1. Introduction
Internal gravity waves (GWs) are prominent dynamic structures that emerge in the atmosphere. If considered as separate physical entities, their life cycle may be divided into three stages. First, their forcing, which is frequently encountered in the troposphere due to processes such as the flow over uneven terrain, convection, and the spontaneous emission by synoptic-scale flow structures. Second, their propagation, which might lead them into upper layers of the atmosphere, where, third, different processes lead to an increased attenuation of GWs until they disappear. This last stage of their life cycle is associated with a drag exerted on the large-scale flow (“large scale” compared to the characteristic wavelength of the GWs), which makes this process relevant to the dynamics of the global circulations in the middle and upper atmosphere (Fritts and Alexander 2003; Kim et al. 2003).
Instabilities are believed to be an important trigger of GW attenuation and there is a vast number of possible mechanisms that are counted among them (Dunkerton 1989; Fritts 1989; Sonmor and Klaassen 1997; Fritts and Alexander 2003; Schlutow 2019). Two of the most thoroughly investigated mechanisms are probably the vertical static instability (VSI; also referred to as convective instability) and the vertical dynamic instability (VDI; e.g., Hodges 1967). The atmospheric state, modified by the presence of GWs, is commonly said to be dynamically or statically unstable, if its Richardson number satisfies Ri < 1/4 or Ri < 0, respectively. A process that almost inevitably leads to one of the aforementioned criteria being satisfied is the vertical propagation of GWs, since their amplitude tends to increase exponentially as the density decreases exponentially. Where the atmospheric state is unstable the GWs are likely to break. In the course of this process, turbulence develops and the larger-scale flow experiences a drag (e.g., Fritts and Alexander 2003).
Many methods to investigate the stability of an atmospheric state may either be classed among the wave-dynamical or the parcel-dynamical stability analyses. In the former class of analyses the atmospheric state, whose stability shall be analyzed, is expanded by wavelike perturbations. Nongrowing (growing) perturbation amplitudes indicate that the state is stable (unstable) with respect to perturbations of the prescribed wave characteristics. The other class of analyses, in contrast, deduces the stability of an atmospheric state at a particular point by investigating the evolution of a displaced air parcel. If it tends to diverge from the point, the state is classed as unstable there (Godson 1950). The VSI, for instance, is commonly identified by parcel-dynamical stability analyses, where the air parcels are displaced in vertical direction (e.g., Holton 2004; Zdunkowski and Bott 2004). As a hand-waving bridging between the two approaches, a displaced parcel could be thought of as delta-function-like perturbation, which in turn may be regarded as composed of an entire spectrum of wave perturbations, all having the same amplitude.
Nonhydrostatic GWs may be associated with significant horizontal gradients, if the vertical gradients are significant, too, since their horizontal and vertical wavelengths are comparable in size (Fritts and Alexander 2003; Kim et al. 2003). Hines (1971, 1988), for instance, investigated, if such horizontal gradients can influence the stability of a GW. He used a variant of the aforementioned parcel-dynamical stability analyses, where the interchange of two parcels, being located relatively close to each other on some specified axis, is considered. The criterion for instability is based on energetic aspects. Static instability is said to be present, if the interchange would result in a reduction of gravitational potential energy, and dynamic instability is made conditional on the reduction of potential energy outweighing the increase of kinetic energy by the interchange of the parcels. The axis on which the two parcels are located initially can have any orientation, so these types of instability are commonly termed slantwise static or dynamic instability (SSI or SDI, respectively; e.g., Marks and Eckermann 1995). Within the framework of this approach, the criteria for VSI and VDI are in a sense special cases of SSI and SDI for a vertically aligned axis. An important result of the SSI of GWs is that for certain regions in a wave train there is always a set of axes for which the criterion for instability is satisfied. So, in contrast to the predictions of VSI, GWs are always unstable, regardless of their amplitudes. If a GW passes through a point, the atmospheric state there changes periodically between being stable and unstable. Therefore, the mere magnitude of the instability is probably not enough to infer the strength of the wave breaking. If the wave period is short compared to the characteristic time scale associated with the instability, the formation of the wave breaking is periodically interrupted at the considered point (Hines 1988; Marks and Eckermann 1995).
The conditions for SSI and SDI depend on the direction of the axis on which the two parcels are located initially. This property might be unfavorable for applications, which only depend on the synopsis if the atmospheric state is unstable at some location or not. To eliminate the dependence on the direction, one can integrate over interchanges along all possible directions. This, however, introduces a number of additional assumptions that might appear more or less arbitrary. Hines (1971), for instance, considered a spherical volume, in which each parcel is interchanged with the parcel diametrically opposite. To avoid such complications, we present a possible alternative to the parcel interchange method, which is based on an extension of the classical vertical-displacement method. To analyze the atmospheric state for instability at some location, one considers the tendency of motion of the parcel there after it being displaced in an arbitrary direction. This results in an equation for the oscillation frequency of the parcel, which depends on the characteristics of the atmospheric state at the location, but does not depend on the displacement. Then imaginary oscillation frequencies (of a certain sign) are associated with an unstable state (Godson 1950; Shutts and Cullen 1987). We call the instability three-dimensional static instability (3DSI), since it can be regarded as an extension of VSI that takes into account the horizontal variation of the atmospheric state. A significant portion of GWs that are stable with respect to VSI, are unstable with respect to 3DSI. In that regard, 3DSI agrees with SSI.
It is common practice to parameterize the unresolved (or subgrid-scale) GWs in atmospheric models, whose characteristic spatial and temporal resolution scales exclude a significant part of the GW spectrum from being resolved. That is, those unresolved processes, which have a significant effect on the resolved flow, are modeled by some effective formulation that depends on properties of the resolved flow (e.g., Gardner 1996; McLandress 1998; Fritts and Alexander 2003; Kim et al. 2003; Medvedev and Yiğit 2019). Ideally, the integration of a GW parameterization into an atmospheric model would bring simulation results closer to the results from a model that resolves most of the GW spectrum or to observations (with respect to some specified quality measures). Here, we mainly focus on the parameterization of the process of GW breaking and the associated effects on the resolved atmospheric state. A common approach is the one introduced by Lindzen (1981), where the GW breaking is made conditional on a static instability. As a consequence of the breaking, the GWs are assumed to saturate to a state of marginal stability (e.g., Lindzen 1981; Warner and McIntyre 2001; Scinocca 2003) or to be obliterated completely (e.g., Alexander and Dunkerton 1999). This in turn leads to a convergence of momentum flux, which is the GW drag on the resolved flow. There are other approaches such as the Doppler-spread parameterization of Hines (1997). Which of the many parameterization approaches performs best with respect to certain quality measures, is still a matter of debate and active research (e.g., Gardner 1996; McLandress 1998; Fritts and Alexander 2003; McLandress and Scinocca 2005; Majdzadeh and Klaassen 2019, hereafter MK19; Plougonven et al. 2020).
The nonlinear process of GW breaking is commonly regarded as an irreversible process (e.g., Medvedev and Klaassen 2003; Shaw and Shepherd 2009; Gassmann and Herzog 2015; Gassmann 2018); therefore, the methods of irreversible thermodynamics (e.g., Herbert 1978; Jou et al. 2001; Zdunkowski and Bott 2003, 2004; Papenfuß 2020) may be a possible tool for the development of a parameterization, and we will use it in this work.
The paper is organized as follows. In section 2 we describe the stability analysis that underlies 3DSI, and apply it to a monochromatic GW in section 3. The ansatz for a parameterization of the effects that are associated with the 3DSI-induced breaking of subgrid-scale GWs can be found in section 4. For two vertical profiles of zonal wind and temperature, we show results for GW drag and dissipative heating rates in section 5, and conclude with a summary in section 6.
2. Parcel-dynamical stability analysis
3. Application to a GW
A nonvanishing coefficient c requires the Exner pressure π to vary in all three directions of an arbitrary normal basis [see Eq. (A8c) in appendix A]. In the case of a single monochromatic plane GW, however, there is variation in ez and k directions only. For this reason, c vanishes in Eq. (21c). If two or more GWs interfere, with wave vectors pointing in different directions, we find c ≠ 0, in general. This, in turn, affects all roots, where
(left),(center) Normalized second and third roots,
Citation: Journal of the Atmospheric Sciences 79, 12; 10.1175/JAS-D-21-0287.1
Ratio of the instability times scale to the GW time scale,
Citation: Journal of the Atmospheric Sciences 79, 12; 10.1175/JAS-D-21-0287.1
In summary, the three-dimensional stability analysis predicts instability for a significantly larger area in k space than does the vertical stability analysis. There is no significant horizontal stratification in the atmosphere, which could counteract destabilizing gradients associated with GWs (e.g., Sonmor and Klaassen 1997). The differences between the predictions of the two stability analyses may be particularly relevant to nonhydrostatic GWs, whose horizontal and vertical wavelengths are of comparable scale (Kim et al. 2003). We emphasize that the above results depend on the assumptions on which the parcel-dynamical stability analysis is based. Another stability model, which is founded on different assumptions, may come to more or less different conclusions. Nevertheless, the above findings are generally in line with the results of the slantwise static stability analysis of Hines (1971, 1988). The author also found that the condition for vertical static instability [
4. Parameterization of GW breaking due to 3DSI
Approaches to the parameterization problem may be based on decomposing the atmospheric state into a resolved (or gridscale) part and unresolved (or subgrid-scale) deviations. This may be written
Note that 〈E〉 on the lhs of Eq. (26), is not solely the energy of the subsystem resolved atmosphere E0, but the energy of the overall system. By writing the fundamental relation in the form
In case of the GW model, which we use in this work, the fundamental relation (26) is separable insofar as
In a next step, the differential of Eq. (26) will be combined with the budget equations of
a. Outline of the approach
GW breaking is commonly considered an irreversible process (e.g., Medvedev and Klaassen 2003; Shaw and Shepherd 2009; Gassmann and Herzog 2015). Therefore, the methods of irreversible thermodynamics are one possible tool for its parameterization. This approach may be divided into the following tasks (cf. Herbert 1978; Peixoto et al. 1991; Jou et al. 2001; Zdunkowski and Bott 2004; Liu et al. 2011; Gassmann and Herzog 2015; Papenfuß 2020):4
-
Make oneself familiar with the (thermo)dynamic properties of the system at hand and express them in terms of the Gibbs equation.
-
Set up budget equations for the extensive quantities contained by the system, and separate irreversible from reversible processes.
-
The extensive quantity entropy is not conserved, but can be produced by irreversible processes. Combine all budget equations with the Gibbs equation to derive an expression for the source of entropy.
-
From this expression, correlating thermodynamic forces and fluxes can be defined. The correlation is expressed in terms of phenomenological equations.
-
If the irreversible processes are of moderate magnitude, the phenomenological equations may be linearized. The coefficients of the linear relations are the transport coefficients.
-
Determine closures for the transport coefficients.
These tasks will serve as a guideline in the following.
b. Gibbs equation of the resolved atmosphere
c. Gibbs equation of the unresolved GWs
d. Combined Gibbs equation, budget equations, and the source of entropy
Turbulence is probably an important player in the context of GW breaking, but its integration into the framework of irreversible thermodynamics is a complex problem (e.g., Blackadar 1955; Zdunkowski and Bott 2003). We will omit an explicit contribution of turbulence to the Gibbs equation, Eq. (42), since that goes beyond the scope of this work.5
e. Phenomenological equations
f. Closures for the transport coefficients
According to Eq. (51b), the second law requires the transport coefficients Lϵ,j and
5. Proof of concept
To get an idea of how the parameterization of GW breaking due to 3DSI acts, we will consider a relatively simple setup and apply a number of simplifications to the governing equations.
Our testing consists in computing offline drag and heating rates for fixed atmospheric state profiles. Online tests with a general circulation model are not part of this work. We use the same profiles as MK19. These are representative zonal averages of zonal wind and temperature at 50° south for summer (January) and winter (June) from CIRA data (Fleming et al. 1990); see Fig. 3 (and Fig. 2 of MK19). For further parameters of the setup, we also closely follow MK19. We use nϕ = 2 azimuths (i.e., east and west directions), a launch level of zl = 17 km and a launch flux density per azimuth of Fl = 7.2 × 10−4 Pa. Apart from that, we set nh = nz = 100, λh,max = 50 km, λz,min = 100 m, λz,max = 20 km, and Kϵ = Kζ = 1, m = 5 for the parameters of Eqs. (55). The layer thickness of the vertical grid is 1 km.
Profiles of zonal averages of (left) zonal wind and (right) temperature at 50° south for summer (January, solid lines) and winter (June, dashed lines) from CIRA data.
Citation: Journal of the Atmospheric Sciences 79, 12; 10.1175/JAS-D-21-0287.1
The wave action density at launch level
For purposes of comparison, we also compute the GW drag for the case that the GW breaking is assumed to be triggered by VSI instead of 3DSI. To achieve this, we set each component of the deviation stability tensor
Before testing the full CIRA profiles, we apply the scheme to the summer temperature profile plus zero zonal wind (u0 = 0 m s−1). East- and westward GW drag profiles for this case are shown in Fig. 4 (left panel, blue solid lines). They cancel each other exactly, due to the azimuthal symmetry of the wind profile and the launched GW spectrum. The total GW heating rate profiles according to Eqs. (58) and (59) are shown in Fig. 5. The dissipative heating, (58), is accumulative and does not vanish. The alternative frictional heating, (59), is zero in this (admittedly highly idealized) case. The drag profiles can be compared to Fig. 5 of MK19, where the authors show GW drag profiles for the Hines (1997) scheme, on the one hand, and for the Warner and McIntyre (2001) and Scinocca (2003) scheme (WMS, hereafter), on the other hand. In contrast to the present scheme, which is based on a discrete spectrum of GW packets, the Hines and WMS schemes are based on a continuous GW spectrum. Apart from that, the latter two schemes differ in many respects from each other (see the aforementioned literature for more details). The drag profiles in Fig. 4 (left panel) are more similar to the WMS than to the Hines drag profiles, both in shape and in magnitude. The Hines profiles exhibit an oscillatory structure and are about 10 times larger in magnitude.
(left) Vertical GW drag profiles for an atmosphere at rest, and the CIRA zonal wind and temperature profiles, shown in Fig. 3, for (center) summer and (right) winter. In the left panel eastward (positive) and westward (negative) GW drag profiles are plotted separately. The center and right panels show the total GW drag (east- plus westward drag). Results for the three-dimensional static instability (3DSI) and vertical static instability (VSI) are shown in blue and red, respectively. Solid lines represent the case Kϵ = Kζ = 1, dotted lines represent the case Kϵ = 1, Kζ = 0. Drag magnitudes below 30 km are comparatively small and therefore not plotted. The red solid line in the middle panel peaks at 25 m s−1 day−1, but we cut at 12 m s−1 day−1 to show the other lines more clearly.
Citation: Journal of the Atmospheric Sciences 79, 12; 10.1175/JAS-D-21-0287.1
Vertical profiles of the total GW heating rates corresponding to the GW drag profiles in Fig. 4. Solid lines are heating rates according to Eq. (58), dashed lines according to Eq. (59).
Citation: Journal of the Atmospheric Sciences 79, 12; 10.1175/JAS-D-21-0287.1
Vertical profiles of the total (east- plus westward) GW drag for the full CIRA profiles are shown in Fig. 4 (middle and right panels, blue solid lines). They can be compared to Fig. 6 of MK19, where corresponding profiles for the Hines and the WMS schemes are shown. Both schemes yield predominantly westward drag in the winter case and predominantly eastward drag in the summer case. The results shown in Fig. 4 are in line with this characteristic. In terms of the shape of the drag profile, the WMS scheme produces a relatively smooth profile, with drag spread over a broad altitude range. In contrast, the Hines scheme produces a relatively localized forcing of a rather oscillatory shape (see MK19 for a detailed analysis). The drag profiles in Fig. 4 lie in between. The shape is significantly less smooth than in case of the WMS scheme, but not as erratic as in case of the Hines scheme. The strength of the drag in both cases is comparable to what the WMS scheme produces. The drag of the Hines scheme is about two orders of magnitude stronger than the WMS drag. A number of discrepancies between the Hines and WMS schemes, on the one hand, and the present scheme, on the other hand, might have two main reasons. First, a continuous GW spectrum forms the (theoretical) basis of their schemes, whereas we use a discrete spectrum. Second, the present scheme describes the GW dissipation by explicit sinks in the GW budget equations, (56), whereas the other two schemes use an “instantaneous” adjustment to a saturation or reference spectrum, in one form or another. Future investigations have to reveal the specific reasons.
The drag based on VSI is of comparable magnitude as the 3DSI drag, in particular in the atmosphere-at-rest case. This is mainly due to the azimuthal symmetry of the discrete spectrum. For even nϕ, the off-diagonal components of
Heating rates according to Eqs. (58) and (59) are shown in Fig. 5 (middle and right panels), but only for 3DSI. The dissipative heating (58) and the frictional heating (59) are of comparable magnitude, at least at altitudes above about 70 km. Whereas the dissipative heating is exclusively positive, there are altitude ranges, where the frictional heating would cool the atmosphere. If the frictional heating would be used as a model for dissipative heating, this would be an undesirable behavior. Here, however, we can only speak for the present scheme. If the drag is in opposite direction to the wind, the frictional heating is positive, and this seems to be the case for a broad altitude range for both the Hines and the WMS schemes (cf. Figs. 2 and 6 of MK19). The magnitude of the heating rate profiles for the summer and winter cases is comparable to heating rates presented by Medvedev and Klaassen (2003) and Becker (2004).
The aim of this proof of concept was to give an idea of the characteristics of the present scheme. In summary, it can produce drag profiles that are similar to the profiles of the WMS scheme, both in shape and in magnitude. Using 3DSI instead of VSI can make a difference for the drag. In general, the 3DSI profiles are smoother than the VSI profiles. Future studies, in which the present scheme is implemented in a general circulation model, have to reveal, if the present scheme can compete with established schemes.
In terms of computational costs, a parameterization scheme based on a discrete spectrum of GWs that attempts to be in line with the principles of irreversible thermodynamics, can likely not compete with established schemes such as the Hines and WMS schemes. The WMS scheme, for instance, is based on the continuous spectrum, (60), and through an analytical integration over the
Finally, there is another aspect of GW parameterizations pointed out, for instance, by Amemiya and Sato (2016) and MK19. Although shape and strength of the GW drag may be quite different for two GW schemes under the same atmospheric conditions, the response of a general circulation model to them may be relatively similar. The resolved model atmosphere appears to show the tendency of compensating different forcings. So two different schemes may result in similar wind statistics. This adjustment effect concerns at least momentum tendencies, which can cancel each other. Dissipative heating, in contrast, is accumulative. What this means in terms of the response behavior of a general circulation model is probably less clear than in case of the GW drag.
6. Summary
Vertical static instability (VSI) is a common indicator for the onset of gravity wave (GW) breaking in the atmosphere, according to which instability sets in, if the vertical gradient of potential temperature of the atmospheric state modulated by the presence of a GW becomes negative. Horizontal variations, whose magnitude can be significant for nonhydrostatic GWs, are ignored in this analysis. Hines (1971, 1988) proposed an alternative approach that accounts for horizontal variations and he found that this analysis predicts an unconditional instability, if GWs are present. This instability is called slantwise static instability (SSI). Some aspects of the analysis method of Hines make, however, the use of its results in certain areas of application difficult, for instance, in a parameterization of the GW breaking that could be triggered by the instability. In this work, we show that the method of Godson (1950) and Shutts and Cullen (1987) may be an alternative. It is related to the method of Hines, and similar to SSI, it predicts a significantly larger range of instability than VSI. We call this instability 3DSI to distinguish it from SSI. In terms of their mathematical expressions, it appears more straightforward to incorporate 3DSI into the development of a parameterization than SSI. We use this fact to make a suggestion for such a parameterization.
To meet the irreversible character of the process of GW breaking, we apply the methods of irreversible thermodynamics, which are embedded in the Gibbs formalism of dynamics. In this way, the parameterization does not only satisfy the second law of thermodynamics, but it can also be made consistent with the conservation of energy and further (non-)conservation principles, if required. In addition to an expression for the GW drag, we obtain an expression for the heating rate associated with the GW breaking, since the production of entropy is at the heart of the approach.
We develop the parameterization for a discrete spectrum of GWs. Following general practice, we apply the steady-state and columnar approximations. Offline computations of GW drag and heating rates are performed for two vertical zonal wind and temperature profiles from CIRA data. One profile is representative for summer, the other for winter conditions. The same profiles where use by MK19 to analyze the GW drag from the Hines (1997) scheme, on the one hand, and the WMS scheme, on the other hand. So our results can be compared with their findings. The GW drag produced by the present scheme is less smooth than the drag produced by the WMS scheme, but still less oscillatory than the drag from the Hines scheme. For both the summer and winter cases the strength of the drag is comparable to the result of the WMS scheme, but significantly less than the result of the Hines scheme. In addition, we show that replacing 3DSI by VSI in the present scheme can change the drag significantly. Although the strength of the drag is of the same magnitude in both cases, the 3DSI drag profiles are significantly smoother than the VSI drag profiles. The dissipative heating rate produced by the present scheme is always positive and in line with the second law of thermodynamics. It is shown that the frictional heating, which follows directly from the GW drag, is not necessarily positive and should be used with care in the case that it is employed for the parameterization of dissipative heating. Future investigations with the present scheme implemented in a general circulation model have to show whether it can compete with established schemes.
In fact, we closely follow the work of Ertel et al. (1941) in a large part of the parcel stability analysis, but no English translation is available, as far as we know.
Expressed in terms of pressure p and density ρ, the stability tensor reads
As an alternative to this approximation, the factor
In fact, we took the task list from Meixner (1960), but no English translation is available, as far as we know.
We point, however, to the inclusive approach to turbulence modeling, where turbulent dynamics are integrated as part of the thermodynamics (e.g., Sievers 1984). If one would adopt this approach, the terms ρ0(T0ds0/dt − p0dυ0/dt) on the rhs of Eq. (42) would cover both thermodynamics and the dynamics of turbulence. The ideal gas law (29c) and the caloric equation of state (29b) provide, however, only first-order (or lower) approximations with respect to turbulence (e.g., Blackadar 1955; Herbert and Kucharski 1998).
The momentum flux tensor is not symmetric, since
A simple example may illustrate this more clearly. Consider a system with two irreversible scalar flux densities Fα and Fβ, and the corresponding thermodynamic forces Xα and Xβ, so that Tσs = −FαXα − FβXβ. Linearizing the phenomenological equations Fα/β = Fα/β(Xα,Xβ) about thermodynamic equilibrium Xα/β = 0 yields Fα ≈ (∂Fα/∂Xα)Xα + (∂Fα/∂Xβ)Xβ = −LααXα − LαβXβ and Fβ ≈ (∂Fβ/∂Xα)Xα + (∂Fβ/∂Xβ)Xβ = −LβαXα − LββXβ. The L coefficients are called transport coefficients. In addition, (−LααXα) and (−LββXβ) are called the direct effects (for their coefficients, we would use the abbreviations Lαα → Lα and Lββ → Lβ), and (−LαβXβ) and (−LβαXα) are the cross effects (see, e.g., Onsager 1931).
Consider a coarsely resolved general circulation model, used for climate projections, for instance, where
However, this also depends on the quality criterion defined to measure the “adequacy” of the ansatz.
We could arrive at this expression, if we neglect an explicit contribution of the GWs to the total energy budget (42), but retain the pseudomomentum flux in Eq. (45), this time, however, not as part of the reversible, but rather as part of the irreversible momentum flux. Then, the dissipation, (47), reads
Acknowledgments.
The authors thank the German Research Foundation (DFG) for partial support through the research unit Multiscale Dynamics of Gravity Waves (MS-GWaves) and through Grant ZA 268/10-2. We greatly appreciate many helpful discussions with Dr. Young-Ha Kim and Dr. Gergely Bölöni. Their valuable suggestions led to significant improvements of the manuscript. We are indebted to the anonymous reviewers for very helpful comments.
Data availability statement.
The scripts and the source code that were used to produce the results, which are shown in the figures of this work, can be made available upon request. Please contact the corresponding author.
APPENDIX A
The Determinant of a Tensor
Here, we state the determinant of a second-order tensor of the form
APPENDIX B
Divergence of the Product of a Tensor and a Vector
APPENDIX C
Expansion of a Second-Order Tensor in Deviatoric, Spherical, and Skew-Symmetric Parts
APPENDIX D
List of Symbols
t, x |
Time and position vector |
x, y, z |
Zonal (west–east), meridional (south–north), and vertical (bottom-up) components of position |
ex, ey, ez |
Unit vectors in zonal, meridional, and vertical directions |
d/dt, ∂/∂t |
Lagrangian (or material) and Eulerian time derivatives |
∇ = ∂/∂x |
Nabla operator |
∇∇ = ∂2/∂x∂x |
Second-order Hessian tensor |
Second-order identity tensor | |
Potential of apparent gravity (true gravity plus centrifugal acceleration) | |
g = gez |
Mean gravitational acceleration at sea level |
Ω, |
Vector and second-order tensor of Earth’s angular velocity |
cp, cυ, R |
Specific heat capacities at constant pressure and at constant volume, gas constant of dry air |
T, p, ρ, θ, π, v |
Temperature, pressure, density, potential temperature, Exner pressure, and velocity (or wind) vector |
u, u, υ, w |
Horizontal part of the velocity vector v, and zonal, meridional, and vertical components of v |
δ(·) |
Difference of a quantity between perturbed and unperturbed air parcel |
Second-order stability tensor | |
Squared buoyancy and Brunt–Väisälä frequencies | |
f = 2Ωsin(φ) |
Coriolis parameter at latitude φ |
(·)00 |
Constant reference value of a quantity |
(·)0 |
Hydrostatic, atmospheric background state or resolved (gridscale) quantities, respectively |
(·)′ |
Perturbation of a quantity associated with a GW, or unresolved (subgrid-scale) quantity, respectively |
i, Re{·}, Im{·} |
Imaginary unit, real, and complex parts of a complex quantity |
Complex oscillation or wave amplitude quantity | |
Oscillation frequency of a perturbed (displaced) parcel | |
a, b, c |
Coefficients of the characteristic equation for |
k, ∇k = ∂/∂x |
Wave vector and derivative with respect to the same |
kx, ky, kz |
Zonal, meridional, and vertical components of wave vector |
λh, λz |
Horizontal and vertical wavelength |
ω, |
Extrinsic and intrinsic GW frequencies |
Second-order GW dispersion relation tensor | |
Extrinsic and intrinsic GW group velocities | |
Intrinsic GW period, time scale of instability, and | |
E, m, P, V, S |
Energy, mass, momentum, volume, and entropy |
e, p, υ, s |
Mass-specific energy, momentum, volume, and entropy |
Energy, wave action, and pseudomomentum densities of a (unresolved) GW packet | |
Wave action phase space density | |
xg, dg/dt |
Position of a GW packet and GW packet–fixed time derivative |
Vectorial and tensorial densities of reversible fluxes | |
Vectorial and tensorial densities of irreversible fluxes | |
σ, σ |
Scalar and vectorial densities of irreversible sources |
Scalar, vectorial, second-, third-, and fourth-order tensorial transport coefficients of linearized phenomenological equations | |
(·)T, adj(·), det(·), tr(·) |
Transpose, adjugate, determinant, and trace of a second-order tensor |
sym(·), skw(·), dev(·), sph(·) |
Symmetric, skew-symmetric, deviatoric, and spherical (or isotropic) parts of a second-order tensor |
(·) · · (·) |
Double scalar product of two tensors |
i, j, k, l, m |
Indices |
REFERENCES
Achatz, U., B. Ribstein, F. Senf, and R. Klein, 2017: The interaction between synoptic-scale balanced flow and a finite-amplitude mesoscale wave field throughout all atmospheric layers: Weak and moderately strong stratification. Quart. J. Roy. Meteor. Soc., 143, 342–361, https://doi.org/10.1002/qj.2926.
Akmaev, R. A., 2007: On the energetics of mean-flow interactions with thermally dissipating gravity waves. J. Geophys. Res., 112, D11125, https://doi.org/10.1029/2006JD007908.
Alexander, M. J., and T. J. Dunkerton, 1999: A spectral parameterization of mean-flow forcing due to breaking gravity waves. J. Atmos. Sci., 56, 4167–4182, https://doi.org/10.1175/1520-0469(1999)056<4167:ASPOMF>2.0.CO;2.
Amemiya, A., and K. Sato, 2016: A new gravity wave parameterization including three-dimensional propagation. J. Meteor. Soc. Japan, 94, 237–256, https://doi.org/10.2151/jmsj.2016-013.
Amiramjadi, M., A. R. Mohebalhojeh, M. Mirzaei, C. Zülicke, and R. Plougonven, 2020: The spatiotemporal variability of nonorographic gravity wave energy and relation to its source functions. Mon. Wea. Rev., 148, 4837–4857, https://doi.org/10.1175/MWR-D-20-0195.1.
Becker, E., 2004: Direct heating rates associated with gravity wave saturation. J. Atmos. Sol.-Terr. Phys., 66, 683–696, https://doi.org/10.1016/j.jastp.2004.01.019.
Becker, E., and G. Schmitz, 2002: Energy deposition and turbulent dissipation owing to gravity waves in the mesosphere. J. Atmos. Sci., 59, 54–68, https://doi.org/10.1175/1520-0469(2002)059<0054:EDATDO>2.0.CO;2.
Blackadar, A. K., 1955: Extension of the laws of thermodynamics to turbulent systems. J. Atmos. Sci., 12, 165–175, https://doi.org/10.1175/1520-0469(1955)012<0165:EOTLOT>2.0.CO;2.
Bölöni, G., B. Ribstein, J. Muraschko, C. Sgoff, J. Wei, and U. Achatz, 2016: The interaction between atmospheric gravity waves and large-scale flows: An efficient description beyond the nonacceleration paradigm. J. Atmos. Sci., 73, 4833–4852, https://doi.org/10.1175/JAS-D-16-0069.1.
Bölöni, G., Y.-H. Kim, S. Borchert, and U. Achatz, 2021: Toward transient subgrid-scale gravity wave representation in atmospheric models. Part I: Propagation model including nondissipative direct wave–mean-flow interactions. J. Atmos. Sci., 78, 1317–1338, https://doi.org/10.1175/JAS-D-20-0065.1.
Bretherton, F. P., 1966: The propagation of groups of internal gravity waves in a shear flow. Quart. J. Roy. Meteor. Soc., 92, 466–480, https://doi.org/10.1002/qj.49709239403.
Bronshtein, I. N., K. A. Semendyayev, G. Musiol, and H. Mühlig, 1990: Handbook of Mathematics. 5th ed. Springer-Verlag, 1164 pp.
Dunkerton, T. J., 1989: Theory of internal gravity wave saturation. Pure Appl. Geophys., 130, 373–397, https://doi.org/10.1007/BF00874465.
Eckermann, S. D., 2011: Explicitly stochastic parameterization of nonorographic gravity wave drag. J. Atmos. Sci., 68, 1749–1765, https://doi.org/10.1175/2011JAS3684.1.
Ertel, H., J.-J. Jaw, and S.-Z. Li, 1941: Tensorielle Theorie der Stabilität. Meteor. Z., 58, 389–392.
Falk, G., F. Herrmann, and G. B. Schmid, 1983: Energy forms or energy carriers? Amer. J. Phys., 51, 1074–1077, https://doi.org/10.1119/1.13340.
Fleming, E. L., S. Chandra, J. J. Barnett, and M. Corney, 1990: Zonal mean temperature, pressure, zonal wind and geopotential height as functions of latitude. Adv. Space Res., 10, 11–59, https://doi.org/10.1016/0273-1177(90)90386-E.
Fritts, D. C., 1989: A review of gravity wave saturation processes, effects, and variability in the middle atmosphere. Pure Appl. Geophys., 130, 343–371, https://doi.org/10.1007/BF00874464.
Fritts, D. C., and M. J. Alexander, 2003: Gravity wave dynamics and effects in the middle atmosphere. Rev. Geophys., 41, 1003, https://doi.org/10.1029/2001RG000106.
Gardner, C. S., 1996: Testing theories of atmospheric gravity wave saturation and dissipation. J. Atmos. Terr. Phys., 58, 1575–1589, https://doi.org/10.1016/0021-9169(96)00027-X.
Gassmann, A., 2018: Entropy production due to subgrid-scale thermal fluxes with application to breaking gravity waves. Quart. J. Roy. Meteor. Soc., 144, 499–510, https://doi.org/10.1002/qj.3221.
Gassmann, A., and H.-J. Herzog, 2015: How is local material entropy production represented in a numerical model? Quart. J. Roy. Meteor. Soc., 141, 854–869, https://doi.org/10.1002/qj.2404.
Godson, W. L., 1950: Generalized criteria for dynamic instability. J. Atmos. Sci., 7, 268–278, https://doi.org/10.1175/1520-0469(1950)007<0268:GCFDI>2.0.CO;2.
Hasha, A., O. Bühler, and J. Scinocca, 2008: Gravity wave refraction by three-dimensional varying winds and the global transport of angular momentum. J. Atmos. Sci., 65, 2892–2906, https://doi.org/10.1175/2007JAS2561.1.
Hauf, T., and H. Höller, 1987: Entropy and potential temperature. J. Atmos. Sci., 44, 2887–2901, https://doi.org/10.1175/1520-0469(1987)044<2887:EAPT>2.0.CO;2.
Henyey, F. S., 1986: Strange semiclassical phenomena for the equation ∇ 2 ∂ t 2 ϕ + a ( ∂ x 2 + ∂ y 2 ) ϕ + b ∂ ϕ 2 = 0, describing waves in stratified fluids. The Physics of Phase Space: Nonlinear Dynamics and Chaos, Geometric Quantization, and Wigner Function, Y. S. Kim and W. W. Zachary, Eds., Vol. 278, Springer, 322–327, https://doi.org/10.1007/3-540-17894-5_374.
Henyey, F. S., and N. Pomphrey, 1983: Eikonal description of internal wave interactions: A non-diffusive picture of “induced diffusion.” Dyn. Atmos. Oceans, 7, 189–219, https://doi.org/10.1016/0377-0265(83)90005-2.
Herbert, F., 1978: The tensor structure of constitutive equations for linear atmospheric heat and momentum exchange with axisymmetric coefficients. Deutscher Wetterdienst Rep. 145, 42 pp.
Herbert, F., and F. Kucharski, 1998: An internal energy theorem for the atmosphere and its association with turbulent (potential) temperature variances. Geophys. Res. Lett., 25, 1185–1188, https://doi.org/10.1029/98GL00914.
Hertzog, A., C. Souprayen, and A. Hauchecorne, 2002: Eikonal simulations for the formation and the maintenance of atmospheric gravity wave spectra. J. Geophys. Res., 107, 4145, https://doi.org/10.1029/2001JD000815.
Hines, C. O., 1971: Generalizations of the Richardson criterion for the onset of atmospheric turbulence. Quart. J. Roy. Meteor. Soc., 97, 429–439, https://doi.org/10.1002/qj.49709741405.
Hines, C. O., 1988: Generation of turbulence by atmospheric gravity waves. J. Atmos. Sci., 45, 1269–1278, https://doi.org/10.1175/1520-0469(1988)045<1269:GOTBAG>2.0.CO;2.
Hines, C. O., 1997: Doppler-spread parameterization of gravity-waves momentum deposition in the middle atmosphere. Part 1: Basic formulation. J. Atmos. Sol.-Terr. Phys., 59, 371–386, https://doi.org/10.1016/S1364-6826(96)00079-X.
Hodges, R. R., Jr., 1967: Generation of turbulence in the upper atmosphere by internal gravity waves. J. Geophys. Res., 72, 3455–3458, https://doi.org/10.1029/JZ072i013p03455.
Holton, J. R., 2004: An Introduction to Dynamic Meteorology. 4th ed. Elsevier, 553 pp.
Job, G., and F. Herrmann, 2006: Chemical potential—A quantity in search of recognition. Eur. J. Phys., 27, 353–371, https://doi.org/10.1088/0143-0807/27/2/018.
Jones, W. L., 1969: Ray tracing for internal gravity waves. J. Geophys. Res., 74, 2028–2033, https://doi.org/10.1029/JB074i008p02028.
Jones, W. L., 1971: Energy-momentum tensor for linearized waves in material media. Rev. Geophys., 9, 917–952, https://doi.org/10.1029/RG009i004p00917.
Jou, D., J. Casas-Vázquez, and G. Lebon, 2001: Extended Irreversible Thermodynamics. 3rd ed. Springer-Verlag, 463 pp.
Kim, Y.-H., G. Bölöni, S. Borchert, H.-Y. Chun, and U. Achatz, 2021: Toward transient subgrid-scale gravity wave representation in atmospheric models. Part II: Wave intermittency simulated with convective sources. J. Atmos. Sci., 78, 1339–1357, https://doi.org/10.1175/JAS-D-20-0066.1.
Kim, Y.-J., S. D. Eckermann, and H.-Y. Chun, 2003: An overview of the past, present and future of gravity-wave drag parametrization for numerical climate and weather prediction models. Atmos.–Ocean, 41, 65–98, https://doi.org/10.3137/ao.410105.
Lebedev, L. P., M. J. Cloud, and V. A. Eremeyev, 2010: Tensor Analysis with Applications in Mechanics. 2nd ed. World Scientific Publishing, 380 pp.
Lindzen, R. S., 1981: Turbulence and stress owing to gravity wave and tidal breakdown. J. Geophys. Res., 86, 9707–9714, https://doi.org/10.1029/JC086iC10p09707.
Liu, Y., C. Liu, and D. Wang, 2011: Understanding atmospheric behaviour in terms of entropy: A review of applications of the second law of thermodynamics to meteorology. Entropy, 13, 211–240, https://doi.org/10.3390/e13010211.
Majdzadeh, M., and G. P. Klaassen, 2019: An analysis of the Hines and Warner–McIntyre–Scinocca non-orographic gravity wave drag parametrizations. Quart. J. Roy. Meteor. Soc., 145, 2308–2334, https://doi.org/10.1002/qj.3559.
Marks, C. J., and S. D. Eckermann, 1995: A three-dimensional nonhydrostatic ray-tracing model for gravity waves: Formulation and preliminary results for the middle atmosphere. J. Atmos. Sci., 52, 1959–1984, https://doi.org/10.1175/1520-0469(1995)052<1959:ATDNRT>2.0.CO;2.
McLandress, C., 1998: On the importance of gravity waves in the middle atmosphere and their parameterization in general circulation models. J. Atmos. Sol.-Terr. Phys., 60, 1357–1383, https://doi.org/10.1016/S1364-6826(98)00061-3.
McLandress, C., and J. F. Scinocca, 2005: The GCM response to current parameterizations of nonorographic gravity wave drag. J. Atmos. Sci., 62, 2394–2413, https://doi.org/10.1175/JAS3483.1.
Medvedev, A. S., and G. P. Klaassen, 2003: Thermal effects of saturating gravity waves in the atmosphere. J. Geophys. Res., 108, 4040, https://doi.org/10.1029/2002JD002504.
Medvedev, A. S., and E. Yiğit, 2019: Gravity waves in planetary atmospheres: Their effects and parameterization in global circulation models. Atmosphere, 10, 531, https://doi.org/10.3390/atmos10090531.
Meixner, J., 1960: Die Thermodynamik irreversibler Prozesse. Phys. Blätter, 16, 506–511, https://doi.org/10.1002/phbl.19600161003.
Muraschko, J., M. D. Fruman, U. Achatz, S. Hickel, and Y. Toledo, 2015: On the application of the Wentzel–Kramer–Brillouin theory for the simulation of the weakly nonlinear dynamics of gravity waves. Quart. J. Roy. Meteor. Soc., 141, 676–697, https://doi.org/10.1002/qj.2381.
Olbers, D., and C. Eden, 2013: A global model for the diapycnal diffusivity induced by internal gravity waves. J. Phys. Oceanogr., 43, 1759–1779, https://doi.org/10.1175/JPO-D-12-0207.1.
Onsager, L., 1931: Reciprocal relations in irreversible processes. I. Phys. Rev., 37, 405–426, https://doi.org/10.1103/PhysRev.37.405.
Papenfuß, C., 2020: Continuum Thermodynamics and Constitutive Theory. 1st ed. Springer, 238 pp.
Peixoto, J. P., A. H. Oort, M. D. Almeida, and A. Tomé, 1991: Entropy budget of the atmosphere. J. Geophys. Res., 96, 10 981–10 988, https://doi.org/10.1029/91JD00721.
Plougonven, R., A. de la Cámara, A. Hertzog, and F. Lott, 2020: How does knowledge of atmospheric gravity waves guide their parameterizations? Quart. J. Roy. Meteor. Soc., 146, 1529–1543, https://doi.org/10.1002/qj.3732.
Schlutow, M., 2019: Modulational stability of nonlinear saturated gravity waves. J. Atmos. Sci., 76, 3327–3336, https://doi.org/10.1175/JAS-D-19-0065.1.
Scinocca, J. F., 2003: An accurate spectral nonorographic gravity wave drag parameterization for general circulation models. J. Atmos. Sci., 60, 667–682, https://doi.org/10.1175/1520-0469(2003)060<0667:AASNGW>2.0.CO;2.
Shaw, T. A., and T. G. Shepherd, 2009: A theoretical framework for energy and momentum consistency in subgrid-scale parameterization for climate models. J. Atmos. Sci., 66, 3095–3114, https://doi.org/10.1175/2009JAS3051.1.
Shutts, G. J., and M. J. P. Cullen, 1987: Parcel stability and its relation to semigeostrophic theory. J. Atmos. Sci., 44, 1318–1330, https://doi.org/10.1175/1520-0469(1987)044<1318:PSAIRT>2.0.CO;2.
Sievers, U., 1984: The turbulent atmosphere and the inclusive system of model equations. Beitr. Phys. Atmos., 57, 324–345.
Sihvola, A., 1999: Electromagnetic Mixing Formulas and Applications. Institution of Engineering and Technology, 296 pp.
Snively, J. B., and V. P. Pasko, 2008: Excitation of ducted gravity waves in the lower thermosphere by tropospheric sources. J. Geophys. Res., 113, A06303, https://doi.org/10.1029/2007JA012693.
Song, I.-S., and H.-Y. Chun, 2008: A Lagrangian spectral parameterization of gravity wave drag induced by cumulus convection. J. Atmos. Sci., 65, 1204–1224, https://doi.org/10.1175/2007JAS2369.1.
Song, I.-S., C. Lee, H.-Y. Chun, J.-H. Kim, G. Jee, B.-G. Song, and J. T. Bacmeister, 2020: Propagation of gravity waves and its effects on pseudomomentum flux in a sudden stratospheric warming event. Atmos. Chem. Phys., 20, 7617–7644, https://doi.org/10.5194/acp-20-7617-2020.
Sonmor, L. J., and G. P. Klaassen, 1997: Toward a unified theory of gravity wave stability. J. Atmos. Sci., 54, 2655–2680, https://doi.org/10.1175/1520-0469(1997)054<2655:TAUTOG>2.0.CO;2.
Souprayen, C., J. Vanneste, and A. H. A. Hauchecorne, 2001: Atmospheric gravity wave spectra: A stochastic approach. J. Geophys. Res., 106, 24 071–24 086, https://doi.org/10.1029/2001JD900043.
Stephan, C. C., H. Schmidt, C. Zülicke, and V. Matthias, 2020: Oblique gravity wave propagation during sudden stratospheric warmings. J. Geophys. Res. Atmos., 125, e2019JD031528, https://doi.org/10.1029/2019JD031528.
Strube, C., P. Preusse, M. Ern, and M. Riese, 2021: Propagation paths and source distributions of resolved gravity waves in ECMWF-IFS analysis fields around the southern polar night jet. Atmos. Chem. Phys., 21, 18 641–18 668, https://doi.org/10.5194/acp-21-18641-2021.
Viúdez, A., 2005: The vorticity–velocity gradient cofactor tensor and the material invariant of the semigeostrophic theory. J. Atmos. Sci., 62, 2294–2301, https://doi.org/10.1175/JAS3482.1.
Walters, D., and Coauthors, 2019: The Met Office Unified Model Global Atmosphere 7.0/7.1 and JULES Global Land 7.0 configurations. Geosci. Model Dev., 12, 1909–1963, https://doi.org/10.5194/gmd-12-1909-2019.
Warner, C. D., and M. E. McIntyre, 2001: An ultrasimple spectral parameterization for nonorographic gravity waves. J. Atmos. Sci., 58, 1837–1857, https://doi.org/10.1175/1520-0469(2001)058<1837:AUSPFN>2.0.CO;2.
Wei, J., G. Bölöni, and U. Achatz, 2019: Efficient modeling of the interaction of mesoscale gravity waves with unbalanced large-scale flows: Pseudomomentum-flux convergence versus direct approach. J. Atmos. Sci., 76, 2715–2738, https://doi.org/10.1175/JAS-D-18-0337.1.
Wilson, E. B., 1929: Vector Analysis: A Text-Book for the Use of Students of Mathematics and Physics, Founded upon the Lectures of J. Willard Gibbs. Yale University Press, 480 pp.
Zdunkowski, W., and A. Bott, 2003: Dynamics of the Atmosphere: A Course in Theoretical Meteorology. Cambridge University Press, 738 pp.
Zdunkowski, W., and A. Bott, 2004: Thermodynamics of the Atmosphere: A Course in Theoretical Meteorology. Cambridge University Press, 251 pp.