1. Introduction
Ogura and Phillips (1962) derived the original anelastic model through systematic formal asymptotics using the flow Mach number as the expansion parameter. Their goal was to derive a set of model equations that would simultaneously represent internal gravity waves and the effects of advection while suppressing any sound modes. Such a model would not only cast the notion that compressibility plays only a subordinate role in the majority of atmospheric flow phenomena in systematic mathematical terms, but it would also lend itself to numerical integration without the necessity of handling the fast yet unimportant acoustic modes through computationally cumbersome numerical means. Such “soundproof” models may also be considered as conceptually important building blocks in a model hierarchy for analyses of fundamental processes of weather and climate (Held 2005).
By design, the soundproof models should be able to address deep atmospheres vertically covering a typical pressure scale height hsc ∼ 10 km or more, and nonhydrostatic flow regimes corresponding to horizontal scales down to 10 km or less (cf. Bannon 1996). Thus, the characteristic vertical and horizontal length scales for the design regime of these models are comparable to hsc. As can be seen in Table 1, the characteristic acoustic time scale tac is small, on the order of O(ε), relative to the advection time tadv, whereas the time scale for internal waves, tint (i.e., the inverse of the Brunt–Väisälä frequency N), is on the order of O{ε[(hsc/θ)dθ/dz]−1/2}, where θ is the potential temperature. As a consequence, if we follow Ogura and Phillips (1962) and construct an asymptotic single time scale model that resolves the advection time scale and includes internal waves at the same time, we would have to adopt a weak stratification so that (hsc/θ)dθ/dz = O(ε2).
For typical flow Mach numbers of


The presence of multiple scales in the soundproof models is, nevertheless, an issue because both the spatial structures and frequencies of internal waves featured by the soundproof models only approximate those represented by the full compressible flow equations. As a consequence, there are two necessary conditions for the validity of the soundproof models over the targeted advective time scales:
- (a) the spatial structures of corresponding internal wave eigenmodes of the soundproof and compressible systems should be asymptotically close as ε → 0, and
- (b) the accumulation of phase differences between such soundproof and compressible internal waves should remain asymptotically small at least over the advective time scale.
- (i) compare the internal wave eigenmode structures of the compressible Euler equations and selected soundproof models;
- (ii) assess the approximation errors due to “soundproofing” for both the spatial eigenmodes and the associated frequencies in terms of the Mach number; and
- (iii) demonstrate, as our main result, that internal wave solutions of the soundproof and compressible models remain asymptotically close for t = O(tadv) for sufficiently weak stratification. Specifically, for both Lipps and Hemler’s and Durran’s soundproof models, the corresponding bound on the stratification isThis corresponds to realistic stratifications with
over 10–15 km.
The rest of the paper is organized as follows. In section 2 we summarize the model equations to be studied. In section 3 we introduce a new set of variables that explicitly reveal the multiscale nature of fully compressible flows within the regime of stratifications from (1). In section 4, using formal asymptotic analysis and vertical mode decompositions, we compare the vertical internal wave eigenmodes and eigenfrequencies for the pseudoincompressible and the Lipps and Hemler anelastic models with those of the compressible equations and show that they are asymptotically close as long as (hsc/θ) dθ/dz = O(εμ) for any μ > 0. In that section we also assess the time it takes compressible and soundproof internal waves to accumulate leading-order deviations of their phases because of these differences in the dispersion relations, and this will lead to the abovementioned principal result in (3). In section 5 we draw conclusions and provide an outlook for future work.
2. Compressible and soundproof model equations
The exposition in this section of the three sets of model equations to be analyzed subsequently closely follows Klein (2009). Here, we restrict our considerations to flows under gravity, but without Coriolis effects and nonresolved-scale closures, and present consistent dimensionless forms of the compressible Euler equations and of two soundproof models.
a. Compressible Euler equations





b. Pseudoincompressible model



c. Anelastic model





3. Scaled variables














Besides the tendencies of temporal change, there are three groups of terms in (16): the terms multiplied by ε−ν induce internal waves, the terms multiplied by ε−1 represent the acoustic modes, and the terms on the right-hand side cover all nonlinearities. In fact, all terms on the left-hand sides are linear in the unknowns. Notice that all terms on the right are nonsingular as ε → 0; that is, they are O(εα) with α ≥ 0. This clean Mach number scaling of acoustic, internal wave, and nonlinear (advective) terms justifies in hindsight the choice ν = 1 − μ/2 introduced earlier.






We observe that the potential temperature transport equations are in agreement between all three models. This was to be expected since in the present adiabatic setting this equation reduces to a simple advection equation. The momentum equations of the compressible and pseudoincompressible models are in complete agreement, whereas the anelastic model’s momentum equation lacks the respective last terms on the left and right from (16b) or (17b) that combine to yield εμ−1(
The only difference between the compressible Euler equations from (16) and the pseudoincompressible model is found in the Exner pressure evolution equation (16c), which becomes the pseudoincompressible divergence constraint in (17c). The anelastic divergence constraint in (18c) again differs from the pseudoincompressible one through an additional term involving the background potential temperature stratification.
4. Internal gravity waves
a. Gravity wave scaling



We observe that in the gravity wave scaling all differences between the compressible model on the one hand and both of the soundproof models on the other hand are O(εμ) or smaller (i.e., at least on the order of the stratification strength). At leading order in ε, all models agree from a formal scaling perspective, although switching off the pressure tendency by letting A = 0 fundamentally changes the mathematical type of the equations from strictly hyperbolic to mixed hyperbolic–elliptic. We will demonstrate below through formal asymptotics that this, nevertheless, affects the internal gravity wave solutions only weakly. Between the pseudoincompressible and anelastic systems there is no such singular switch, however, so that their solutions will differ only by O(εμ) at least on internal wave time scales with ϑ = O(1).
b. The constraint on the stratification




Another indication that at the threshold of μ = ⅔, the dynamics change nontrivially arises as follows. When μ = ⅔ we have ν = 1 − μ/2 = μ, so that the leading nonlinearities on the rhs of (19), which are O(εν), become comparable to the perturbation terms of the linearized system on the lhs of (19), which are O(εμ). Thus, for μ ≤ ⅔ any perturbation analysis of internal waves in compressible flows that go beyond the leading-order solution must necessarily account for nonlinear effects.
Note that there is no noticeable transition or change in the structure of the linear eigenmodes and eigenvalues considered in the next section as μ decreases below the threshold of μ = ⅔. The importance of this threshold is associated entirely with the more subtle effects just explained.
The present estimates rely on the linearized equations. However, since all three models considered feature the same leading nonlinearities represented by the nonlinear advection of potential temperature and velocity in (19a) and (19b) [see the terms O(εν)], we expect asymptotic agreement of the solutions over advective time scales as long as the fast linearized dynamics do not already lead to leading-order deviations between the model results (i.e., as long as μ > ⅔). A mathematically rigorous proof of the validity of the fully nonlinear pseudoincompressible, and possibly the anelastic, models over advective time scales is a work in progress.
c. Vertical mode decomposition and the Sturm–Liouville eigenvalue problem






- (i) There is a sequence of eigenvalues and associated eigenfunctions,
, with 0 < Λ00 < Λ10 … , and Λk0 → ∞ as k → ∞. - (ii) The
form an orthonormal basis of a Hilbert space of functions f:[0, H] → ℝ with scalar product . Note that the scalar product and, thus, the Hilbert space are independent of the horizontal wavenumber λ. - (iii) The vertical mode number k equals the number of zeroes of the associated eigenmodes on the open interval 0 < z < H (i.e., excluding the boundary points). Thus, k = 0 represents the leading, vertically nonoscillatory mode.

d. Asymptotics for the compressible internal wave modes

Notice that there is a set of eigenvalues with Λ = 1/ω2 = O(εμ) that correspond to the system’s high-frequency acoustic modes. Those will not be considered further in this paper.







This determines the first-order perturbations in terms of εμ from (28) and (29). For a forthcoming companion paper, two of the authors are currently working on a rigorous proof that the remainders are actually O(ε2μ) as indicated in (28). A remark is in order. If Λj(p, q, r) is a simple eigenvalue of a Sturm–Liouville operator L(p, q, r) on [0, 1]—that is, if there exists a unique eigenfunction Wj such that L(p, q, r)Wj = −(pW ′j)′ + qWj = rΛjWj with Wj(0) = Wj(1) = 0—then Λj(p, q, r) depends analytically on the functions p and q in a neighborhood of the coefficients. The derivative of the eigenvalue Λj and eigenvector Wj are given by the expressions in (32) and (33) (see Kato 1995; Kong and Zettl 1996).
e. Examples


We observe that the relative deviation of the eigenvalues between the soundproof and compressible cases is surprisingly small in practice. According to our previous analysis, we would expect deviations of the same order of magnitude as the relative vertical variation of the potential temperature, which in the present case is εμ ∼ 0.1. Yet, the maximum relative deviation between the eigenvalues is less than 0.02 (in modulus) for the cases documented in Fig. 2 for mode number k = 0, and it decreases rapidly for larger k. The situation is very similar for other horizontal wavenumbers (not shown).
The deviations in the vertical structure functions are similarly small as demonstrated in an exemplary fashion by the differences in the vertical velocity structure functions,
f. The long-wave limit



5. Conclusions
In this paper we have addressed the formal asymptotics of weakly compressible atmospheric flows involving three asymptotically different time scales for sound, internal waves, and advection. Both the pseudoincompressible and a particular anelastic model yield very good approximations to the linearized internal wave dynamics in a compressible flow for realistic background stratifications and on length scales comparable to the pressure and density scale heights. These soundproof models should be applicable for stratification strengths (hsc/
A number of important open questions remain to be addressed, such as (i) Could either of the soundproof models be also justified even for (hsc/
R. K. thanks the Johns Hopkins University and the U.S. National Center for Atmospheric Research for hosting him during his 2009 sabbatical leave, the Wolfgang Pauli Institute at Wirtschaftsuniversität Wien for their generous hospitality during an intense week of joint research with D. B., Dr. Veerle Ledoux from Gent University for providing the open source Sturm–Liouville eigen-problem solver “MATSLISE,” Deutsche Forschungsgemeinschaft for partial support through the MetStröm Priority Research Program (SPP 1276), and through Grant KL 611/14. U. A. and R. K. both thank the Leibniz-Gemeinschaft (WGL) for partial support within their PAKT program. D. B. acknowledges support from the ANR-08-BLAN-0301-01 project. O. K. and R. K. thank Alexander-von-Humboldt Stiftung for partial support of this work through their Friedrich-Wilhelm-Bessel prize program. P. K. S. acknowledges partial support by the DOE Award DE-FG02-08ER64535.
REFERENCES
Almgren, A. S., , J. B. Bell, , C. A. Rendleman, , and M. Zingale, 2006: Low Mach number modeling of type Ia supernovae. I. Hydrodynamics. Astrophys. J., 637 , 922–936.
Bannon, P. R., 1996: On the anelastic approximation for a compressible atmosphere. J. Atmos. Sci., 53 , 3618–3628.
Davies, T., , A. Staniforth, , N. Wood, , and J. Thuburn, 2003: Validity of anelastic and other equation sets as inferred from normal-mode analysis. Quart. J. Roy. Meteor. Soc., 129 , 2761–2775.
Durran, D. R., 1989: Improving the anelastic approximation. J. Atmos. Sci., 46 , 1453–1461.
Dutton, J. A., , and G. H. Fichtl, 1969: Approximate equations of motion for gases and liquids. J. Atmos. Sci., 26 , 241–254.
Held, I. M., 2005: The gap between simulation and understanding in climate modeling. Bull. Amer. Meteor. Soc., 86 , 1609–1614.
Kato, T., 1995: Pertubation Theory for Linear Operators. Springer Verlag, 619 pp.
Klein, R., 2009: Asymptotics, structure, and integration of sound-proof atmospheric flow equations. Theor. Comput. Fluid Dyn., 23 , 161–195.
Kong, Q., , and A. Zettl, 1996: Eigenvalues of regular Sturm–Liouville problems. J. Differ. Equations, 131 , 1–19.
Lipps, F., , and R. Hemler, 1982: A scale analysis of deep moist convection and some related numerical calculations. J. Atmos. Sci., 39 , 2192–2210.
Ogura, Y., , and N. A. Phillips, 1962: Scale analysis of deep and shallow convection in the atmosphere. J. Atmos. Sci., 19 , 173–179.
Prusa, J., , P. Smolarkiewicz, , and A. Wyszogrodzki, 2008: EULAG, a computational model for multiscale flows. Comput. Fluids, 37 , 1193–1207.
Smolarkiewicz, P. K., , and A. Dörnbrack, 2008: Conservative integrals of adiabatic Durran’s equations. Int. J. Numer. Methods Fluids, 56 , 1513–1519.
Zettl, A., 2005: Sturm–Liouville Theory. Mathematical Surveys & Monographs Series, Vol. 121, American Mathematical Society, 328 pp.
APPENDIX
Derivation of the Sturm–Liouville Eq. (23)




















(left) Sample potential temperature distribution, and (right) the resulting dimensionless vertical distributions of
Citation: Journal of the Atmospheric Sciences 67, 10; 10.1175/2010JAS3490.1

Comparison between the soundproof and first-order accurate approximations to the compressible internal wave eigenvalues for the background state from Fig. 1 and horizontal wavenumbers |λ| = (a) 0.5, (b) 2.0, and (c) 8.0. The approximate compressible eigenvalues and eigenmodes are defined through the perturbed Sturm–Liouville problem in (35). The graph shows relative differences of the eigenvalues (Λj0 − Λj1)/Λj0.
Citation: Journal of the Atmospheric Sciences 67, 10; 10.1175/2010JAS3490.1

(left) Vertical velocity structure functions and (right) deviations between the soundproof and compressible modes for the same case as in Fig. 2 and (top) mode number k = 10, horizontal wavenumber |λ| = 0.5 and (bottom) mode number k = 0, horizontal wavenumber |λ| = 8.
Citation: Journal of the Atmospheric Sciences 67, 10; 10.1175/2010JAS3490.1
Characteristic inverse time scales.

Notice that μ − 1 + ν = 2(1 − ν) − 1 + ν = 1 − ν.