1. Introduction
The Madden–Julian oscillation (MJO) is a planetary-scale equatorial disturbance that dominates tropical variability on intraseasonal time scales. The disturbance is typified by a zonal dipole pattern in convective heating and cooling that moves eastward at a phase speed of ∼5 m s−1. The heating and cooling extends through the depth of the troposphere and drives horizontal divergence and convergence at upper levels (∼200 hPa) and convergence and divergence below, respectively (Kiladis et al. 2005, and references therein). The upper-tropospheric component of the MJO’s circulation has a much larger meridional extent than its lower-tropospheric component and is marked by pronounced off-equatorial cyclonic and anticyclonic Rossby gyres whose centers lie in the subtropics (Knutson and Weickmann 1987; Rui and Wang 1990; Kiladis and Weickmann 1992; Hendon and Salby 1994; Kiladis et al. 2005).
The Rossby gyres are thought to be a result of interaction between the convectively forced divergent flow with the basic-state vorticity gradient, known as the “Rossby wave source” (Sardeshmukh and Hoskins 1988). Wintertime MJO composites reveal that these gyres form on the southern flank of the subtropical westerly jet, move eastward in tandem with the MJO convection, and are most pronounced in the Indo-Pacific sector—a region where both MJO convective activity and the subtropical jet are found to be the strongest in the boreal winter (Adames and Wallace 2014). The relative location of the MJO-induced Rossby gyres with respect to the climatological background flow affects extratropical teleconnection patterns that influence global weather on subseasonal-to-seasonal time scales (Liebmann and Hartmann 1984; Weickmann et al. 1985; Lau and Lau 1986; Lau and Phillips 1986; Knutson and Weickmann 1987; Ferranti et al. 1990; Hoskins and Ambrizzi 1993; Jin and Hoskins 1995; Hsu 1996; Matthews et al. 2004; Lin et al. 2010; Seo and Lee 2017; Tseng et al. 2019; Hall et al. 2020). However, the connection between the MJO and extratropics is not just in one direction.
There have been several different studies indicating that extratropical variability has an important influence on the MJO. Among the earliest is the study by Straus and Lindzen (2000), who documented a strong coherence between slow eastward-propagating circulation signals in the subtropical upper troposphere and the MJO zonal winds in the tropics. Although they attributed the subtropical low-frequency variability to planetary-scale baroclinic instability (Frederiksen and Frederiksen 1997), the baroclinic generation of extratropical long waves is not well understood, and remains an active area of research (Hsieh et al. 2021; Moon et al. 2022). On the modeling side, Lin et al. (2007) used a dry atmospheric model with a wintertime basic state and showed that an MJO-like disturbance (in the form of a slow planetary-scale Kelvin wave with 15 m s−1 phase speed) can be generated in the Eastern Hemisphere in response to an imposed subtropical forcing. Ray and Zhang (2010) also performed experiments using a tropical channel model and were able to initiate an MJO event by including extratropical influences via lateral boundary conditions. Subsequently, Ray and Li (2013) performed mechanism denial experiments and showed that they could eliminate the MJO by suppressing extratropical waves. A potential issue with that study, however, was later identified by Ma and Kuang (2016), who performed more carefully designed experiments showing that the MJO “can exist without extratropical influence,” provided the basic state is maintained. At the same time, there are some competing MJO theories based on the dynamics of Rossby vortices that implicitly include extratropical influences on the MJO (Yano and Tribbia 2017; Rostami and Zeitlin 2019; Hayashi and Itoh 2017).
Such disparate findings and viewpoints are reflective of our general lack of understanding of how extratropical circulations interact with the MJO. Nevertheless, there is broad agreement that the subtropical jet and attendant Rossby gyres are important for providing a complete dynamical description of the phenomenon. While many studies have primarily focused on the forcing of subtropical circulations by the MJO (Schwendike et al. 2021, and references therein), here we focus on the opposite side of the coin, namely, how does the presence and strength of a subtropical jet affect the MJO?
Recently Tulich and Kiladis (2021, hereafter TK21) explored the impact of jet structure on the MJO and convectively coupled Kelvin waves in aquaplanet experiments performed using the superparameterized Weather Research and Forecasting Model (SP-WRF). They prescribed zonally symmetric sea surface temperature and nudged the subtropics toward a desired wind profile and found considerable weakening of the MJO signal when the model’s Indo-Pacific (IPAC) subtropical jet was weakened by 25% (Fig. 1; see TK21 for details). Although the sophisticated SP-WRF modeling setup produced a reasonably realistic MJO, the model complexity masked the precise pathway by which the jet controlled the simulated MJO strength.
To disentangle the feedback mechanism from the subtropics to the tropics, here we use a dry spherical shallow water model with variable jet speeds and perform two types of forcing experiments, namely, MJO-like thermal forcing at the equator and MJO induced gyre-like vorticity forcing in the subtropics. We then use a steady-state vorticity budget to show how the Rossby mode generated by each type of forcing experiments influences the Kelvin-mode divergence as a function of subtropical jet speed.
The paper is organized as follows. In section 2, we provide model details and outline the analytical approach for decomposing the model divergence into Matsuno–Gill modes and dynamical quantities from the vorticity budget. Section 3 describes the results of the steady-state model response for different jet speeds in response to thermal forcing and vorticity forcing experiments. Finally, in section 4 we discuss and summarize our results.
2. Methods
a. Model setup
The question of how the background flow structure affects the model’s steady-state response to an imposed MJO-like forcing can be addressed in at least two different ways. The first (termed “method 1”) is to run the model through separate “spinup” and “forcing” stages. During the spinup stage, a stable subtropical jet is first generated by raising the zonally symmetric topography, i.e., Ho is increased from 0 → Hmax. By day 50, the model reaches an equilibrium and Hmax determines the maximum jet speed, Ujet. During the subsequent forcing stage, the MJO-like forcing is switched on and the model is run further to a steady-state equilibrium, which is typically reached in 200 days. While this technique has become standard in the literature (Kraucunas and Hartmann 2007; Bao and Hartmann 2014; Monteiro et al. 2014), it can be time consuming when considering a large number of different Ujet profiles.
A more efficient way of probing the effects of changes in Ujet (termed “method 2”) is to effectively combine the spinup and forcing stages. Specifically, the model is initialized with a resting basic state (Ho = 0) and subjected to a steady external forcing. Then over 600 days, Ujet is gradually increased by slowly raising the zonally symmetric topography, i.e., Ho is gradually increased from 0 to 3500 m allowing Ujet to span from 0 to 78 m s−1, while being in quasi-steady state
In this paper, we mainly rely on method 2 to examine how the model responds to an imposed MJO-like forcing under a wide range of Ujet values. We ensure that a quasi-steady state is maintained throughout this process by checking that at every time step the tendency term in the momentum budget is very small relative to the remaining terms (not shown). We also performed a few runs using method 1, to ensure results are similar to those obtained using method 2 (details in appendix B).
1) Description of the background state
The specified background state is hemispherically symmetric with zero-mean winds at the equator, as an idealization of Earth’s upper-tropospheric zonal-mean circulation. Figures 2a–c plot the model’s steady-state zonal-mean horizontal winds (
In our model, as
2) External forcing
b. Analytical approach
1) Modal decomposition
2) Vorticity budget decomposition
Note that in Eq. (8) the denominator,
3. Results
a. Steady-state response to thermal forcing and variable jet speed
We first focus on the impact of the subtropical jet on the MJO’s thermally forced circulation in the upper troposphere with the fluid depth heq set at 500 m and Ujet ranging from 0 to 78 m s−1. Setting heq = 500 m ensures that the model runs stably for a large range of jet values and the tropical fluid depth stays close to the real world (200–500 m), with the precise value depending inversely on the strength of the jet (Fig. 2c).
1) Subtropical response
Figure 3 shows the steady-state eddy geopotential and wind anomalies excited by the stationary MJO-like thermal forcing for different subtropical jet speeds. The steady-state circulation obtained from method 2 is comparable to method 1 except for some minor differences which do not affect our overall results (see appendix B and Fig. B1 for a detailed comparison between methods 1 and 2). In the familiar case where Ujet = 0 (Fig. 3a), the positive part of the forcing induces a classic Gill-like pattern, consisting of a stationary Kelvin wave to the east and equatorial Rossby wave to the west of the mass source or heating region (Gill 1980). This same Rossby–Kelvin pattern is also excited by the mass sink or cooling region, but with opposite sign. As the jet speed increases, the equatorial Rossby wave response amplifies and shifts poleward, while the overall stationary wave pattern becomes meridionally tilted (shown for Ujet = 14 m s−1 in Fig. 3b). The tilted structure of the Rossby response (Fig. 3b) gives the impression of a teleconnection pattern due to weak waveguiding effect of the subtropical jet (Ambrizzi et al. 1995; Branstator and Teng 2017). In contrast, for the stronger jets, i.e., Ujet ≥ 28 m s−1, the equatorial Rossby waves transform into prominent subtropical gyres that are meridionally trapped due to stronger waveguiding effect and are advected eastward with respect to the forcing (Figs. 3c–f). See online supplementary Fig. S1 for more details.
This systematic shift from an equatorial waveguide to a wider subtropical stationary wave pattern due to imposed changes in background jet strength was first reported by Monteiro et al. (2014), using a similar shallow water model setup. In addition to those authors’ findings, we observe an interesting threshold behavior that has not been previously documented. Qualitatively, the overall strength of the subtropical gyres, as measured by their maximum magnitude of geopotential anomalies, is seen to increase monotonically for Ujet ≈ 0–42 m s−1 while it decreases for Ujet ≈ 42–70 m s−1.
An important difference between the model used here versus that of Monteiro et al. (2014) is in terms of the formulation of the geopotential tendency equation. Specifically, while those authors assumed a linearized geopotential flux, i.e., ϕeq(∇ ⋅ v) [see Eq. (3) in their supplementary material], here we include the full geopotential flux term, i.e., ∇ ⋅ (ϕv) [Eq. (3)]. As shown later, this difference has important implications for the divergent part of the eddy response in the tropics, whose dependence on jet speed is documented below.
2) Tropical response
It is worth mentioning here that Eq. (9) is a statement of the weak-temperature gradient (WTG) approximation for a shallow water system [see Eq. (4) in Sobel et al. 2001]. Therefore, we define the quantity,
Globally, since the imposed net mass source is zero, the area-averaged eddy divergence must also be zero. However, in the vicinity of the forcing region the eddy divergence shows a clear jet speed dependence, despite the thermal forcing being constant. To capture the local amplification of the eddy divergence anomalies in the tropics, Fig. 5a shows the root-mean-square of
To emphasize the threshold behavior of the divergent part of the response, we define a critical jet speed, Uc = 46 m s−1 at which δRMS reaches its peak value (black vertical line in Fig. 5a). The precise value of Uc is found to depend on the specified gravity wave speed
To break down the response of
To summarize, the jet-speed dependence of the divergent part of the response to an imposed MJO-like thermal forcing exhibits two distinct regimes: (i) a “weak-jet” regime (Ujet < Uc) where the deviation between actual divergence and WTG divergence near the forcing region (δRMS) grows with the increase in jet speed mainly due to stronger amplification of Kelvin divergence and (ii) a “strong-jet” regime (Ujet > Uc) where the deviation (δRMS) is reduced and becomes negative with increasing jet speed mainly due to a reduction in Kelvin divergence, despite the increased contribution by the higher-order Matsuno modes. In both jet regimes, the WTG divergence anomaly projects onto the steady-state forced Kelvin mode, as opposed to free Kelvin waves whose phase speed can be considered as infinite under the WTG approximation (Bretherton and Sobel 2003; Ahmed et al. 2021).
3) Weak-jet regime
To identify the key dynamical processes regulating the jet-speed dependence of the model’s divergence response, we decompose the eddy divergence from the steady-state vorticity budget to reflect contributions from the Sverdrup effect, the Hadley cell effect, and jet advection [Eq. (8)]. Figure 6 shows the divergence decomposition for the weak-jet cases using Eq. (8). As expected from the steady-state mass balance, eddy divergence at the forcing region is positive over the heat source and negative over the heat sink (Figs. 6a-i–d-i). In the absence of a jet, the local eddy divergence at the forcing region is primarily balanced by the Sverdrup effect (Fig. 6a-i) and has no contribution from the Hadley cell effect or jet advection. As the jet speed strengthens, the Sverdrup effect also strengthens and amplifies the local eddy divergence (Figs. 6a-ii–d-ii), particularly in the eastern flank of the forcing region due to zonally advected subtropical gyres (Fig. 3). With the increase in jet speed, the Hadley cell effect increases but has relatively weaker magnitude (Figs. 6a-iii–d-iii) and the jet advection counteracts the increase in local eddy divergence although its effect is only strong at latitudes poleward of the forcing region (Figs. 6a-iv–d-iv).
Figure 7 captures the change in divergence/convergence and the change in Sverdrup effect for Ujet = 34 m s−1 (Fig. 6d) minus the Ujet = 0 m s−1 (Fig. 6a) decomposed into individual tropical modes as in Eq. (11). We find that near the heat source (heat sink) the increase in divergence (convergence) is primarily due to amplification of the Kelvin mode (Fig. 7a-i), while anomalies for the Rossby and higher-order Matsuno modes are negligible near the forcing (Figs. 7a-ii,a-iii). At the same time, the Sverdrup change is dominated by the Rossby mode and has an amplifying effect on eddy divergence/convergence at the forcing region (notice the same signs in Figs. 7b-ii,a-i). There is little to no Sverdrup effect from the Kelvin and higher-order Matsuno modes (Figs. 7b-iii,b-i).
The weak-jet regime may be the most relevant for Earth’s upper troposphere since the zonal-mean subtropical jet is rarely found to be any stronger than ∼30–40 m s−1. This relevance points to a possible jet–MJO feedback mechanism which can be summarized as follows. As long as the jet speed is below a critical value (Ujet < Uc), a stronger jet leads to a stronger MJO-forced subtropical Rossby mode, which by the Sverdrup effect, amplifies the equatorial Kelvin mode and hence, MJO convective heating.
4) Strong-jet regime
For jet speeds greater than the critical value (Ujet > Uc), we see a regime shift in the role of dynamical processes that affect the local eddy divergence at the forcing region. Figure 8 shows the divergence decomposition for the strong-jet cases using Eq. (8). Again, eddy divergence at the forcing region is positive over the heat source and negative over the heat sink (Figs. 8a-i–d-i). In contrast to the weak-jet cases, the Hadley cell effect plays the most important role in amplifying the local eddy divergence (Figs. 8a-iii–d-iii) while the Sverdrup effect attenuates it (Figs. 8a-ii–d-ii). For all cases, jet advection has an almost negligible role on the local divergence; rather, its effect is only strong outside of the forcing region (Figs. 6a-iv–d-iv).
In the strong-jet regime, while the subtropical Rossby mode is quite pronounced (Figs. 3e,f and A2e,f), the tropical divergence associated with the MJO is dominated by higher-order Matsuno modes rather than the Kelvin mode (Fig. 5b). The strong-jet regime may be relevant for climate change scenarios or other planetary systems where the subtropical jet and the Hadley cell could be stronger than what is presently observed on Earth. For even stronger jet speeds (Ujet > 80 m s−1) the model becomes nonlinear and unstable, which may indicate another regime transition toward equatorial superrotation (Showman and Polvani 2011; Potter et al. 2014; Zurita-Gotor and Held 2018).
To stay relevant to Earth’s atmosphere, here we focus on the weak-jet regime (Ujet < Uc), where a stronger jet amplifies both the subtropical Rossby mode and the equatorial Kelvin mode. Further decomposition of the eddy divergence from the vorticity budget reveals that the Kelvin divergence and Rossby winds are linked to one another via the Sverdrup effect. This result leads naturally to the question of whether a remotely forced Rossby mode can amplify the Kelvin mode in the absence of tropical heat source?
b. Steady-state response to vorticity forcing and variable jet speed
To answer the above, we performed a vorticity forcing experiment, involving a sequence of quasi-steady states produced using method 2 for a range of jet speeds, under no thermal forcing and a stationary vorticity forcing in the subtropics resembling the quadrupole Rossby gyres associated with the MJO. The vorticity forcing is specified by imposing the subtropical cyclonic and anticyclonic vortices derived from one of the steady states in the thermal forcing experiment, namely, that with Ujet = 40 m s−1 [see section 2a(2) and Eq. (5) for details].
Figure 9 shows the steady-state eddy geopotential and wind anomalies excited in response to the same vorticity forcing under different subtropical jet speeds. In the case of no jet, the vorticity forcing induces a strong local response in the subtropics and negligible response in the tropics (θ < 15°N/S) (Fig. 9a). When a jet is present, the same vorticity forcing induces a remote tropical response that gets stronger with increasing jet speed, as well as a local subtropical response that weakens proportionately, indicating a transfer of energy from the subtropics to the tropics (Figs. 9b–f). For Ujet = 56 and 70 m s−1, the tropical response acquires a well-defined Kelvin structure indicated by the same phase of zonal winds (
Figure 10 shows the strength of steady-state tropical divergence anomalies (measured by
To identify the key dynamical processes behind the jet-speed dependence of the model’s divergent response, we focus on the Sverdrup effect [Eq. (10)] of the vorticity forcing experiment. In the weak-jet regime (i.e., Ujet < 58 m s−1), other divergent sources [from Eq. (8)], namely, Hadley cell effect, jet advection, and vorticity forcing are also present but they do not dominate tropical divergence (not shown). Figure 11 shows the change in divergence/convergence and the change in Sverdrup effect for Ujet = 56 m s−1 (Fig. 9e) minus the Ujet = 0 m s−1 (Fig. 9a) decomposed into individual tropical modes as in Eq. (11). At the equatorial region, we find that an increase in divergence (convergence) is primarily due to an amplification of the Kelvin mode (Fig. 11a-i) while the divergence/convergence from Rossby and higher-order Matsuno modes are weaker near the equator (Figs. 11a-ii,a-iii). At the same time, the Sverdrup change is dominated by the Rossby mode and has an amplifying effect on eddy divergence/convergence at the equator (see the same signs in Figs. 11b-ii and a-i). There is little to no Sverdrup effect from the Kelvin and higher-order Matsuno modes (Figs. 7b-iii,b-i).
In summary, our results confirm that in the absence of equatorial thermal forcing, subtropical Rossby gyres are able to induce a shear-mediated Kelvin response, which increases in strength as the jet speed increases up to a critical value. It should be kept in mind that these findings do not imply the Kelvin-mode component of the MJO is produced solely by subtropical Rossby gyres. Rather, the point is that the Kelvin- and Rossby-mode components of the MJO are closely linked to one another via the Sverdrup effect. If the subtropical Rossby gyres are strengthened by external processes [for, e.g., via low-frequency variability in the extratropics (Lin et al. 2007; Lin and Brunet 2011) or via changes in the subtropical jet due to Arctic warming (Barnes and Screen 2015)], then their effect will be felt by the Kelvin-mode circulation component of the MJO, potentially leading to strengthening of MJO convection, in accordance with the idealized MJO simulations of TK21. In other words, the subtropical Rossby gyres are coupled to the Kelvin mode not just by convective heating but also by the zonal-mean meridional wind shear. It is the latter feature of the basic state that enables the Rossby gyres to act as a “Kelvin wave source” for the tropics, in much the same way that tropical heating can act as “Rossby wave source” for the extratropics (Sardeshmukh and Hoskins 1988).
Interestingly, the shear-mediated coupling is the strongest when Ujet is approximately equal to the dry gravity wave speed in the tropics (Figs. 5b and 10b) indicating a phase speed resonance between the Doppler-shifted subtropical Rossby gyres and the equatorial Kelvin mode. This behavior is consistent with some other studies that have emphasized the idea of resonant coupling between midlatitudes and the tropics (Majda and Biello 2003; Wang and Xie 1996; Hoskins and Yang 2000; Cheng et al. 2022).
4. Discussion and conclusions
a. Potential implications on the vertical structure and amplitude of the MJO
Based on our modeling result, we hypothesize two feedback mechanisms by which the subtropics can affect the MJO’s horizontal circulation. The first is governed by WTG balance and implies that if the tropical static stability and/or vertical heating profile respond to changes in the subtropical jet strength, then the MJO’s thermally forced divergence anomalies (
The second feedback mechanism involves deviations from strict WTG balance in the form of shear-mediated coupling between Kelvin and Rossby waves, via the “Sverdrup effect.” The Sverdrup effect is the most important mechanism that couples subtropics to a deep tropical Kelvin response and is most strongly felt when the jet speed is between 40 and 60 m s−1. Our results imply that there is a critical jet speed at which the Kelvin-mode divergence of the MJO is maximized due to the impact of the mean flow on subtropical eddies. We speculate that our current climate may be operating in the weak-jet regime where the mean subtropical jet is weaker than the critical value, but approaches the critical limit during boreal winter. Long-term climate variability (longer than the intraseasonal time scales) may alter the relative jet speed with respect to its critical value and affect the MJO’s convective amplitude via Rossby–Kelvin feedback. While beyond the scope of this paper, we plan to test this conjecture using general circulation model (GCM) experiments in a future work.
In reality, both feedback mechanisms may be operating simultaneously and could provide a conceptual framework for understanding 1) the MJO’s response to quasi-biennial oscillation (QBO) phases via changes in subtropical jet speed (Garfinkel and Hartmann 2011a,b; Gray et al. 2018; Martin et al. 2021), 2) the MJO’s response to different climate change scenarios (Carlson and Caballero 2016), and 3) the cause of MJO biases in global climate models (Ahn et al. 2020). However, it is important to remember that the conclusions drawn here are based on a highly simplified model with no moisture or cloud radiative processes where heating patterns associated with the MJO is imposed. For instance, we do not know how much of the convective outflow generated by the upper-level feedback couples to the low-level MJO convergence/divergence. Depending on the strength of the vertical coupling, it may have different effects on vertical motion, cloud distribution and moisture feedback, which may in turn affect the phase speed and the amplitude of the MJO.
b. Linear versus nonlinear MJO dynamics
It should also be noted that the forced circulations studied here were derived by running a nonlinear shallow water model in a stable linear regime. This approach is physically meaningful, because MJO composites from ERA5 dataset reveal that the zonal momentum budget of the MJO during boreal winter is dominated by the linear advection terms (not shown), in accordance with several other observational studies (e.g., Lin et al. 2005; Sakaeda and Roundy 2014, 2015).
However, our approach conflicts with several dry MJO theories that describe the MJO either as a nonlinear phenomenon driven by extratropical forcing (Wedi and Smolarkiewicz 2010; Yano and Tribbia 2017; Rostami and Zeitlin 2020) or a heavily damped Kelvin wave with no role for Rossby waves (Kim and Zhang 2021). TK21 also highlighted the role of nonlinear momentum fluxes on the MJO, but they did not evaluate the impact of the linear terms. The effects of nonlinearities might be important for the transient (onset or decay stage) or in moist feedback processes of the MJO. However, the problem of MJO maintenance can be simply explained on the basis of linear dynamics. The linear Rossby–Kelvin feedback mechanism in the lower troposphere may also aid in the eastward propagation of the MJO as noted by Hayashi and Itoh (2017) and might also explain why the MJO tends to be stronger in the Indo-Pacific region, where the subtropical jet is the strongest and closest to the equator during wintertime. The Rossby–Kelvin coupling also plays an important role in meridional moisture advection by the MJO as recently noted by Berrington et al. (2022).
Note that our conclusions about linear Rossby–Kelvin interaction is derived from simple modal decomposition that makes the β-plane approximation on a sphere [see section 2b(1)]. These results could be improved in future by using more sophisticated technique such as the spherical normal mode decomposition as shown by some recent MJO studies (Kosovelj et al. 2019; Kitsios et al. 2019; Franzke et al. 2019).
c. Concluding remarks
Previous studies on MJO dynamics have shown that moisture, cloud radiation, and boundary layer processes (Zhang et al. 2020, and references therein) play a crucial role in MJO’s initiation and propagation, which we consider as given. Here we focus on the impact of the wintertime subtropical jet in the Indo-Pacific region which sits just north of the MJO dipole and creates substantially strong upper-level horizontal shear for equatorial convective systems. The mean-flow interaction between MJO convection and the jet gives rise to planetary-scale Rossby gyres in the subtropical upper troposphere, which forms an integral part of the MJO’s circulation (Sardeshmukh and Hoskins 1988; Adames and Wallace 2014; Monteiro et al. 2014). The question of whether these forced Rossby gyres and jet structure have any subsequent feedback onto the tropics is much less understood. Recently TK21 found considerable weakening of MJO-like signals in idealized SP-WRF calculations when the zonal-mean zonal jet was weakened by 25%, while other parameters like static stability and surface temperature were kept constant.
To understand this result, we used a dry spherical shallow water model to examine how the divergent part of its response to an MJO-like thermal forcing is affected by the presence and strength of an imposed subtropical jet. Results showed a positive correlation between equatorial divergence/convergence and subtropical jet speed, but with two different regimes of behavior (weak jet versus strong jet). In the weak-jet regime, the MJO-induced divergence is amplified due to the “Sverdrup effect,” while in the strong-jet regime, the divergence amplifies due to the “Hadley cell effect.”
To leading order, the divergence induced by the forcing was seen to be well explained by WTG balance (
Regardless of the simplicity of the model setup, our results point to the potentially important feedback mechanisms by which the presence of a subtropical jet can affect the MJO’s structure and amplitude. Future developments of MJO theory should therefore consider the role of the upper-tropospheric subtropical background flow on the precise nature of the phenomenon.
Acknowledgments.
We thank Maria Gehne, Nedjeljka Žagar, Joseph Biello, and one anonymous reviewer for their comments and suggestions which substantially improved the clarity of the paper. P. Barpanda is grateful to Joy Monteiro for his help in setting up the shallow water model and thanks Brandon Wolding and Yuan-Ming Cheng for helpful discussions during the project. This research was supported in part by NOAA Cooperative Agreement NA22OAR4320151 and National Science Foundation (NSF) through Award AGS-1839741. P. Barpanda also acknowledges support from NOAA and CIRES for the postdoc opportunity.
Data availability statement.
The ERA5 data can be accessed through the ECMWF website (https://www.ecmwf.int/en/forecasts/datasets/reanalysis-datasets/era5). The scripts for running shallow water model experiments can be accessed from https://github.com/Pragallva/MJO-waves-mean-flow.
Footnotes
Original code is downloaded from https://nschaeff.bitbucket.io/shtns/shallow_water_8py-example.html and modified for the experiments.
APPENDIX A
Meridional Mode Decomposition
Our shallow water system is neither on a β plane nor does it have a resting basic state (spherical model with a background horizontal shear). However, the final steady-state solutions can be approximated as linear superposition of Matsuno modes up to meridional truncation number N. This is the same as Galerkin method of discretization and has been successfully used in reanalysis dataset for identifying equatorial waves (Yang et al. 2003; Gehne and Kleeman 2012; Knippertz et al. 2022; Haertel 2022). We take the same approach for decomposing our steady-state shallow water model response into Matsuno modes as described below.
Each of the coefficients in Eqs. (A2)–(A4) is determined by projecting the nth-order PCF on to q, r, and υ, respectively, and using the orthogonality relation for PCF functions. For example,
Once all the mode coefficients (qn+1, υn, rn−1) are determined, the eigenvector for each mode in terms of winds and geopotential can be expressed as follows:
-
Kelvin:
-
Rossby:
-
Higher order:
For the same set of thermal forcing experiments documented in the main text (Fig. 3), Figs. A1–A3 show the decomposition of steady-state response into individual Kelvin, Rossby, and higher-order Matsuno modes as outlined in section 2b(1).
APPENDIX B
Steady-State Model Response Using Method 1
Here we highlight our results from an ensemble of six steady-state shallow water model experiments, each initialized with different jet speeds (0, 14, 28, 42, 56, 70 m s−1) but forced with the same MJO-like thermal forcing (see a detailed description of method 1 in section 2a). The steady-state circulation anomalies from six method 1 runs (Fig. B1) are very similar to those obtained from a single quasi-steady state run using method 2 (Fig. 3). Method 1 also captures the threshold behavior of the subtropical gyres as noted in the main text. Although there are some differences in the magnitude of the responses and the peak location of the gyres, the errors are negligible and do not affect the main conclusions of the paper as evidenced by method 1 results shown in Fig. 5a.
APPENDIX C
Sensitivity of Kelvin Divergence to Changing Equivalent Depth and Forcing Phase Speeds
Here we explore the sensitivity of our experiments to changing equivalent depths, which is a measure of effective static stability in the atmosphere and an imposed thermal forcing moving at different forcing phase speed, cf. Figure C1 captures results from several thermal forcing experiments showing root-mean-square of Kelvin divergence (averaged between 10°S and 10°N) for a wide range of jet speeds for two equivalent depths (heq = 200 and 500 m) and 3 different forcing phase speeds (cf = 0, 5, and 15 m s−1) but with a fixed heating amplitude. Note that the critical jet speed (Uc) is not a constant; rather, it is lowered for smaller gravity wave speed (
REFERENCES
Adames, A. F., and J. M. Wallace, 2014: 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.
Ahmed, F., J. D. Neelin, and Á. F. Adames, 2021: Quasi-equilibrium and weak temperature gradient balances in an equatorial beta-plane model. J. Atmos. Sci., 78, 209–227, https://doi.org/10.1175/JAS-D-20-0184.1.
Ahn, M.-S., and Coauthors, 2020: MJO propagation across the Maritime Continent: Are CMIP6 models better than CMIP5 models? Geophys. Res. Lett., 47, e2020GL087250, https://doi.org/10.1029/2020GL087250.
Ambrizzi, T., B. J. Hoskins, and H.-H. Hsu, 1995: Rossby wave propagation and teleconnection patterns in the austral winter. J. Atmos. Sci., 52, 3661–3672, https://doi.org/10.1175/1520-0469(1995)052<3661:RWPATP>2.0.CO;2.
Bao, J., V. Dixit, and S. C. Sherwood, 2022: Zonal temperature gradients in the tropical free troposphere. J. Climate, 35, 7937–7948, https://doi.org/10.1175/JCLI-D-22-0145.1.
Bao, M., and D. L. Hartmann, 2014: The response to MJO-like forcing in a nonlinear shallow-water model. Geophys. Res. Lett., 41, 1322–1328, https://doi.org/10.1002/2013GL057683.
Barnes, E. A., and J. A. Screen, 2015: The impact of Arctic warming on the midlatitude jet-stream: Can it? Has it? Will it? Wiley Interdiscip. Rev.: Climate Change, 6, 277–286, https://doi.org/10.1002/wcc.337.
Berrington, A. H., N. Sakaeda, J. Dias, and G. N. Kiladis, 2022: Relationships between the eastward propagation of the Madden–Julian oscillation and its circulation structure. J. Geophys. Res. Atmos., 127, e2021JD035806, https://doi.org/10.1029/2021JD035806.
Betts, A. K., 1982: Saturation point analysis of moist convective overturning. J. Atmos. Sci., 39, 1484–1505, https://doi.org/10.1175/1520-0469(1982)039<1484:SPAOMC>2.0.CO;2.
Branstator, G., and H. Teng, 2017: Tropospheric waveguide teleconnections and their seasonality. J. Atmos. Sci., 74, 1513–1532, https://doi.org/10.1175/JAS-D-16-0305.1.
Bretherton, C. S., and A. H. Sobel, 2003: The Gill model and the weak temperature gradient approximation. J. Atmos. Sci., 60, 451–460, https://doi.org/10.1175/1520-0469(2003)060<0451:TGMATW>2.0.CO;2.
Carlson, H., and R. Caballero, 2016: Enhanced MJO and transition to superrotation in warm climates. J. Adv. Model. Earth Syst., 8, 304–318, https://doi.org/10.1002/2015MS000615.
Castanheira, J. M., and C. A. F. Marques, 2021: The equatorial wave skeleton of the Madden–Julian oscillation. Quart. J. Roy. Meteor. Soc., 147, 3778–3788, https://doi.org/10.1002/qj.4156.
Chao, W. C., 1987: On the origin of the tropical intraseasonal oscillation. J. Atmos. Sci., 44, 1940–1949, https://doi.org/10.1175/1520-0469(1987)044<1940:OTOOTT>2.0.CO;2.
Cheng, Y.-M., S. Tulich, G. N. Kiladis, and J. Dias, 2022: Two extratropical pathways to forcing tropical convective disturbances. J. Climate, 35, 6587–6609, https://doi.org/10.1175/JCLI-D-22-0171.1.
Ferranti, L., T. N. Palmer, F. Molteni, and E. Klinker, 1990: Tropical–extratropical interaction associated with the 30–60 day oscillation and its impact on medium and extended range prediction. J. Atmos. Sci., 47, 2177–2199, https://doi.org/10.1175/1520-0469(1990)047<2177:TEIAWT>2.0.CO;2.
Franzke, C. L. E., D. Jelic, S. Lee, and S. B. Feldstein, 2019: Systematic decomposition of the MJO and its Northern Hemispheric extratropical response into Rossby and inertio-gravity components. Quart. J. Roy. Meteor. Soc., 145, 1147–1164, https://doi.org/10.1002/qj.3484.
Frederiksen, J. S., and C. S. Frederiksen, 1997: Mechanisms of the formation of intraseasonal oscillations and Australian monsoon disturbances: The roles of convection, barotropic and baroclinic instability. Contrib. Atmos. Phys., 70, 39–56.
Fulton, S. R., and W. H. Schubert, 1985: Vertical normal mode transforms: Theory and application. Mon. Wea. Rev., 113, 647–658, https://doi.org/10.1175/1520-0493(1985)113<0647:VNMTTA>2.0.CO;2.
Garfinkel, C. I., and D. L. Hartmann, 2011a: The influence of the quasi-biennial oscillation on the troposphere in winter in a hierarchy of models. Part I: Simplified dry GCMs. J. Atmos. Sci., 68, 1273–1289, https://doi.org/10.1175/2011JAS3665.1.
Garfinkel, C. I., and D. L. Hartmann, 2011b: The influence of the quasi-biennial oscillation on the troposphere in winter in a hierarchy of models. Part II: Perpetual winter WACCM runs. J. Atmos. Sci., 68, 2026–2041, https://doi.org/10.1175/2011JAS3702.1.
Gehne, M., and R. Kleeman, 2012: Spectral analysis of tropical atmospheric dynamical variables using a linear shallow-water modal decomposition. J. Atmos. Sci., 69, 2300–2316, https://doi.org/10.1175/JAS-D-10-05008.1.
Gill, A. E., 1980: Some simple solutions for heat-induced tropical circulation. Quart. J. Roy. Meteor. Soc., 106, 447–462, https://doi.org/10.1002/qj.49710644905.
Gray, L. J., J. A. Anstey, Y. Kawatani, H. Lu, S. Osprey, and V. Schenzinger, 2018: Surface impacts of the quasi biennial oscillation. Atmos. Chem. Phys., 18, 8227–8247, https://doi.org/10.5194/acp-18-8227-2018.
Haertel, P., 2022: Kelvin and Rossby wave contributions to the mechanisms of the Madden-Julian oscillation. Geosciences, 12, 314, https://doi.org/10.3390/geosciences12090314.
Hall, N. M. J., H. H. Le, and S. Leroux, 2020: The extratropical response to a developing MJO: Forecast and climate simulations with the DREAM model. Climate Dyn., 55, 813–829, https://doi.org/10.1007/s00382-020-05299-y.
Hayashi, M., and H. Itoh, 2017: A new mechanism of the slow eastward propagation of unstable disturbances with convection in the tropics: Implications for the MJO. J. Atmos. Sci., 74, 3749–3769, https://doi.org/10.1175/JAS-D-16-0300.1.
Hendon, H. H., and M. L. Salby, 1994: The life cycle of the Madden–Julian oscillation. J. Atmos. Sci., 51, 2225–2237, https://doi.org/10.1175/1520-0469(1994)051<2225:TLCOTM>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., and G.-Y. Yang, 2000: The equatorial response to higher-latitude forcing. J. Atmos. Sci., 57, 1197–1213, https://doi.org/10.1175/1520-0469(2000)057<1197:TERTHL>2.0.CO;2.
Hsieh, T.-L., C.-Y. Chang, I. M. Held, and P. Zurita-Gotor, 2021: Nonlinear generation of long waves and the reversal of eddy momentum fluxes in a two-layer quasigeostrophic model. J. Atmos. Sci., 78, 3525–3536, https://doi.org/10.1175/JAS-D-20-0368.1.
Hsu, H.-H., 1996: Global view of the intraseasonal oscillation during northern winter. J. Climate, 9, 2386–2406, https://doi.org/10.1175/1520-0442(1996)009<2386:GVOTIO>2.0.CO;2.
Jiang, X., and Coauthors, 2020: Fifty years of research on the Madden–Julian oscillation: Recent progress, challenges, and perspectives. J. Geophys. Res. Atmos., 125, e2019JD030911, https://doi.org/10.1029/2019JD030911.
Jin, F., and B. J. Hoskins, 1995: The direct response to tropical heating in a baroclinic atmosphere. J. Atmos. Sci., 52, 307–319, https://doi.org/10.1175/1520-0469(1995)052<0307:TDRTTH>2.0.CO;2.
Kasahara, A., 1980: Effect of zonal flows on the free oscillations of a barotropic atmosphere. J. Atmos. Sci., 37, 917–929, https://doi.org/10.1175/1520-0469(1980)037<0917:EOZFOT>2.0.CO;2.
Kiladis, G. N., and K. M. Weickmann, 1992: Circulation anomalies associated with tropical convection during northern winter. Mon. Wea. Rev., 120, 1900–1923, https://doi.org/10.1175/1520-0493(1992)120<1900:CAAWTC>2.0.CO;2.
Kiladis, G. N., K. H. Straub, and P. T. Haertel, 2005: Zonal and vertical structure of the Madden–Julian oscillation. J. Atmos. Sci., 62, 2790–2809, https://doi.org/10.1175/JAS3520.1.
Kiladis, G. N., M. C. Wheeler, P. T. Haertel, K. H. Straub, and P. E. Roundy, 2009: Convectively coupled equatorial waves. Rev. Geophys., 47, RG2003, https://doi.org/10.1029/2008RG000266.
Kim, J.-E., and C. Zhang, 2021: Core dynamics of the MJO. J. Atmos. Sci., 78, 229–248, https://doi.org/10.1175/JAS-D-20-0193.1.
Kitsios, V., T. J. O’Kane, and N. Žagar, 2019: A reduced-order representation of the Madden–Julian oscillation based on reanalyzed normal mode coherences. J. Atmos. Sci., 76, 2463–2480, https://doi.org/10.1175/JAS-D-18-0197.1.
Knippertz, P., and Coauthors, 2022: The intricacies of identifying equatorial waves. Quart. J. Roy. Meteor. Soc., 148, 2814–2852, https://doi.org/10.1002/qj.4338.
Knutson, T. R., and K. M. Weickmann, 1987: 30–60 day atmospheric oscillations: Composite life cycles of convection and circulation anomalies. Mon. Wea. Rev., 115, 1407–1436, https://doi.org/10.1175/1520-0493(1987)115<1407:DAOCLC>2.0.CO;2.
Kosovelj, K., F. Kucharski, F. Molteni, and N. Žagar, 2019: Modal decomposition of the global response to tropical heating perturbations resembling MJO. J. Atmos. Sci., 76, 1457–1469, https://doi.org/10.1175/JAS-D-18-0203.1.
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.
Lau, K.-M., and T. J. Phillips, 1986: Coherent fluctuations of extratropical geopotential height and tropical convection in intraseasonal time scales. J. Atmos. Sci., 43, 1164–1181, https://doi.org/10.1175/1520-0469(1986)043<1164:CFOFGH>2.0.CO;2.
Lau, N.-C., and K.-M. Lau, 1986: The structure and propagation of intraseasonal oscillations appearing in a GFDL general circulation model. J. Atmos. Sci., 43, 2023–2047, https://doi.org/10.1175/1520-0469(1986)043<2023:TSAPOI>2.0.CO;2.
Liebmann, B., and D. L. Hartmann, 1984: An observational study of tropical–midlatitude interaction on intraseasonal time scales during winter. J. Atmos. Sci., 41, 3333–3350, https://doi.org/10.1175/1520-0469(1984)041<3333:AOSOTI>2.0.CO;2.
Lin, H., and G. Brunet, 2011: Impact of the North Atlantic Oscillation on the forecast skill of the Madden-Julian oscillation. Geophys. Res. Lett., 38, L02802, https://doi.org/10.1029/2010GL046131.
Lin, H., G. Brunet, and J. Derome, 2007: Intraseasonal variability in a dry atmospheric model. J. Atmos. Sci., 64, 2422–2441, https://doi.org/10.1175/JAS3955.1.
Lin, H., G. Brunet, and R. Mo, 2010: Impact of the Madden–Julian oscillation on wintertime precipitation in Canada. Mon. Wea. Rev., 138, 3822–3839, https://doi.org/10.1175/2010MWR3363.1.
Lin, J.-L., M. Zhang, and B. Mapes, 2005: Zonal momentum budget of the Madden–Julian oscillation: The source and strength of equivalent linear damping. J. Atmos. Sci., 62, 2172–2188, https://doi.org/10.1175/JAS3471.1.
Ma, D., and Z. Kuang, 2016: A mechanism-denial study on the Madden–Julian oscillation with reduced interference from mean state changes. Geophys. Res. Lett., 43, 2989–2997, https://doi.org/10.1002/2016GL067702.
Majda, A. J., and J. A. Biello, 2003: The nonlinear interaction of barotropic and equatorial baroclinic Rossby waves. J. Atmos. Sci., 60, 1809–1821, https://doi.org/10.1175/1520-0469(2003)060<1809:TNIOBA>2.0.CO;2.
Maloney, E. D., and D. L. Hartmann, 1998: Frictional moisture convergence in a composite life cycle of the Madden–Julian oscillation. J. Climate, 11, 2387–2403, https://doi.org/10.1175/1520-0442(1998)011<2387:FMCIAC>2.0.CO;2.
Martin, Z., C. Orbe, S. Wang, and A. Sobel, 2021: The MJO–QBO relationship in a GCM with stratospheric nudging. J. Climate, 34, 4603–4624, https://doi.org/10.1175/JCLI-D-20-0636.1.
Matsuno, T., 1966: Quasi-geostrophic motions in the equatorial area. J. Meteor. Soc. Japan, 44, 25–43, https://doi.org/10.2151/jmsj1965.44.1_25.
Matthews, A. J., B. J. Hoskins, and M. Masutani, 2004: The global response to tropical heating in the Madden–Julian oscillation during the northern winter. Quart. J. Roy. Meteor. Soc., 130, 1991–2011, https://doi.org/10.1256/qj.02.123.
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.
Moon, W., G. E. Manucharyan, and H. A. Dijkstra, 2022: Baroclinic instability and large-scale wave propagation in a planetary-scale atmosphere. Quart. J. Roy. Meteor. Soc., 148, 809–825, https://doi.org/10.1002/qj.4232.
Potter, S. F., G. K. Vallis, and J. L. Mitchell, 2014: Spontaneous superrotation and the role of Kelvin waves in an idealized dry GCM. J. Atmos. Sci., 71, 596–614, https://doi.org/10.1175/JAS-D-13-0150.1.
Ray, P., and C. Zhang, 2010: A case study of the mechanics of extratropical influence on the initiation of the Madden–Julian oscillation. J. Atmos. Sci., 67, 515–528, https://doi.org/10.1175/2009JAS3059.1.
Ray, P., and T. Li, 2013: Relative roles of circumnavigating waves and extratropics on the MJO and its relationship with the mean state. J. Atmos. Sci., 70, 876–893, https://doi.org/10.1175/JAS-D-12-0153.1.
Rostami, M., and V. Zeitlin, 2019: Eastward-moving convection-enhanced modons in shallow water in the equatorial tangent plane. Phys. Fluids, 31, 021701, https://doi.org/10.1063/1.5080415.
Rostami, M., and V. Zeitlin, 2020: Can geostrophic adjustment of baroclinic disturbances in the tropical atmosphere explain MJO events? Quart. J. Roy. Meteor. Soc., 146, 3998–4013, https://doi.org/10.1002/qj.3884.
Rui, H., and B. Wang, 1990: Development characteristics and dynamic structure of tropical intraseasonal convection anomalies. J. Atmos. Sci., 47, 357–379, https://doi.org/10.1175/1520-0469(1990)047<0357:DCADSO>2.0.CO;2.
Sakaeda, N., and P. E. Roundy, 2014: The role of interactions between multiscale circulations on the observed zonally averaged zonal wind variability associated with the Madden–Julian oscillation. J. Atmos. Sci., 71, 3816–3836, https://doi.org/10.1175/JAS-D-13-0304.1.
Sakaeda, N., and P. E. Roundy, 2015: The development of upper-tropospheric wind over the Western Hemisphere in association with MJO convective initiation. J. Atmos. Sci., 72, 3138–3160, https://doi.org/10.1175/JAS-D-14-0293.1.
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.
Schwendike, J., G. J. Berry, K. Fodor, and M. J. Reeder, 2021: On the relationship between the Madden–Julian oscillation and the Hadley and Walker circulations. J. Geophys. Res. Atmos., 126, e2019JD032117, https://doi.org/10.1029/2019JD032117.
Seo, K.-H., and H.-J. Lee, 2017: Mechanisms for a PNA-like teleconnection pattern in response to the MJO. J. Atmos. Sci., 74, 1767–1781, https://doi.org/10.1175/JAS-D-16-0343.1.