1. Introduction
The vertical velocity w, w = dz/dt, is a key variable associated with the conversion of the available potential energy to kinetic energy in the global atmosphere. Vertical motions define clouds and precipitation associated with day-to-day weather, the vertical energy propagation by internal gravity waves, and transport of trace constituents. The vertical velocity is needed to represent wind stresses in the horizontal momentum equations in order to compute vertical momentum fluxes from unresolved motions (e.g., Liu 2019). However, w is not an observed quantity of the global observing system. Sporadic observations of vertical velocity make evident the missing variance by the models, at least locally (e.g., Dörnbrack et al. 2018).
The vertical velocity computed by Eq. (1) or Eq. (2) contains signatures of both the Rossby and gravity wave dynamics, of resolved and parameterized physical processes as well as numerical effects. The separation of the governing dynamics and various processes is difficult, especially in the tropics where a frequency gap between the gravity and Rossby dynamical regimes, which is present in the middle latitudes, disappears. But, even in the extratropics the two regimes coexist at a large part of subsynoptic scales where the divergent-dominated dynamics increasingly projects on the inertia–gravity waves (e.g., Žagar et al. 2017), requiring a somewhat arbitrary cutoff scale for the computation of gravity wave momentum fluxes.
We derive herein a new method for linear decomposition of the vertical velocity and associated kinetic energy spectra for the two main dynamical regimes in the hydrostatic atmosphere. The method provides zonal wavenumber kinetic energy spectra of vertical motions as a function of latitude and altitude (pressure level) and is demonstrated using the ERA5 data (Hersbach et al. 2020). It leads the way to a more accurate computation of vertical momentum fluxes in the spherical atmosphere.
Traditionally, the vertical velocity associated with the Rossby wave and inertia–gravity (IG; or gravity) wave dynamics has been studied by computing each of the dynamical components independently of each other. For example, in the middle latitudes, synoptic-scale vertical velocities coupled with ageostrophic motions in baroclinic Rossby waves can be estimated using the quasigeostrophic omega equation (e.g., Hoskins et al. 1978; Stepanyuk et al. 2017). At mesoscales, vertical motions are largely due to internal gravity waves generated by processes such as interaction of the flow with orography, surface and boundary layer processes, tropospheric moist convection, frontogenesis, imbalances of synoptic jets and wave–wave interactions (e.g., Fritts and Nastrom 1992; Fritts and Alexander 2003). Associated vertical velocities can then be quantified using the polarization equations from linear wave theory on the f plane (e.g., Nappo 2002).
A more common way to diagnose internal gravity waves in hydrostatic numerical simulations is using information derived from the divergence field (e.g., Dörnbrack et al. 2018). While suitable for extratropical mesoscale processes, this approach is less informative in the tropics, where the large-scale divergence field includes a mix of dynamical modes, i.e., the Kelvin and mixed Rossby–gravity (MRG) waves on top of a spectrum of inertia–gravity waves. Not only are vertical velocities for each of these modes poorly known, their horizontal velocities are also just as reliable as the methods that are used to filter the waves from the observed or simulated circulation (Knippertz et al. 2022). The Kelvin and MRG waves, together with IG waves, constitute the non-Rossby part of the discrete, linear mode spectrum for the stratified, rotating atmosphere bounded at the top, and their horizontal wavenumber spectra have been analyzed in several studies in recent years (Žagar et al. 2009b; Stephan et al. 2021; Žagar et al. 2022).
The analytical solutions of the Rossby and non-Rossby modes on the sphere, i.e., eigensolutions of the linearized primitive equations, are known as the normal-mode functions (NMF; e.g., Kasahara 2020). As we show in this paper, the normal-mode framework provides also vertical velocities associated with the Rossby and non-Rossby waves in the spherical atmosphere. The involved hydrostatic framework is considered suitable even for model simulations at horizontal resolutions as high as a few kilometers (e.g., Craig and Selz 2018; Dueben et al. 2020).
The decomposition of both horizontal and vertical motions in terms of the Rossby and non-Rossby modes within the same framework provides a consistent comparison of the zonal wavenumber horizontal and vertical energy spectra for the two regimes. The former has been extensively studied including a transition from the −3 power law to −5/3 power law at scales between 1000 and 500 km (e.g., Nastrom and Gage 1985; Rodda and Harlander 2020, and references therein). In contrast, the spectrum of kinetic energy of vertical motions, i.e., the vertical kinetic energy (VKE) spectrum, is poorly known. Observations are few and their spatial coverage is limited (e.g., Bacmeister et al. 1996; Schumann 2019; Dörnbrack et al. 2022). Aircraft observations reveal a local maximum in the VKE at scales between 10 and 100 km, depending on the regional forcing (Dörnbrack et al. 2022, their Fig. B2).
The large-scale part of the VKE spectrum is available only from numerical model simulations. High-resolution nonhydrostatic simulations suggest a nearly flat spectrum of the VKE in the zonal wavenumber domain (e.g., Terasaki et al. 2009; Müller et al. 2018) whereas the spherical harmonics decomposition (global horizontal wavenumber) reveals two maxima, one at synoptic scales and the other near the effective resolutions of the models and dependent on convection modeling (e.g., Skamarock et al. 2014; Polichtchouk et al. 2022; Morfa and Stephan 2023). The interpretation of the global horizontal wavenumber VKE spectra is challenging not least because the spherical harmonics decomposition does not distinguish latitudinal variation in the VKE spectra. A new analytical derivation presented in this paper couples the VKE spectra with dynamics of horizontal motions. This is made possible by a new approach to the computation of the zonal wavenumber spectra of the kinetic energy of horizontal velocities [i.e., horizontal kinetic energy (HKE) spectra] in the NMF framework. The application of the new developments to ERA5 data exposes significant latitudinal variations, i.e., anisotropy of both the HKE and VKE spectra.
The paper consists of four sections, including the introduction that serves as the first section. Section 2 derives, for the first time, the regime-dependent kinetic energy spectra of vertical motions in the pressure coordinate system within a multimodal framework and alongside it, the latitudinal HKE spectra. The validation and application of the new framework to the ECMWF ERA5 data are presented in section 3. Conclusions and outlook are given in section 4.
2. Regime-dependent vertical motions and their kinetic energy spectra
We first derive equations for vertical motions associated with the Rossby and non-Rossby modes in the hydrostatic atmosphere using the normal-mode framework. This is followed by the derivation of the kinetic energy spectrum of vertical velocity and discussion of spectral slopes. Finally, we derive the zonal wavenumber kinetic energy spectrum of horizontal velocity in modal framework building upon the existing, global three-dimensional energy spectra from the MODES software (Žagar et al. 2015) and associated theory that is reviewed in Kasahara (2020) and Tanaka and Žagar (2020).
a. Normal-mode decomposition of the global horizontal motions
For k > 0, the eigensolutions are organized in two distinct groups depending on their frequencies: predominantly divergent inertia–gravity modes that propagate eastward or westward (EIG and WIG modes, respectively), and westward-propagating rotational waves of the Rossby–Haurwitz type. Another subscript, which would denote the three wave species of the normal modes (Rossby, EIG, and WIG waves), does not appear explicitly as it is absorbed within the meridional modal index n. The fastest eastward mode, which corresponds to the n = 0 EIG solution on the sphere, is the Kelvin wave, whereas the fastest westward-propagating rotational mode is the MRG wave (n = 0 Rossby mode in Swarztrauber and Kasahara 1985). For k = 0, all rotational modes have zero frequency while the eastward-propagating and westward-propagating inertia–gravity modes have frequencies of the opposite sign and the same magnitude that are assigned to the eastward and westward propagations, respectively.
b. Discrete three-dimensional solutions
For a stably stratified atmosphere represented in terms of M layers, there are M eigenvalues D of (8), D1 > D2 > … > Dm > … > DM. For every Dm, a number of discrete wave solutions of (12) in terms of Hough harmonics is defined by the maximal number of waves along a latitude circle and the meridional truncations of the Hough functions
c. Frequency relationships for the Rossby and gravity waves
We shall make use of these scaling laws in the derivation of the scaling laws for the vertical kinetic energy in the next subsection and in the discussion of the results.
With respect to Eq. (22), it is necessary to address the question to what extent the dispersion relationships on the sphere, derived for the case D → ∞, apply to frequencies associated with Ds ranging between a few meters and 10 km, as computed for the atmosphere with a finite depth. This question is addressed in appendix A, where we demonstrate that the dispersion relationship (23) for large D is an excellent approximation for numerically computed IG frequencies for many equivalent depths and meridional modes. On the other hand, the relationship (22) provides a good approximation for nearly all Rossby frequencies of the first vertical mode (D1 ≈ 10 km), for the mesoscale range of wavenumbers with D > 200 m and for the waves with large k and equivalent depths D > 10 m. Furthermore, we show that the frequencies of all Rossby modes can be well approximated by a more elaborate formula, structurally similar to the β-plane relationship (25), implying similar scaling and asymptotic behavior.
d. Modal decomposition of the vertical velocity
e. Zonal wavenumber kinetic energy spectrum of vertical motions
Scaling laws for the frequency, the total mechanical energy I (sum of the kinetic energy of horizontal motions and available potential energy), and the vertical kinetic energy (KE) E as a function of the zonal wavenumber k. The wavenumber index κ is used for the large-scale tropical flows where both k and meridional mode index n are small.
The −5 power law for VKE is associated with the geostrophic wind divergence which is due to the β term and is proportional to υgβ/f. On the other hand, ageostrophic motions, a part of quasigeostrophic turbulence, which is characterized by the −3 power law, project on the IG modes. This can be compared with a common assumption that the VKE is proportional to the square of the horizontal wind divergence that leads to a −1 power law for
The large-scale tropical flows and the two special tropical wave solutions are given a separate treatment in Table 1. For scales with k < 7, Žagar et al. (2017) showed that global energy spectra of horizontal motions associated with non-Rossby modes approximately follow a −1 slope. Involving the frequency scaling discussed in the previous subsection and a nearly white spectrum of the planetary-scale horizontal circulation in tropics suggests a somewhat steeper
f. Latitude- and altitude-dependent kinetic energy spectra of horizontal motions
3. Decomposition of the vertical velocity and kinetic energy in the ECMWF system
a. MODES implementation and validation
Referring to Žagar et al. (2015), the computation of ω within the MODES package required the implementation of the pressure vertical coordinate (Kasahara 1984; Tanaka 1985). The default version of MODES is in the terrain-following σ coordinate, but the evaluation of vertical velocity in the σ system involves the term dependent on the surface pressure which is cumbersome to use. MODES is thus complemented by an option for the pressure system.
The computation of ω in MODES is based on the following steps. First, we compute the meridionally dependent part of Eq. (39),
If the vertical structure functions Gm(p) are constructed analytically then Eq. (36) involves the evaluation of derivatives ∂Gm(p)/∂p as shown by Tanaka and Yatagai (2000). They compared the spherical and Hough harmonics expansions for the computation of ω and showed that the Hough harmonics expansion provides a suitable representation of the pressure vertical velocity.
An example of our retrieval of ω(φ, p) using MODES is shown in Fig. 1 in comparison with the ω field retrieved from the ECMWF ERA5 archive and decomposed along the same latitude using the 1D FFT in Python. The two fields are in very good agreement but not identical for several reasons. There are differences between the model-level pressure at points along the latitude and a constant pressure level used in the MODES expansion. Other factors causing the differences, especially in the lower troposphere and in the polar regions are the vertical and meridional truncations, respectively. For the vertical truncation, M < J always due to a rapid reduction in the values of equivalent depth (Žagar et al. 2009a). More smoothed fields of ω from MODES are expected also due to the truncated input fields. Nevertheless, the reconstruction of the ω field by the new method is very good and certainly sufficiently accurate for the intended decomposition, as confirmed by statistics of root-mean-square differences between the two fields (not shown).
Comparison of the pressure vertical velocity ω from ERA5 (red lines) and ω reconstructed by MODES (blue lines) at pressure level 75.2 hPa, along 60°S at 1200 UTC 11 Aug 2018. ERA5 ω is decomposed using the 1D FFT in Python. (a) Power spectra; (b) ω(λ).
Citation: Journal of the Atmospheric Sciences 80, 11; 10.1175/JAS-D-23-0090.1
The horizontal and vertical structures of regime-decomposed ω are illustrated in Figs. 2 and 3. Figure 2 shows vertical motions illustrative of internal gravity wave above the Antarctic Peninsula that projects onto the westward-propagating inertia–gravity modes. Note that the large-scale part of the IG modes, associated with the gradient wind balance and orography, is filtered out by removing k < 7. The Rossby wave component, ωR, has large scales and two orders of magnitude smaller amplitude. The upward and downward motions are found superimposed on the front and rear side of the large-scale wave in the horizontal circulation, respectively, as expected from the geostrophic wind divergence, ∇ ⋅ Vg = −υgβ/f. The vertical cross section of the retrieved ω over the Antarctic Peninsula in Fig. 3 shows that data corroborate expectations from the decomposition as in Fig. 1. A part of the standing signal over the topography projects on both eastward and westward modes whereas the propagating wave is made of the WIG modes. We do not intend here a detailed examination of relatively high-resolution features over orography, as it would require decomposition at multiple time steps during the day. Features of ω in the lower troposphere are expected to somewhat deviate from ERA5 due to the vertical and horizontal truncation and lower resolution of MODES, similar to Fig. 1.
Pressure vertical velocity ω at level near 75 hPa at 1200 UTC 11 Aug 2018. (a) ω from ERA5 data, (b) ω derived by MODES that is further split into (c) Rossby modes (ROT), (d) inertia–gravity (IG) modes, (e) eastward IG (EIG), and (f) westward IG (WIG). The IG modes in (d)–(f) are filtered for k > 7. Small-scale features of EIG and WIG modes sum up to zero over the Andes between 20° and 35°S. The contours in (c) are the horizontal wind speed (in m s−1). The Rossby mode ω is multiplied by a factor of 100.
Citation: Journal of the Atmospheric Sciences 80, 11; 10.1175/JAS-D-23-0090.1
As in Fig. 2, but the vertical cross section along 70°S across the Antarctic peninsula.
Citation: Journal of the Atmospheric Sciences 80, 11; 10.1175/JAS-D-23-0090.1
The ω decomposition is motivated by the need to quantify vertical momentum fluxes associated with equatorial non-Rossby waves. This requires decomposition of both horizontal and vertical velocities. Having decomposed ω, quantification of the vertical momentum can be performed and it is a subject of follow-on papers. Here we discuss the decomposition of the VKE spectra into the Rossby and non-Rossby modes. The VKE spectra are analyzed for August 2018 using ERA5 data (Hersbach et al. 2020) once per day at 1200 UTC. To remain close to the levels at which dynamical fields were computed, the pressure levels are assigned as the average pressure of the 137 model levels from the definition of the hybrid sigma–pressure coordinate. The horizontal grid is the regular Gaussian grid with 1280 points along the latitude circle with 320 circles between the equator and pole. The numerical truncations are K = 350 zonal wavenumbers, and R = 600 meridional modes including equal numbers of the Rossby and both types of inertia–gravity modes (RR = RE = RW = 200). The number of vertical modes M = 60 suffices to represent most of variance in the lower troposphere while providing an accurate representation of the middle atmosphere.
By definition of the Rossby modes in the linear normal-mode function decomposition, ageostrophic motions in the extratropical troposphere will project on the inertia–gravity modes. In addition, the gradient wind balance in the polar stratosphere will partly project on the IG modes. How are these properties presented in the HKE and VKE spectra at different latitudes and pressure levels? The HKE and VKE is computed along every latitude circle and averaged in bins of 10° latitude width for every level over all days for August 2018. Vertical averaging is carried out following the identification of gross properties of the spectra as discussed in the next section.
b. Horizontal kinetic energy spectra for August 2018
The horizontal kinetic energy spectra EH,
The horizontal kinetic energy (HKE) spectra averaged over latitude belts and altitude bands for the ERA5 data in August 2018. Dashed lines show theoretical slopes with power laws −3, −5/3, −1, and 0, with 0 denoting the white spectrum. Dotted lines are empirical fit for average energy spectrum between 200 and 400 hPa for zonal wavenumbers k = 20–80. (left) Total HKE; (center) the Rossby part,
Citation: Journal of the Atmospheric Sciences 80, 11; 10.1175/JAS-D-23-0090.1
The Rossby and non-Rossby spectra in Fig. 4 show the extent to which the ERA5 data on average deviate from the theoretical power laws for the inertial range in different latitude belts and altitudes. In particular, Fig. 4 is informative about differences between the upper troposphere and the stratospheric levels which on average have large weight in the 3D spectra due to associated equivalent depths [Eq. (42)]. For this reason, and because of the presence of the available potential energy in the 3D spectrum Ik, the non-Rossby total 3D energy spectrum for August 2018 (not shown) does not show a prominent energy maximum at synoptic scales, although such maxima are clearly seen in daily energy spectra of the ECMWF model analyses and forecasts on the MODES web page.1
In addition to lines of the theoretical power laws (dashed lines), Fig. 4 contains empirical fits (as dotted lines) of the spectra averaged over model levels between 200 and 400 hPa for k = 20–80. Even though individual samples may largely deviate from the theoretical power laws (Dörnbrack et al. 2022), it is informative to compare slopes of average spectra with theoretical expectations. Overall, the HKE spectra in the upper troposphere suggest an insufficient Rossby wave variance in the extratropical SH and a lack of the non-Rossby wave variance at all latitudes in ERA5 in August 2018. On the other hand, the upper troposphere in the tropics and in the subtropical summer hemisphere (Figs. 4b,c) appear more balanced, especially the deep tropics. This is seen as a shallower than −3 slope of the HKE spectra at 20 < k < 200 in Fig. 4c2. On the other hand, the steepest non-Rossby spectrum is found in the subtropical upper troposphere (Fig. 4b3). At all latitudes, stratospheric subsynoptic
c. Vertical kinetic energy spectra for August 2018
Now we discuss the VKE spectra averaged over the same latitude belts as the HKE spectra but for a few more vertical layers to show gradual changes in vertical velocities in ERA5 data. Many physical processes and numerical effects in the model are likely to manifest in the structure of the VKE spectra in the tropics and the middle and high latitudes as discussed in what follows.
1) Extratropics
The VKE spectra for middle latitudes are shown in Fig. 5 and can be compared with the high-latitude spectra in Fig. 6 and subtropical belts in Fig. 7. The first notable feature is the redness of the Rossby spectra,
Kinetic energy spectra of vertical motions (VKE) per unit mass in the midlatitude belt between (a),(c) 30° and 60°S and (b),(d) 30° and 60°N, averaged for August 2018. (a),(b) Rossby mode VKE and (c),(d) IG modes VKE. The bottom spectrum (thick blue line) in each panel is the average over levels between 10 and 30 hPa, and the spectra above belong to layers lower in the atmosphere as defined in the legend. Black lines show power laws discussed in section 2.
Citation: Journal of the Atmospheric Sciences 80, 11; 10.1175/JAS-D-23-0090.1
As in Fig. 5, but for the latitude belt (a),(c) 60°–80°S and (b),(d) 60°–80°N.
Citation: Journal of the Atmospheric Sciences 80, 11; 10.1175/JAS-D-23-0090.1
As in Fig. 5, but for the latitude belt (a),(c) 10°–30°S and (b),(d) 10°–30°N.
Citation: Journal of the Atmospheric Sciences 80, 11; 10.1175/JAS-D-23-0090.1
The IG VKE spectra vary more distinctly than the Rossby spectra not only vertically but also latitudinally (summer versus winter of August 2018). First, note a remarkable maximum in the VKE spectra in Fig. 5c at k = 3 in the SH midlatitude stratosphere shifting to synoptic scales k = 7–8 in the troposphere where baroclinic Rossby waves dynamics creates intense ageostrophic circulation in the winter and Rossby waves penetrate the stratosphere. The maximum is present also in the SH high-latitude lower stratosphere (Fig. 6c) and SH subtropical troposphere (Fig. 7c), but absent in the high-latitude troposphere (Figs. 6c,d) and in the subtropical stratosphere (Fig. 7c), because Rossby wave activity is either weak or waves do not propagate vertically. Similar synoptic-scale peaks in VKE spectra are seen in spherical harmonics decomposition of simulated vertical velocity by kilometer-scale models, both hydrostatic (Polichtchouk et al. 2022) and nonhydrostatic (Skamarock et al. 2014; Morfa and Stephan 2023) including a shift to larger scales in the stratosphere.
The tropospheric IG VKE spectra have slopes between −1 and −1/3 for a range of synoptic and subsynoptic scales 10 < k < 100 in both hemispheres. In agreement with more shallow
Finally, we find the expected positive slope of the VKE spectra associated with gravity waves [Eq. (45)] in the SH stratosphere in midlatitudes (Fig. 5c). This “ideal” gravity wave VKE spectrum with a slope close to 1/3 starts at about k = 10 and extends toward k = 60. For comparison, the
2) Tropics
The presentation of the tropical VKE spectra in Fig. 8 is somewhat different from the extratropical case because of the presence of the two special equatorial modes, the Kelvin and MRG wave. The non-Rossby spectra (the sum of the IG, Kelvin and MRG modes, Fig. 8a) are almost indistinguishable from the total spectra (not shown) but also from the IG spectra (Fig. 8c). This is because the VKE spectra for the Kelvin and MRG waves (Fig. 9) have one or two orders of magnitude less VKE than the IG modes at every k.
As in Fig. 5, but for the tropical belt between 10°S and 10°N. (a) non-Rossby (IG, MRG, and Kelvin) modes, (b) Rossby modes, (c) IG modes, (d) EIG, and WIG modes.
Citation: Journal of the Atmospheric Sciences 80, 11; 10.1175/JAS-D-23-0090.1
As in Fig. 8, but for the (a) Kelvin waves and (b) MRG waves.
Citation: Journal of the Atmospheric Sciences 80, 11; 10.1175/JAS-D-23-0090.1
The Rossby spectra,
The
In the stratosphere and within the tropical tropopause layer, the westward IG modes dominate at planetary scales k = 1–2 (spectra drawn with dashed lines in Fig. 8d) and are significantly smaller at other wavenumbers. This can be related to the quasi-biennial oscillation (QBO) in August 2018 being in its easterly phase and the horizontal eigenstructures derived with respect to the state of rest. The role of the phase of the QBO on the VKE spectra and the vertical momentum fluxes is a subject of future work.
The VKE spectra for the Kelvin wave and the MRG wave are shown in Fig. 9. For the Kelvin wave, the VKE increases going from 10 to 300 hPa, and there is little change of VKE below this level. Overall, the Kelvin wave VKE distribution is similar to that of IG modes except that amplitudes are smaller and the spectra at subsynoptic scales are steeper and noisy. On the other hand, the MRG VKE spectra are similar to those for the Rossby modes. Both Kelvin and MRG spectra approximately follow the predicted sloped of −1 and −5, respectively, over a range of scales between k = 10 and k = 100. Even if the VKEs for the IG modes and the Kelvin waves were of similar amplitudes, a difference in the slope alone accounts for a significant difference in vertical velocity and thus momentum fluxes. This can be illustrated by a single zonal wavenumber, for example, k = 20. The ratio between VKEs for power laws −1 (Kelvin wave) and −1/3 (IG modes) gives about 14% of
4. Conclusions and outlook
We derived expressions for vertical velocity associated with the Rossby–Haurwitz waves and inertia–gravity (IG), Kelvin, and mixed Rossby–gravity waves for a hydrostatic atmosphere. The applied linear decomposition, based on the normal-mode functions, projects ageostrophic motions on the inertia–gravity modes. As a result, the IG modes contain both high-frequency signals and low-frequency components that result from the balance between the linear and nonlinear terms of the prognostic equations (e.g., Errico 1984; Warn and Menard 1986; Ko et al. 1989; Tribbia 2020). A further differentiation between the linear and nonlinear balance, which would decompose IG modes into ageostrophic component coupled with the quasigeostrophic dynamics and gravity waves, is beyond the scope of the present work, and will be the subject of future work.
The new decomposition method provides the vertical velocity along latitude circles at pressure levels. It is implemented in the MODES software (Žagar et al. 2015) and applied to the ERA5 data. The spectra of the kinetic energy of vertical velocity (VKE) are computed for August 2018 and discussed in relation with the KE spectra of the horizontal motions (HKE spectra) for the same latitude belts. The intended application to longer datasets (e.g., the complete ERA5 period) will provide further insight and quantify vertical velocities and associated momentum fluxes due to the Kelvin, MRG and IG modes in driving low-frequency variability such as the QBO. As an example, our method enables decomposition of the vertical momentum fluxes in the gray zone of the zonal wavenumbers k ≈ 7–40 where unbalanced dynamics gradually grows comparable and eventually exceeds in amplitude balanced dynamics.
The analytically derived expression coupling the limit VKE spectrum, EV, and the total mechanical energy (kinetic plus available potential energy) spectrum, TE, for every modal component, states that EV ∝ ν2TE, where ν is the normal mode frequency. Invoking frequency relationships for the Rossby and inertia–gravity modes and the k−3 and k−5/3 power laws for the EH within the inertial range, the VKE spectra of the Rossby–Haurwitz and inertia–gravity waves follow k−5 and k1/3 power laws, respectively. At the planetary scales in the tropics, which constitute much of the divergent circulation, both Rossby and IG VKE spectra are nearly white. The Kelvin and mixed Rossby–gravity waves are expected to follow k−1 and k−5 power laws, respectively. However, the linear IG modes include also ageostrophic circulation, which is coupled with the divergence and expected to follow a k−1 power law, and the level to which the internal gravity waves make the observed k−5/3 power law of the horizontal kinetic energy depends on the latitude (season), altitude and, maybe most of all, model characteristics.
The derived VKE spectra for August 2018 ERA5 data approximately follow the expected −5 slope for the Rossby modes up to k ≈ 150 depending on the latitude belt, altitude, and season. The MRG wave VKE spectra are similar to the Rossby spectra. The VKE spectra associated with the IG modes reveal a synoptic-scale VKE peak in the winter hemisphere (SH August) associated with quasigeostrophic ageostrophic motions. The peak moves to planetary scales going from the upper troposphere toward the upper stratosphere due to a strong attenuation of the Rossby waves with altitude (Charney and Drazin 1961). A part of the midlatitude winter stratosphere VKE spectra between k ≈ 10 and k ≈ 60 follows the “true” gravity wave spectrum with a power law of k1/3. The spectrum becomes shallower near the tropopause and takes slopes from −1 to −2/3 near 200 hPa level and lower in the troposphere. More shallow IG spectra are found in the NH hemisphere troposphere (summer season) and in the tropics, with slopes closer to −2/3 and −1/3, respectively, for 10 < k < 200. The slopes of VKE spectra are in a qualitative agreement with the HKE spectra that are overall steeper than expected for both Rossby and inertia–gravity wave regime. An exception is the Rossby wave horizontal kinetic energy spectrum in the summer hemisphere tropics that is significantly shallower than −3. The Kelvin waves VKE spectra follow the expected −1 power law implying a significantly smaller amplitude of their vertical velocities compared to the IG modes in the tropics.
The derived 1/3 limit spectra for the VKE in hydrostatic atmosphere suggest that the VKE spectra with slopes 2/3 or even greater in kilometer-scale models (Morfa and Stephan 2023) are either due to nonhydrostatic processes or particular modeling solutions. The latter is supported by similar VKE spectra in the hydrostatic ECMWF model (Polichtchouk et al. 2022) and by an apparent lack of resolution convergence in both hydrostatic and nonhydrostatic models (Skamarock et al. 2014; Polichtchouk et al. 2022). Using the new method, the hydrostatic component of vertical velocity in nonhydrostatic models can be computed to identify nonhydrostatic effects on the VKE spectra and momentum fluxes. This additionally provides a basis for an improved strategy for evaluating kilometer-scale models, i.e., a subject for further research.
Acknowledgments.
This paper is a contribution to the Collaborative Research Centre TRR 181 “Energy Transfers in Atmosphere and Ocean” funded by the Deutsche Forschungsgemeinschaft (DFG; German Research Foundation), Project 274762653. Ž. Zaplotnik was supported by the Slovenian Research Agency (ARRS), Grant J1-9431 and Program P1-0188. We thank Andreas Dörnbrack, Inna Polichtchouk, and Frank Lunkeit for the discussions, and them and Richard Blender for reading the paper. We are very grateful for the insightful comments by four reviewers.
Data availability statement.
The ERA5 data are available from the Copernicus Programme, via https://climate.copernicus.eu/climate-reanalysis. The default version of the MODES software is available via http://modes.cen.uni-hamburg.de. Outputs of the new decomposition are available on request.
APPENDIX A
Dispersion Relationship for the Rossby and Inertia–Gravity Waves on the Sphere
Computing frequencies of Rossby and gravity waves on the sphere amounts to finding eigenvalues of a symmetric pentadiagonal matrix (Swarztrauber 1984). Solving this problem in full generality is beyond reach of currently available analytical methods. Nevertheless, useful scaling laws and asymptotics can be obtained numerically.
To this end we compute frequencies of the waves for equivalent depths representative of Earth’s atmosphere using MODES software at T170 resolution and compare them to the limit γ → ∞ frequencies (22) and (23) for the Rossby and gravity waves, respectively (Fig. A1). We observe that (22) provides an excellent approximation for IG dispersion relationship on the sphere for all scales and equivalent depths. We omitted plotting the curves for the west-propagating IG modes in the right panel of Fig. A1 as for n > 0 they are nearly identical to those of EIG with exception of sign reversal.
Dispersion curves of the (left) Rossby and (right) inertia–gravity modes on the sphere for different equivalent depths and meridional modes compared to their respective γ → ∞ limit curves (23) and (22). Dotted lines in the left panel correspond to the fitted Rossby dispersion curves (A1).
Citation: Journal of the Atmospheric Sciences 80, 11; 10.1175/JAS-D-23-0090.1
The picture is more complicated for the Rossby modes. First, limit D → ∞ Rossby–Haurwitz (RH) frequencies (23) match those of barotropic Rossby waves (D ≈ 10 km) outside of large-scale region (n < 5, k ≤ 5). Second, these limit RH dispersion curves approximate well those of small-scale Rossby waves (κ → ∞) regardless of equivalent depth. The accuracy of approximation improves with increase of equivalent depth and decrease of horizontal scale so that for D > 200 m the match is good throughout atmospheric mesoscales.
Dependence of nondimensional Rossby mode frequencies on the sphere on the (left) zonal and (right) meridional scales for D = 10 km, 200 m, and 10 m.
Citation: Journal of the Atmospheric Sciences 80, 11; 10.1175/JAS-D-23-0090.1
The dependence on the meridional scale is markedly different from the ones expected from both (23) and β-plane dispersion relationships with no universal limits at both ends of the spectrum (Fig. A2, right panel). For small meridional scales, the slopes fall into the [−2, −1] range, depending on zonal wavenumber and equivalent depth, with large D producing n−2 curves. The slope gets progressively shallower as equivalent depth and zonal length scale decrease. Less can be concluded for the planetary scales, where the asymptotics is the function of k and D, with nonmonotone dependence on both.
APPENDIX B
Derivation of the Limit Spectra for the Vertical Kinetic Energy
Vertical kinetic energy (VKE) for a random date in August 2018 along latitude circle 15°N at 171 hPa level. The Rossby (denoted R) and non-Rossby (denoted nR) VKE (full lines) is compared to three other quadratic quantities involved in the derivation of the limit VKE spectra (B7). All dashed curves are scaled so that their values at the zonal wavenumber k = 1 match that of the VKE. The constant parameter c is thus different for each curve.
Citation: Journal of the Atmospheric Sciences 80, 11; 10.1175/JAS-D-23-0090.1
Vertical kinetic energy and divergence spectra
Comparing (B9) with (B14), we observe that the former approach provides a much sharper bound on the Rossby VKE spectral slope than estimating it from divergence. This is because the dispersion relation for Rossby waves involved in (B7) is for the barotropic Rossby–Haurwitz waves with divergence due to the beta term. Quasigeostrophic turbulence with IS ∝ k−3 involves ageostrophic motions that cause divergence but in linear decomposition project on the IG modes. On the other hand, both approaches yield identical limit IG spectra.
REFERENCES
Bacmeister, J. T., S. D. Eckermann, P. A. Newman, L. Lait, K. R. Chan, M. Loewenstein, M. H. Proffitt, and B. L. Gary, 1996: Stratospheric horizontal wavenumber spectra of winds, potential temperature, and atmospheric tracers observed by high-altitude aircraft. J. Geophys. Res., 101, 9441–9470, https://doi.org/10.1029/95JD03835.
Bergman, J. W., and M. L. Salby, 1994: Equatorial wave activity derived from fluctuations in observed convection. J. Atmos. Sci., 51, 3791–3806, https://doi.org/10.1175/1520-0469(1994)051<3791:EWADFF>2.0.CO;2.
Boyd, J. P., and C. Zhou, 2008: Kelvin waves in the nonlinear shallow water equations on the sphere: Nonlinear travelling waves and the corner wave bifurcation. J. Fluid Mech., 617, 187–205, https://doi.org/10.1017/S0022112008003959.
Charney, J. G., and P. G. Drazin, 1961: Propagation of planetary-scale disturbances from the lower into the upper atmosphere. J. Geophys. Res., 66, 83–109, https://doi.org/10.1029/JZ066i001p00083.
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.
Craig, G. C., and T. Selz, 2018: Mesoscale dynamical regimes in the midlatitudes. Geophys. Res. Lett., 45, 410–417, https://doi.org/10.1002/2017GL076174.
Dörnbrack, A., and Coauthors, 2018: Gravity waves excited during a minor sudden stratospheric warming. Atmos. Chem. Phys., 18, 12 915–12 931, https://doi.org/10.5194/acp-18-12915-2018.
Dörnbrack, A., P. Bechtold, and U. Schumann, 2022: High-resolution aircraft observations of turbulence and waves in the free atmosphere and comparison with global model predictions. J. Geophys. Res. Atmos., 127, e2022JD036654, https://doi.org/https://doi.org/10.1029/2022JD036654.
Dueben, P. D., N. Wedi, S. Saarinen, and C. Zeman, 2020: Global simulations of the atmosphere at 1.45 km grid-spacing with the Integrated Forecasting System. J. Meteor. Soc. Japan, 98, 551–572, https://doi.org/10.2151/jmsj.2020-016.
Errico, R. M., 1984: The dynamical balance of a general circulation model. Mon. Wea. Rev., 112, 2439–2454, https://doi.org/10.1175/1520-0493(1984)112<2439:TDBOAG>2.0.CO;2.
Fritts, D. C., and G. D. Nastrom, 1992: Sources of mesoscale variability of gravity waves. Part II: Frontal, convective, and jet stream excitation. J. Atmos. Sci., 49, 111–127, https://doi.org/10.1175/1520-0469(1992)049<0111:SOMVOG>2.0.CO;2.
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.
Gill, A. E., 1982: Atmosphere–Ocean Dynamics. Academic Press, 662 pp.
Haurwitz, B., 1940: The motion of atmospheric disturbances. J. Mar. Res., 3, 35–50.
Hersbach, H., and Coauthors, 2020: The ERA5 global reanalysis. Quart. J. Roy. Meteor. Soc., 146, 1999–2049, https://doi.org/10.1002/qj.3803.
Hoskins, B. J., I. Draghici, and H. C. Davies, 1978: A new look at the ω-equation. Quart. J. Roy. Meteor. Soc., 104, 31–38, https://doi.org/10.1002/qj.49710443903.
Hough, S. S., 1898: On the application of harmonic analysis to the dynamical theory of the tides.—Part II. On the general integration of Laplace’s dynamical equations. Philos. Trans. Roy. Soc., A191, 139–185, https://doi.org/10.1098/rsta.1898.0005.
Kasahara, A., 1978: Further studies on a spectral model of the global barotropic primitive equations with Hough harmonic expansions. J. Atmos. Sci., 35, 2043–2051, https://doi.org/10.1175/1520-0469(1978)035<2043:FSOASM>2.0.CO;2.
Kasahara, A., 1984: The linear response of a stratified global atmosphere to a tropical thermal forcing. J. Atmos. Sci., 41, 2217–2237, https://doi.org/10.1175/1520-0469(1984)041<2217:TLROAS>2.0.CO;2.
Kasahara, A., 2020: 3D normal mode functions (NMFs) of a global baroclinic atmospheric model. Modal View of Atmospheric Variability: Applications of Normal-Mode Function Decomposition in Weather and Climate Research, N. Žagar and J. Tribbia, Eds., Mathematics of Planet Earth Series, Vol. 8, Springer, 1–62.
Knippertz, P., and Coauthors, 2022: The intricacies of identifying equatorial waves. Quart. J. Roy. Meteor. Soc., 148, 2814–2852, https://doi.org/10.1002/qj.4338.
Ko, S. D., J. J. Tribbia, and J. P. Boyd, 1989: Energetics analysis of a multilevel global spectral model. Part I: Balanced energy and transient energy. Mon. Wea. Rev., 117, 1941–1953, https://doi.org/10.1175/1520-0493(1989)117<1941:EAOAMG>2.0.CO;2.
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.
Liu, H.-L., 2019: Quantifying gravity wave forcing using scale invariance. Nat. Commun., 10, 2605, https://doi.org/10.1038/s41467-019-10527-z.
Longuet-Higgins, M. S., 1968: The eigenfunctions of Laplace’s tidal equations over a sphere. Philos. Trans. Roy. Soc., A262, 511–607, https://doi.org/10.1098/rsta.1968.0003.
Matsuno, T., 1966: Quasi-geostrophic motions in the equatorial area. J. Meteor. Soc. Japan, 44, 25–43, https://doi.org/10.2151/jmsj1965.44.1_25.
Morfa, Y. A., and C. C. Stephan, 2023: The relationship between horizontal and vertical velocity wavenumber spectra in global storm-resolving simulations. J. Atmos. Sci., 80, 1087–1105, https://doi.org/10.1175/JAS-D-22-0105.1.
Müller, S. K., E. Manzini, M. Giorgetta, K. Sato, and T. Nasuno, 2018: Convectively generated gravity waves in high resolution models of tropical dynamics. J. Adv. Model. Earth Syst., 10, 2564–2588, https://doi.org/10.1029/2018MS001390.
Nappo, C. J., 2002: An Introduction to Atmospheric Gravity Waves. International Geophysics Series, Vol. 85, Academic Press, 276 pp.
Nastrom, G. D., and K. S. Gage, 1985: A climatology of atmospheric wavenumber spectra of wind and temperature 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.
Paldor, N., I. Fouxon, O. Shamir, and C. I. Garfinkel, 2018: The mixed Rossby–gravity wave on the spherical Earth. Quart. J. Roy. Meteor. Soc., 144, 1820–1830, https://doi.org/10.1002/qj.3354.
Polichtchouk, I., N. Wedi, and Y.-H. Kim, 2022: Resolved gravity waves in the tropical stratosphere: Impact of horizontal resolution and deep convection parametrization. Quart. J. Roy. Meteor. Soc., 148, 233–251, https://doi.org/10.1002/qj.4202.
Rodda, C., and U. Harlander, 2020: Transition from geostrophic flows to inertia–gravity waves in the spectrum of a differentially heated rotating annulus experiment. J. Atmos. Sci., 77, 2793–2806, https://doi.org/10.1175/JAS-D-20-0033.1.
Schumann, U., 2019: The horizontal spectrum of vertical velocities near the tropopause from global to gravity wave scales. J. Atmos. Sci., 76, 3847–3862, https://doi.org/10.1175/JAS-D-19-0160.1.
Simmons, A. J., and D. M. Burridge, 1981: An energy and angular-momentum conserving vertical finite-difference scheme and hybrid vertical coordinates. Mon. Wea. Rev., 109, 758–766, https://doi.org/10.1175/1520-0493(1981)109<0758:AEAAMC>2.0.CO;2.
Skamarock, W. C., S.-H. Park, J. B. Klemp, and C. Snyder, 2014: Atmospheric kinetic energy spectra from global high-resolution nonhydrostatic simulation. J. Atmos. Sci., 71, 4369–4381, https://doi.org/10.1175/JAS-D-14-0114.1.
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.
Stepanyuk, O., J. Räisänen, V. A. Sinclair, and H. Järvinen, 2017: Factors affecting atmospheric vertical motions as analyzed with a generalized omega equation and the OpenIFS model. Tellus, 69A, 1271563, https://doi.org/10.1080/16000870.2016.1271563.
Stephan, C. C., N. Žagar, and T. G. Shepherd, 2021: Waves and coherent flows in the tropical atmosphere: New opportunities, old challenges. Quart. J. Roy. Meteor. Soc., 147, 2597–2624, https://doi.org/10.1002/qj.4109.
Swarztrauber, P. N., 1984: FFT algorithms for vector computers. Parallel Comput., 1, 45–63, https://doi.org/10.1016/S0167-8191(84)90413-7.
Swarztrauber, P. N., and A. Kasahara, 1985: The vector harmonic analysis of Laplace’s tidal equations. SIAM J. Sci. Stat. Comput., 6, 464–491, https://doi.org/10.1137/0906033.
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.
Tanaka, H., and A. Yatagai, 2000: Comparative study of vertical motions in the global atmosphere evaluated by various kinematic schemes. J. Meteor. Soc. Japan, 78, 289–298, https://doi.org/10.2151/jmsj1965.78.3_289.
Tanaka, H., and N. Žagar, 2020: 3D modal variability and energy transformations on the sphere. Modal View of Atmospheric Variability: Applications of Normal-Mode Function Decomposition in Weather and Climate Research, N. Žagar and J. Tribbia, Eds., Mathematics of Planet Earth Series, Vol. 8, Springer, 121–179.
Taylor, G. I., 1936: The oscillations of the atmosphere. Proc. Roy. Soc. London, 156A, 318–326, https://doi.org/10.1098/rspa.1936.0150.
Terasaki, K., H. L. Tanaka, and M. Satoh, 2009: Characteristics of the kinetic energy spectrum of NICAM model atmosphere. SOLA, 5, 180–183, https://doi.org/10.2151/sola.2009-046.
Tribbia, J., 2020: Normal mode functions and initialization. Modal View of Atmospheric Variability: Applications of Normal-Mode Function Decomposition in Weather and Climate Research, N. Žagar and J. Tribbia, Eds., Mathematics of Planet Earth Series, Vol. 8, Springer, 63–78.
Warn, T., and R. Menard, 1986: Nonlinear balance and gravity-inertial wave saturation in a simple atmospheric model. Tellus, 38A, 285–294, https://doi.org/10.3402/tellusa.v38i4.11719.
Žagar, N., J. Tribbia, J. L. Anderson, and K. Raeder, 2009a:Uncertainties of estimates of inertia–gravity energy in the atmosphere. Part I: Intercomparison of four analysis systems. Mon. Wea. Rev., 137, 3837–3857, https://doi.org/10.1175/2009MWR2815.1; Corrigendum, 138, 2476–2477, https://doi.org/10.1175/2010MWR3256.1.
Žagar, N., J. Tribbia, J. L. Anderson, and K. Raeder, 2009b: Uncertainties of estimates of inertia–gravity energy in the atmosphere. Part II: Large-scale equatorial waves. Mon. Wea. Rev., 137, 3858–3873, https://doi.org/10.1175/2009MWR2816.1; Corrigendum, 138, 2476–2477, https://doi.org/10.1175/2010MWR3256.1.
Žagar, N., A. Kasahara, K. Terasaki, J. Tribbia, and H. Tanaka, 2015: Normal-mode function representation of global 3D data sets: Open-access software for the atmospheric research community. Geosci. Model Dev., 8, 1169–1195, https://doi.org/10.5194/gmd-8-1169-2015.
Žagar, N., D. Jelić, M. Blaauw, and P. Bechtold, 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.
Žagar, N., F. Lunkeit, F. Sielmann, and W. Xiao, 2022: Three-dimensional structure of the equatorial Kelvin wave: Vertical structure functions, equivalent depths, and frequency and wavenumber spectra. J. Climate, 35, 2209–2230, https://doi.org/10.1175/JCLI-D-21-0342.1.