1. Introduction
Dima et al. (2005) noted that the observed climatological-mean stationary wave in the tropics bears a striking degree of hemispheric symmetry during both solstice seasons, in spite of the strong seasonality of the eddy forcing associated with the shift of the convective heating into the summer hemisphere. They proposed that the propagation of the eddies from the summer to the winter hemisphere might be important for setting up this symmetric pattern. Their hypothesis is supported by the strong seasonality of tropical eddy momentum fluxes, which are directed from the winter to the summer hemisphere (indicating propagation in the opposite direction for waves with westerly pseudomomentum) during both solstice seasons. This seasonal cross-equatorial propagation cancels out in the annual mean, producing much smaller annually averaged eddy momentum fluxes.
Despite its important impacts, the interhemispheric propagation of Rossby waves is not fully understood. Classical Rossby wave propagation theory (Hoskins and Karoly 1981) predicts no meridional propagation through boreal summer easterlies. During boreal winter, it has been argued that meridional propagation may be possible across westerly ducts, particularly over the eastern Pacific (Webster and Holton 1982; Hoskins and Ambrizzi 1993). However, the classical propagation conditions are modified in the presence of significant zonal-mean meridional flow (Schneider and Watterson 1984), a scenario of some relevance to the tropics due to the strong cross-equatorial Hadley cell. Li et al. (2015) show that this meridional flow plays an important role for interhemispheric Rossby wave propagation, consistent with the arguments of Schneider and Watterson (1984).
In a study very relevant to this work, Kraucunas and Hartmann (2007) used a simple shallow-water model to investigate what factors are important for the observed structure of the tropical climatological-mean stationary wave and the determination of the cross-equatorial eddy momentum flux. They found that the winter-hemisphere response to summer-hemisphere forcing is significantly enhanced by the Hadley cell flow, a sensitivity that they attributed to the impact of the mean meridional flow on propagation suggested by the work of Schneider and Watterson (1984). They also used a simple barotropic, beta-plane model to illustrate this effect analytically.
However, because all these ideas on meridional propagation are based on the analysis of the barotropic or quasigeostrophic shallow-water vorticity equations, they should only apply in principle to the rotational momentum flux
Tropical eddy momentum fluxes arise from spatial correlations between divergent eddy meridional velocities
This is the question that we aim to address with our study, which is organized as follows. Section 2 introduces the data and some definitions, and provides a quick overview of the main results in Part I. Section 3 uses a forced, homogeneous, linear model to predict the rotational response to divergent forcing, and shows that this model can reproduce quite well the observed, seasonally varying eddy momentum flux when it is forced by the observed divergent flow. Section 4 analyzes in some detail the dynamical determinants of the eddy momentum flux response in the simple model using the vorticity balance and section 5 investigates the relevance of that analysis in observations. We finish with some concluding remarks in section 6.
2. Data and conventions
We use in this paper the same ERA-Interim data (Dee et al. 2011) of Part I, which span the period 1979–2016. Data are defined on pressure levels at 2.5° resolution and the diagnostics presented in this paper represent an upper-troposphere average integrated over the 150–300-hPa layer. The horizontal velocities are decomposed into their rotational and divergent components (denoted with r and d subscripts, respectively), obtained by inversion of the vorticity and divergence fields. We also use monthly mean GPCP precipitation data (Adler et al. 2003) for some diagnostics.
Eddies are defined as differences from the zonal mean but we will focus in this paper on the stationary eddy momentum fluxes, which were shown to dominate the momentum transport in Part I. Because of the strong eddy momentum flux seasonality, however, this stationary circulation must be defined seasonally. We thus define a seasonal cycle of stationary eddy momentum transport computing for each calendar month the zonal mean of the products of the 37-yr monthly mean eddy velocity components. This procedure eliminates the submonthly and interannual variability but some of the slow intraseasonal variability may be contaminated into the seasonal cycle due to the limited record. However, Lee (1999) showed that the transient tropical eddy momentum flux is dominated by the interannual variability and Part I shows that the MJO contribution to the eddy momentum transport is an order of magnitude smaller than the stationary wave contribution.
Figure 1 shows the seasonal cycle of the rotational (

Rotational (solid lines) and divergent (dashed lines) components of the stationary cross-equatorial eddy momentum flux. The black lines show the total flux, the red lines the k = 1 contributions and the blue lines the contributions by wavenumbers k = 2–3.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1

Rotational (solid lines) and divergent (dashed lines) components of the stationary cross-equatorial eddy momentum flux. The black lines show the total flux, the red lines the k = 1 contributions and the blue lines the contributions by wavenumbers k = 2–3.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
Rotational (solid lines) and divergent (dashed lines) components of the stationary cross-equatorial eddy momentum flux. The black lines show the total flux, the red lines the k = 1 contributions and the blue lines the contributions by wavenumbers k = 2–3.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
We will focus in this paper on the determination of the dominant k = 1 component of the stationary
3. A forced linear model














However, a theory of unforced propagation is unlikely to explain the observed tropical momentum fluxes in the strongly convecting terrestrial atmosphere. As discussed in Part I, the divergent momentum fluxes are associated with a region of strong momentum convergence/eddy generation adjacent to the equator in the summer hemisphere (see Figs. 7b and 8b in Part I), pointing to the important role of the forcing for the determination of these fluxes. Motivated by this, we study the forced
As will be shown in the next section, the closure of the vorticity balance can be very different for eddies of different meridional scales. Our focus in this paper will be on the gravest meridional modes, which we will show to dominate the momentum transport. Aiming to uncouple the meridional modes, we use a homogeneous formulation with constant U, V, and




We solve Eq. (4) for the dominant zonal wavenumber k = 1 as follows. For each calendar month, we define U and V as the climatological zonal-mean zonal and meridional winds averaged over the equatorial band
Use of an effective beta improves the quantitative agreement with observations by increasing the amplitude of the response but the results are not very sensitive to the other modeling choices. In contrast, the results are fairly sensitive to the viscosity coefficient ν, chosen to damp the resonant wavenumber
This procedure produces the

Seasonal cycle of k = 1 stationary eddy momentum flux
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1

Seasonal cycle of k = 1 stationary eddy momentum flux
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
Seasonal cycle of k = 1 stationary eddy momentum flux
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
To investigate the relative importance of vortex stretching and meridional divergent advection for the model’s response, we have repeated the analysis forcing the model with each of the terms on the right-hand side of Eq. (4) in isolation. Figures 2c and 2d show that the two forcings produce a response of the same sign (a momentum flux from the winter to the summer hemisphere) but the model’s response to stretching is significantly larger. Section 4b discusses in more detail the dynamics of the two responses.
Likewise, we can investigate the role played by changes in the eddy forcing and in the basic state for the seasonal cycle of the eddy momentum flux in our model. The left panels of Fig. 3 show that the mean tropical winds play virtually no role for the determination of the eddy momentum flux, as the results when fixing U and V to their climatological DJF or JJA values and when using zero winds are very similar to those with seasonally dependent winds. This implies that the seasonal dependence of the eddy momentum flux in the model must be due to seasonal changes in the eddy forcing, as indeed confirmed by the right panels of Fig. 3. We provide an interpretation of these results in the next section.

(left) Model’s sensitivity to prescribed mean winds U and V: (top) climatological DJF winds, (middle) climatological JJA winds, and (bottom)
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1

(left) Model’s sensitivity to prescribed mean winds U and V: (top) climatological DJF winds, (middle) climatological JJA winds, and (bottom)
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
(left) Model’s sensitivity to prescribed mean winds U and V: (top) climatological DJF winds, (middle) climatological JJA winds, and (bottom)
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
4. Dynamical analysis
a. Natural frequency and dominant meridional scales





As discussed in section 3, the asymmetry in Rossby’s dispersion relation due to the meridional Hadley flow V plays an important role in theories of (unforced) Rossby wave propagation in the tropics (Li et al. 2015). A related argument is that with easterly or weak westerly U, modes with
Figure 4 shows the structure of

Natural frequency
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1

Natural frequency
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
Natural frequency
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
The model’s success suggests that these large meridional scales are also primarily responsible for the observed momentum transport. To assess the scales of dominant momentum transport in model and observations, we have decomposed the meridionally averaged momentum flux between 20°S and 20°N (shown in Fig. 5a) into contributions by different meridional wavenumbers. This metric is a sensible measure of the tropical momentum flux, as the momentum flux is nearly one signed during both solstice seasons (cf. Fig. 2). To compute the contributions of each meridional wavenumber to this metric, we expand

(a) Eddy momentum flux, integrated between 20°S and 20°N in observations (blue) and the model (red). The dashed lines show the mean flux computed using only modes with
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1

(a) Eddy momentum flux, integrated between 20°S and 20°N in observations (blue) and the model (red). The dashed lines show the mean flux computed using only modes with
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
(a) Eddy momentum flux, integrated between 20°S and 20°N in observations (blue) and the model (red). The dashed lines show the mean flux computed using only modes with
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
Figure 5b shows that the observed momentum flux is dominated by long waves, especially during JJA. During DJF, shorter-scale contributions to the eddy momentum flux are found around
b. Divergent forcing and momentum flux direction
As noted in Part I, while the sign of the rotational momentum flux is determined by the tilt of the streamlines, the sign of the divergent momentum flux depends instead on the phase relation between the rotational and the divergent flow. When the rotational flow is simply forced by the divergent flow as assumed in this section, this phase relation is determined by the closure of the vorticity balance. In this section, we investigate what aspects of the divergent eddy forcing and the mean state determine the direction of the divergent momentum flux.






















Sketch illustrating the rotational circulation forced by prescribed divergent forcing and the associated momentum flux
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1

Sketch illustrating the rotational circulation forced by prescribed divergent forcing and the associated momentum flux
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
Sketch illustrating the rotational circulation forced by prescribed divergent forcing and the associated momentum flux
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
The configuration in Fig. 6a (Fig. 6b) is associated with a southward (northward) divergent momentum flux
We next discuss the
It is thus key to take into account the meridional f structure to obtain a nonzero momentum flux response. Figure 6d shows the response to the forcing
With nonconstant f, it is not possible to find a simple expression for the momentum flux similar to Eq. (7). A divergent mode with meridional wavenumber l does not force a rotational flow of the same scale and the pseudocospectrum has a complicated structure (Fig. 5e). The analysis is simplified if we can assume that the vorticity tendency is dominated by β (or more generally, if we can neglect meridional derivatives). In that limit, it is shown in appendix B that for a slowly varying divergence field the response to vortex stretching approaches the response to divergent meridional advection (and is thus determined by the tilt of the divergence phase lines). However, the two responses can differ significantly when the divergence field has meridional structure.
5. Closure of the vorticity balance and divergence phase tilt in observations
Although the simple arguments presented in the previous section can help explain the sensitivity of the idealized model, their actual relevance for the observed momentum flux is a bit more questionable given the crude assumptions of the simple model. In particular, our arguments rely heavily on (a very simplified version of) the vorticity balance that may not be appropriate for observations. The closure of the observed tropical vorticity balance is subject to big errors even in modern reanalysis products (e.g., Yang and Hoskins 2017), as it has been known for a long time that transience and nonlinearity play an important role for this balance (Sardeshmukh and Hoskins 1985).
We contend that the aforementioned difficulties in closing the tropical vorticity balance are mainly due to the small scales, those most affected by nonlinearity, while the large scales found to dominate the meridional eddy momentum transport can be reasonably understood using the idealized vorticity closure equation, Eq. (4). In support of this argument, we show below that the simple relations between the divergent forcing and the rotational response that constrain the sensitivity of the eddy momentum flux in the model are also at work for the observed JJA and DJF flows.
Focusing on JJA first, Fig. 7a shows the k = 1 eddy precipitation field (shading), upper-level (150–300-hPa average) divergence (contours), and upper-level divergent wind vectors. Although the divergent wind is evidently related to the precipitation and divergence fields, it is also apparent that the latter have finer spatial structure than the winds, as expected for a differenced field. This is consistent with our findings that the eddy momentum flux is primarily determined by the gravest meridional modes of the divergence field. As shown in Fig. 7b, the coarse-grained

JJA climatology of k = 1 (a) eddy precipitation (color shading), eddy upper-level divergence (contours), and eddy divergent velocity (blue vectors); (b) coarse-grained
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1

JJA climatology of k = 1 (a) eddy precipitation (color shading), eddy upper-level divergence (contours), and eddy divergent velocity (blue vectors); (b) coarse-grained
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
JJA climatology of k = 1 (a) eddy precipitation (color shading), eddy upper-level divergence (contours), and eddy divergent velocity (blue vectors); (b) coarse-grained
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
The shading in Fig. 7c shows the net vorticity forcing by the (nonfiltered) divergence field. We emphasize with filled red circles (blue diamonds) in this figure the approximate locations of maximum positive (negative) vorticity forcing in both hemispheres. For a divergence field with
On the other hand, the shading in Fig. 7d shows the rotational tendency, linearized about the time- and zonal-mean flow as on the left-hand side of Eq. (2) [note that we took into account the meridional structure of the basic-state
Figure 8 shows a similar analysis during DJF. The negative eddy momentum flux during this season is consistent with the northeast-to-southwest tilt

As in Fig. 7, but for DJF.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1

As in Fig. 7, but for DJF.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
As in Fig. 7, but for DJF.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
6. Concluding remarks
In this paper we have analyzed the mechanisms coupling the divergent and rotational circulations aiming to understand the determination of the observed tropical eddy momentum flux, associated with correlations between the divergent eddy meridional velocity and the rotational eddy zonal velocity. We showed that a simple homogeneous, linear model with uniform basic-state winds can reproduce quite well the observed momentum flux when the rotational circulation is forced by the observed divergent flow, which suggests that the eddy momentum flux can be understood as a response to the divergent forcing. We found our simple model to display very little sensitivity to changes in the mean state, in contrast to previous studies showing significant eddy momentum flux modulation by the Hadley cell (Kraucunas and Hartmann 2007). In the context of a forced linear model, the sensitivity to the basic state is associated with changes in the natural frequency of the system and a possible transition to resonance of the sort described by Arnold et al. (2012). The inevitable nonlinearity near the resonant scale cannot be resolved by our simple model, which uses strong scale-sensitive damping to prevent resonant behavior. We would expect more sensitivity of the solution to the basic-state winds in a nonlinear model that can resolve these and shorter scales.
As changes in the basic state have little bearing on the determination of the eddy momentum flux, the single most important factor affecting the seasonal reversal of the momentum flux in our model is the seasonal variability in the divergent forcing. We found that vortex stretching and divergent beta advection produce responses with the same sign—a momentum flux directed from the winter to the summer hemisphere—but the former is about twice as large. We can understand these responses by noting that for the long waves that dominate the momentum transport the natural frequency
The important role played by the divergence phase tilt underscores a crucial limitation of our analysis: the use of a prescribed divergence field. It is now well recognized that the heating is not independent of the circulation in the tropics. Our results provide an eloquent validation to this axiom: even though the mean heating is thought to be mainly determined by the boundary conditions (Hoskins et al. 1999), we cannot think of any obvious reason why the divergence field should tilt in opposite meridional directions during both solstice seasons. It seems more likely that the seasonal reversal of the phase tilt in observations should be associated with changes in meridional wave propagation as the wave source moves into the summer hemisphere (Dima et al. 2005). Because Rossby waves have westerly pseudomomentum, they transport easterly momentum as they propagate and converge momentum into their source region. In the extratropics the eddy momentum flux is dominated by the rotational flow, so the streamfunction field must tilt westward with latitude away from the source region to produce eddy momentum convergence into that latitude. Along similar lines, we speculate that the observed eastward phase tilt of the divergence field moving from the summer hemisphere to the winter hemisphere is linked to the requirement that the divergent momentum flux, dominant in the tropics, converges momentum into the source region. As Figs. 7b and 8b show, this phase tilt has important implications for the large-scale distribution of precipitation in the tropics.
Under this perspective, what we have described here as a forced rotational response could perhaps be better regarded as coupled rotational–divergent propagation. It is unclear what a theory for this propagation might look like, but we expect it to be different from the divergent extension of the classical rotational theory discussed in appendix A due to the important role of heating. In this scenario of an internally determined divergence field, it is plausible that the divergence phase tilt could be affected by the mean meridional flow (e.g., through its impact on meridional propagation), which would explain our different conclusions from Kraucunas and Hartmann (2007) on the sensitivity of the eddy momentum fluxes to changes in the Hadley cell. It would be of interest to investigate what factors determine the divergence phase tilt in an idealized moist model similar to that of Shaw (2014), in which the divergence field is internally determined.
Although we have focused on the dominant stationary eddy momentum transport in this paper, similar constraints apply to propagating tropical modes in the corotating reference system, with implications for the latitudinal distribution of precipitation. A good example is the MJO. As shown in Part I, the MJO also produces divergent eddy momentum transport into the Northern (Southern) Hemisphere during JJA (DJF), even if this transport is much weaker than that by the climatological stationary wave (Lee 1999). The left panels of Fig. 9 show regressions of upper-level divergence and

(top) NDJFM, (middle) JJAS, and (bottom) full-year MJO regressions of (left) upper-level divergence and
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1

(top) NDJFM, (middle) JJAS, and (bottom) full-year MJO regressions of (left) upper-level divergence and
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
(top) NDJFM, (middle) JJAS, and (bottom) full-year MJO regressions of (left) upper-level divergence and
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0304.1
To conclude, we note that the arguments presented in this paper are based on an inviscid closure of the vorticity balance. In this limit, we have shown that for long zonal waves the streamfunction minimum is shifted a quarter wavelength to the west of the maximum positive vorticity forcing. In the presence of friction we expect this phase shift to be reduced, with the vorticity maximum–streamfunction minimum catching up in phase with the vorticity forcing in the frictionally dominated limit. Friction may be due to cumulus mixing or represent the damping effect of transience and nonlinearity on vorticity. Whatever the source, friction is an essential ingredient to the Gill (1980) problem and is thought to be important for the tropical circulation more generally. Friction is responsible for instance for preventing the westward and eastward spreading of the Rossby and Kelvin components of the Gill response and confining zonally the response to localized heating. We have shown that friction is also necessary to prevent resonance in the presence of a strong Hadley cell. However, our results suggest that friction may only be needed at the small scales, at least during JJA, when the inviscid, large-scale vorticity balance can be closed linearly to reasonable accuracy.
Acknowledgments
We thank three anonymous reviewers for suggestions that improved the focus and the clarity of this manuscript. P.Z.-G. acknowledges financial support by Grant CGL2015-72259-EXP from the State Research Agency of Spain.
APPENDIX A
Divergent Eddy Momentum Flux in Unforced Propagation
Section 3 reviews the theory of unforced, barotropic Rossby wave propagation in the presence of a basic state with both zonal and meridional components:
























At first sight, this result seems consistent with the observations. Comparing the solid and dashed black lines in Fig. 1, we can see that the rotational and divergent contributions to the cross-equatorial stationary eddy momentum flux have the same sign through the seasonal cycle (see also Figs. 7 and 8 in Part I). However, while the divergent flux is strongly dominated by the k = 1 component (red dashed line), the rotational momentum flux by this wavenumber (red solid line) is negligible for all seasons but DJF, when its sign is opposite to that of
As discussed in Part I, while the rotational momentum fluxes have an extratropical origin (they diverge through the tropics), the divergent momentum fluxes are associated with a region of strong momentum convergence/eddy generation adjacent to the equator in the summer hemisphere. The implication is that one cannot regard
APPENDIX B
Divergent Momentum Flux Response When the Rotational Tendency Is Dominated by β



























If
REFERENCES
Adames, Á. F., and J. M. Wallace, 2014a: Three-dimensional structure and evolution of the MJO and its relation to the mean flow. J. Atmos. Sci., 71, 2007–2026, https://doi.org/10.1175/JAS-D-13-0254.1.
Adames, Á. F., and J. M. Wallace, 2014b: Three-dimensional structure and evolution of the vertical velocity and divergence fields in the MJO. J. Atmos. Sci., 71, 4661–4681, https://doi.org/10.1175/JAS-D-14-0091.1.
Adames, Á. F., J. M. Wallace, and J. M. Monteiro, 2016: Seasonality of the structure and propagation characteristics of the MJO. J. Atmos. Sci., 73, 3511–3526, https://doi.org/10.1175/JAS-D-15-0232.1.
Adler, R. F., and Coauthors, 2003: The version-2 Global Precipitation Climatology Project (GPCP) monthly precipitation analysis (1979–present). J. Hydrometeor., 4, 1147–1167, https://doi.org/10.1175/1525-7541(2003)004<1147:TVGPCP>2.0.CO;2.
Arnold, N. P., E. Tziperman, and B. Farrell, 2012: Abrupt transition to strong superrotation driven by equatorial wave resonance in an idealized GCM. J. Atmos. Sci., 69, 626–640, https://doi.org/10.1175/JAS-D-11-0136.1.
Dee, D. P., and Coauthors, 2011: The Era-Interim reanalysis: Configuration and performance of the data assimilation system. Quart. J. Roy. Meteor. Soc., 137, 553–597, https://doi.org/10.1002/qj.828.
Dima, I. M., J. M. Wallace, and I. Kraucunas, 2005: Tropical zonal momentum balance in the NCEP reanalyses. J. Atmos. Sci., 62, 2499–2513, https://doi.org/10.1175/JAS3486.1.
Gill, A., 1980: Some simple solutions for heat-induced tropical circulation. Quart. J. Roy. Meteor. Soc., 106, 447–462, https://doi.org/10.1002/qj.49710644905.
Hoskins, B. J., and D. J. Karoly, 1981: The steady linear response of a spherical atmosphere to thermal and orographic forcing. J. Atmos. Sci., 38, 1179–1196, https://doi.org/10.1175/1520-0469(1981)038<1179:TSLROA>2.0.CO;2.
Hoskins, B. J., and T. Ambrizzi, 1993: Rossby wave propagation on a realistic longitudinally varying flow. J. Atmos. Sci., 50, 1661–1671, https://doi.org/10.1175/1520-0469(1993)050<1661:RWPOAR>2.0.CO;2.
Hoskins, B. J., R. Neale, M. Rodwell, and G.-Y. Yang, 1999: Aspects of the large-scale tropical atmospheric circulation. Tellus, 51B, 33–44, https://doi.org/10.3402/tellusb.v51i1.16258.
Kraucunas, I., and D. L. Hartmann, 2007: Tropical stationary waves in a nonlinear shallow-water model with realistic basic states. J. Atmos. Sci., 64, 2540–2557, https://doi.org/10.1175/JAS3920.1.
Lee, S., 1999: Why are the climatological zonal winds easterly in the equatorial upper troposphere? J. Atmos. Sci., 56, 1353–1363, https://doi.org/10.1175/1520-0469(1999)056<1353:WATCZW>2.0.CO;2.
Li, Y., J. Li, F. F. Jin, and S. Zhao, 2015: Interhemispheric propagation of stationary Rossby waves in a horizontally nonuniform background flow. J. Atmos. Sci., 72, 3233–3256, https://doi.org/10.1175/JAS-D-14-0239.1.
Monteiro, J. M., A. F. Adames, J. M. Wallace, and J. S. Sukhatme, 2014: Interpreting the upper level structure of the Madden-Julian oscillation. Geophys. Res. Lett., 41, 9158–9165, https://doi.org/10.1002/2014GL062518.
Sardeshmukh, P. D., and B. J. Hoskins, 1985: Vorticity balances in the tropics during the 1982-83 El Niño-Southern Oscillation event. Quart. J. Roy. Meteor. Soc., 111, 261–278, https://doi.org/10.1256/smsqj.46801.
Sardeshmukh, P. D., and B. J. Hoskins, 1988: The generation of global rotational flow by steady idealized tropical divergence. J. Atmos. Sci., 45, 1228–1251, https://doi.org/10.1175/1520-0469(1988)045<1228:TGOGRF>2.0.CO;2.
Schneider, E. K., and I. G. Watterson, 1984: Stationary Rossby wave propagation through easterly layers. J. Atmos. Sci., 41, 2069–2083, https://doi.org/10.1175/1520-0469(1984)041<2069:SRWPTE>2.0.CO;2.
Schubert, W. H., R. K. Taft, and L. G. Silvers, 2009: Shallow water quasi-geostrophic theory on the sphere. J. Adv. Model. Earth Syst., 1 (2), https://doi.org/10.3894/JAMES.2009.1.2.
Shaw, T. A., 2014: On the role of planetary-scale waves in the abrupt seasonal transition of the Northern Hemisphere general circulation. J. Atmos. Sci., 71, 1724–1746, https://doi.org/10.1175/JAS-D-13-0137.1.
Showman, A. P., and L. M. Polvani, 2011: Equatorial superrotation on tidally locked exoplanets. Astrophys. J., 738, 71, https://doi.org/10.1088/0004-637X/738/1/71.
Webster, P. J., and J. R. Holton, 1982: Cross-equatorial response to middle-latitude forcing in a zonally varying basic state. J. Atmos. Sci., 39, 722–733, https://doi.org/10.1175/1520-0469(1982)039<0722:CERTML>2.0.CO;2.
Yang, G.-Y., and B. J. Hoskins, 2017: The equivalent barotropic structure of waves in the tropical atmosphere in the Western Hemisphere. J. Atmos. Sci., 74, 1689–1704, https://doi.org/10.1175/JAS-D-16-0267.1.
Zurita-Gotor, P., 2019: The role of the divergent circulation for large-scale eddy momentum transport in the tropics. Part I: Observations. J. Atmos. Sci., 76, 1125–1144, https://doi.org/10.1175/JAS-D-18-0297.1.
Zurita-Gotor, P., and P. Álvarez-Zapatero, 2018: Coupled interannual variability of the Hadley and Ferrel cells. J. Climate, 31, 4757–4773, https://doi.org/10.1175/JCLI-D-17-0752.1.
Zurita-Gotor, P., and I. M. Held, 2018: The finite-amplitude evolution of mixed Kelvin–Rossby wave instability and equatorial superrotation in a shallow-water model and an idealized GCM. J. Atmos. Sci., 75, 2299–2316, https://doi.org/10.1175/JAS-D-17-0386.1.
The fact that the divergent momentum flux is sensitive to the range of variability of f, rather than to its mean value, is key to the impact of vortex stretching for the model’s response in spite of the small values of f near the equator. These arguments also suggest that the impact of vortex stretching will be larger for waves of longer meridional scales.
Note that only the divergence field is coarse grained in Figs. 7 and 8, but not the vorticity forcing or the rotational response. The filtered fields are in better agreement (not shown).
For the different MJO phases, these regressions are computed as linear combinations of the regressions on the leading velocity potential EOFs defining the MJO (Adames and Wallace 2014a; see also Part I).