1. Introduction
Atmospheric waves play a central role in the middle atmospheric circulation. Planetary-scale Rossby waves (PWs) arising from large-scale topography and the land–sea thermal contrast predominate stratospheric phenomena, whereas in the upper stratosphere and mesosphere, smaller-scale gravity waves (GWs) forced by small-scale topography, convection, and jet/front systems play a key role (Achatz et al. 2024; Andrews et al. 1987; Kim et al. 2003). While the relative importance of these waves has long been of particular interest (Butchart et al. 2011; Kang et al. 2020), the interactions between them and their consequent impacts are complex and, therefore, not easily examined. However, as cautioned by Cohen et al. (2013), analyzing the individual impacts of PWs and GWs without considering their mutual interactions could result in misunderstandings of their roles and the associated middle atmospheric phenomena.
By inducing horizontal variations in large-scale flow, PWs cause selective transmission of GWs into the middle atmosphere and localized GW breaking (Dunkerton and Butchart 1984; Schoeberl and Strobel 1984). The transformed Eulerian-mean equations (Andrews et al. 1987) demonstrate that PWs can also influence GWs by modifying the zonal-mean flow. A well-documented example is the onset of sudden stratospheric warming (SSW) caused by PWs and subsequent critical-level filtering of westward-propagating GWs (Matsuno 1971). This mechanism operates in the same manner when GWs exert impact on PWs. GWs can modulate the propagation and dissipation of PWs by changing the zonal-mean flow (Cohen et al. 2014; Yoo and Chun 2023) or can even cause in situ PW generation by triggering barotropic/baroclinic (BT/BC) instability (Cohen et al. 2013; Sato and Nomoto 2015).
The direct impact of GWs on PWs is also worth noting, but this mechanism has received relatively little attention. The excitation and selective transmission of GWs collectively lead to longitudinal variations in GW drag (GWD). It can, in turn, generate PWs by acting as a nonconservative source for linearized disturbance quasigeostrophic potential vorticity (QGPV; Andrews et al. 1987). Holton (1984) first examined this mechanism in the context of a semispectral model implementing an orographically induced GW (OGW) parameterization. In the absence of preexisting PWs during the Northern Hemisphere (NH) winter, OGWs originating from the zonally asymmetric topography excited PWs in the mesosphere due to the localized nature of OGW drag (OGWD). Taking PWs forced from the lower boundary source into account, McLandress and McFarlane (1993, MM93) extended Holton’s work using the quasi-linear QGPV model with the OGWD parameterization of McFarlane (1987). By conducting sensitivity tests comparing scenarios with and without nonconservative GWD forcing, they verified the role of localized OGWD in generating PWs. The resultant PWs varied with the phase relation between the OGWD-induced PWs and PWs forced from the lower boundary source. Despite substantial changes in the individual forcing of OGWs and the resultant PWs, the total forcing remains nearly constant—an interesting phenomenon known as compensation, which has become a topic of study for several authors nearly 20 years after MM93 (e.g., Cohen et al. 2013, 2014; McLandress et al. 2012; Sigmond and Shepherd 2014; Yoo and Chun 2023).
Smith (1996) first identified the actual wind disturbances observed in high-resolution Doppler images as attributable to nonzonal GWD in the NH winter mesosphere. In this case, the selective filtering of the GW spectrum by underlying stratospheric PWs causes zonal asymmetries in GWD, and thus, the associated quasi-stationary PWs (SPWs) in the upper mesosphere are negatively correlated with those in the stratosphere. By simulating these waves with a fully nonlinear three-dimensional general circulation model (GCM), Smith (2003) demonstrated that the GWD-associated PWs tend to dominate the upward-propagating Rossby waves above the zero zonal wind. Lieberman et al. (2013) provided supportive satellite observations for these SPWs, while they explained the underlying mechanism in relation to the ageostrophic flow induced by the zonal disturbances of GWD.
After theoretical studies, the impact of the GWD-induced PWs has become recognized. Following the onset of SSW events, selectively filtered GWs excite PWs in the mesosphere and lower thermosphere (MLT), thereby contributing to variability (Liu and Roble 2002) and temperature changes in the MLT by modifying the meridional circulation (Yamashita et al. 2010). Song et al. (2020) suggested that the strong amplitude PWs of zonal wavenumber (ZWN) 2 during the SSW occurring in 2009 were partially related to the downward-propagating PWs generated in situ from zonally asymmetric GWD in the lower mesosphere. By examining the unique latitudinal double peaks of SPWs in the austral winter, Lu et al. (2018) demonstrated the contribution of the GWD-induced PWs in the NH to the secondary peak through interhemispheric propagation. Meanwhile, Šácha et al. (2016) focused on the influence of strong and localized stratospheric GWD, attributable to the orographic GWs, on the formation of PWs, Brewer–Dobson circulation, and SSW based on a mechanistic model. Following Šácha et al. (2016), Samtleben et al. (2019, 2020a,b) investigated the influence of localized OGWD on SPWs, depending on the longitudinal and latitudinal positions of OGWD and the number of localized GWD regions, based on the primitive equation models.
Although theoretical, numerical, and observational studies have identified spatially varying GWD as an in situ source of PWs, a fundamental question remains unanswered: How do the properties of the in situ excited PWs vary with the GWD configuration? While MM93 identified one attribute—the phase of zonal asymmetries in GWD with respect to PWs forced from below—the influence of varying GWD asymmetries, including magnitude and ZWN of asymmetries, has not been fully investigated. In addition, previous studies have primarily examined steady-state PWs, attained through sufficient interactions among GWs, PWs, and the zonal-mean flow. In such cases, the simultaneous occurrence of GW–PW–mean-flow interactions is likely, thereby modifying the properties of the GWD-originated PWs. Focusing on the steady state also limits exploring the evolution of PWs originating from the nonzonal GWD, a factor identified as important in determining the characteristics of the GWD-induced PWs in this study.
To bridge this gap in understanding and gain a more comprehensive insight into the mechanism by which spatial variations in GWD excite PWs, we revisit this issue using a fully nonlinear primitive equation GCM extended into the lower thermosphere, named the System for Whole Atmosphere Dynamics (SWAD) Research (Song 2023). Unlike the simple linearized QG models providing a clear causality of this mechanism in prior studies, primitive equation GCMs can simulate nonlinear wave–mean-flow interaction on the globe without a limitation in the latitudinal width at which the QG assumption is valid. However, the use of primitive equation models makes it challenging to separate the individual mechanisms of the generation, propagation, and superposition of PWs and wave–mean-flow interaction, particularly when the processes are correlated and feedback on each other. Therefore, we establish a highly idealized but efficient framework with SWAD capable of isolating in situ PW generation due to the zonally asymmetric GWD processes. In this idealized GCM (IGCM) framework, potential PW sources other than GWD are excluded by incorporating a baroclinic adjustment scheme to establish a zonal-mean field that excludes instability-related PWs (BT/BC instability in this study) and a flat-bottom setup to eliminate stationary PW sources. In addition, by employing a mechanical relaxation approach referred to as nudging (Hitchcock and Haynes 2014), the zonal-mean state of the entire atmosphere is constrained without changes, thereby preventing its interaction with waves and the emergence of instability throughout integration. Since nudging is applied only to the zonal-mean component of the model, the zonally asymmetric component evolves unconstrained. This approach enables us to examine PWs exclusively generated by the zonally varying GWD in a realistic NH winter atmosphere without requiring a long integration.
The details of the idealized framework and a demonstration of its attainment for the intended purpose of this study are presented in section 2. In section 3, we explore the characteristics of PWs exclusively generated by the zonally asymmetric GWD, their temporal evolution, and their dependence on GWD configuration. Section 4 compares the results in this study with those in previous studies, and section 5 provides a summary of our findings.
2. A modeling framework and analysis method
a. Model description and configuration
The present study employs SWAD, which is developed for exploring the atmospheric dynamical processes from the troposphere to the lower thermosphere. This modeling system adopts the semi-implicit, semi-Lagrangian dynamical core (Williamson and Olson 1994) implemented in the Community Atmosphere Model, version 3 (CAM3; Collins et al. 2004), of the National Centre for Atmospheric Research (NCAR). The spherical harmonic (SPH) expansion, on which the semi-implicit, semi-Lagrangian dynamical core is based, allows for preserving the zonal-mean field while enabling the wave components to evolve freely. This model solves the dry and hydrostatic primitive equations in a vorticity–divergence form in the hybrid vertical coordinate using the SPH transform method for horizontal discretization, a two-time-level semi-implicit scheme for time integration, and a semi-Lagrangian transport scheme for advection (see section 3.2 of the CAM3 documentation for details). The horizontal resolution is T42, which can resolve up to ZWN 42 on the Gaussian latitudinal grid at an approximate 2.8° interval, and 66 vertical levels are employed from the surface to an altitude of approximately 130 km. The modeling system is driven by radiative heating and cooling computed by relaxing temperature toward a Held–Suarez-type radiative equilibrium temperature which is extended into the lower thermosphere (Held and Suarez 1994; Polvani and Kushner 2002). This model incorporates a GWD parameterization identical to that used in NASA’s GEOS-5 and GEOS-6 models (Molod et al. 2012, 2015). Vertical diffusion processes are also implemented for the planetary boundary layer (Held and Suarez 1994) below about 700 hPa and molecular diffusion (Banks and Kockarts 1973) above 0.5 hPa. Perpetual January and July runs exhibit the typical structure of the zonal-mean fields and meridional circulations. Moreover, middle atmospheric variabilities, such as SSWs and quasi-biennial oscillations, are simulated reasonably well, although SSW frequency and periods of quasi-biennial oscillations depend on some tunable factors in the upper troposphere of the winter polar and equatorial regions, respectively (Song 2023).
The primary aim of this study is to investigate the characteristics of PWs generated by zonally asymmetric GWD in the absence of any other sources. To eliminate major stationary sources of PWs (e.g., topography and land–sea thermal contrast), the lower boundary of the modeling system is set as a flat bottom with zero geopotential height. In addition, during integration, another possible in situ source, BT/BC instability, is removed from the zonal-mean state of the entire atmosphere by employing a nudging approach elaborated on in the subsequent section. In an idealized framework, henceforth, the aforementioned physical processes (radiation scheme, GWD, and vertical diffusion parameterizations) are disabled in our simulations. To avoid potentially harmful impacts from the top boundary, the analysis is restricted to the middle atmosphere below an altitude of 90 km. In each simulation, the model is integrated for 30 days with a time step of 600 s and daily averaged outputs are used for analysis.
b. Nudging of the zonal-mean field without BT/BC instability
To prevent the emergence of BT/BC instability during integration, we first establish a barotropically and baroclinically stable NH winter zonal-mean flow by using the zonally averaged global ground-to-space (G2S) data on 1 January 2020, considered representative of the NH winter condition. Figure 1 describes the sequential stages of the process. Vertically continuous G2S atmospheric profiles are constructed by fitting B-spline curves to reanalysis data, including NASA’s Modern-Era Retrospective Analysis for Research and Applications (MERRA-2; Gelaro et al. 2017) and the fifth major global reanalysis produced by ECMWF (ERA5; Hersbach et al. 2020) for below the lower mesosphere and empirical model results for the upper atmosphere; see Song et al. (2018) for details.
Figure 1d depicts the stable
Nudging for certain parts or components of the general circulation has been employed in several contexts to examine its effects on the other regions or components (Hoskins et al. 2012; Simpson et al. 2011). Our approach is similar to that of Simpson et al. (2011), which relaxed the zonal-mean component of stratospheric circulation toward the seasonally varying model climatology in the stratosphere and above to quantify the influence of stratospheric variability on tropospheric annular-mode time scales. However, the nudging technique of the present study, used to constrain the zonal-mean state without BT/BC instability, differs from that of Simpson et al. (2011) in certain aspects. First, we apply relaxation throughout the atmosphere, not limiting it to the stratosphere and above as in Simpson et al. (2011). Accordingly, second, our approach does not necessitate additional physical processes to construct the general circulation, while in Simpson et al. (2011), the comprehensive GCM was integrated with the seasonally varying climatological sea surface temperatures, sea ice, and greenhouse gases fixed at 1990s values. Simpson’s method can induce spurious responses in waves through anomalous meridional circulation and artificial sponge-layer feedback (Hitchcock and Haynes 2014). Nudging the entire atmosphere without any longitudinal disturbances in the present study eliminates the likelihood of such spurious responses. Another distinction in our approach is nudging the zonal-mean temperature toward T0(z) without latitudinal variation.
Before introducing the zonally asymmetric GWD, a simulation controlled exclusively by nudging without GWD is conducted to verify its performance. We refer to this experiment as “stable
Summary of the configurations for the different experimental simulations.
The complete absence of any longitudinally varying forcings results in the absence of waves, producing a GHP amplitude of zero (Fig. 2c). However, it is noteworthy that this waveless state persists only up to approximately 60 days of integration (not shown). Beyond this time, small-scale waves emerge from the tropospheric jets and propagate upward and downward. The excitation of small-scale waves from the tropospheric jet stream consistently occurs across multiple simulations with a number of different setups (e.g., horizontal diffusion, Rayleigh damping, and nudging time scale), featuring variations solely in the timing of the appearance. However, given that nudging effectively constrains the background state from day 15, long-time integration is considered unnecessary. Henceforth, experiments are carried out for only 30 days.
c. Prescribed GWD
The absence of waves in StaNOGWD confirms that the nudging approach does not initiate any zonal disturbances during the selected analysis period (30 days). Therefore, in subsequent simulations where zonally varying GWD is included, the in situ generation of waves is solely attributed to the inherent zonal asymmetries in GWD.
To elucidate the characteristics of PWs depending on various GWD configurations, a series of experiments with different GWD configurations, which is listed in Table 1, is conducted. As illustrated in the longitude–latitude cross section of XGWD at the peak altitude (0.2 hPa) in Fig. 3b, XGWD is set to default with a ZWN 1 structure (n = 1). Therefore, the control simulation (“Sta1GWD100”) is run with XGWD having a ZWN 1 structure with X0 = 100 m s−1 day−1 under the stable background state, where the naming convention is given by “Nudging field + ZWN + GWD + X0.” For “Sta1GWD200,” X0 is double that of ZWN 1, while “Sta2GWD100” is a run with ZWN 2 (n = 2) and the default X0 (Fig. 3c). To examine PW generation based on the linearized disturbance QGPV equation, an additional run with 1/100 of X0 is conducted using the GWD of ZWN 1, and then, the results are multiplied by 100. This approach, so-called quasi-linear simulation, is necessary to examine the effects of nonlinearity in the results of fully nonlinear numerical simulations. Finally, a simulation (Uns1GWD100) where the zonal-mean zonal wind is nudged toward the unstable condition, which is shown in Fig. 1a, is conducted.
In all the experiments, prescribed XGWD is initiated on day 15, coinciding with the successful nudging of the background field toward the stable state. To minimize noise induced by the initial adjustment process, XGWD, which is initially zero, increases exponentially for 7 days, attaining its ultimate steady value on day 22.
d. Analysis of PWs
3. Results
As this is the first result using the new framework, we closely examine PWs generated in situ by the idealized GWD in the control simulation (Sta1GWD100). Primarily, we verify whether the background state has been maintained as a stable condition in Sta1GWD100. Figure 4 presents latitude–height cross sections of the zonal-mean zonal wind and its deviation from the stable field shown in Fig. 1d, at days 15, 22, and 27. At day 15,
a. PW amplitude
Analyzing the characteristics of PWs begins with their GHP amplitude. This is conducted by decomposing the resolved waves into ZWN 1, ZWN 2, and ZWN > 2 components at days 15, 22, and 27 (Fig. 5). Unlike in StaNOGWD, PWs manifest in Sta1GWD100 above the region where the idealized GWD is introduced. The PWs exhibit predominant ZWN 1 structure, equivalent to that of GWD. We also observe weak amplifications in waves having ZWNs greater than 1. This is likely associated with nonlinear wave–wave interactions, and the verification of this will be explored in section 3d. Throughout the period of increasing GWD (days 15–22), PWs are not only reinforced but also extend into the upper mesosphere of the SH. Once GWD reaches a peak (from day 22), PWs attain a maximum GHP amplitude of about 700 m and settle into a quasi-steady state with minor fluctuations (please refer to the Hovmöller diagram in Fig. S2). It is interesting to see that PWs are also found in the SH upper mesosphere. This is likely due to equatorward propagation of these waves into the SH, which will be discussed further in sections 3c and 4c.
b. Linearized disturbance QGPV arguments
Even within the GWD area, the amplification of PWs occurs at two distinct centers located at 30° and 60°N (Fig. 5). These regions do not align with the maximum GWD regions located at 45°N but correspond to the meridional boundaries of the peak. This is associated with the nonconservative GWD forcing Z′, taking the form of the meridional gradient of the idealized XGWD. Figure 6 shows the amplitude of Z′ and EP fluxes along with EPFD of PWs on latitude–height cross sections. The XGWD (=X′) with a maximum at 45°N results in two peaks of Z′ at the areas of its most significant latitudinal gradient, specifically 30°–40°N and 50°–60°N (Fig. 6a). These regions coincide with the two major PW GHP amplification locations observed in Fig. 5. In addition, due to the tangential relation of Z′ with respect to latitudes [Eq. (8):
Figures 7a–i present a series of polar stereographic projections depicting q′ development over 6 h from day 15 and the contributions by each term in Eq. (11) calculated from the results of the Sta1GWD100 simulation. The initially unperturbed q′ (Fig. 7a) manifests an obvious ZWN 1 pattern after 6 h (Fig. 7b). During this period, both
The initial behavior of the q′ evolution undergoes a notable change with time. Figures 7j–r illustrate q′ development at day 22. A remarkable distinction is the eastward movement of q′ (Fig. 7j) by approximately 90° from its state on day 15 (Fig. 7b). This is associated with the zonal advection of q′. Tied to the amplifying
Considering the nonnegligible magnitude of the Residual term in both the initial (Fig. 7i) and quasi-steady (Fig. 7r) states, its potential sources are worth noting. First, the linearized disturbance QGPV equation [Eq. (7)], despite being modified to include an ageostrophic term (the isallobaric contribution to υ′, which is
c. Influences of in situ generated PWs
The analysis of EP flux and EPFD in Fig. 6 also provides insights into the propagation and influences of the GWD-induced PWs. Poleward and equatorward EP fluxes propagating away from the equatorward and poleward peaks of Z′, respectively, converge above the maximum GWD area, depositing a localized negative EPFD of about 10 m s−1 day−1. Prominent positive EPFD at the equatorward peak of Z′ and negative EPFD above the maximum GWD are responsible for the acceleration and deceleration of the zonal-mean zonal wind, respectively, in Fig. 4b, though the magnitude is significantly diminished due to the mitigating effect of nudging. The influence of the in situ PWs is not confined to their origin. Following westerlies extended into the SH upper mesosphere, these PWs propagate across the equator. A similar feature was previously identified in some satellite observation studies (e.g., Garcia et al. 2005; Forbes et al. 2002) as well as in the numerical study by Smith (2003), where SPWs from the winter hemisphere propagate toward the summer hemisphere along westerlies in the mesosphere tropics. Along the path toward the SH mesosphere above the zero-wind line, these waves not only initiate zonal disturbances (Fig. 4) but also induce negative EPFD due to their decreasing amplitude with distance from the source region. This interhemispheric propagation property aligns with the theoretically inferred characteristic of quasi-stationary PWs which propagate under westerlies weaker than the critical Rossby speed (Charney and Drazin 1961). Although interhemispheric propagation could be contingent on the prevailing zonal wind condition, this feature indicates a possible contribution of the GWD-generated PWs in the NH to PWs in the SH mesosphere.
d. Characteristics of the in situ PWs induced by different GWD configurations
This section delves into a comparative analysis of the characteristics of PWs dependent on the properties of zonally asymmetric GWD. Figure 8 illustrates the amplitude of PWs across the simulations with varying zonal asymmetries in GWD. In the simulation Sta1GWD200 (Fig. 8b), PW amplitude is intensified to about twice the levels seen in Sta1GWD100, in concordance with the doubled peak value of GWD. This indicates that the GWD-induced PWs have magnitudes in proportion to the peak magnitude of asymmetries in GWD. The amplification is not restricted to PWs of ZWN 1 but extends into the higher ZWN ones arising from the nonlinear process. Regarding this, we speculate whether the degree of the nonlinear interaction increases with the magnitude of zonal asymmetries in GWD. To verify this, an additional experiment, Sta1GWD1, in which the peak magnitude of GWD is reduced by a factor of 100 compared to that of Sta1GWD100, is conducted. In this simulation, the PW amplitude is recovered by multiplying by 100 from the model result. As evidenced in Fig. 8c, the amplitude of PWs with ZWNs larger than 1, indicative of the nonlinear interaction, almost disappears. This result confirms our speculation that as the peak magnitude of the GWD increases, the amplitude of shorter wavelength PWs is enhanced by nonlinear interactions.
In the simulation with ZWN 2 GWD (Sta2GWD100; Fig. 8d), PWs of ZWN 2 (PW2) are excited as expected. However, although the GWD magnitude is identical to that in Sta1GWD100, the amplitude of PW2 is reduced to approximately a quarter that of the ZWN 1 PWs seen in Sta1GWD100, with a maximum of only ∼150 m. The decrease in PW amplitude with increasing ZWN follows the relationship q′ ∝ ∇2Φ′ ∝ − k2Φ′, implying that the Z′-enhanced q′ induces PWs with GHP amplitudes inversely proportional to the square of their ZWN. This indicates that larger-scale asymmetries in GWD tend to generate PWs with more significant amplitude and influence.
The ZWN dependency is also identified in the time required to achieve the stationary state: In Sta2GWD100, the stationary state is attained in roughly half the time (3 days) taken in Sta1GWD100 (6 days). Close inspection of the q′ evolution provides insight into the underlying cause. As discussed above with Fig. 7, Z′ initially causes in-phase changes in q′, thereby inducing the two advection terms shifted 90° relative to that of Z′. Phase difference between Z′ and subsequent Advection (the sum of two advection terms), which is dominated by the zonal advection, triggers the positive feedback, which moves the resultant q′ and Advection eastward until when Advection balances with Z′, achieving the stationary state. Concerning this relation, we compare the phase relationship between Advection and Z′ along with their combined effect in Sta2GWD100 to those in Sta1GWD100, as shown in Fig. 9. The initial phase difference and resultant positive feedback occur consistently in Sta2GWD100. However, aligning with the normal mode method, the doubled ZWN in Sta2GDW100 induces a phase difference of only 45° at the initial state (Fig. 9c). Accordingly, under the same zonal-mean zonal wind condition, Advection reaches the phase opposite that of Z′ in about 3 days, faster than the approximate 6 days required in Sta1GWD100.
Additional experiments varying the meridional location and width of the zonally asymmetric GWD demonstrate the dependence of in situ excited PWs on these factors (Text S2, Table S1, and Figs. S4 and S5). As GWD shifts toward lower latitudes, the opposite-phase PWs originating from the lower- and higher-latitude Z′ peaks exhibit an increasing difference in their GHP amplitude (Fig. S4c). The degree of cancellation between these two opposite-phase PWs decreases with decreasing latitudes. Consequently, PWs at higher-latitude Z′ peak tend to strengthen with decreasing latitude. This phenomenon also appears in simulations varying meridional width of GWD (Fig. S5). With the widening meridional range of GWD, the lower and higher Z′ locations become more distant from each other (Fig. S5b). Furthermore, the difference in GHP amplitude between PWs at these two Z′ peaks also increases. Attributed to both factors, the offsetting of opposite-phase PWs tends to decrease as the meridional width increases. Therefore, PWs at higher-latitude peak intensify with increasing meridional width. Note that the steady-state balance in the linearized disturbance QGPV equations remains consistent across the different meridional locations and width of GWD, thereby not accounting for the latitudinal dependence in GHP amplitude (not shown).
e. Influences of instability on the in situ PW generation by GWD
Focusing on the exclusive role of GWD in exciting PWs without another in situ source, BT/BC instability, may prompt the question “How does in situ PW excitation by GWD alter in the presence of instability?” This question is addressed by performing an additional experiment, Uns1GWD100, in which the zonal-mean atmosphere is constrained toward the G2S wind and temperature data including BT/BC instability (Figs. 1a,b). Figure 10 compares the nudged zonal-mean zonal wind and properties of PWs in the Sta1GWD100 and Uns1GWD100 simulations. In the simulation with an unstable background state, PW amplitudes are notably enhanced. While quantifying the contributions of GWD and instability to the overall increase in PW amplitudes seen in the Uns1GWD100 is not straightforward, we can qualify the influence of instability when PWs are forced by the zonally asymmetric GWD. At the higher-latitude GWD source region (60°N at a 65–70 km altitude), instabilities are concurrent with the enhanced PW amplitudes, suggesting a role of instability in amplifying PWs initiated by GWD. In addition, outside the GWD region in the polar mesosphere (60°–90°N at altitudes of 70–90 km; Fig. 10e), the intensification of PWs with ZWNs larger than 1 is remarkable (Figs. 10c,d) and is accompanied by substantial positive EPFD (Fig. 10e). This suggests an additional but important effect of instability: the strengthening of nonlinear wave–wave interactions. We hypothesize that this nonlinear interaction likely aligns with the findings of Hartmann (1983), which showed that barotropic instability coupled with the preexisting ZWN 1 PWs destabilizes waves having shorter wavelength. To confirm this, an analysis examining unstable wave growth rates under corresponding background conditions is necessary. However, this is beyond the scope of this study.
4. Discussion
In this study, in situ PWs generated exclusively by nonconservative GWD forcing are examined with an idealized GCM framework using a fully nonlinear primitive equation GCM extended from the surface to the lower thermosphere. This enables us to examine the characteristics of PWs with respect to the varying GWD as well as their evolution toward a steady state. Therefore, it is worth comparing the GWD-induced PWs in this study with those from previous modeling and observational research. The waves being compared in this section are predominantly SPWs generated by longitudinal variability in GWD during the NH winter condition.
a. PW amplitude
Regarding the increase in perturbations with height due to decreasing air density, we compare PW amplitudes taking note of their altitudinal region. Furthermore, we also make an effort to note the magnitude of the zonal asymmetries and meridional width of GWD depicted in previous studies, which are crucial determinants of nonconservative GWD forcing. Quasi-steady-state PWs in this study have a GHP amplitude in a range of 500–700 m above their generation altitude (60 km). This magnitude is comparable to that of the OGWD-induced PWs at the 71-km altitude found by Holton (1984). This is likely attributed to three reasons: (i) Although Holton (1984) parameterized OGWs forced by the idealized ZWN 1 topography, OGWD had similar characteristics to our idealized GWD, showing zonal asymmetries of 100 m s−1 day−1 between 30° and 70°N at the 71-km altitude. Moreover, (ii) the absence of PWs originating from the lower boundary sources and (iii) the extinction of barotropic and inertial instability in Holton (1983), excluding other PW sources, are consistent with our approach. MM93 had put forth a similar argument that the concurrent presence of PWs forced from below and the GWD-induced PWs in their simulation led to larger amplitude mesospheric PWs than those seen in Holton (1984). While the GWD-induced PWs in Smith (2003) exhibited an amplitude of 200–300 m within an altitude of 75–85 km, about half of that in the present study, this can likely be attributed to a smaller GWD asymmetry (∼25 m s−1 day−1) in Smith (2003).
Another important finding in Smith (2003) is that GWD-induced PWs in the upper mesosphere (around 80 km altitude) take precedence over the upward-propagating Rossby waves due to the presence of zonal wind reversals from westerlies to easterlies. To examine this possibility, we conducted an additional experiment (not shown) by prescribing the idealized topography devised by Gerber and Polvani (2009) as a source for stationary PWs within 30°–60°N at a lower boundary, wherein the same stable zonal wind nudging is performed. Although the elevation of the topography is lowered to 1.5 km (from the original 3 km) in this simulation, the orographically induced PWs have amplitudes (700–1000 m) greater than those of the GWD-induced waves. This is likely due to the westerlies extending up to 90 km without reversal, enabling the continuous amplification of the stationary waves propagating upward without being filtered.
b. Wind responses to the zonally asymmetric GWD
By analyzing the linearized disturbance QGPV equation, this study elucidates the initial excitation and subsequent growth of the GWD-induced PWs toward a steady state: Initially, nonconservative GWD forcing generates PWs by determining the q′ tendency. Within 7 days, a quasi-steady state is achieved as the nonconservative GWD forcing balances with the dominant zonal advection of q′, which has an opposite phase but comparable magnitude, following its eastward movement driven by the zonal-mean westerlies. The initial feature is consistent with the generation of eastward-propagating mesospheric PWs seen on 18 January 2009 by Song et al. (2020) using the MERRA-2 reanalysis data, which was attributed to the q′ tendency determined by nonconservative GWD forcing (see Fig. 8 in their study). The steady-state balance, on the other hand, aligns with the findings in MM93 and Smith (2003) that the nonconservative GWD forcing was primarily balanced by the zonal advection of q′. Even in a steady state, in Holton (1984), the curl of GWD was balanced by the meridional advection of planetary vorticity.
Despite varying factors balancing the nonconservative GWD forcing in these studies, the mesospheric wind response to a localized GWD showed a consistent downstream shift of the wind minimum relative to the strongest negative drag position (here, eastward). Holton (1984) observed the minimum zonal wind downstream (to the east) of the strongest OGWD by 90°, while a smaller shift, 30°–40°, was noted by MM93. Additional details regarding the wind response were provided by Smith (1996) using a simple barotropic model. Smith (1996) reported that during and immediately after a strong GWD pulse, the minimum wind coincides with the strongest drag, whereas it shifts downstream and persists there until another pulse or the decay of longitudinal variability. Considering the prominent eastward advection of q′ in the present study, similar behavior is expected. This is confirmed in Fig. 11, which depicts the zonally varying GWD and response of
Meanwhile, Lieberman et al. (2013) suggested that the zonal and meridional GWD perturbations induce the meridional
c. Influences of the in situ PWs
The GWD-induced PWs exhibit divergent fluxes away from their source regions and subsequent dissipation along their propagation, providing positive and negative EPFD, respectively (Fig. 6). A similar divergent feature of EP fluxes, directed upward and downward away from the source level (71 km altitude), was noted by Holton (1984). As previously mentioned, the clear discernment of local PW generation by GWD in Holton (1984) was attributable to the absence of upward-propagating PWs forced from below in his ideal simulation. Similar features also manifested in the numerical simulation of MM93 (refer to Figs. 7 and 8 of MM93, which compare EPFD between cases A and B), where PWs forced from the tropopause existed simultaneously. In Case A, which incorporated the curl of GWD, positive EPFD emerged above the significant nonconservative GWD forcing area (40°–50°N), accompanied by negative EPFD in the adjacent higher latitude. While not explicitly stated in MM93, it is reasonable to attribute the positive EPFD to the in situ PW generation. In Case B, excluding the curl of GWD led to a marked reduction in the midlatitude positive EPFD and higher-latitude negative EPFD. This demonstrated that the GWD-induced PWs emanating away from the source region (40°–50°N) contributed to a substantial portion of the negative EPFD along with the upward-propagating PWs. These findings emphasize the ability of GWD-induced PWs to not only deposit positive forcing locally at their origin but also exert negative forcing over a wide area through dissipation.
Related to changes in the PW forcing driven by the curl of GWD, MM93 also observed an interesting phenomenon: compensation between GW and PW forcings. In the framework enabling the interaction among GWs, PWs, and zonal-mean flow in MM93, reduced negative EPFD accelerates the zonal-mean westerlies, thereby negatively enhancing the parameterized OGWD [see Eqs. (1) and (2) in MM93]. Although the value of the negative ΔOGWD equivalent to that of the positive ΔEPFD remains elusive, the compensation resulted in the sum of the two forcings being nearly identical in cases A and B, as depicted in Fig. 10 of MM93. Our EPFD results, which are consistent with those in MM93, support a potential for compensation, suggesting the mechanism of in situ PW generation by zonally asymmetric GWD as an addition to the three compensation mechanisms between parameterized GWs and resolved waves outlined in Cohen et al. (2014): (i) a stability constraint, (ii) a potential vorticity mixing constraint, and (iii) refractive index modification. It is also conceivable that the Cohen et al. (2014) compensation interactions driven by zonal-mean GWD could be influenced by the changes in EPFD attributed to the longitudinal variability in GWD (Šácha et al. 2016). While Cohen et al. (2013) showed that the impact of zonal asymmetries in GWD on the compensation associated with instability is insignificant in the NH winter stratosphere, its importance is expected to rise notably in the mesosphere, where GWD magnitudes and asymmetries increase.
The influence of GWD-induced PWs originating in the NH is not confined to that hemisphere but extends into the SH as the waves propagate across the equator in the westerly waveguide (Fig. 4). This is a consistent phenomenon identified in previous observational and numerical studies (e.g., Forbes et al. 2002; Garcia et al. 2005; Smith 2003), suggesting that the interhemispheric propagation of GWD-induced SPWs from the winter hemisphere is a plausible explanation for SPWs observed in the summer mesosphere (Garcia et al. 2005; Wang et al. 2000). These waves also deposit the negative EPFD in the SH MLT, where positive GWD predominates in the real atmosphere (see Fig. 2 of Yasui et al. 2021). The introduction of a negative momentum anomaly in the SH summer MLT is worth discussing in the context of the interhemispheric coupling of the middle atmospheric circulation (Becker et al. 2004). Körnich and Becker (2010) proposed a plausible mechanism for interhemispheric coupling in association with GW forcing: An increase in negative Rossby wave forcing in the winter stratosphere decelerates the westerlies, which, in turn, facilitates the propagation of eastward-propagating GWs into the winter mesosphere, thereby reducing the dominant negative GWD in that region. The positive GWD anomaly weakens the meridional circulation toward the winter pole, leading to a warm anomaly in the tropical mesosphere. Responding to this tropical warm anomaly via thermal wind relation, the easterlies in the summer upper stratosphere and lower mesosphere are weakened, causing a downward shift in the eastward GWD. A westward GWD anomaly at the altitude of dominant eastward GWD in the summer MLT weakens the meridional circulation toward the equator, leading to a warm anomaly in the summer polar mesosphere. Recently, Yasui et al. (2021), based on growing evidence for a significant influence of in situ excited waves in the MLT, demonstrated that quasi-2-day PWs and secondary GWs, spontaneously generated by destabilized easterlies in the summer stratosphere, propagate upward and deposit westward momentum in the summer MLT, thereby inducing a warm anomaly in the summer polar MLT. The instability-associated quasi-2-day waves were also identified in Lieberman et al. (2021) based on high-altitude observations and observation-driven modeling, while the destabilization of the summer easterly jet was induced by the inertial instability in the lower latitude, resulting from anomalously strong PW breaking in the winter stratosphere. Considering earlier findings of in situ PW generation resulting from enhanced zonal asymmetries in GW fluxes, as filtered by intensified stratospheric PWs during the onset of SSW events (Song et al. 2020) coupled with the interhemispheric propagation toward the SH MLT observed in this study, there is an apparent possibility that GWD-induced PWs in the NH propagate across the equator and deposit negative momentum in the SH, thereby contributing to the warm anomaly. This hypothesis is worth further exploration in elucidating the mechanism of interhemispheric coupling.
5. Summary
A comprehensive understanding of middle atmospheric phenomena requires an improved insight into the mutual interactions between the two primary drivers, PWs and GWs. While numerous observational and modeling studies have revealed various interaction mechanisms among PWs, GWs, and the mean flow, the influence of longitudinally varying GWD on PWs has received relatively less attention, thereby leaving the essential question of how these PWs vary with GWD characteristics. Furthermore, within the limited body of research on this subject, earlier numerical investigations were primarily conducted using simplified models based on QG equations and focused on the steady state of this interaction. Accordingly, in some of these studies, the coexistence of other PW sources and wave–mean-flow interactions could have obscured the exclusive effect of zonally asymmetric GWD on PWs. Therefore, returning to an idealized modeling approach with an advanced GCM, we revisit this issue aiming to (i) enhance our understanding of GWD-induced PW properties and (ii) explore the evolution of this interaction mechanism until reaching the steady state.
In pursuit of our objective, this study devises an effective idealized framework that excludes all potential sources of PWs besides zonally varying GWD and constrains alterations in the zonal-mean flow. Specifically, along with a flat-bottom setup eliminating stationary PW sources, a baroclinic adjustment scheme is implemented to construct a stable zonal-mean field without BT/BC instability. To restrict the occurrence of instability and other interaction mechanisms related to zonal-mean flow modulation, the zonally symmetric component of the entire atmosphere is nudged toward the stable flow state during integration. Without realistic topography, physical relaxation, and GWD parameterization, this nudging technique effectively represents a stable yet realistic zonal-mean structure in the middle atmosphere during a short integration and maintains this state throughout integration. In this framework, an idealized zonally asymmetric GWD, directly introduced in the upper stratosphere/lower mesosphere of the NH midlatitudes, serves as the sole in situ origin of PWs.
The GWD-induced PWs have substantial amplitude in the mesosphere and propagate across the equator following westerlies extended into the SH. Considerable positive EPFD at their generation location and negative EPFD along their subsequent propagation result in the zonal wind changes beyond the nudging. These waves also exhibit temporal evolution in response to the nonconservative GWD forcing before attaining a steady state. Additional experiments introducing different GWD configurations reveal the characteristics of the induced PWs that depend on the properties of the GWD. The PWs amplify in proportion to the magnitude of asymmetries in GWD. Widening the meridional extent of GWD and changing the meridional location of GWD toward lower latitudes also lead to an increase in PW amplitudes. Both the amplitude of the in situ excited PWs and the time for achieving the steady state are inversely correlated with the ZWN of the GWD—a previously unreported phenomenon to the best of our knowledge. Utilizing the fully nonlinear model, we also identify an increasing likelihood of nonlinear wave–wave interactions with increasing magnitudes of the GWD asymmetries. The dependency of in situ excited PWs on ZWN of GWD and nonlinear wave–wave interactions necessitate further investigation and validation under more realistic atmospheric conditions.
Acknowledgments.
This work was supported by a National Research Foundation of Korea (NRF) Grant funded by the South Korean government (MSIT) (2021R1A2C100710212). The third author is supported by the Yonsei University Research Fund of 2023 (2023-22-0095). We thank Anne Smith and R. S. Liebermann for their helpful discussions in the early stage of this work.
Data availability statement.
The SWAD is available upon request to the third author, In-Sun Song.
REFERENCES
Achatz, U., and Coauthors, 2024: Atmospheric gravity waves: Processes and parameterization. J. Atmos. Sci., 81, 237–262, https://doi.org/10.1175/JAS-D-23-0210.1.
Adams, J. C., 1989: MUDPACK: Multigrid portable FORTRAN software for the efficient solution of linear elliptic partial differential equations. Appl. Math. Comput., 34, 113–146, https://doi.org/10.1016/0096-3003(89)90010-6.
Andrews, D. G., J. R. Holton, and C. B. Leovy, 1987: Middle Atmosphere Dynamics. International Geophysics Series, Vol. 40, Academic Press, 489 pp.
Banks, P. M., and G. Kockarts, 1973: Aeronomy, Part B. Academic Press, 372 pp.
Becker, E., A. Müllemann, F.-J. Lübken, H. Körnich, P. Hoffmann, and M. Rapp, 2004: High Rossby-wave activity in austral winter 2002: Modulation of the general circulation of the MLT during the MaCWAVE/MIDAS northern summer program. Geophys. Res. Lett., 31, L24S03, https://doi.org/10.1029/2004GL019615.
Butchart, N., and Coauthors, 2011: Multimodel climate and variability of the stratosphere. J. Geophys. Res., 116, D05102, https://doi.org/10.1029/2010JD014995.
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.
Cohen, N. Y., E. P. Gerber, and O. Bühler, 2013: Compensation between resolved and unresolved wave driving in the stratosphere: Implications for downward control. J. Atmos. Sci., 70, 3780–3798, https://doi.org/10.1175/JAS-D-12-0346.1.
Cohen, N. Y., E. P. Gerber, and O. Bühler, 2014: What drives the Brewer–Dobson circulation? J. Atmos. Sci., 71, 3837–3855, https://doi.org/10.1175/JAS-D-14-0021.1.
Collins, W. D., and Coauthors, 2004: Description of the NCAR Community Atmosphere Model (CAM 3.0). NCAR Tech. Note NCAR/TN-464+STR, 226 pp., https://doi.org/10.5065/D63N21CH.
Dunkerton, T. J., and N. Butchart, 1984: Propagation and selective transmission of internal gravity waves in a sudden warming. J. Atmos. Sci., 41, 1443–1460, https://doi.org/10.1175/1520-0469(1984)041<1443:PASTOI>2.0.CO;2.
Forbes, J. M., X. Zhang, W. Ward, and E. R. Talaat, 2002: Climatological features of mesosphere and lower thermosphere stationary planetary waves within ±40° latitude. J. Geophys. Res., 107, ACL 1–1–ACL 1–14, https://doi.org/10.1029/2001JD001232.
Garcia, R. R., R. Lieberman, J. M. Russell III, and M. G. Mlynczak, 2005: Large-scale waves in the mesosphere and lower thermosphere observed by SABER. J. Atmos. Sci., 62, 4384–4399, https://doi.org/10.1175/JAS3612.1.
Gelaro, R., and Coauthors, 2017: The Modern-Era Retrospective Analysis for Research and Applications, version 2 (MERRA-2). J. Climate, 30, 5419–5454, https://doi.org/10.1175/JCLI-D-16-0758.1.
Gerber, E. P., and L. M. Polvani, 2009: Stratosphere–troposphere coupling in a relatively simple AGCM: The importance of stratospheric variability. J. Climate, 22, 1920–1933, https://doi.org/10.1175/2008JCLI2548.1.
Hartmann, D. L., 1983: Barotropic instability of the polar night jet stream. J. Atmos. Sci., 40, 817–835, https://doi.org/10.1175/1520-0469(1983)040<0817:BIOTPN>2.0.CO;2.
Held, I. M., and M. J. Suarez, 1994: A proposal for the intercomparison of the dynamical cores of atmospheric general circulation models. Bull. Amer. Meteor. Soc., 75, 1825–1830, https://doi.org/10.1175/1520-0477(1994)075<1825:APFTIO>2.0.CO;2.
Hersbach, H., and Coauthors, 2020: The ERA5 global reanalysis. Quart. J. Roy. Meteor. Soc., 146, 1999–2049, https://doi.org/10.1002/qj.3803.
Hitchcock, P., and P. H. Haynes, 2014: Zonally symmetric adjustment in the presence of artificial relaxation. J. Atmos. Sci., 71, 4349–4368, https://doi.org/10.1175/JAS-D-14-0013.1.
Holton, J. R., 1983: The influence of gravity wave breaking on the general circulation of the middle atmosphere. J. Atmos. Sci., 40, 2497–2507, https://doi.org/10.1175/1520-0469(1983)040<2497:TIOGWB>2.0.CO;2.
Holton, J. R., 1984: The generation of mesospheric planetary waves by zonally asymmetric gravity wave breaking. J. Atmos. Sci., 41, 3427–3430, https://doi.org/10.1175/1520-0469(1984)041<3427:TGOMPW>2.0.CO;2.
Hoskins, B., R. Fonseca, M. Blackburn, and T. Jung, 2012: Relaxing the tropics to an ‘observed’ state: Analysis using a simple Baroclinic model. Quart. J. Roy. Meteor. Soc., 138, 1618–1626, https://doi.org/10.1002/qj.1881.
Kang, M. J., H. Y. Chun, and B. G. Song, 2020: Contributions of convective and orographic gravity waves to the Brewer–Dobson circulation estimated from NCEP CFSR. J. Atmos. Sci., 77, 981–1000, https://doi.org/10.1175/JAS-D-19-0177.1.
Kim, Y.-J., S. D. Eckermann, and H.-Y. Chun, 2003: An overview of the past, present and future of gravity‐wave drag parametrization for numerical climate and weather prediction models. Atmos.–Ocean, 41, 65–98, https://doi.org/10.3137/ao.410105.
Körnich, H., and E. Becker, 2010: A simple model for the interhemispheric coupling of the middle atmosphere circulation. Adv. Space Res., 45, 661–668, https://doi.org/10.1016/j.asr.2009.11.001.
Lieberman, R. S., D. M. Riggin, and D. E. Siskind, 2013: Stationary waves in the wintertime mesosphere: Evidence for gravity wave filtering by stratospheric planetary waves. J. Geophys. Res. Atmos., 118, 3139–3149, https://doi.org/10.1002/jgrd.50319.
Lieberman, R. S., J. France, D. A. Ortland, and S. D. Eckermann, 2021: The role of inertial instability in cross-hemispheric coupling. J. Atmos. Sci., 78, 1113–1127, https://doi.org/10.1175/JAS-D-20-0119.1.
Liu, H.-L., and R. G. Roble, 2002: A study of a self‐generated stratospheric sudden warming and its mesospheric–lower thermospheric impacts using the coupled TIME‐GCM/CCM3. J. Geophys. Res., 107, 4695, https://doi.org/10.1029/2001JD001533.
Lu, X., H. Wu, J. Oberheide, H. L. Liu, and J. M. McInerney, 2018: Latitudinal double‐peak structure of stationary planetary wave 1 in the austral winter middle atmosphere and its possible generation mechanism. J. Geophys. Res. Atmos., 123, 11 551–11 568, https://doi.org/10.1029/2018JD029172.
Matsuno, T., 1970: Vertical propagation of stationary planetary waves in the winter Northern Hemisphere. J. Atmos. Sci., 27, 871–883, https://doi.org/10.1175/1520-0469(1970)027<0871:VPOSPW>2.0.CO;2.
Matsuno, T., 1971: A dynamical model of the stratospheric sudden warming. J. Atmos. Sci., 28, 1479–1494, https://doi.org/10.1175/1520-0469(1971)028<1479:ADMOTS>2.0.CO;2.
McFarlane, N. A., 1987: The effect of orographically excited gravity wave drag on the general circulation of the lower stratosphere and troposphere. J. Atmos. Sci., 44, 1775–1800, https://doi.org/10.1175/1520-0469(1987)044<1775:TEOOEG>2.0.CO;2.
McLandress, C., and N. A. McFarlane, 1993: Interactions between orographic gravity wave drag and forced stationary planetary waves in the winter Northern Hemisphere middle atmosphere. J. Atmos. Sci., 50, 1966–1990, https://doi.org/10.1175/1520-0469(1993)050<1966:IBOGWD>2.0.CO;2.
McLandress, C., T. G. Shepherd, S. Polavarapu, and S. R. Beagley, 2012: Is missing orographic gravity wave drag near 60°S the cause of the stratospheric zonal wind biases in chemistry–climate models? J. Atmos. Sci., 69, 802–818, https://doi.org/10.1175/JAS-D-11-0159.1.
Molod, A., L. Takacs, M. Suarez, J. Bacmeister, I. S. Song, and A. Eichmann, 2012: The GEOS-5 atmospheric general circulation model: Mean climate and development from MERRA to Fortuna. NASA/TM-2012-104606, Vol. 28, 124 pp., https://ntrs.nasa.gov/api/citations/20120011790/downloads/20120011790.pdf.
Molod, A., L. Takacs, M. Suarez, and J. Bacmeister, 2015: Development of the GEOS-5 atmospheric general circulation model: Evolution from MERRA to MERRA2. Geosci. Model Dev., 8, 1339–1356, https://doi.org/10.5194/gmd-8-1339-2015.
Palmer, T. N., 1982: Properties of the Eliassen-Palm flux for planetary scale motions. J. Atmos. Sci., 39, 992–997, https://doi.org/10.1175/1520-0469(1982)039<0992:POTEPF>2.0.CO;2.
Polvani, L. M., and P. J. Kushner, 2002: Tropospheric response to stratospheric perturbations in a relatively simple general circulation model. Geophys. Res. Lett., 29, 1114, https://doi.org/10.1029/2001GL014284.
Šácha, P., F. Lilienthal, C. Jacobi, and P. Pišoft, 2016: Influence of the spatial distribution of gravity wave activity on the middle atmospheric dynamics. Atmos. Chem. Phys., 16, 15 755–15 775, https://doi.org/10.5194/acp-16-15755-2016.
Samtleben, N., C. Jacobi, P. Pišoft, P. Šácha, and A. Kuchař, 2019: Effect of latitudinally displaced gravity wave forcing in the lower stratosphere on the polar vortex stability. Ann. Geophys., 37, 507–523, https://doi.org/10.5194/angeo-37-507-2019.
Samtleben, N., A. Kuchař, P. Šácha, P. Pišoft, and C. Jacobi, 2020a: Impact of local gravity wave forcing in the lower stratosphere on the polar vortex stability: Effect of longitudinal displacement. Ann. Geophys., 38, 95–108, https://doi.org/10.5194/angeo-38-95-2020.
Samtleben, N., A. Kuchař, P. Šácha, P. Pišoft, and C. Jacobi, 2020b: Mutual interference of local gravity wave forcings in the stratosphere. Atmosphere, 11, 1249, https://doi.org/10.3390/atmos11111249.
Sato, K., and M. Nomoto, 2015: Gravity wave–induced anomalous potential vorticity gradient generating planetary waves in the winter mesosphere. J. Atmos. Sci., 72, 3609–3624, https://doi.org/10.1175/JAS-D-15-0046.1.
Schoeberl, M. R., and Strobel, D. F., 1984: Nonzonal gravity wave breaking in the winter mesosphere. Dynamics of the Middle Atmosphere, J. R. Holton and T. Matsuno, Eds., Terra Scientific, 45–64.
Sigmond, M., and T. G. Shepherd, 2014: Compensation between resolved wave driving and parameterized orographic gravity wave driving of the Brewer–Dobson circulation and its response to climate change. J. Climate, 27, 5601–5610, https://doi.org/10.1175/JCLI-D-13-00644.1.
Simpson, I. R., P. Hitchcock, T. G. Shepherd, and J. F. Scinocca, 2011: Stratospheric variability and tropospheric annular‐mode timescales. Geophys. Res. Lett., 38, L20806, https://doi.org/10.1029/2011GL049304.
Smith, A. K., 1996: Longitudinal variations in mesospheric winds: Evidence for gravity wave filtering by planetary waves. J. Atmos. Sci., 53, 1156–1173, https://doi.org/10.1175/1520-0469(1996)053<1156:LVIMWE>2.0.CO;2.
Smith, A. K., 1997: Stationary planetary waves in upper mesospheric winds. J. Atmos. Sci., 54, 2129–2145, https://doi.org/10.1175/1520-0469(1997)054<2129:SPWIUM>2.0.CO;2.
Smith, A. K., 2003: The origin of stationary planetary waves in the upper mesosphere. J. Atmos. Sci., 60, 3033–3041, https://doi.org/10.1175/1520-0469(2003)060<3033:TOOSPW>2.0.CO;2.
Song, B.-G., H.-Y. Chun, and I.-S. Song, 2020: Role of gravity waves in a vortex-split sudden stratospheric warming in January 2009. J. Atmos. Sci., 77, 3321–3342, https://doi.org/10.1175/JAS-D-20-0039.1.
Song, I.-S., 2023: A modeling system for whole atmosphere dynamics researches. 60th Korean Meteorological Society 2023 Spring Conf., Busan, South Korea, KMA, 115, https://www.dbpia.co.kr/journal/articleDetail?nodeId=NODE11426525.
Song, I.-S., H.-Y. Chun, G. Jee, S.-Y. Kim, J. Kim, Y.-H. Kim, and M. A. Taylor, 2018: Dynamic initialization for whole atmospheric global modeling. J. Adv. Model. Earth Syst., 10, 2096–2120, https://doi.org/10.1029/2017MS001213.
Wang, D. Y., W. E. Ward, G. G. Shepherd, and D.-L. Wu, 2000: Stationary planetary waves inferred from WINDII wind data taken within altitudes 90–120 km during 1991–96. J. Atmos. Sci., 57, 1906–1918, https://doi.org/10.1175/1520-0469(2000)057<1906:SPWIFW>2.0.CO;2.
Williamson, D. L., and J. G. Olson, 1994: Climate simulations with a semi-Lagrangian version of the NCAR Community Climate Model. Mon. Wea. Rev., 122, 1594–1610, https://doi.org/10.1175/1520-0493(1994)122<1594:CSWASL>2.0.CO;2.
Yamashita, C., H.-L. Liu, and X. Chu, 2010: Responses of mesosphere and lower thermosphere temperatures to gravity wave forcing during stratospheric sudden warming. Geophys. Res. Lett., 37, L09803, https://doi.org/10.1029/2009GL042351.
Yasui, R., K. Sato, and Y. Miyoshi, 2021: Roles of Rossby waves, Rossby–gravity waves, and gravity waves generated in the middle atmosphere for interhemispheric coupling. J. Atmos. Sci., 78, 3867–3888, https://doi.org/10.1175/JAS-D-21-0045.1.
Yoo, J.-H., and H.-Y. Chun, 2023: Compensation between resolved wave forcing and parameterized orographic gravity wave drag in the Northern Hemisphere winter stratosphere revealed in NCEP CFS reanalysis data. J. Atmos. Sci., 80, 487–499, https://doi.org/10.1175/JAS-D-22-0102.1.