1. Introduction
This paper is concerned with the dynamics of meridional eddy momentum transport in the tropical troposphere. While early paradigms of tropical momentum transport envisioned a nearly inviscid Hadley cell decelerated by Rossby waves of extratropical origin, Lee (1999) showed that in fact, the westerly eddy forcing compensates the deceleration by the off-equatorial Hadley cell in the deep tropics. As this eddy forcing plays a key role in theories of equatorial superrotation (Hide 1969), the maintenance of the tropical momentum budget has received much attention in idealized studies of planetary atmospheres (Showman and Polvani 2011; Pinto and Mitchell 2014; Laraia and Schneider 2015). It has also been suggested that the terrestrial atmosphere might transition to a superrotating state in a much warmer climate (Caballero and Huber 2010).
Dima et al. (2005) performed a detailed analysis of the tropical momentum budget in the terrestrial atmosphere and found the eddy forcing to be dominated by the stationary meridional momentum convergence. They proposed that the observed momentum transport could be associated with the atmospheric response to asymmetric tropical heating as in the classical Gill (1980) problem, though they did not study the dynamics of this response in detail. There is also modeling evidence from idealized studies (Suarez and Duffy 1992; Saravanan 1993; Kraucunas and Hartmann 2005) that asymmetric tropical heating produces a westerly acceleration in the deep tropics and can lead to superrotation when the Hadley cell upwelling is not too far off the equator (Kraucunas and Hartmann 2005). In all these studies, the posited mechanism for the tropical acceleration is the meridional propagation of Rossby waves forced by the heating, similar to the dynamics of the extratropical eddy-driven jet. An alternative perspective was provided by Showman and Polvani (2011), who studied the dynamics of equatorial acceleration in Gill-like settings motivated by the superrotation of tidally locked planets. These authors noted some problems with the classical paradigm of jet driving by meridionally propagating Rossby waves in the tropics, and proposed an alternative mechanism based on the interaction between the Kelvin and Rossby components of the Matsuno–Gill response. Zurita-Gotor and Held (2018) have recently argued that the same mechanism could drive the superrotation of nonconvecting atmospheres at large thermal Rossby number, except that the Kelvin and Rossby waves would be forced by an ageostrophic instability in this case (Iga and Matsuda 2005; Wang and Mitchell 2014).
A fundamental concept in our understanding of wave–mean flow interaction is the link between eddy dissipation and mean-flow acceleration (Andrews and McIntyre 1976). Zurita-Gotor and Held (2018) identify two different routes to eddy dissipation in planetary atmospheres that can lead to the spinup of eddy-driven jets. The first mechanism involves dissipation by breaking vorticity waves (i.e., vorticity mixing along isentropic surfaces). This produces an easterly acceleration over the breaking region (if the vorticity gradient is positive), which is compensated by westerly acceleration over the source region when these two regions are distinct. This mechanism is thought to be responsible for the spinup of the extratropical jet (Vallis 2006) and may also give rise to tropical eddy-driven jets in the presence of a robust tropical vorticity source (Suhas et al. 2017). Alternatively, the dissipation driving the mean flow change could be diabatic. The latter appears to be the underlying mechanism for superrotation in the simulations of Showman and Polvani (2011) and Zurita-Gotor and Held (2018), in which the tropical vorticity fluxes are weak and/or negative, and the westerly acceleration is imparted by vertical (cross isentropic) momentum advection instead.
The weakness of the Rossby wave source in the deep tropics (Sardeshmukh and Hoskins 1988) and the diabatic character of the tropical circulation suggest that the second mechanism might also be more relevant for eddy momentum forcing in the terrestrial atmosphere. Our study aims to test this conjecture and to investigate its implications for eddy momentum transport in the tropical troposphere more broadly. We will show that since the divergent flow is an important component of tropical eddy momentum transport, the mechanism of meridional momentum transport is different from that in the extratropics. We will also investigate the role of the different components of the climatological-mean tropical circulation in the momentum transport and characterize the momentum transport by the Madden–Julian oscillation (MJO). We present a comprehensive description of the observed tropical momentum flux in this paper while a companion study (Zurita-Gotor 2019, manuscript submitted to J. Atmos. Sci., hereafter Part II) will investigate the dynamical determinants of this momentum flux.
This paper is organized as follows. Section 2 introduces the data and methodology used. Section 3 provides some context by reviewing the seasonally varying tropical momentum balance and the impact of eddy forcing for angular momentum conservation and superrotation. Section 4 performs a meridional–vertical decomposition of the eddy momentum forcing using the advective and flux forms. It is shown that while meridional propagation dominates, the momentum mixing occurs primarily in the vertical near the equator. Section 5 discusses two different mechanisms that can produce meridional eddy momentum convergence, which are relevant for tropical and extratropical eddy-driven jets, respectively. Section 6 analyzes the dominant contributions to the meridional eddy momentum flux, and section 7 focuses on the MJO. We close with a brief summary and some discussion in section 8.
2. Data and methods
We use in this study four-times daily gridded wind data (u, υ, ω) at 2.5° horizontal resolution from ERA-Interim (Dee et al. 2011) for the 37-yr period spanning from 1979 to 2016. For some of our analyses, the horizontal wind field (u, υ) is decomposed into its nondivergent (rotational) and irrotational (divergent) components, which will be denoted using r and d subscripts in the following. The rotational wind is defined in terms of the streamfunction ψ, obtained by inversion of the vorticity field ξ, and the divergent wind is defined in terms of the velocity potential χ, obtained by inversion of the divergence field D.
Eddy components, denoted with primes, are defined as deviations from zonal averages (indicated by overbars), and eddy fluxes are computed by zonally averaging the products of the daily mean eddy anomalies. Seasonal cycles of individual variables and eddy fluxes are defined in terms of their daily climatologies over the 37-yr record. For display purposes, this daily climatology is smoothed using a 20-day running mean.
For each calendar month, we define the stationary eddy component of a variable as the 37-yr monthly mean eddy climatology for that variable. Seasonal cycles of stationary eddy fluxes are then computed by zonally averaging the products of these stationary eddy components for each month. We also define seasonal cycles in the transient eddy fluxes as the differences between the seasonal cycles in the full eddy fluxes (averaged for each calendar month) and in the stationary eddy fluxes. These transient eddy fluxes include variability at both submonthly and interannual time scales, but some of the slow intraseasonal variability may be contaminated into the seasonal variability due to the reduced record. This could affect in particular the MJO variability, with periods on the order of a month.
To investigate the impact of the MJO in more detail, we use regression analysis based on the MJO indices of Adames and Wallace (2014). These authors define the MJO cycle in terms of the two leading global EOFs of the velocity potential difference between 150 and 850 hPa, computed after removing the seasonal cycle and low-frequency (T > 120 days) variability from the data. The two leading EOFs (with similar explained variances) represent structures in near quadrature, and the corresponding principal components are very highly correlated at lags of about a week, as expected for a propagating mode of variability. Regressions for the different MJO phases are computed as linear combinations of the regressions on the two standardized principal components characterizing the in-quadrature MJO phases. We perform independent analyses for the extended winter [November–March (NDJFM)] and summer [June–September (JJAS)] seasons following Adames et al. (2016), and assess statistical significance using a conservative decorrelation time scale of 10 days to estimate the number of degrees of freedom (Adames and Wallace 2014).
3. The seasonally varying tropical momentum balance
Figure 1a shows the seasonal cycle of the net [meridional plus vertical; see Eq. (3) in the next section] upper-troposphere zonal-mean eddy acceleration, vertically averaged between 300 and 150 hPa. As noted by Dima et al. (2005), tropical eddies transport momentum from the winter to the summer hemisphere against the Hadley cell meridional flow (illustrated with arrows). The eddy forcing in Fig. 1a includes contributions by both transient and stationary eddies. The stationary component dominates the full eddy forcing during DJF but the transient eddy forcing is also important during JJA (Fig. 1b).

(a) Seasonal cycle of net eddy acceleration (m s−1 day−1), with the zero contour shown with a thick black dotted line. The red line shows the vorticity equator (latitude with
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

(a) Seasonal cycle of net eddy acceleration (m s−1 day−1), with the zero contour shown with a thick black dotted line. The red line shows the vorticity equator (latitude with
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
(a) Seasonal cycle of net eddy acceleration (m s−1 day−1), with the zero contour shown with a thick black dotted line. The red line shows the vorticity equator (latitude with
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1




The above balance requires that the meridional momentum flux is maximized near the “vorticity equator,” which we define as the latitude where





Consistent with the arguments of Lindzen and Hou (1988), the equatorial winds are easterly during most of the year, with the exception of weak westerlies during boreal winter (Fig. 1d, black line). Nevertheless, the westerly acceleration by the eddy forcing over the summer hemisphere significantly weakens the easterlies that would be observed if angular momentum were conserved. We can get an indication of the impact of the eddy forcing on angular momentum conservation by comparing the observed wind with the wind that would be observed if angular momentum were conserved. The red line in Fig. 1d shows the equatorial wind that would be observed if the Hadley cell conserved the planetary angular momentum at the latitude of maximum zonal-mean upwelling, while the blue line shows the same using the mean planetary momentum over the latitudes of zonal-mean tropical upwelling (weighted by the magnitude of this upwelling). It is apparent that the seasonal migration of the upwelling plays an important role for the seasonal cycle of the equatorial wind, even though the Hadley cell is far from angular momentum conserving.
Discussions about the plausibility of terrestrial superrotation usually revolve around the annual-mean equatorial momentum balance. In this context, the analysis of Lee (1999) shows that the main deceleration term on the equator arises from temporal correlations in the seasonal cycles of
4. Meridional–vertical decomposition of the eddy forcing: Advective and flux forms




(a) Full eddy forcing over the 300–150-hPa tropospheric layer as in Fig. 1a and its decomposition into (b) meridional and (c) vertical eddy momentum convergence, and (d) meridional eddy vorticity flux and (e) eddy vertical advection. (f),(g) The decomposition of the meridional eddy momentum convergence in (b) into rotational and divergent contributions, as defined in the text. For clarity, the daily climatologies have been smoothed using a 20-day running mean. The thick black dotted lines mark the zero contours and all panels use the same color bar (m s−1 day−1).
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

(a) Full eddy forcing over the 300–150-hPa tropospheric layer as in Fig. 1a and its decomposition into (b) meridional and (c) vertical eddy momentum convergence, and (d) meridional eddy vorticity flux and (e) eddy vertical advection. (f),(g) The decomposition of the meridional eddy momentum convergence in (b) into rotational and divergent contributions, as defined in the text. For clarity, the daily climatologies have been smoothed using a 20-day running mean. The thick black dotted lines mark the zero contours and all panels use the same color bar (m s−1 day−1).
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
(a) Full eddy forcing over the 300–150-hPa tropospheric layer as in Fig. 1a and its decomposition into (b) meridional and (c) vertical eddy momentum convergence, and (d) meridional eddy vorticity flux and (e) eddy vertical advection. (f),(g) The decomposition of the meridional eddy momentum convergence in (b) into rotational and divergent contributions, as defined in the text. For clarity, the daily climatologies have been smoothed using a 20-day running mean. The thick black dotted lines mark the zero contours and all panels use the same color bar (m s−1 day−1).
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

For the seasons indicated, vertical profiles of (first row) full eddy acceleration and its decomposition into (second row) meridional and (third row) vertical eddy momentum convergence, and (fourth row) meridional eddy vorticity flux and (fifth row) eddy vertical advection. Units are m s−1 day−1, with the zero contour in thick black dotted line.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

For the seasons indicated, vertical profiles of (first row) full eddy acceleration and its decomposition into (second row) meridional and (third row) vertical eddy momentum convergence, and (fourth row) meridional eddy vorticity flux and (fifth row) eddy vertical advection. Units are m s−1 day−1, with the zero contour in thick black dotted line.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
For the seasons indicated, vertical profiles of (first row) full eddy acceleration and its decomposition into (second row) meridional and (third row) vertical eddy momentum convergence, and (fourth row) meridional eddy vorticity flux and (fifth row) eddy vertical advection. Units are m s−1 day−1, with the zero contour in thick black dotted line.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
Based on the flux formulation, the 300–150-hPa eddy acceleration is dominated by the meridional momentum flux convergence (Figs. 2b,c), consistent with the results of Dima et al. (2005). However, the partition is somewhat sensitive to the vertical levels used. As shown in Fig. 3, the vertical momentum flux convergence is weaker than the meridional convergence but not entirely negligible. The vertical convergence partially balances the meridional convergence, strongly peaked at upper levels, and spreads the eddy forcing over a deep layer, especially during DJF. Dima et al. (2005) described a similar momentum flux pattern, which is ubiquitous in comprehensive and idealized models (e.g., Kraucunas and Hartmann 2005; Lutsko 2018; Caballero and Carlson 2018), although the degree of meridional–vertical cancellation differs.
However, the decomposition is quite different when an advective formulation is used [rightmost equality in Eq. (3)]. The meridional advection term is written here as a meridional eddy vorticity flux1 so as to emphasize the vorticity dynamics. As noted in the introduction, the classical paradigm for eddy-driven jets is based on the notion that Rossby waves transport easterly momentum as they propagate (Held 1975). In the extratropics, Rossby waves generated by baroclinic instability propagate equatorward and produce downgradient (negative) vorticity fluxes when they break in the subtropics. These are compensated by positive (upgradient) vorticity fluxes over the source region, which impart a westerly acceleration in the midlatitudes (Vallis 2006). A similar mechanism has been proposed to account for the westerly eddy forcing in the tropics, except that Rossby waves are presumed to be generated by asymmetric tropical heating (see, e.g., Saravanan 1993).
The advective decomposition of the eddy forcing (Figs. 2d,e) is qualitatively consistent with this picture during JJA, which shows a broad NH region with upgradient vorticity fluxes extending all the way to the equator. In contrast, the SH region with upgradient vorticity flux is weak and spatially confined during DJF, and it does not reach the equator. During this season, the westerly acceleration at the equator is due instead to vertical momentum advection (Fig. 2e). Similar results are found for both equinox seasons so that, except for the JJA case discussed above, the tropical vorticity fluxes are generally negative (more consistent with a Rossby wave sink than with a Rossby wave source) and the westerly acceleration is imparted by the vertical advection term. Figure 3 shows that this term can also capture the deep vertical structure of the full eddy forcing.


Motivated by the relation between wave propagation and Reynolds stresses [Andrews et al. 1987; note that the baroclinic contribution to the vertical Eliassen–Palm (EP) flux is much smaller than
The dominance of vertical advection and mixing is connected to the fact that the eddy dissipation driving the mean-flow acceleration in the tropics is diabatic, in contrast with the mechanical dissipation (irreversible Rossby wave breaking) that is key to the spinup of the extratropical jet (Vallis 2006, p. 489). As discussed by Zurita-Gotor and Held (2018), in the presence of a positive-definite vorticity gradient, a westerly eddy-driven acceleration requires either a vorticity source or cross-isentropic momentum advection. In the tropics, vertical advection provides a good approximation to cross-isentropic advection:
5. Rotational and divergent contributions to the meridional eddy momentum flux convergence




(a) Annual-mean eddy meridional momentum convergence (m s−1 day−1) and contributions by (b) total vorticity flux, (c)
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

(a) Annual-mean eddy meridional momentum convergence (m s−1 day−1) and contributions by (b) total vorticity flux, (c)
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
(a) Annual-mean eddy meridional momentum convergence (m s−1 day−1) and contributions by (b) total vorticity flux, (c)
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
Equation (4) implies that the familiar association between precipitation/upper-level divergence and anomalous upper-level tropical easterlies is associated with momentum transport onto the equator. Figure 5 illustrates the mechanism involved in this meridional momentum transport. A negative

Sketch illustrating the mechanism of eddy momentum convergence by the divergent circulation (wide orange arrows). When
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

Sketch illustrating the mechanism of eddy momentum convergence by the divergent circulation (wide orange arrows). When
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
Sketch illustrating the mechanism of eddy momentum convergence by the divergent circulation (wide orange arrows). When
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1










As will be shown below, the first of the two components on the right-hand side of Eq. (6) strongly dominates:



To conclude, Fig. 6 shows the partition of the meridional momentum flux convergence for the two solstice seasons, with results similar to those presented above. As for the annual mean, the equatorial convergence is dominated by the divergent momentum flux in both seasons, though the rotational component has a small contribution during JJA. The role of this term is nevertheless much less important than suggested by the seasonal cycle of the full eddy vorticity flux (the dominant JJA forcing; cf. Fig. 2d), since a significant fraction of this vorticity flux is in fact associated with

As in Figs. 4a, 4d, and 4f, but for (a),(c),(e) DJF and (b),(d),(f) JJA.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

As in Figs. 4a, 4d, and 4f, but for (a),(c),(e) DJF and (b),(d),(f) JJA.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
As in Figs. 4a, 4d, and 4f, but for (a),(c),(e) DJF and (b),(d),(f) JJA.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
6. Analysis of the divergent momentum flux
We have shown above that eddy momentum flux convergence in the tropics is dominated by the divergent momentum flux. However, the same is not necessarily true for the momentum flux, because the above diagnostics could also be compatible with a cross-equatorial rotational momentum flux connecting the two subtropical hemispheres, with weak convergence in the deep tropics. Motivated by recent findings on such an interhemispheric teleconnection (Ji et al. 2014; Li et al. 2015), Figs. 7 and 8 show the partitioning of the tropical momentum flux into its rotational and divergent components during DJF and JJA, respectively.

During DJF, (a) rotational (blue) and divergent (red; the
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

During DJF, (a) rotational (blue) and divergent (red; the
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
During DJF, (a) rotational (blue) and divergent (red; the
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

As in Fig. 7, but for JJA.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

As in Fig. 7, but for JJA.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
As in Fig. 7, but for JJA.
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
During both solstice seasons, wave propagation is directed from the summer to the winter hemisphere (Figs. 7a, 8a). The cross-equatorial rotational (blue) and divergent (red) momentum fluxes are of the same sign, but the latter are more than twice as large on the equator. We can also see that the divergent eddy momentum flux is strongly dominated by the
In contrast, the rotational momentum flux is associated in both hemispheres and seasons with waves generated at higher latitudes. During DJF the rotational momentum flux diverges throughout the tropics but during JJA there is a narrow band of (weak) eddy momentum flux convergence slightly to the north of the equator. In the deep tropics the rotational momentum flux is larger in the summer than in the winter hemisphere but the contrary is true poleward of about 15°. The latter is consistent with the seasonal cycle in extratropical eddy generation, which gives rise to stronger Rossby wave propagation into the winter subtropics. However, because the rotational wave flux does not penetrate deep into the tropics in this hemisphere, close to the equator propagation from the summer hemisphere dominates. This may reflect the asymmetry in propagation introduced by the Hadley cell flow (Schneider and Watterson 1984; Li et al. 2015), or it could be due to the stronger Rossby wave forcing in the summer tropics.
The bottom panels of Figs. 7 and 8 show with dashed lines the rotational and divergent momentum fluxes and flux convergence due to the stationary eddies alone. Dima et al. (2005) already noted the dominance of the stationary momentum flux in the tropics but it is interesting to perform the analysis for the rotational and divergent components separately. As expected, the stationary rotational momentum flux is larger in the Northern Hemisphere than in the Southern Hemisphere, particularly during DJF, when extratropical planetary waves propagating into the tropics are known to force the Hadley cell (Zurita-Gotor and Álvarez-Zapatero 2018). In contrast, the stationary divergent momentum flux has a very similar structure to the full divergent momentum flux: it peaks near the equator in the summer hemisphere and decays away rapidly on the winter side of the equator. It is apparent that the stationary wave component dominates the divergent momentum flux, especially during DJF.
Figures 9a and 9b show the climatological-mean upper-level eddy divergence

(a) DJF climatology of upper-level (150–300-hPa average) eddy divergence (shading; day−1) and
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

(a) DJF climatology of upper-level (150–300-hPa average) eddy divergence (shading; day−1) and
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
(a) DJF climatology of upper-level (150–300-hPa average) eddy divergence (shading; day−1) and
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
Figure 10 describes the zonal structure contributing to the stationary momentum transport in more detail. The top two panels show the seasonal cycles of the upper-troposphere, monthly-mean

(left) Annual cycle of the monthly mean eddy wind components (top)
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

(left) Annual cycle of the monthly mean eddy wind components (top)
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
(left) Annual cycle of the monthly mean eddy wind components (top)
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
Although one can capture the latitudinal structure and much of the amplitude of the zonal-mean
7. MJO eddy momentum transport
We showed above that the stationary wave contribution dominates the divergent eddy momentum flux in the tropics (consistent with Dima et al. 2005), with some seasonal differences. The transient contribution is negligible during DJF (Fig. 7d) but nearly as large as the stationary contribution during JJA (Fig. 8d; see also Figs. 1a,b). Lee (1999) used spectral analysis (in annual data) to show that the transient eddy momentum flux is dominated by interannual variability in the tropics, with the MJO playing a much smaller role. However, the MJO impact on momentum transport may be enhanced in a warmer climate and potentially lead to superrotation (Lee 1999; Caballero and Huber 2010). Motivated by this, we investigate in this section the dynamics of the MJO momentum transport by means of regression analysis using the diagnostics of Adames and Wallace (2014).
The top three panels of Fig. 11 show regressions of upper-level divergence (shading) and

Regression of upper-level (150–300-hPa average) divergence (shading; day−1) and
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

Regression of upper-level (150–300-hPa average) divergence (shading; day−1) and
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
Regression of upper-level (150–300-hPa average) divergence (shading; day−1) and
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
Figure 12a shows the evolution of the NDJFM MJO-driven meridional eddy momentum flux through an MJO period, calculated by averaging zonally the products of the zonal and meridional wind anomalies for the different MJO phases (phases are expressed here in terms of the longitude of maximum

(a) MJO momentum flux (zonal mean of the product of the regressed MJO velocity anomalies) as a function of the MJO phase, expressed in terms of the longitude of maximum velocity potential. (b) As in (a), but for the divergent contribution
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

(a) MJO momentum flux (zonal mean of the product of the regressed MJO velocity anomalies) as a function of the MJO phase, expressed in terms of the longitude of maximum velocity potential. (b) As in (a), but for the divergent contribution
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
(a) MJO momentum flux (zonal mean of the product of the regressed MJO velocity anomalies) as a function of the MJO phase, expressed in terms of the longitude of maximum velocity potential. (b) As in (a), but for the divergent contribution
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
On the other hand, Grise and Thompson (2012) have shown that tropical eddy momentum transport is significantly modulated by the MJO, with the regressed eddy momentum flux exceeding 3 m2 s−2 (cf. their Fig. 9). Although these anomalies are larger than the MJO fluxes in Figs. 12a and 12b, they produce no net momentum transport when averaged over an MJO period. This is illustrated in Fig. 12e, which shows the regression of the daily eddy momentum flux on the MJO as a function of phase for the NDJFM season. We can see that tropical eddy momentum transport is enhanced when the
8. Summary and concluding remarks
In this work, we have analyzed the role played by the divergent circulation for tropical eddy momentum transport. Although the importance of the divergent flow is of course well recognized in the tropics, the implications for meridional eddy momentum transport appear not to have been investigated before. Since the role of tropical eddies in the determination of the climatological-mean tropical winds has received little attention until relatively recently (Lee 1999; Dima et al. 2005), the dynamics of tropical eddy momentum transport has not been studied in as much depth as other aspects of the tropical circulation.
The important role of the divergent momentum transport is a consequence of the diabatic character of the tropical circulation. Although wave propagation is predominantly meridional (implying dominance of meridional eddy momentum fluxes), the bulk of the forcing/dissipation is associated with cross-isentropic momentum transport and vertical mixing. This entails a very different mechanism of meridional eddy momentum flux convergence from the extratropical jet, associated with mass convergence over sectors with anomalous westerlies and mass divergence over sectors with anomalous easterlies, rather than with vorticity fluxes. This mechanism is similar to that described by Zurita-Gotor and Held (2018) for coupled Kelvin–Rossby instability, except that the cross-isentropic momentum transport is presumably due to convective heating in the terrestrial atmosphere.
The dominant component of the eddy momentum flux in the tropics is associated with correlations between the rotational zonal velocity and the divergent meridional velocity:
Because the tropical momentum flux involves both rotational and divergent anomalies, one cannot infer the sign of the eddy momentum flux from the meridional tilt of the isohypses alone—the full velocity vectors need to be considered. The sign of the divergent eddy momentum flux is instead determined by the relative phases of the divergent and rotational anomalies, which requires some coupling between them. We can envision two simple, idealized models for this coupling. First, one could have nearly balanced, remotely forced rotational waves that develop a secondary divergent circulation through adjustment to balance as they propagate into the tropics, similar to the ageostrophic circulation of extratropical Rossby waves. This could be a reasonable model for weakly convecting atmospheres. Alternatively (and more plausibly for the strongly convecting Earth), we might envision a diabatically forced divergent circulation generating Rossby waves through vortex stretching and/or meridional advection by the divergent component of the flow (Sardeshmukh and Hoskins 1988).
The key question in this latter scenario is what determines the phase relation between the divergent forcing and the rotational response. Assuming that the forced rotational flow will adjust to balance the vorticity forcing by the divergent flow, the answer will depend on the closure of the vorticity balance. For instance, consider the simple thought scenario of a cross-equatorial southward Hadley circulation with uniform meridional velocity V at upper levels, that is locally enhanced by anomalous southward flow

(a) Sketch illustrating two alternative closures of the eddy vorticity balance when the zonal-mean Hadley circulation V is locally enhanced over some longitudes by the anomalous flow
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1

(a) Sketch illustrating two alternative closures of the eddy vorticity balance when the zonal-mean Hadley circulation V is locally enhanced over some longitudes by the anomalous flow
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
(a) Sketch illustrating two alternative closures of the eddy vorticity balance when the zonal-mean Hadley circulation V is locally enhanced over some longitudes by the anomalous flow
Citation: Journal of the Atmospheric Sciences 76, 4; 10.1175/JAS-D-18-0297.1
On the other hand, if we assume that V is so strong that the vorticity tendency is dominated by meridional vorticity advection by the Hadley flow and the beta term is negligible,
We conclude with a caveat. As most previous studies (e.g., Dima et al. 2005), we followed in this work the conventional approach of decomposing the circulation into zonal-mean and eddy components. Given our findings that the eddy momentum transport is dominated by the gravest zonal component and can be largely understood as a zonal modulation of the symmetric Hadley cell, an alternative model of a nonzonal Hadley circulation (Schneider 1987; Hsu and Plumb 2000) may be preferable.
Acknowledgments
We are grateful to Sukyoung Lee and two anonymous reviewers for suggestions that improved this manuscript. P.Z.-G. acknowledges financial support by Grant CGL2015-72259-EXP by the State Research Agency of Spain. This work benefited from helpful discussions with I. Held during a series of summer visits to Princeton funded by NSF Grant AGS-1733818.
REFERENCES
Adames, Á. F., and J. M. Wallace, 2014: 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, 2017: On the tropical atmospheric signature of El Niño. J. Atmos. Sci., 74, 1923–1939, https://doi.org/10.1175/JAS-D-16-0309.1.
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.
Andrews, D. G., and M. E. McIntyre, 1976: Planetary waves in horizontal and vertical shear: The generalized Eliassen-Palm relation and the 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., J. R. Holton, and C. B. Leovy, 1987: Middle Atmosphere Dynamics. International Geophysics Series, Vol. 40, Academic Press, 489 pp.
Bordoni, S., and T. Schneider, 2008: Monsoons as eddy-mediated regime transitions of the tropical overturning circulation. Nat. Geosci., 1, 515–519, https://doi.org/10.1038/ngeo248.
Caballero, R., and M. Huber, 2010: Spontaneous transition to superrotation in warm climates simulated by CAM3. Geophys. Res. Lett., 37, L11701, https://doi.org/10.1029/2010GL043468.
Caballero, R., and H. Carlson, 2018: Surface superrotation. J. Atmos. Sci., 75, 3671–3689, https://doi.org/10.1175/JAS-D-18-0076.1.
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.
Grise, K. M., and D. W. Thompson, 2012: Equatorial planetary waves and their signature in atmospheric variability. J. Atmos. Sci., 69, 857–874, https://doi.org/10.1175/JAS-D-11-0123.1.
Harnik, N., E. Heifetz, O. Umurhan, and F. Lott, 2008: A buoyancy–vorticity wave interaction approach to stratified shear flow. J. Atmos. Sci., 65, 2615–2630, https://doi.org/10.1175/2007JAS2610.1.
Held, I. M., 1975: Momentum transport by quasi-geostrophic eddies. J. Atmos. Sci., 32, 1494–1497, https://doi.org/10.1175/1520-0469(1975)032<1494:MTBQGE>2.0.CO;2.
Hide, R., 1969: Dynamics of the atmospheres of the major planets with an appendix on the viscous boundary layer at the rigid bounding surface of an electrically-conducting rotating fluid in the presence of a magnetic field. J. Atmos. Sci., 26, 841–853, https://doi.org/10.1175/1520-0469(1969)026<0841:DOTAOT>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.
Hsu, C. J., and R. A. Plumb, 2000: Nonaxisymmetric thermally driven circulations and upper-tropospheric monsoon dynamics. J. Atmos. Sci., 57, 1255–1276, https://doi.org/10.1175/1520-0469(2000)057<1255:NTDCAU>2.0.CO;2.
Iga, S. I., and Y. Matsuda, 2005: Shear instability in a shallow water model with implications for the Venus atmosphere. J. Atmos. Sci., 62, 2514–2527, https://doi.org/10.1175/JAS3484.1.
Ji, X., J. D. Neelin, S.-K. Lee, and C. R. Mechoso, 2014: Interhemispheric teleconnections from tropical heat sources in intermediate and simple models. J. Climate, 27, 684–697, https://doi.org/10.1175/JCLI-D-13-00017.1.
Kraucunas, I., and D. L. Hartmann, 2005: Equatorial superrotation and the factors controlling the zonal-mean zonal winds in the tropical upper troposphere. J. Atmos. Sci., 62, 371–389, https://doi.org/10.1175/JAS-3365.1.
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.
Laraia, A. L., and T. Schneider, 2015: Superrotation in terrestrial atmospheres. J. Atmos. Sci., 72, 4281–4296, https://doi.org/10.1175/JAS-D-15-0030.1.
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.
Lindzen, R. S., and A. V. Hou, 1988: Hadley circulations for zonally averaged heating centered off the equator. J. Atmos. Sci., 45, 2416–2427, https://doi.org/10.1175/1520-0469(1988)045<2416:HCFZAH>2.0.CO;2.
Lutsko, N. J., 2018: The response of an idealized atmosphere to localized tropical heating: Superrotation and the breakdown of linear theory. J. Atmos. Sci., 75, 3–20, https://doi.org/10.1175/JAS-D-17-0192.1.
Pinto, J. R. D., and J. L. Mitchell, 2014: Atmospheric superrotation in an idealized GCM: Parameter dependence of the eddy response. Icarus, 238, 93–109, https://doi.org/10.1016/j.icarus.2014.04.036.
Saravanan, R., 1993: Equatorial superrotation and maintenance of the general circulation in two-level models. J. Atmos. Sci., 50, 1211–1227, https://doi.org/10.1175/1520-0469(1993)050<1211:ESAMOT>2.0.CO;2.
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., 1987: A simplified model of the modified Hadley circulation. J. Atmos. Sci., 44, 3311–3328, https://doi.org/10.1175/1520-0469(1987)044<3311:ASMOTM>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.
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.
Suarez, M. J., and D. G. Duffy, 1992: Terrestrial superrotation: A bifurcation of the general circulation. J. Atmos. Sci., 49, 1541–1554, https://doi.org/10.1175/1520-0469(1992)049<1541:TSABOT>2.0.CO;2.
Suhas, D., J. Sukhatme, and J. M. Monteiro, 2017: Tropical vorticity forcing and superrotation in the spherical shallow-water equations. Quart. J. Roy. Meteor. Soc., 143, 957–965, https://doi.org/10.1002/qj.2979.
Vallis, G. K., 2006: Atmospheric and Oceanic Fluid Dynamics. Cambridge University Press, 745 pp.
Wang, P., and J. L. Mitchell, 2014: Planetary ageostrophic instability leads to superrotation. Geophys. Res. Lett., 41, 4118–4126, https://doi.org/10.1002/2014GL060345.
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 vertical advection term can likewise be written as a vertical flux of y vorticity, but we see little advantage to using a vorticity description in this case because y vorticity is not conserved for 2D motions in the x–z plane. See Harnik et al. (2008) for a description of the coupled gravity–vorticity dynamics.
To keep the notation simple, we do not use any special symbol to denote stationary wave components. Eddy (primed) anomalies will represent climatological-mean values for the remainder of this section.