1. Introduction
In the decades since Nastrom and Gage (1985) first described the shape of the atmospheric kinetic energy (KE) spectrum, the mystery of this spectrum has motivated fundamental research in several areas of geophysical fluid dynamics. The −3 spectral slope at synoptic scales is well explained by quasigeostrophic (QG) turbulence theory (Charney 1971), but the dynamical mechanisms behind the mesoscale −5/3 slope, which extends from scales of O(100) to O(1) km, is still an active area of research. While some early work focused on the possibility of an inverse energy cascade through the mesoscale driven by convection (Gage 1979; Lilly 1983), it now appears that the energy cascade is downscale at sub-100-km scales (Lindborg and Cho 2001). Different mechanisms have been proposed for such a cascade, including inertia–gravity waves (IGWs; Dewan 1979; VanZandt 1982; Bartello 1995), anisotropic stratified turbulence (Lindborg 2006; Riley and Lindborg 2008), and balanced dynamics near the tropopause (Tulloch and Smith 2006). Nevertheless, a complete picture of the mesoscale cascade process remains elusive. Indeed, it is possible that more than one mechanism is important, or that different mechanisms act at different levels or scales. In addition to being a fundamental problem in geophysical turbulence, this question has important practical implications: the mesoscale cascade connects the large, energy-containing scales with microscales, where turbulent dissipation occurs, and is not fully resolved by atmospheric models. Parameterizations of the cascade should be consistent with its physics (e.g., Shutts 2005; Schaefer-Rolffs and Becker 2013).
A basic question about the atmospheric energy spectrum, which must be answered before the various proposed theories can be evaluated, is whether it is dominated by IGWs and/or geostrophic vortices. Several studies have attempted to decompose spectra from aircraft data into wave and vortex contributions, mainly using the Helmholtz decomposition of the horizontal velocity field. This decomposition separates the KE into horizontally rotational and divergent kinetic energy (RKE and DKE, respectively), which are loosely attributed to geostrophic vortices and IGWs. This attribution is not exact, even in a linear sense, since IGWs have some rotational motion, especially at low frequencies. Furthermore, at large scales and low latitudes, Rossby waves have some divergent motion. Nonlinearities complicate matters further, since some of the horizontal divergence is balanced, even on an f-plane (e.g., Kafiabad and Bartello 2016). Nevertheless, the Helmholtz decomposition is useful and straightforward. Helmholtz-decomposed KE spectra from aircraft data show that, while the KE spectrum is dominated by the RKE in the troposphere (Cho et al. 1999; Lindborg 2007), the level of DKE is not negligible in the troposphere, and actually dominates over the RKE in the lower stratosphere (Callies et al. 2016; Li and Lindborg 2018). Furthermore, Callies et al. (2016) used the Helmholtz-decomposed KE spectra, along with potential energy (PE) spectra from temperature data, to construct linear IGW energy spectra, confirming that wave energy dominates the total mesoscale energy in the lower stratosphere. These studies suggest that the mesoscale spectrum may be explained by IGW theories in the lower stratosphere, but not in the upper troposphere.
Most atmospheric models are capable of reproducing at least the qualitative shape of the atmospheric spectrum when run at sufficiently high resolution (e.g., Koshyk and Hamilton 2001; Skamarock 2004; Hamilton et al. 2008; Skamarock et al. 2014). Such models can be used to address physical questions about the mesoscale cascade. For example, numerical studies have confirmed that the spectral flux is downscale below scales of 100 km (Augier and Lindborg 2013). Other studies have shown that, while the lower-stratospheric KE spectrum is dominated by DKE consistent with IGWs, the spectral energy budget is strongly influenced by vertical wave fluxes in addition to downscale transfer (Waite and Snyder 2009; Peng et al. 2015a). Nevertheless, there are some differences between models, such as the relatively low fraction of DKE in Hamilton et al. (2008) compared to Skamarock and Klemp (2008). Some of these differences may be due to vertical resolution (Waite 2016; Skamarock et al. 2019) and physical parameterizations (Malardel and Wedi 2016), both of which can affect the simulated spectra.
For spectral analysis from aircraft data, spectra are computed at (approximately) constant height by necessity, and energy decompositions are restricted to horizontal, not vertical, structure. However, analysis of model output allows for more flexibility. An alternative spectral decomposition for model data can be formulated with linear normal modes. The equations of motion can be linearized around a reference state, and solutions to the equations can be expressed as a sum of the orthogonal normal modes of the system (e.g., Kasahara and Puri 1981). By orthogonality, the total domain-averaged energy can be expressed as a sum of modal energies. This approach has several advantages over the usual procedure of analyzing spectra independently at different model levels. First, unlike the Helmholtz decomposition, the normal modes fully separate linear IGWs and geostrophic vortices; there is no ambiguity, at least in a linear sense, about whether the RKE is due to geostrophic vortices or low-frequency IGWs. Second, the vertical structure of the IGWs is built into the decomposition. Third, since potential energy is considered, the normal modes decompose the total mechanical energy, not just the kinetic energy. Finally, since the modes are orthogonal, they allow for a complete decomposition of the total energy into geostrophic vortices and IGWs at every horizontal and vertical scale.
Normal modes have been used to investigate the mesoscale spectrum in both idealized simulations and more realistic model data. Bartello (1995) used normal modes to decompose energy spectra in homogeneous rotating–stratified turbulence, and found that IGW energy cascades downscale with a shallow spectrum due to catalytic wave–vortex interactions. The normal mode approach is now commonly employed in idealized studies of rotating–stratified turbulence (e.g., Kitamura and Matsuda 2010; Deusebio et al. 2013; Herbert et al. 2016). In a more realistic context, Terasaki et al. (2011) and Žagar et al. (2017) used normal modes to decompose the energy spectrum in global reanalysis data. They found that, like Bartello (1995), the IGW spectrum is shallower than the geostrophic spectrum at all scales; indeed, the IGW spectral slope is around −5/3 across a wide range of length scales. At sufficiently small scales, the shallower IGW spectrum crosses the steeper geostrophic spectrum, and the total energy is dominated by IGWs. In the idealized homogeneous studies, the normal mode structure is Fourier in the horizontal and vertical; by contrast, with global data, the horizontal structure are Hough modes, while the vertical structure is found by solving the associated eigenvalue problem (e.g., Kasahara and Puri 1981).
In this paper, we use normal modes on a periodic f-plane to analyze the spectrum in an idealized mesoscale simulation driven by baroclinic instability. Baroclinic waves have been used in several studies for idealized investigations of the mesoscale spectrum (Waite and Snyder 2009; Peng et al. 2015a,b). Despite the simple setup, these simulations yield many realistic mesoscale features, including DKE spectra with a −5/3 slope. These simulations, while idealized, are able to capture the basic mechanism by which IGWs energy is transferred downscale with a −5/3 spectrum. In dry simulations, the amplitude of the DKE spectrum is not large enough to significantly modify the KE spectrum in the troposphere, but it does in the stratosphere, where the total kinetic energy shallows under the dominance of the DKE (Waite and Snyder 2009). The addition of moist physics (Waite and Snyder 2013; Peng et al. 2015a,b) or topography and surface fluxes (Menchaca and Durran 2019) enhances the DKE spectrum and, as a result, the mesoscale KE spectrum.
Idealized baroclinic instability offers a test case of intermediate complexity in which to investigate the role of IGWs in the mesoscale spectrum. Compared to homogeneous rotating–stratified turbulence, this approach uses more realistic vertical structure and large-scale vortices. On the other hand, the geometry of the f-plane allows for the use of Fourier modes to represent the horizontal structure, which are considerably simpler than Hough modes. This setup allows for a careful analysis of the normal mode spectra and comparison between the normal mode and Helmholtz decompositions. The rest of this paper is organized as follows. In section 2, we review the equations for the vertical and horizontal modes. In section 3, we outline the numerical test case and model. Results are presented in section 4, and conclusions are given in section 5.
2. Modes
a. Separation of variables
b. Vertical modes
c. Horizontal modes
In summary, the full three-dimensional dynamical fields can decomposed, first into vertical modes, and then into horizontal modes. For each vertical mode n and each horizontal wave vector k, there are three modes: one geostrophic mode (with amplitude
d. Reconstruction and energetics
e. Relationship between modal and Helmholtz decomposition
This analysis suggests that, for sufficiently small horizontal scales with
3. Numerical approach
a. Initial conditions
Parameters used in the initialization and simulation.



Initial mean vertical profile of the (a) horizontal mean temperature and (b) Brunt–Väisälä frequency.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

Initial mean vertical profile of the (a) horizontal mean temperature and (b) Brunt–Väisälä frequency.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
Initial mean vertical profile of the (a) horizontal mean temperature and (b) Brunt–Väisälä frequency.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
To use periodic horizontal boundary conditions, which simplifies the analysis of the horizontal normal modes, we extend the jet meridionally to construct a doubly periodic double jet. The domain is extended from

Meridional slice of initial velocity (colored; m s−1) and potential temperature (gray contours; interval: 10 K) after the domain has been extended to be doubly periodic.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

Meridional slice of initial velocity (colored; m s−1) and potential temperature (gray contours; interval: 10 K) after the domain has been extended to be doubly periodic.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
Meridional slice of initial velocity (colored; m s−1) and potential temperature (gray contours; interval: 10 K) after the domain has been extended to be doubly periodic.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
Finally, to initialize baroclinic instability, each jet is perturbed with a two-dimensional Gaussian potential temperature bubble of amplitude 4 K and half-width 600 km (Ullrich et al. 2015) centered at
b. Model setup
Simulations are performed with the Advanced Research dynamical core of the Weather Research and Forecasting (WRF, version 3.7.1) Model (Skamarock et al. 2008), which solves the equations for a compressible, nonhydrostatic atmosphere. The domain is a square doubly periodic f-plane with size L × L. The horizontal grid spacing is Δx = Δy = 5 km. The domain depth is 25 km, with 100 levels evenly spaced in the basic-state pressure. The grid spacing Δz therefore increases with height from 84 m near the surface, to 260 m at z = 10 km, to 2 km near the top. The basic-state pressure at the model surface and lid are ps = 1017.5 hPa and pt = 7.0 hPa. The time step is Δt = 30 s.
Since these are idealized simulations, most physical parameterizations are turned off. Simulations are dry and there are no surface fluxes, radiative transfer, or boundary layer mixing. A diffusive damping layer with dimensionless coefficient 0.05 is employed within 5 km of the upper boundary. The advection scheme is fifth order in the horizontal and third order in the vertical; the weak numerical viscosity from the upwind-biased fluxes is sufficient to suppress gridscale noise, and no additional eddy viscosity is imposed (e.g., Wicker and Skamarock 2002). Simulations were run for 16 days with fields output every 2 h.
c. Vertical modes
To analyze the simulation output in terms of the normal modes, we need the set of vertical modes Zn(p) and equivalent depths hn associated with the background state and numerical grid of our simulation. We compute the modes from the initial basic-state
Figure 3 shows the computed equivalent depths hn and deformation wavenumber

(a),(c) Equivalent depths hn and (b),(d) deformation wavenumbers
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

(a),(c) Equivalent depths hn and (b),(d) deformation wavenumbers
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
(a),(c) Equivalent depths hn and (b),(d) deformation wavenumbers
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

Vertical structure of select vertical modes Zn(p).
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

Vertical structure of select vertical modes Zn(p).
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
Vertical structure of select vertical modes Zn(p).
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
4. Results
a. Simulation overview
Figure 5 shows the evolution of the potential temperature in the lowest grid cell and 500-hPa vertical vorticity and horizontal divergence at t = 0, 4, 8, 12, and 16 days. At the initial time, the jet structure and temperature bubble are clearly visible; since the initial jet is in geostrophic balance, there is no velocity divergence. By 4 days, a single baroclinic wave is growing on each jet, accompanied by a small band of inertia–gravity waves in the divergence field. By 8 days, two mature baroclinic waves are present on each jet, with clearly defined cold regions, cyclonic vorticity, and embedded inertia–gravity waves. Waves are visible above the cold fronts and inside the cyclones. By 12 days and beyond, the baroclinic instability appears to have saturated, instabilities are developing on the surface potential temperature filaments, and waves fill the entire domain.

Horizontal slices of (left) potential temperature (K) in the lowest grid cell (z ≈ 42 m), and (center) 500-hPa vertical vorticity, and (right) horizontal divergence at t = (a)–(c) 0, (d)–(f) 4, (g)–(i) 8, (j)–(l) 12, and (m)–(o) 16 days. Vorticity and divergence are normalized by f.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

Horizontal slices of (left) potential temperature (K) in the lowest grid cell (z ≈ 42 m), and (center) 500-hPa vertical vorticity, and (right) horizontal divergence at t = (a)–(c) 0, (d)–(f) 4, (g)–(i) 8, (j)–(l) 12, and (m)–(o) 16 days. Vorticity and divergence are normalized by f.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
Horizontal slices of (left) potential temperature (K) in the lowest grid cell (z ≈ 42 m), and (center) 500-hPa vertical vorticity, and (right) horizontal divergence at t = (a)–(c) 0, (d)–(f) 4, (g)–(i) 8, (j)–(l) 12, and (m)–(o) 16 days. Vorticity and divergence are normalized by f.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
b. Vertically averaged kinetic energy spectra
Figure 6 shows vertically averaged spectra of kinetic, rotational kinetic, and divergent kinetic energy at a selection of times. By averaging spectra of energy per unit mass on model levels, the average spectra are proportional to energy per unit volume using the basic-state density (e.g., Waite and Snyder 2009). The kinetic energy spectrum (Fig. 6a) grows and fills out as the baroclinic instability develops. The spectrum saturates at around t ≈ 12 days with a slope of −2.6 (unless otherwise stated, spectral slopes are measured between k = 10 and 100). The spectrum falls off rapidly beyond k ≈ 200 due to the numerical dissipation.

Vertically averaged spectra of (a) kinetic energy, (b) rotational kinetic energy, and (c) divergent kinetic energy at t = 2, 4, 6, 8, 12, and 16 days. The straight reference lines show the best-fit slope between k = 10 and 100 at t = 16 days: (a) −2.6, (b) −2.8, and (c) −1.9. Here and in other spectra figures, a factor of Δk is included so that the y axis shows energy, not energy spectral density.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

Vertically averaged spectra of (a) kinetic energy, (b) rotational kinetic energy, and (c) divergent kinetic energy at t = 2, 4, 6, 8, 12, and 16 days. The straight reference lines show the best-fit slope between k = 10 and 100 at t = 16 days: (a) −2.6, (b) −2.8, and (c) −1.9. Here and in other spectra figures, a factor of Δk is included so that the y axis shows energy, not energy spectral density.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
Vertically averaged spectra of (a) kinetic energy, (b) rotational kinetic energy, and (c) divergent kinetic energy at t = 2, 4, 6, 8, 12, and 16 days. The straight reference lines show the best-fit slope between k = 10 and 100 at t = 16 days: (a) −2.6, (b) −2.8, and (c) −1.9. Here and in other spectra figures, a factor of Δk is included so that the y axis shows energy, not energy spectral density.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
The RKE spectrum (Fig. 6b) resembles the total kinetic energy spectrum, and saturates with a slightly steeper slope of −2.8. After 12 days, the RKE spectrum experiences a slight decrease in energy at large k. By contrast, the shallower DKE spectrum, which has a slope of −1.9, continues to grow slowly in amplitude until the end of the simulation. The kinetic energy spectrum is the sum of the steeper RKE spectrum and shallower DKE spectrum; however, as in the dry baroclinic wave simulations of Waite and Snyder (2009), the amplitude of the DKE spectrum is not high enough to influence the total kinetic energy, which is dominated by the steeper RKE spectrum across all resolved wavenumbers. No mesoscale shallowing occurs in the vertically integrated KE spectrum. In what follows, we focus on t = 16 days, since the DKE spectrum has the largest amplitude at this time.
Energy spectra computed on individual model levels are shown in Fig. 7. At 850 hPa (Fig. 7a), the spectra resemble the vertical average, which is dominated by the lower troposphere: the RKE spectrum is steep, the DKE spectrum is shallow, and the total KE spectrum resembles the RKE over most wavenumbers. In fact, the vertically integrated KE spectrum (gray dashed curve in Fig. 7a) is nearly indistinguishable from the 850 hPa KE spectrum. At 500 hPa (Fig. 7b), the RKE and DKE spectra cross around k ≈ 60, beyond which the DKE dominates and the kinetic energy spectrum shallows slightly. This transition is even more pronounced at 200 hPa, where the RKE and DKE spectra cross around k ≈ 25 (Fig. 7c). Therefore, while the KE spectra in the lower stratosphere show clear mesoscale shallowing dominated by DKE (as found in other idealized baroclinic wave studies, for example, Waite and Snyder 2009), the vertically averaged spectra, which are dominated by the lower troposphere, do not.

Spectra of kinetic, rotational kinetic, and divergent kinetic energy at (a) 850, (b) 500, and (c) 200 hPa at t = 16 days. Spectra are computed on model levels, and pressures correspond to basic state. In (a), the vertically averaged KE spectrum is also shown by the gray dashed line for reference, and is almost entirely hidden by KE and RKE (solid thick and thin black curves).
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

Spectra of kinetic, rotational kinetic, and divergent kinetic energy at (a) 850, (b) 500, and (c) 200 hPa at t = 16 days. Spectra are computed on model levels, and pressures correspond to basic state. In (a), the vertically averaged KE spectrum is also shown by the gray dashed line for reference, and is almost entirely hidden by KE and RKE (solid thick and thin black curves).
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
Spectra of kinetic, rotational kinetic, and divergent kinetic energy at (a) 850, (b) 500, and (c) 200 hPa at t = 16 days. Spectra are computed on model levels, and pressures correspond to basic state. In (a), the vertically averaged KE spectrum is also shown by the gray dashed line for reference, and is almost entirely hidden by KE and RKE (solid thick and thin black curves).
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
c. Modal spectra
Figure 8 shows the geostrophic and ageostrophic energy spectra for a selection of vertical modes at t = 16 days. For small mode numbers, the geostrophic energy dominates at all horizontal scales, and both the geostrophic and ageostrophic energy spectra are very steep, with slopes of around −4 (the spectral slopes are plotted against n in Fig. 9). Indeed, for n ≤ 3, the ageostrophic energy spectrum is actually steeper than the geostrophic. As n increases, both spectra get shallower, and the ageostrophic spectra are consistently shallower than the geostrophic. For example, at n = 10, the geostrophic and ageostrophic slopes are −3.3 and −2.6; for n = 20, they are −3.0 and −2.0; and for n = 30, they are −2.5 and −1.4. As a result of their different slopes, the shallower ageostrophic spectrum intersects the steeper geostrophic spectrum for n ≥ 5 and exceeds the GE at large k. The geostrophic energy spectral slope shallows to around −2 by n = 40 and stays there for larger n, while the ageostrophic spectrum gets increasingly shallow for larger n (see below).

Spectra of geostrophic (solid) and ageostrophic (dashed) energy at t = 16 days for a selection of modes, as labeled (as in Fig. 3). The black reference curves have slopes of −5/3 and −3, and the gray vertical line indicates the Rossby deformation wavenumber
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

Spectra of geostrophic (solid) and ageostrophic (dashed) energy at t = 16 days for a selection of modes, as labeled (as in Fig. 3). The black reference curves have slopes of −5/3 and −3, and the gray vertical line indicates the Rossby deformation wavenumber
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
Spectra of geostrophic (solid) and ageostrophic (dashed) energy at t = 16 days for a selection of modes, as labeled (as in Fig. 3). The black reference curves have slopes of −5/3 and −3, and the gray vertical line indicates the Rossby deformation wavenumber
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

Mesoscale slope of the geostrophic, ageostrophic, rotational kinetic, and divergent kinetic energy spectra at t = 16 days. Slopes are measured for 10 ≤ k ≤ 100.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

Mesoscale slope of the geostrophic, ageostrophic, rotational kinetic, and divergent kinetic energy spectra at t = 16 days. Slopes are measured for 10 ≤ k ≤ 100.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
Mesoscale slope of the geostrophic, ageostrophic, rotational kinetic, and divergent kinetic energy spectra at t = 16 days. Slopes are measured for 10 ≤ k ≤ 100.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
The geostrophic energy spectrum is peaked at small wavenumbers for all n. The largest-scale geostrophic vortices project onto small k at all n. By contrast, the peak in the ageostrophic spectrum moves to larger wavenumbers with increasing n. The wavenumber peak appears to occur near the deformation wavenumber, which is indicated by the vertical lines in Fig. 8. We can measure the peak by finding the wavenumber of maximum kEA(k, n) for each n; this wavenumber is plotted along with kR in Fig. 10. The actual peak wavenumbers are indeed clustered around kR for n ≳ 10. These spectra imply that inertia–gravity waves with increasing n, and therefore decreasing vertical scale, are characterized by a correspondingly small horizontal scale that is given by the deformation scale. For n ≳ 50, the peak occurs near the numerical dissipation range; it cannot increase further, and no spectral power law is resolved at larger k.

The wavenumber of the peak in ageostrophic energy (black circles) and the Rossby deformation wavenumber (solid) plotted against n.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

The wavenumber of the peak in ageostrophic energy (black circles) and the Rossby deformation wavenumber (solid) plotted against n.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
The wavenumber of the peak in ageostrophic energy (black circles) and the Rossby deformation wavenumber (solid) plotted against n.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
Generally, the amplitudes of the spectra in Fig. 8, and the total amount of energy in each vertical mode, decrease with increasing n. The geostrophic and ageostrophic energy in each vertical mode are plotted in Fig. 11a. Most of the energy in the system is geostrophic energy in the lowest few modes, associated with the large-scale baroclinic wave, as expected. The geostrophic energy is peaked at n = 3 and dominates over the ageostrophic energy for all n [cf. Terasaki and Tanaka (2007), who similarly found a peak in the total energy at mode 4]. Interestingly, both the geostrophic and ageostrophic energy per mode decreases like n−3 for n ≳ 10, which is reminiscent of the −3 vertical wavenumber spectrum associated with QG turbulence. The mesoscale energy in each vertical mode is computed by restricting to k ≥ 10 (Fig. 11b). The mesoscale geostrophic energy is peaked at n = 3, like the total geostrophic energy, which is dominated by large horizontal scales. By contrast, the mesoscale ageostrophic energy is contained in intermediate vertical modes, and peaked at n ≈ 9.

Total geostrophic (solid) and ageostrophic (dashed) energy in each vertical mode, computed using (a) all horizontal scales and (b) mesoscales (k ≥ 10). The solid reference line has a slope of −3.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

Total geostrophic (solid) and ageostrophic (dashed) energy in each vertical mode, computed using (a) all horizontal scales and (b) mesoscales (k ≥ 10). The solid reference line has a slope of −3.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
Total geostrophic (solid) and ageostrophic (dashed) energy in each vertical mode, computed using (a) all horizontal scales and (b) mesoscales (k ≥ 10). The solid reference line has a slope of −3.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
d. Helmholtz decomposition
The left column of Fig. 12 shows the geostrophic, rotational kinetic, and potential energy spectra for modes n = 1, 10, 20, 40, and 80. Equations (38) and (42) suggest that the geostrophic energy spectrum should be well approximated by the rotational kinetic energy spectrum for

(left) Spectra of geostrophic, rotational kinetic, and potential energy. (right) Spectra of ageostrophic, divergent kinetic, divergent kinetic plus potential, and kinetic energy. All spectra are at t = 16 days for a selection of modes, as labeled. The black reference curves have slopes of −5/3 and −3, and the gray vertical line indicates the Rossby deformation radius
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

(left) Spectra of geostrophic, rotational kinetic, and potential energy. (right) Spectra of ageostrophic, divergent kinetic, divergent kinetic plus potential, and kinetic energy. All spectra are at t = 16 days for a selection of modes, as labeled. The black reference curves have slopes of −5/3 and −3, and the gray vertical line indicates the Rossby deformation radius
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
(left) Spectra of geostrophic, rotational kinetic, and potential energy. (right) Spectra of ageostrophic, divergent kinetic, divergent kinetic plus potential, and kinetic energy. All spectra are at t = 16 days for a selection of modes, as labeled. The black reference curves have slopes of −5/3 and −3, and the gray vertical line indicates the Rossby deformation radius
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
However, the transition happens at k larger than

(a) The ratio of the potential and rotational kinetic energy to the geostrophic energy and (b) the ratio of the kinetic and the potential plus divergent kinetic energy to the ageostrophic energy. In both panels, n = 20 and the thick vertical gray line shows
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

(a) The ratio of the potential and rotational kinetic energy to the geostrophic energy and (b) the ratio of the kinetic and the potential plus divergent kinetic energy to the ageostrophic energy. In both panels, n = 20 and the thick vertical gray line shows
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
(a) The ratio of the potential and rotational kinetic energy to the geostrophic energy and (b) the ratio of the kinetic and the potential plus divergent kinetic energy to the ageostrophic energy. In both panels, n = 20 and the thick vertical gray line shows
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
The right column of Fig. 12 shows the ageostrophic, divergent kinetic, divergent kinetic plus potential, and kinetic energy spectra for the same modes as in the left column. As was the case for the geostrophic energy, the ageostrophic energy spectra follow the approximations in Eqs. (43) and (46) at large and small wavenumbers: the ageostrophic energy is approximated by the potential plus divergent kinetic energy for
However, as was the case for the geostrophic energy spectra, the transition between these approximation regimes occurs at a wavenumber beyond
e. Vertically averaged geostrophic and ageostrophic spectra
While the modal spectra allow for a detailed comparison of the geostrophic/ageostrophic and Helmholtz-decomposed spectra, most model spectral analysis considers vertically averaged (over layers if not the whole domain), not modal, spectra. Figure 14 shows modal spectra summed over all vertical modes, corresponding to a vertical average over the whole domain [cf. Eq. (34)]. The geostrophic and rotational kinetic energy spectra agree well over all k. There is slightly more GE than RKE at intermediate wavenumbers, but overall, the amplitudes and slopes are very similar: the mesoscale slopes of the geostrophic and rotational kinetic energy spectra are −3.0 and −2.8, respectively. As discussed above, the RKE spectrum is expected to approximate the GE spectrum for k larger than the Rossby deformation wavenumber; since the vertically averaged spectra are dominated by small n (Fig. 11), which have relatively small

Spectra of geostrophic, ageostrophic, rotational kinetic, and divergent kinetic energy summed over all n.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1

Spectra of geostrophic, ageostrophic, rotational kinetic, and divergent kinetic energy summed over all n.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
Spectra of geostrophic, ageostrophic, rotational kinetic, and divergent kinetic energy summed over all n.
Citation: Journal of the Atmospheric Sciences 77, 3; 10.1175/JAS-D-19-0146.1
The ageostrophic and divergent kinetic energy spectra have similar slopes but different amplitudes over all k. The mesoscale slopes are −2.1 and −1.9, respectively, for the AE and DKE spectra. The vertically averaged ageostrophic spectrum is shallower than the ageostrophic spectra of the dominant low vertical modes (Fig. 8); for example, the ageostrophic spectral slopes are −3.9, −4.8, and −3.3 for n = 1, 2, and 4, which are all steeper than −2.1. The shallowness of the vertically averaged ageostrophic spectrum seems to be due to the higher n modes, the spectra of which are peaked at
5. Conclusions
Energy spectra in idealized baroclinic wave simulations were investigated using linear normal modes. The usual approach for analyzing energy spectra in atmospheric models is to consider kinetic energy spectra, often Helmholtz-decomposed into rotational and divergent KE, at different vertical levels. The normal mode approach is different: it decomposes the total domain-averaged mechanical energy into geostrophic and ageostrophic energy at every horizontal and vertical mode. In the idealized simulations performed here, the horizontal mode structure is Fourier and the vertical mode structure is found by solving the appropriate eigenvalue problem.
The vertically averaged KE spectra have a simple and familiar power-law structure: the RKE has a slope of around −2.8 and the DKE has a shallower slope of around −1.9, which are close to the −3 and −5/3 reported elsewhere (e.g., Waite and Snyder 2009; Peng et al. 2015a). However, the amplitude of the DKE spectrum is too small to influence the kinetic energy spectrum, even at small scales. The KE spectral slope is −2.6 down to the (numerical) dissipation scale and no transition to a shallower mesoscale spectrum is apparent. As found in previous studies of idealized dry baroclinic waves (e.g., Waite and Snyder 2009), there is a transition to a shallower spectrum in the lower stratosphere, but not in the troposphere; since the vertically averaged spectra are dominated by lower levels, this shallowing is not apparent in the vertical average.
The normal mode decomposition gives an interesting perspective into how the energy is partitioned into geostrophic and ageostrophic (IGW) motion at different horizontal and vertical scales. For small vertical mode numbers, corresponding to deep vertical structures, both the GE and AE spectra are very steep, and the GE energy dominates at all k. However, for increasing n, the AE spectrum shallows while the GE spectrum remains steep; for n ≥ 5, the GE and AE spectra cross and the small horizontal scales are dominated by ageostrophic energy. Overall, both the GE and AE spectra get shallower with increasing n. The GE spectrum, which is always peaked at large scales, appears to be approaching a slope of around −2. By contrast, the AE spectra are peaked around the Rossby deformation wavenumber
Most of the energy in the system, both GE and AE, is contained in the lowest vertical modes, which have the largest vertical scales. Indeed, the modal energy decreases with increasing n like n−3, similar to the Fourier spectrum expected for homogeneous QG turbulence. Since the energy in every vertical mode is dominated by the largest scales, it is perhaps not surprising that the vertical modal spectrum follows the QG scaling. However, when only the mesoscale (k ≥ 10) is considered, there is a significant difference between geostrophic and ageostrophic energy. The mesoscale geostrophic energy continues to be dominated by the smallest n, but the ageostrophic energy is peaked at n = 9, which has a midtropospheric vertical scale of a few kilometers.
A detailed comparison between the normal mode and Helmholtz decompositions was given. We showed that, for
The similarity between the normal mode and Helmholtz decomposed spectral slopes is promising. The Helmholtz decomposition is simple and straightforward to employ for both aircraft and model data, and these results shown here suggest that the RKE and DKE spectral slopes are indeed good approximations to the more physically meaningful geostrophic and ageostrophic modal spectra. However, caution and more work are required before extrapolating these idealized results to more comprehensive models. In particular, simulations and reanalysis in larger or global domains, in which the Rossby deformation scales of the dominant vertical modes are smaller than the domain size, would likely show more of a difference between the normal mode and Helmholtz spectral slopes at large scales. Rossby wave velocity divergence associated with meridionally varying Coriolis parameter is also missing in the f-plane setup considered here. In addition, physical mechanisms that are neglected here are known to amplify the DKE spectra, and will therefore have a similar effect on the AE: these include moist physics (Waite and Snyder 2013; Peng et al. 2015a,b; Sun et al. 2017) and topography and surface fluxes (Menchaca and Durran 2019). Indeed, normal mode spectra from reanalysis data show that the GE and AE spectra do indeed cross at length scales of a few hundred kilometers, leading to a more pronounced mesoscale spectrum that what was found here (Terasaki et al. 2011; Žagar et al. 2017). The effects of these physical processes on the normal mode spectra, and in particular on the amplitude of the ageostrophic (IGW) energy, requires further investigation.
Acknowledgments
Model data used in this study are available upon request to the corresponding author. This work benefited from comments by Francis Poulin, Marek Stastna, and two anonymous referees. Computational resources were provided by the Shared Hierarchical Academic Research Computing Network (SHARCNET) and Compute/Calcul Canada. Funding from the Natural Sciences and Engineering Research Council of Canada (Grant RGPIN-386456-2015) is gratefully acknowledged.
REFERENCES
Augier, P., and E. Lindborg, 2013: A new formulation of the spectral energy budget of the atmosphere, with application to two high-resolution general circulation models. J. Atmos. Sci., 70, 2293–2308, https://doi.org/10.1175/JAS-D-12-0281.1.
Bartello, P., 1995: Geostrophic adjustment and inverse cascades in rotating stratified turbulence. J. Atmos. Sci., 52, 4410–4428, https://doi.org/10.1175/1520-0469(1995)052<4410:GAAICI>2.0.CO;2.
Callies, J., O. Bühler, and R. Ferrari, 2016: The dynamics of mesoscale winds in the upper troposphere and lower stratosphere. J. Atmos. Sci., 73, 4853–4872, https://doi.org/10.1175/JAS-D-16-0108.1.
Charney, J. G., 1971: Geostrophic turbulence. J. Atmos. Sci., 28, 1087–1095, https://doi.org/10.1175/1520-0469(1971)028<1087:GT>2.0.CO;2.
Cho, J. Y. N., R. E. Newell, and J. D. Barrick, 1999: Horizontal wavenumber spectra of winds, temperature, and trace gases during the Pacific Exploratory Missions: 2. Gravity waves, quasi-two-dimensional turbulence, and vortical modes. J. Geophys. Res., 104, 16 297–16 308, https://doi.org/10.1029/1999JD900068.
Cohn, S. E., and D. P. Dee, 1989: An analysis of the vertical structure equation for arbitrary thermal profiles. Quart. J. Roy. Meteor. Soc., 115, 143–171, https://doi.org/10.1002/qj.49711548508.
Daley, R., 1991: Atmospheric Data Analysis. Cambridge University Press, 457 pp.
Deusebio, E., A. Vallgren, and E. Lindborg, 2013: The route to dissipation in strongly stratified and rotating flows. J. Fluid Mech., 720, 66–103, https://doi.org/10.1017/jfm.2012.611.
Dewan, E. M., 1979: Stratospheric wave spectra resembling turbulence. Science, 204, 832–835, https://doi.org/10.1126/science.204.4395.832.
Gage, K. S., 1979: Evidence for a k−5/3 law inertial range in mesoscale two-dimensional turbulence. J. Atmos. Sci., 36, 1950–1954, https://doi.org/10.1175/1520-0469(1979)036<1950:EFALIR>2.0.CO;2.
Hamilton, K., Y. O. Takahashi, and W. Ohfuchi, 2008: Mesoscale spectrum of atmospheric motions investigated in a very fine resolution global general circulation model. J. Geophys. Res., 113, D18110, https://doi.org/10.1029/2008JD009785.
Herbert, C., R. Marino, D. Rosenberg, and A. Pouquet, 2016: Waves and vortices in the inverse cascade regime of stratified turbulence with or without rotation. J. Fluid Mech., 806, 165–204, https://doi.org/10.1017/jfm.2016.581.
Kafiabad, H. A., and P. Bartello, 2016: Balance dynamics in rotating stratified turbulence. J. Fluid Mech., 795, 914–949, https://doi.org/10.1017/jfm.2016.164.
Kasahara, A., 1976: Normal modes of ultralong waves in the atmosphere. Mon. Wea. Rev., 104, 669–690, https://doi.org/10.1175/1520-0493(1976)104<0669:NMOUWI>2.0.CO;2.
Kasahara, A., and K. Puri, 1981: Spectral representation of three-dimensional global data by expansion in normal mode functions. Mon. Wea. Rev., 109, 37–51, https://doi.org/10.1175/1520-0493(1981)109<0037:SROTDG>2.0.CO;2.
Kasahara, A., and J.-H. Qian, 2000: Normal modes of a global nonhydrostatic atmospheric model. Mon. Wea. Rev., 128, 3357–3375, https://doi.org/10.1175/1520-0493(2000)128<3357:NMOAGN>2.0.CO;2.
Kitamura, Y., and Y. Matsuda, 2010: Energy cascade processes in rotating stratified turbulence with application to the atmospheric mesoscale. J. Geophys. Res., 115, D11104, https://doi.org/10.1029/2009JD012368.
Koshyk, J. N., and K. Hamilton, 2001: The horizontal kinetic energy spectrum and spectral budget simulated by a high-resolution troposphere-stratosphere-mesosphere GCM. J. Atmos. Sci., 58, 329–348, https://doi.org/10.1175/1520-0469(2001)058<0329:THKESA>2.0.CO;2.
Li, Q., and E. Lindborg, 2018: Weakly or strongly nonlinear mesoscale dynamics close to the tropopause. J. Atmos. Sci., 75, 1215–1229, https://doi.org/10.1175/JAS-D-17-0063.1.
Lilly, D. K., 1983: Stratified turbulence and the mesoscale variability of the atmosphere. J. Atmos. Sci., 40, 749–761, https://doi.org/10.1175/1520-0469(1983)040<0749:STATMV>2.0.CO;2.
Lindborg, E., 2006: The energy cascade in a strongly stratified fluid. J. Fluid Mech., 550, 207–242, https://doi.org/10.1017/S0022112005008128.
Lindborg, E., 2007: Horizontal wavenumber spectra of vertical vorticity and horizontal divergence in the upper troposphere and lower stratosphere. J. Atmos. Sci., 64, 1017–1025, https://doi.org/10.1175/JAS3864.1.
Lindborg, E., and J. Y. N. Cho, 2001: Horizontal velocity structure functions in the upper troposphere and lower stratosphere: 2. Theoretical considerations. J. Geophys. Res., 106, 10 233–10 241, https://doi.org/10.1029/2000JD900815.
Malardel, S., and N. P. Wedi, 2016: How does subgrid-scale parametrization influence nonlinear spectral energy fluxes in global NWP models? J. Geophys. Res. Atmos., 121, 5395–5410, https://doi.org/10.1002/2015JD023970.
Menchaca, M. Q., and D. R. Durran, 2019: The influence of gravity waves on the slope of the kinetic-energy spectrum in simulations of idealized midlatitude cyclones. J. Atmos. Sci., 76, 2103–2122, https://doi.org/10.1175/JAS-D-18-0329.1.
Nastrom, G. D., and K. S. Gage, 1985: A climatology of atmospheric wavenumber spectra observed by commercial aircraft. J. Atmos. Sci., 42, 950–960, https://doi.org/10.1175/1520-0469(1985)042<0950:ACOAWS>2.0.CO;2.
Peng, J., L. Zhang, and J. Guan, 2015a: Applications of a moist nonhydrostatic formulation of the spectral energy budget to baroclinic waves. Part I: The lower-stratospheric energy spectra. J. Atmos. Sci., 72, 2090–2108, https://doi.org/10.1175/JAS-D-14-0306.1.
Peng, J., L. Zhang, and J. Guan, 2015b: Applications of a moist nonhydrostatic formulation of the spectral energy budget to baroclinic waves. Part II: The upper-tropospheric energy spectra. J. Atmos. Sci., 72, 3923–3939, https://doi.org/10.1175/JAS-D-14-0359.1.
Riley, J. J., and E. Lindborg, 2008: Stratified turbulence: A possible interpretation of some geophysical turbulence measurements. J. Atmos. Sci., 65, 2416–2424, https://doi.org/10.1175/2007JAS2455.1.
Schaefer-Rolffs, U., and E. Becker, 2013: Horizontal momentum diffusion in GCMs using the dynamic Smagorinsky model. Mon. Wea. Rev., 141, 887–899, https://doi.org/10.1175/MWR-D-12-00101.1.
Shutts, G., 2005: A kinetic energy backscatter algorithm for use in ensemble prediction systems. Quart. J. Roy. Meteor. Soc., 131, 3079–3102, https://doi.org/10.1256/qj.04.106.
Skamarock, W. C., 2004: Evaluating mesoscale NWP models using kinetic energy spectra. Mon. Wea. Rev., 132, 3019–3032, https://doi.org/10.1175/MWR2830.1.
Skamarock, W. C., and J. B. Klemp, 2008: A time-split nonhydrostatic atmospheric model for weather research and forecasting applications. J. Comput. Phys., 227, 3465–3485, https://doi.org/10.1016/j.jcp.2007.01.037.
Skamarock, W. C., and Coauthors, 2008: A description of the Advanced Research WRF version 3. NCAR Tech. Note NCAR/TN-475+STR, 113 pp., https://doi.org/10.5065/D68S4MVH.
Skamarock, W. C., S.-H. Park, J. B. Klemp, and C. Snyder, 2014: Atmospheric kinetic energy spectra from global high-resolution nonhydrostatic simulations. J. Atmos. Sci., 71, 4369–4381, https://doi.org/10.1175/JAS-D-14-0114.1.
Skamarock, W. C., C. Snyder, J. B. Klemp, and S.-H. Park, 2019: Vertical resolution requirements in atmospheric simulation. Mon. Wea. Rev., 147, 2641–2656, https://doi.org/10.1175/MWR-D-19-0043.1.
Staniforth, A., M. Béland, and J. Côté, 1985: An analysis of the vertical structure equation in sigma coordinates. Atmos.–Ocean, 23, 323–358, https://doi.org/10.1080/07055900.1985.9649232.
Sun, Y. Q., R. Rotunno, and F. Zhang, 2017: Contributions of moist convection and internal gravity waves to building the atmospheric −5/3 kinetic energy spectra. J. Atmos. Sci., 74, 185–201, https://doi.org/10.1175/JAS-D-16-0097.1.
Tanaka, H., 1985: Global energetics analysis by expansion into three-dimensional normal mode functions during the FGGE winter. J. Meteor. Soc. Japan, 63, 180–200, https://doi.org/10.2151/jmsj1965.63.2_180.
Terasaki, K., and H. L. Tanaka, 2007: An analysis of the 3-D atmospheric energy spectra and interactions using analytical vertical structure functions and two reanalyses. J. Meteor. Soc. Japan, 85, 785–796, https://doi.org/10.2151/jmsj.85.785.
Terasaki, K., H. L. Tanaka, and N. Žagar, 2011: Energy spectra of Rossby and gravity waves. SOLA, 7, 45–48, https://doi.org/10.2151/SOLA.2011-012.
Tulloch, R., and K. S. Smith, 2006: A theory for the atmospheric energy spectrum: Depth-limited temperature anomalies at the tropopause. Proc. Natl. Acad. Sci. USA, 103, 14 690–14 694, https://doi.org/10.1073/pnas.0605494103.
Ullrich, P. A., K. A. Reed, and C. Jablonowski, 2015: Analytical initial conditions and an analysis of baroclinic instability waves in f- and β-plane 3D channel models. Quart. J. Roy. Meteor. Soc., 141, 2972–2988, https://doi.org/10.1002/qj.2583.
VanZandt, T. E., 1982: A universal spectrum of buoyancy waves in the atmosphere. Geophys. Res. Lett., 9, 575–578, https://doi.org/10.1029/GL009i005p00575.
Waite, M. L., 2016: Dependence of model energy spectra on vertical resolution. Mon. Wea. Rev., 144, 1407–1421, https://doi.org/10.1175/MWR-D-15-0316.1.
Waite, M. L., and C. Snyder, 2009: The mesoscale kinetic energy spectrum of a baroclinic life cycle. J. Atmos. Sci., 66, 883–901, https://doi.org/10.1175/2008JAS2829.1.
Waite, M. L., and C. Snyder, 2013: Mesoscale energy spectra of moist baroclinic waves. J. Atmos. Sci., 70, 1242–1256, https://doi.org/10.1175/JAS-D-11-0347.1.
Warn, T., 1986: Statistical mechanical equilibria of the shallow water equations. Tellus, 38A, 1–11, https://doi.org/10.1111/j.1600-0870.1986.tb00448.x.
Wicker, L. J., and W. C. Skamarock, 2002: Time-splitting methods for elastic models using forward time schemes. Mon. Wea. Rev., 130, 2088–2097, https://doi.org/10.1175/1520-0493(2002)130<2088:TSMFEM>2.0.CO;2.
Wiin-Nielsen, A., 1971: On the motion of various vertical modes of transient, very long waves: Part I. Beta plane approximation. Tellus, 23, 87–98, https://doi.org/10.3402/TELLUSA.V23I1.10309.
Žagar, N., K. Terasaki, and H. L. Tanaka, 2012: Impact of the vertical resolution of analysis data on the estimates of large-scale inertio-gravity energy. Mon. Wea. Rev., 140, 2297–2307, https://doi.org/10.1175/MWR-D-11-00103.1.
Žagar, N., D. Jelić, and M. Blaauw, 2017: Energy spectra and inertia–gravity waves in global analyses. J. Atmos. Sci., 74, 2447–2466, https://doi.org/10.1175/JAS-D-16-0341.1.