also parameterize the transfer of energy and momentum between the resolved and subgrid scales. Thus, the relationships between the fluxes derived here are also relevant for such models.
In the context of gravity wave propagation, Lindzen (1973) showed that the thermodynamic energy tendency due to gravity wave dissipation has the form (57b); however, a constant background wind was assumed in the derivation. In the context of gravity wave drag parameterization, the Lindzen (1981) parameterization does not represent the thermodynamic energy tendency according to (57b), although Becker and Schmitz (2002) extended Lindzen’s parameterization to include the correct tendency [see their Eq. (15)]. In the Hines (1997) parameterization, an analogous thermodynamic energy tendency is used; however, it is scaled by a fudge factor Φ5 with a suggested range of 1 ≤ Φ5 ≤ 3. According to the current analysis, Φ5 must equal unity to ensure consistency between energy and momentum conservation. In all the above studies, the thermodynamic tendency is only applied in the dissipation (wave breaking) region; the present framework implies that it should be also applied in the generation region (and be associated with the wave sources).
Under the time scale separation assumption of our framework, which is that typically made in the parameterization of subgrid scales in climate models, the mesoscale is forced to be statistically stationary on the planetary scale. This leads to the wave-activity conservation laws being diagnostic relations [e.g., (41) and (51)]. Some climate models include turbulent kinetic energy equations that are prognostic. To include prognostic effects on the mesoscale, the theory would need to incorporate an intermittency factor to preserve the scalings on the planetary scale.
A weakness of our framework is that the mesoscale dynamics are assumed to be quasi-linear, in the sense that they are described by the small-amplitude form of the wave-activity conservation laws. This would seem to limit the applicability of the framework (e.g., convection and boundary layer turbulence are clearly nonlinear processes). The introduction of an intermittency factor should also make it possible to allow a nonlinear mesoscale without disturbing the balances on the planetary scale, and the Hamiltonian framework ensures that the mesoscale wave-activity conservation laws can be extended to finite amplitude. [Extension to moist dynamics should also be possible; see, e.g., Bannon (2003).] Unfortunately, in this case the direct connection between the vertical wave-activity fluxes and the planetary-scale equations is lost because there is an additional term in the vertical fluxes involving advection of the wave-activity density (Scinocca and Shepherd 1992). (This problem also arises for the classical theory of wave–mean-flow interaction involving the Eliassen–Palm wave activity; see Andrews et al. 1987.) Since the planetary-scale equations remain the same in the case of a nonlinear mesoscale, our framework still provides a general understanding of the transfers of energy and momentum between the resolved and subgrid scales in this case. In particular, every pseudomomentum flux must be accompanied by a pseudoenergy flux. Thus, the parameterizations of convective momentum transport of Schneider and Lindzen (1976) and Gregory et al. (1997), which neglect the energetics, are in error according to (57a). Similarly, parameterizations of the turbulent dissipation of resolved-scale kinetic energy in the boundary layer that neglect the associated dissipative heating are also in error. Computing the actual importance of these errors in realistic applications is the subject of future investigation.
Acknowledgments
This research has been supported by the Natural Sciences and Engineering Research Council of Canada, in part through a Canada Graduate Scholarship to the first author, and by the Canadian Foundation for Climate and Atmospheric Sciences. The authors are grateful to Dr. N. A. McFarlane and Dr. E. Becker for many helpful discussions and to Dr. R. Klein and two anonymous reviewers for helping to improve the manuscript. The first author also acknowledges support from the Canadian Meteorological and Oceanographic Society and Zonta International.
REFERENCES
Andrews, D. G., J. R. Holton, and C. B. Leovy, 1987: Middle Atmosphere Dynamics. Academic Press, 489 pp.
Bannon, P. R., 2003: Hamiltonian description of idealized binary geophysical fluids. J. Atmos. Sci., 60 , 2809–2819.
Batchelor, G. K., 1967: An Introduction to Fluid Dynamics. Cambridge University Press, 616 pp.
Becker, E., 2003: Frictional heating in global climate models. Mon. Wea. Rev., 131 , 508–520.
Becker, E., and G. Schmitz, 2002: Energy deposition and turbulent dissipation owing to gravity waves in the mesosphere. J. Atmos. Sci., 59 , 54–68.
Boville, B. A., and C. S. Bretherton, 2003: Heating and kinetic energy dissipation in the NCAR community atmosphere model. J. Climate, 16 , 3877–3887.
Bretherton, F. P., 1966: The propagation of groups of internal gravity waves in a shear flow. Quart. J. Roy. Meteor. Soc., 92 , 466–480.
Burkhardt, U., and E. Becker, 2006: A consistent diffusion–dissipation parameterization in the ECHAM climate model. Mon. Wea. Rev., 134 , 1194–1204.
Charney, J. G., and P. G. Drazin, 1961: Propagation of planetary-scale disturbances from the lower into the upper atmosphere. J. Geophys. Res., 66 , 83–109.
Eliassen, A., and E. Palm, 1961: On the transfer of energy by mountain waves. Geofys. Publ., 22 , 94–101.
Gregory, D., R. Kershaw, and P. M. Inness, 1997: Parameterization of momentum transport by convection. II: Tests in single column and general circulation models. Quart. J. Roy. Meteor. Soc., 123 , 1153–1183.
Hines, C. O., 1997: Doppler-spread parameterization of gravity wave momentum dissipation in the middle atmosphere. Part 1: Basic formulation. J. Atmos. Sol.-Terr. Phys., 59 , 371–386.
Hines, C. O., and C. A. Reddy, 1967: On the propagation of atmospheric gravity waves through regions of wind shear. J. Geophys. Res., 72 , 1015–1034.
Kevorkian, J., and J. D. Cole, 1981: Perturbation Methods in Applied Mathematics. Applied Mathematics Sciences Series, Vol. 34, Springer-Verlag, 558 pp.
Klein, R., 2000: Asymptotic analyses for atmospheric flows and the construction of asymptotically adaptive numerical methods. Z. Angew. Math. Mech., 80 , 765–777.
Lindzen, R. S., 1973: Wave–mean flow interactions in the upper atmosphere. Bound.-Layer Meteor., 4 , 327–343.
Lindzen, R. S., 1981: Turbulence and stress owing to gravity wave and tidal breakdown. J. Geophys. Res., 86 , 9707–9714.
Lindzen, R. S., 1990: Dynamics in Atmospheric Physics. Cambridge University Press, 310 pp.
Lipps, F. B., and R. S. Hemler, 1982: A scale analysis of deep moist convection and some related numerical calculations. J. Atmos. Sci., 39 , 2192–2210.
Lorenz, E. N., 1967: The Nature and Theory of the General Circulation of the Atmosphere. World Meteorological Organization, 161 pp.
Majda, A. J., and R. Klein, 2003: Systematic multiscale models for the tropics. J. Atmos. Sci., 60 , 393–408.
Schneider, E. K., and R. S. Lindzen, 1976: A discussion of the parameterization of momentum exchange by cumulus convection. J. Geophys. Res., 81 , 3158–3160.
Scinocca, J. F., and T. G. Shepherd, 1992: Nonlinear wave-activity conservation laws and Hamiltonian structure for the two-dimensional anelastic equations. J. Atmos. Sci., 49 , 5–28.
Shaw, T. A., and T. G. Shepherd, 2007: Angular momentum conservation and gravity wave drag parameterization: Implications for climate models. J. Atmos. Sci., 64 , 190–203.
Shaw, T. A., and T. G. Shepherd, 2008: Wave-activity conservation laws for the three-dimensional anelastic and Boussinesq equations with a horizontally homogeneous background flow. J. Fluid Mech., 594 , 493–506.
Shepherd, T. G., 1990: Symmetries, conservation laws, and Hamiltonian structure in geophysical fluid dynamics. Advances in Geophysics, Vol. 32, Academic Press, 297–338.
Shepherd, T. G., 2003: Hamiltonian dynamics. Encyclopedia of Atmospheric Sciences, J. R. Holton, Ed., Academic Press, 929–938.
Vallis, G. K., 2006: Atmospheric and Oceanic Fluid Dynamics: Fundamentals and Large-Scale Circulation. Cambridge University Press, 745 pp.
Zeytounian, R., 1990: Asymptotic Modeling of Atmospheric Flows. Springer-Verlag, 396 pp.
APPENDIX A
Compressible Wave-Activity Conservation Laws





APPENDIX B
Anelastic Wave-Activity Conservation Laws

Scaling parameters and their values.
In keeping with our motivation, we consider only two time scales. We do not include the O(ϵ−1) time scale in our framework, which is the mesoscale advective time scale. Its neglect does not fundamentally change what is derived in the subsequent sections.
The mesoscale average of (22) contributes to O(ϵ2) according to the mass-weighted average decomposition:
Eliassen and Palm (1961) derived their first and second theorems for the case of steady, two-dimensional gravity waves, assuming conservative dynamics. However, they did not derive conservation laws that relate the vertical fluxes to conserved quantities. Subsequent authors (Bretherton 1966; Hines and Reddy 1967; Lindzen 1973) generalized the Eliassen–Palm theorems to nonsteady disturbances, but they appealed to WKB-type conditions in the vertical (large Richardson number) to define a wave packet with phase velocity c and the associated wave action.