1. Introduction












Many studies have generalized the TEM equations to three dimensions (3D) to examine local wave activity and material transport. Hoskins et al. (1983), Trenberth (1986), and Plumb (1986) extended the TEM equations to 3D by using the time mean instead of the zonal mean under the QG approximation. Takaya and Nakamura (1997, 2001) derived a phase-independent 3D wave activity flux and quasi-Stokes correction applicable to quasi-stationary Rossby waves in zonally asymmetric flow by using the wave activity density and wave energy conservation laws.
Regarding the primitive equation, Miyahara (2006) and Kinoshita et al. (2010) extended the TEM equations to 3D by using the time mean under the assumption that the Coriolis parameter is constant. They used the dispersion relation for the inertia–gravity wave in the terms included in the horizontal component of the 3D quasi-Stokes correction. Thus, their quasi-Stokes correction is applicable only to inertia–gravity waves. Kinoshita and Sato (2013a,b) derived time-mean 3D residual mean flow and wave activity flux applicable to both Rossby waves and gravity waves. Kinoshita and Sato (2014) extended the 3D TEM equations to equatorial waves. Finally, Noda (2010, 2014) formulated a generalized 3D TEM equation for a plane wave under the Wentzel–Kramers–Brillouin (WKB) approximation.
In the ocean dynamics, McDougall and McIntosh (1996, 2001) formulated the temporal-residual-mean (TRM) flow using the conservation equations of mean density and density variance. The TRM flow includes not only the horizontal eddy flux of density but also the effects of advection of density variance by the mean flow corresponding to the second order of quasi-Stokes corrections in the vertical. The horizontal component of the TRM flow is equivalent to the thickness-weighted mean velocity in density coordinates. Plumb and Ferrari (2005) discussed the conception of residual mean circulation on a zonal-mean flow. They focused on the quasi-Stokes streamfunction included in the residual mean circulation and residual buoyancy flux in the TEM buoyancy equation. They argued that the quasi-Stokes streamfunction should be chosen in such a way as to eliminate the skew component of the raw eddy flux. That is, the residual buoyancy flux on the right-hand side of the TEM buoyancy equation depends only on the diapycnal component of the raw buoyancy flux. In this conception, the residual vertical flow and residual buoyancy flux in the TEM buoyancy equation are zero in the adiabatic condition.
Almost all of these previous studies derived the 3D TEM equations by proving that the residual mean flow is equal to the sum of the mean flow and quasi-Stokes correction and that the wave activity flux is a product of the group velocity and wave activity density under the small-amplitude assumption. Relations similar to those expressed by Eq. (1.2) and especially the relation between the residual mean vertical flow and diabatic heating rate were investigated for transient waves in 3D. However, effects of stationary waves in those relations have not been confirmed in 3D.
On the other hand, Sato et al. (2009) examined the horizontal structure of ozone increase processes after a mature stage of the Antarctic ozone hole in spring by using ozonesonde observations at Syowa Station and observation by the satellite-based Improved Limb Atmospheric Spectrometer II (ILAS-II) in 2003. They found that the observed downward transport of ozone has significant longitudinal dependency. They also showed that the longitudinal structure of the downward transport corresponds not to a quasi-stationary planetary wave’s vertical component but to isentropes modified by a quasi-stationary planetary wave, with the amplitude and phase varying on a seasonal time scale. Thus, it appears that dynamical material transport at high latitudes has longitudinal dependency and relates to stationary wave activities.
Based on the above considerations, the purpose of this study is to derive the 3D quasi-residual mean flow to diagnose 3D dynamical material transport including the effects of stationary waves. The obtained 3D quasi-residual mean flow corresponds to the 3D residual mean flow of Plumb (1986) with the tilting effects of potential temperature and potential vorticity surfaces due to stationary waves removed. Then, for transient waves, the conception of skew eddy flux is established. On the other hand, for stationary waves, the conception is established when the zonal mean is applied. Therefore, the derived flow is named “3D quasi-residual mean flow.” The 3D quasi-residual mean flow is balanced with diabatic heating rate and potential vorticity flux. Next, to demonstrate the usefulness of the derived 3D residual mean flow, the dynamical material transport of sulfur hexafluoride (SF6) is examined by using both the traditional and derived residual mean flows.
The rest of the paper is organized as follows. In section 2, the vertical component of the 3D residual mean flow is derived. The horizontal components are then derived in section 3. In section 4, the treatment of gravity wave forcing is discussed. Section 5 shows the usefulness of the derived 3D residual mean flow through the analysis of SF6 transport. A summary and the concluding remarks are given in section 6.
2. 3D quasi-residual mean vertical flow balanced with diabatic heating rate








A schematic figure of zonal–log pressure cross section of the tilt of potential temperature and zonal wind;
Citation: Journal of the Atmospheric Sciences 76, 3; 10.1175/JAS-D-18-0085.1
3. 3D quasi-residual mean horizontal flow balanced with potential vorticity flux


























A schematic diagram of horizontal cross section of effective Coriolis forcing, pressure gradient forcing, and residual mean meridional flow.
Citation: Journal of the Atmospheric Sciences 76, 3; 10.1175/JAS-D-18-0085.1
4. Nonconservative terms and gravity wave forcing








5. Case study
Using the newly derived 3D quasi-residual mean flow, we investigate the 3D structure of dynamical transport of minor constituents. The minor constituent used in this study was SF6, which is industrially produced as an electrical insulator and almost entirely an anthropogenic gas. Moreover, because its lifetime is almost 3200 years, SF6 can be used as a passive tracer. By using the SF6 mixing ratio, a comparison analysis among the 2D residual mean flow, Plumb’s (1986) 3D residual mean flow, and our derived 3D quasi-residual mean flow was conducted.
a. Data
To prepare atmospheric SF6 data, we used the Center for Climate System Research/National Institute for Environmental Studies/Frontier Research Center for Global Change (CCSR/NIES/FRCGC) an atmospheric general circulation model-based chemistry transport model (ACTM) (Numaguti et al. 1997; Ishijima et al. 2010). The horizontal resolution is T42 (corresponding to almost 300 km), and the vertical resolution is 67 sigma-coordinate layers (Ishijima et al. 2010). The model top is at almost 90 km. The analysis period was January 2013, after a spinup run from 2001 to 2012. The sea surface temperature was prescribed by ERA-Interim data. Convection and vertical diffusion schemes based on Arakawa and Schubert (1974) and Moorthi and Suarez (1992) were used for subgrid scales. The model uses orographic gravity wave forcing of McFarlane (1987).
b. Meridional cross section of residual mean flow
First, Fig. 3 shows the monthly mean, zonal-mean, residual mean flow and mass-streamfunction results obtained in this analysis. Comparison between the traditional residual mean flow (Fig. 3a) and that of Plumb (1986) (Fig. 3b) shows indirect circulation in the Northern Hemisphere’s stratospheric polar region around 60°N in Fig. 3b, unlike in Fig. 3a. This difference is because the quasi-Stokes correction of Plumb (1986) includes the effects of only transient waves. Thus, in the region where the transient waves are small, such as the Northern Hemisphere’s stratospheric polar region, the residual mean flow of Plumb (1986) is nearly equal to the Eulerian-mean flow. On the other hand, the quasi-residual mean flow and mass streamfunction in Fig. 3c, obtained using Eqs. (3.2) and (3.5), are almost the same as those in Fig. 3a, since the zonal mean of quasi-Stokes correction derived in this study does include the effects of stationary waves. Note that we confirm that transient waves are dominant below 100 hPa, quasi-stationary waves are dominant above 30 hPa around 60°N, and the stratospheric sudden warming does not occur in this analysis period (not shown).
Meridional cross sections of monthly mean, zonal-mean residual mean flow (arrows) and mass streamfunction (color shading) in January 2013: (a) traditional 2D TEM equations, (b) 3D residual mean flow of Plumb (1986), and (c) 3D quasi-residual mean flow of this study.
Citation: Journal of the Atmospheric Sciences 76, 3; 10.1175/JAS-D-18-0085.1
Figure 4 shows meridional cross sections of the monthly mean, zonal-mean, residual mean vertical flow (Fig. 4a: traditional; Fig. 4b: this study), together with
Meridional cross sections of monthly mean, zonal-mean residual mean vertical flows: (a) traditional 2D TEM equations, (b) this study, and (c)
Citation: Journal of the Atmospheric Sciences 76, 3; 10.1175/JAS-D-18-0085.1
c. Horizontal structure of residual mean flow
Figure 5 shows the 3D quasi-residual mean meridional flows of this study, together with the term
Horizontal maps of (a) meridional components of monthly mean 3D quasi-residual mean flow and (b)
Citation: Journal of the Atmospheric Sciences 76, 3; 10.1175/JAS-D-18-0085.1
Next, Fig. 6 shows horizontal maps of the time-mean vertical flow vertical component of quasi-Stokes correction of this study, residual mean vertical flows of Sato et al. (2013), quasi-residual mean vertical flow, and
Horizontal maps of monthly mean (a) vertical flow, (b) vertical component of quasi-Stokes correction, (c) 3D residual mean vertical flow of Sato et al. (2013), (d) quasi-residual mean vertical flow, and (e)
Citation: Journal of the Atmospheric Sciences 76, 3; 10.1175/JAS-D-18-0085.1
d. Dynamical transport of SF6


Figure 7 shows a meridional cross section of the monthly mean, zonal-mean SF6 and its time derivative. The distribution is almost symmetric at the equator in the stratosphere and large in the Northern Hemisphere troposphere. The time derivative of SF6 is positive in the Southern Hemisphere, and its distribution is along the contour line of monthly mean SF6. The increase, however, is about 1.0 × 10−3 ppt day−1 and thus small.
Meridional cross section of monthly mean, zonal-mean SF6 (contours) and its time derivative (color shading).
Citation: Journal of the Atmospheric Sciences 76, 3; 10.1175/JAS-D-18-0085.1
Next, from the distribution of SF6 and the residual mean flow, the advection of residual mean flow
Meridional cross section of monthly mean, zonal-mean (a) advection of residual mean flow
Citation: Journal of the Atmospheric Sciences 76, 3; 10.1175/JAS-D-18-0085.1







Meridional cross section of monthly mean, zonal-mean (a) advection of 3D quasi-residual mean flow and (b)
Citation: Journal of the Atmospheric Sciences 76, 3; 10.1175/JAS-D-18-0085.1
Figure 10 shows horizontal maps of the monthly mean SF6 (Fig. 10a), the advection of monthly mean 3D quasi-residual mean flow of this study (Fig. 10b), and the term
Horizontal maps of (a) monthly mean SF6 (color shading) and horizontal component of 3D quasi-residual mean flow (arrows), (b) advection of 3D residual mean flow, and (c) the term of
Citation: Journal of the Atmospheric Sciences 76, 3; 10.1175/JAS-D-18-0085.1
Therefore, the 3D quasi-residual mean flow, which includes the quasi-Stokes correction associated with transient and stationary waves and corresponds to the diabatic heating rate and potential vorticity flux in the monthly mean state, satisfies the required characteristics of the traditional 2D residual mean flow. That is, under the condition that effects of source/sink of SF6 and time variation of SF6 with longer than monthly mean are small, the advection by 3D quasi-residual mean flow is balanced with nonlinear, nonconservative effects of disturbances.
6. Concluding remarks
In this study, the 3D quasi-residual mean flow has been extended to investigate dynamical transport associated with both stationary and transient waves. Note that the 3D quasi-residual mean flow was derived as the sum of Plumb’s (1986) residual mean flow and terms due to the effects of stationary waves, which were derived from the time-mean flow and potential temperature. Therefore, this theory is consistent with the previous theory applicable to transient waves. To confirm the usefulness of the derived residual mean flow, we analyzed the dynamical transport of SF6 through an ACTM simulation. The results showed that the zonal-mean 3D quasi-residual mean flow has the same distribution as the traditional 2D residual mean flow. It was also shown that the horizontal structure of the 3D quasi-residual mean flow is balanced with the term
In the future, we will extend this theory to the equatorial region. Almost all theories deriving 3D TEM equations, including the one in this study, are not applicable to the equatorial region, because the Coriolis parameter is included in the denominators of the formulas for Stokes drift. The formulas of Kinoshita and Sato (2014), however, are applicable to the equatorial region but cannot express the effects of stationary waves. Thus, we will derive a 3D quasi-residual mean flow applicable to the equatorial region by combining this study and those formulas.
Acknowledgments
We thank Dr. Anne Smith and anonymous reviewers for providing constructive comments and suggestions. The GFD-DENNOU library was used for drawing figures. This study was supported by a Grant-in-Aid for Young Scientists (B) JP15K21673 and JP16K16186 of the Japan Society for the Promotion of Science (JSPS) and by a Grant-in-Aid for Scientific Research (A) JP25247075 of the Ministry of Education, Culture, Sports and Technology (MEXT), Japan.
REFERENCES
Andrews, D. G., and M. E. McIntyre, 1976: Planetary waves in horizontal and vertical shear: Generalized Eliassen-Palm relation and mean zonal acceleration. J. Atmos. Sci., 33, 2031–2048, https://doi.org/10.1175/1520-0469(1976)033<2031:PWIHAV>2.0.CO;2.
Andrews, D. G., and M. E. McIntyre, 1978: Generalized Eliassen-Palm and Charney-Drazin theorems for waves on axisymmetric mean flows in compressible atmospheres. J. Atmos. Sci., 35, 175–185, https://doi.org/10.1175/1520-0469(1978)035<0175:GEPACD>2.0.CO;2.
Andrews, D. G., J. R. Holton, and C. B. Leovy, 1987: Middle Atmosphere Dynamics. Academic Press, 489 pp.
Arakawa, A., and W. H. Schubert, 1974: Interaction of a cumulus cloud ensemble with the large-scale environment, part I. J. Atmos. Sci., 31, 674–701, https://doi.org/10.1175/1520-0469(1974)031<0674:IOACCE>2.0.CO;2.
Brewer, A. W., 1949: Evidence for a world circulation provided by the measurements of helium and water vapor distribution in the stratosphere. Quart. J. Roy. Meteor. Soc., 75, 351–363, https://doi.org/10.1002/qj.49707532603.
Dobson, G. M. B., 1956: Origin and distribution of the polyatomic molecules in the atmosphere. Proc. Roy. Soc. London, 236A, 187–193, https://doi.org/10.1098/rspa.1956.0127.
Dunkerton, T., 1978: Mean meridional mass motions of the stratosphere and mesosphere. J. Atmos. Sci., 35, 2325–2333, https://doi.org/10.1175/1520-0469(1978)035<2325:OTMMMM>2.0.CO;2.
Edmon, H. J., B. J. Hoskins, and M. E. McIntyre, 1980: Eliassen-Palm cross sections for the troposphere. J. Atmos. Sci., 37, 2600–2616, https://doi.org/10.1175/1520-0469(1980)037<2600:EPCSFT>2.0.CO;2.
Haynes, P. H., C. J. Marks, M. E. McIntyre, T. G. Shepherd, and K. P. Shine, 1991: On the downward control of extratropical diabatic circulations by eddy-induced mean zonal forces. J. Atmos. Sci., 48, 651–679, https://doi.org/10.1175/1520-0469(1991)048<0651:OTCOED>2.0.CO;2.
Hoskins, B. J., I. N. James, and G. H. White, 1983: The shape, propagation and mean-flow interaction of large-scale weather systems. J. Atmos. Sci., 40, 1595–1612, https://doi.org/10.1175/1520-0469(1983)040<1595:TSPAMF>2.0.CO;2.
Ishijima, K., and Coauthors, 2010: Stratospheric influence on the seasonal cycle of nitrous oxide in the troposphere as deduced from aircraft observations and model simulations. J. Geophys. Res., 115, D20308, https://doi.org/10.1029/2009JD013322.
Kinoshita, T., and K. Sato, 2013a: A formulation of three-dimensional residual mean flow applicable both to inertia–gravity waves and to Rossby waves. J. Atmos. Sci., 70, 1577–1602, https://doi.org/10.1175/JAS-D-12-0137.1.
Kinoshita, T., and K. Sato, 2013b: A formulation of unified three-dimensional wave activity flux of inertia–gravity waves and Rossby waves. J. Atmos. Sci., 70, 1603–1615, https://doi.org/10.1175/JAS-D-12-0138.1.
Kinoshita, T., and K. Sato, 2014: A formulation of three-dimensional residual mean flow and wave activity flux applicable to equatorial waves. J. Atmos. Sci., 71, 3427–3438, https://doi.org/10.1175/JAS-D-13-0161.1.
Kinoshita, T., Y. Tomikawa, and K. Sato, 2010: On the three-dimensional residual mean circulation and wave activity flux of the primitive equations. J. Meteor. Soc. Japan, 88, 373–394, https://doi.org/10.2151/jmsj.2010-307.
McDougall, T. J., and P. C. McIntosh, 1996: The temporal-residual-mean velocity. Part I: Derivation and the scalar conservation equations. J. Phys. Oceanogr., 26, 2653–2665, https://doi.org/10.1175/1520-0485(1996)026<2653:TTRMVP>2.0.CO;2.
McDougall, T. J., and P. C. McIntosh, 2001: The temporal-residual-mean velocity. Part II: Isopycnal interpretation and the tracer and momentum equations. J. Phys. Oceanogr., 31, 1222–1246, https://doi.org/10.1175/1520-0485(2001)031<1222:TTRMVP>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 T. G. Shepherd, 2009: Simulated anthropogenic changes in the Brewer–Dobson circulation, including its extension to high latitudes. J. Climate, 22, 1516–1540, https://doi.org/10.1175/2008JCLI2679.1.
Miyahara, S., 2006: A three dimensional wave activity flux applicable to inertio-gravity waves. SOLA, 2, 108–111, https://doi.org/10.2151/sola.2006-028.
Moorthi, S., and M. Z. Suarez, 1992: Relaxed Arakawa–Schubert: A parameterization of moist convection for general circulation models. Mon. Wea. Rev., 120, 978–1002, https://doi.org/10.1175/1520-0493(1992)120<0978:RASAPO>2.0.CO;2.
Noda, A., 2010: A general three-dimensional transformed Eulerian mean formulation. SOLA, 6, 85–88, https://doi.org/10.2151/sola.2010-022.
Noda, A., 2014: Generalized transformed Eulerian mean (GTEM) description for Boussinesq fluids. J. Meteor. Soc. Japan, 92, 411–431, https://doi.org/10.2151/jmsj.2014-501.
Numaguti, A., M. Takahashi, T. Nakajima, and A. Sumi, 1997: Development of CCSR/NIES atmospheric general circulation model. Study on the Climate System and Mass Transport by a Climate Model, Supercomput. Monogr., Vol. 3, National Institute for Environmental Studies, 1–48.
Okamoto, K., K. Sato, and H. Akiyoshi, 2011: A study on the formation and trend of the Brewer-Dobson circulation. J. Geophys. Res., 116, D10117, https://doi.org/10.1029/2010JD014953.
Plumb, R. A., 1986: Three-dimensional propagation of transient quasi-geostrophic eddies and its relationship with the eddy forcing of the time–mean flow. J. Atmos. Sci., 43, 1657–1678, https://doi.org/10.1175/1520-0469(1986)043<1657:TDPOTQ>2.0.CO;2.
Plumb, R. A., and R. Ferrari, 2005: Transformed Eulerian-mean theory. Part I: Nonquasigeostrophic theory for eddies on a zonal-mean flow. J. Phys. Oceanogr., 35, 165–174, https://doi.org/10.1175/JPO-2669.1.
Sato, K., Y. Tomikawa, G. Hashida, T. Yamanouchi, H. Nakajima, and T. Sugita, 2009: Longitudinally dependent ozone increase in the Antarctic polar vortex revealed by balloon and satellite observations. J. Atmos. Sci., 66, 1807–1820, https://doi.org/10.1175/2008JAS2904.1.
Sato, K., T. Kinoshita, and K. Okamoto, 2013: A new method to estimate three-dimensional residual-mean circulation in the middle atmosphere and its application to gravity wave–resolving general circulation model data. J. Atmos. Sci., 70, 3756–3779, https://doi.org/10.1175/JAS-D-12-0352.1.
Takaya, K., and H. Nakamura, 1997: A formulation of a wave-activity flux for stationary Rossby waves on a zonally varying basic flow. Geophys. Res. Lett., 24, 2985–2988, https://doi.org/10.1029/97GL03094.
Takaya, K., and H. Nakamura, 2001: A formulation of a phase-independent wave-activity flux for stationary and migratory quasigeostrophic eddies on a zonally varying basic flow. J. Atmos. Sci., 58, 608–627, https://doi.org/10.1175/1520-0469(2001)058<0608:AFOAPI>2.0.CO;2.
Trenberth, K. E., 1986: An assessment of the impact of transient eddies on the zonal flow during a blocking episode using localized Eliassen–Palm flux diagnostics. J. Atmos. Sci., 43, 2070–2087, https://doi.org/10.1175/1520-0469(1986)043<2070:AAOTIO>2.0.CO;2.