1. Introduction
The Madden–Julian oscillation (MJO) and boreal summer intraseasonal oscillation (BSISO) are the two most prominent tropical intraseasonal modes. The MJO (Zhang 2005) is dominant in northern winter while the BSISO (Kikuchi 2021) is more important in northern summer. These tropical intraseasonal oscillations (ISOs) play large roles in global weather and climate. Since the discovery independently by several authors of the MJO (Xie et al. 1963; Li et al. 2018; Madden and Julian 1971) and the BSISO (Yasunari 1979; Sikka and Gadgil 1980; Krishnamurti and Subrahmanyam 1982), a great deal of effort has been devoted to understanding their dynamics, as well as to numerical modeling, subseasonal prediction, and projection of their behavior under global warming.
Both the MJO and BSISO develop most often in the Indian Ocean (although MJO events in particular can originate at any longitude; e.g., Matthews 2008) and share similar time scales, while their spatial patterns differ markedly. The MJO is largely symmetric about the equator in this region, with twin cyclonic gyres straddling the equator and trailing the main near-equatorial convection clusters (Rui and Wang 1990), a structure sometimes characterized as a “swallowtail” (Zhang and Ling 2012; Kim and Zhang 2021). On the other hand, the BSISO is mostly present in the Northern Hemisphere, with a southeast–northwest-tilted structure in precipitation spanning from South Asia to the western Pacific Ocean. The MJO primarily propagates eastward, while the BSISO propagates both eastward and northward (Lawrence and Webster 2002; Wang et al. 2018). These intraseasonal oscillations share similar time and space scales, but their morphologies differ substantially. The conceptual boundary between the two phenomena in the literature is arbitrary and confusing: depending on how different authors prioritize time scale versus spatial pattern, and which one is of greater interest, the BSISO is treated either as the northern summer incarnation of the MJO or as a distinct phenomenon.
The current understanding of the MJO has converged to some degree on the crucial roles of moisture in shaping its eastward propagation and planetary-scale instability, with competing theories emphasizing different aspects of convection–circulation interaction (Zhang et al. 2020). Studies of the BSISO, however, have followed a different line historically. Several authors have considered its northward propagation largely independent of eastward propagation, and even used zonally symmetric models to understand its dynamics (Webster 1983; Nanjundiah et al. 1992; Jiang et al. 2004; Drbohlav and Wang 2005; Bellon and Sobel 2008a,b). No existing theory seems to be able to simultaneously explain the BSISO’s spatial patterns, temporal and spatial scales, and relationship to the MJO, if any.
The MJO and BSISO both draw energy from the tropical ocean and from the background atmospheric state. Over the Indian Ocean, the large-scale circulation features large-scale ascent in the Maritime Continent and descent over the western Indian Ocean, with low-level mean westerlies and upper-level easterlies completing the local branch of the zonal overturning Walker cell. Consistent with this structure, atmospheric moisture increases eastward over the Indian Ocean, with drier air near the African coast and moister air near the Maritime Continent. The mean meridional circulation changes with the seasons, with ascent and the greatest moisture in the summer hemisphere and a drier descending winter hemisphere. In northern summer in particular, sea surface temperatures in the Bay of Bengal and Arabian Sea become quite warm. Moisture and enthalpy build up near the southern slope of the Tibetan Plateau, and the meridional moisture gradient over the Indian Ocean basin in the Northern Hemisphere becomes very weak or even reverses (i.e., moisture increases with latitude). These mean gradients are relevant to the intraseasonal modes. Recent observational analyses (Jiang et al. 2018; Adames et al. 2016) and numerical modeling (Wang et al. 2021) emphasize the importance of moisture for the BSISO, and suggest that the BSISO, like the MJO, may be thought of as a moisture mode, despite the marked differences in morphology between the two.
Here, we perform a unified analysis of an idealized moist linear model which (we argue) explains some key features of both the MJO and BSISO parsimoniously. In this model, the background meridional moisture gradient controls the morphology of the ISOs. An increase in this gradient, representative of that which occurs from northern summer to northern winter over the Indian Ocean, leads to a transformation in the spatial pattern of the modeled intraseasonal oscillation from a more BSISO-like one to a more MJO-like one.
The rest of this article is structured as follows. Section 2 presents the model and its solutions. Model parameter values are estimated in section 3. Results are discussed in section 4, followed by physical interpretation of this simple model in section 5. A discussion and conclusions are presented in section 6.
2. A linear moist shallow water model
a. Formulation
These are equations for horizontal momentum u and υ, perturbation pressure ϕ, and moisture q, respectively; x and y denote nondimensional distances in longitude and latitude, respectively, with a scaling constant
Model parameters.
-
α, moisture relaxation time scale for precipitation (αq);
-
r, cloud radiative feedback;
-
Γ, nondimensional gross moist stability;
-
-
αu, Rayleigh friction (used only on zonal wind u; its effect on υ is small and thus neglected);
-
αϕ, Newtonian damping for perturbation pressure;
-
-
This linear shallow water model may be thought of as an extension of the classic dry linear equatorial wave model of Matsuno (1966). All the column physical processes associated with convection, clouds, and radiation are packed into six parameters in the q and ϕ equations. If all of these parameters are set to zero, the model reduces to Matsuno’s. Similar moist shallow water equation sets—but generally without horizontal moisture advection—have been analytically solved (e.g., Neelin and Yu 1994; Fuchs and Raymond 2002, 2005, 2017; Sugiyama 2009; Ahmed et al. 2021). For some parameter values, the solutions of these systems display planetary instability on intraseasonal time scales. However, the spatial patterns often lack the off-equatorial cyclonic gyres typical of the observed MJO (Rui and Wang 1990), suggesting that they are really convectively coupled Kelvin waves. Many other studies have obtained solutions by projecting the meridional structure to the first several leading basis functions associated with the dry equation set (e.g., Majda and Stechmann 2009; Sobel and Maloney 2012, 2013; Adames and Kim 2016; Liu and Wang 2016; Stechmann and Hottovy 2020), thus including (sometimes by construction) MJO-like cyclonic gyres, but no BSISO-like solutions were found.
Here, we include both zonal and meridional moisture advection of the basic state moisture by perturbation winds, namely the last two terms in Eq. (4). We make a specific choice for the meridional structure of the basic state moisture that facilitates analysis, while still being consistent with observations at a level of idealization consistent with the model’s other aspects. We do not use the meridional basis functions of the dry system as a basis, instead using the distinct meridional structures that emerge naturally from the moist model. These structures were also identified by Ahmed (2021), who used a very similar model, and whose work we became aware of in the late stages of writing this paper. We find here that these structures are key to explaining the MJO–BSISO differences as a function of the meridional moisture gradient.
Our simplifying assumptions about the horizontal moisture gradients are as follows.
First, we consider zonal advection. Assuming a constant zonal moisture gradient, the zonal advection term is mathematically equivalent to the WISHE term, and they may be combined into a single coefficient,
It might seem questionable that the absolute value of the meridional gradient Qy increases with |y| even at large y. However, assuming boundedness as y → ∞ will lead us to retain only equatorially trapped solutions in which υ decays exponentially with y2. This exponential decay dominates over the linear increase in moisture gradient at large y, and justifies the use of υyQ0 as a reasonable approximation for the moisture meridional advection
b. Reduction to a second-order ODE for υ
Rayleigh friction may be added back to υ by replacing ω in d0 with ωu. A moisture convergence closure is achieved by replacing ωα with iα (i.e.,
c. Wave solutions
The first solution is referred to as the υ = 0 mode. It is analogous to the equatorial Kelvin wave, but modified by zonal moisture advection, and displaying different dispersion c (see Fig. 2 below). The second one contains nth-order Hermite polynomials Hn in standard form (Abramowitz and Stegun 1964), except that the shape factor ξn and the scaling factor ηn are complex numbers to be determined as part of the dispersion relationships for the wave parameters (ω, ηn, ξn). The key difference between the above solution and the traditional ones is in these shape and scaling factors. This functional form was also recently obtained by Ahmed (2021), who showed MJO-like solutions (but no BSISO-like solutions). The main differences between our study and Ahmed (2021) result from the choices of the parameter values. The most important of these is the range of meridional moisture gradients considered, particularly in that we consider very small, zero, or even reversed gradients. Another substantial difference is that our convective closure depends only on moisture, while that of Ahmed (2021) includes temperature dependence as well. The differences in other parameter choices may be quantitatively consequential, but are probably not qualitatively so.
We focus on the n = 1 solution in the present study. This simplifies the problem as we need solve for only two wave parameters, (ω, ξ1), while retaining the essential dynamics. Higher-order Hermite polynomial solutions will be reported elsewhere. The solutions of u, q, and ϕ are symmetric about the equator. While υ is antisymmetric, the use of linear moisture gradient makes the meridional advection term −υyQ0 also symmetric, consistent with the other terms in Eq. (4) for moisture q.
d. The υ = 0 solution
e. The n = 1 dispersion relation
The above equation isolates the term containing Q0 so that we can see it as a perturbation to the dispersion relation for the υ = 0 mode. Another way to look at
The values for Re(ω) obtained from this equation fall well in the intraseasonal range for values of k representing planetary spatial scales. Further neglecting Q0 and Γ leads to the first-order approximation
Higher-order or global asymptotic approximations are possible and may provide further insight, but we do not pursue these in the present study. In practice, we solve ω in Eq. (20) for each wavenumber k numerically as in previous studies. We will return to physical interpretation by considering the relationship of the n = 1 mode to two simpler modes in section 5. The numerical solutions are constrained by the boundary condition in Eq. (12).
f. Detailed n = 1 solution
Equation (11) for
3. Parameter estimates for moist physics
The values of the eight free parameters in this linear system—α, r, Γ,
Next, we estimate the WISHE parameter ε and zonal moisture gradient
(a) Column-integrated zonal MSE advection (Hadv) vs U850 (zonal winds at 850 hPa). (b) Surface latent heat fluxes (LH) vs U850 using the ERA5 dataset. All three variables are averaged over 70°–90°E, 10°S–10°N. A 20–90-day bandpass filtering is applied to all the time series. The gray bars indicate the mean values.
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
According to the estimates above, we use a nondimensional value
(a),(d) Frequency, (b),(e) growth rate, and (c),(f) zonal phase speed for the υ = 0 and n = 1 mode with no meridional moisture gradient, for (left) Q0 = 0 and (right) Q0 = 0.22. Other parameters are α = 0.35, αϕ = 0.01, αu = 0.03,
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
The meridional moisture gradient parameter Q0 is more difficult to estimate. The term
4. Results
a. The n = 1 solutions and meridional moisture gradient
The n = 1 mode has between one and three physical solutions at different zonal wavenumbers k (see full spectrum in Fig. A2). The left column of Fig. 2 shows the intraseasonal branch of the dispersion relation for the n = 1 [Eq. (20)] and υ = 0 [Eq. (18)] modes with zero meridional moisture gradient Q0 = 0. The other parameter values are α = 0.35, αϕ = 0.01, αu = 0.03,
Analogous results for Q0 = 0.22, a value representative of the meridional moisture gradient at y = 1 during northern winter (and 4–5 times the zonal gradient), are displayed in the right column of Fig. 2; the other parameters are the same as for the Q0 = 0 solutions shown on the left. For k ≥ 2, the n = 1 solutions are considered unphysical and not shown, as Re(ξ1) turns negative with equatorward propagation (cy < 0), and the solution grows with y, which is consistent with the Eq. (24) that increasing k leads to smaller Re(ξ1). For smaller k, the n = 1 mode’s growth rate is greater than that of the υ = 0 mode. Comparing the two cases indicates that the meridional moisture gradient further contributes to planetary-scale instability, but stabilizes higher zonal wavenumbers. The zonal phase speed at k = 1 is 10.8 m s−1, about 20% higher than for Q0 = 0. The frequency–wavenumber relationship is different for the two values of Q0; for Q0 = 0.22, the frequency of the n = 1 mode increases approximately linearly with wavenumber, whereas for Q0 = 0 it varies only weakly below k = 2 and decreases with k thereafter, staying closer to the υ = 0 mode.
Figure 3 shows the spatial structures of the pressure, moisture, and wind fields associated with the n = 1 mode at k = 1 for six values of Q0. When Q0 = −0.05 this represents the positive meridional moisture gradient at y = 1 during northern summer, similar in magnitude to the zonal gradient but small compared to the meridional gradient in northern winter. The structures in ϕ and u are symmetric about the equator, and display northwest–southeast tilt for y > 1, and southwest–northeast tilt for y < −1, under small positive or negative Q0 values (Figs. 3a,b). We suggest that this pattern represents an idealized version of the BSISO without hemispheric asymmetry. Analysis of the moist static energy budget will be further discussed to support this claim in the next subsection. As Q0 further increases, the swallowtail-like wave pattern (Zhang and Ling 2012; Kim and Zhang 2021) emerges (Figs. 3c,d). Large Q0 (0.22) leads to nearly a neutral wave pattern at |y| = 2 [where by “neutral” we mean Im(ξ) = 0 as in Matsuno’s solutions], with cyclonic gyres straddling the equator, reminiscent of the classical Gill response to steady equatorial heating. The cyclonic gyres tilt farther eastward for even larger Q0 (Fig. 3f). No zonal wave k = 1 solution exists for Q0 > 0.54. Emanuel (2020; see Fig. 3 therein) also showed an MJO-like neutral wave pattern. This shape can be traced back to the WISHE term [α in a2 of Emanuel’s Eq. (9)] in the temperature equation [Emanuel’s Eq. (4)], which differs from our Q0 parameter. Regardless of this difference, both results indicate that it is a term of the form
Spatial pattern of the n = 1 solution with different values of the magnitude of the meridional moisture gradient Q0 at zonal wavenumber k = 1. Perturbation pressure ϕ (blue and red contours); moisture q: black; solid, 0.7 × maxq (maximum value of q); dashed, −0.7 × maxq. Horizontal wind vectors are in gray. Units are arbitrary; 1.5 zonal waves are shown for clarity. Parameters are as in Fig. 2 except for the different values of Q0.
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
Figure 4 shows the growth rate, zonal phase speed, meridional decay scale, and meridional wavenumber at y = 1 for zonal wave k = 1 as a function of Q0 over the range discussed above. The growth rate (Fig. 4a) varies nonmonotonically with Q0, with a maximum at Q0 = 0.15. The zonal phase speed (Fig. 4b) increases nearly linearly with Q0 from 8 to 12 m s−1 over the range shown. Figure 4c shows the meridional scale Ly = [Re(ξ1)]−1/2, which varies from 2.2 at Q0 = −0.05 to 1.62 at Q0 = 0.3, indicating that the solution becomes increasingly trapped near the equator as Q0 increases. The changing wave tilt with Q0 corresponds to the variations
(a) Growth rate, (b) zonal phase speed, (c) meridional scale [Re(ξ1)]−1/2, and (d) local meridional wavenumber ly at y = 1 for the six cases shown in Fig. 3.
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
The spatial patterns associated with these n = 1 solutions are associated with the two terms in Eq. (26) for ϕ, proportional to
The BSISO-like tilted structures found at large Q0 have relatively fine meridional scale compared to those of the more MJO-like structures found at small Q0. This might be the reason that some other theoretical models have captured the MJO-like pattern but not the BSISO-like pattern: the meridional resolution used in those models, typically n = 1–3 in the expansion in parabolic cylinder functions (e.g., Gill 1980), is inadequate to resolve the finer meridional structures of the BSISO, whereas these arise naturally here even for n = 1 due to the complexity of ξ.
Figure 5 shows the time–latitude diagram of convergence for k = 1. The constant C in Eq. (25a) is chosen to be unity for simplicity. The poleward propagation is symmetric about the equator, and its speed varies as a function of latitude [cf. Eq. (16)]. A bulk estimate of this meridional phase speed is ∼50–100 km day−1 between Y (distance from the equator) = 1000 and 2000 km when Q0 = −0.05 (Fig. 5a) and 0.1 (Fig. 5b). This value is broadly consistent with observations of the BSISO (e.g., Jiang et al. 2004; Wang et al. 2018), as is the poleward propagation in both hemispheres, as shown in Waliser et al. (2009, their Fig. 6). The poleward speed is faster for higher Q0. Wave phase is established at nearly the same time for all latitudes for Q0 = 0.22 (Fig. 5c).
Time–latitude diagram of low-level convergence for Q0 = −0.05, 0.1, and 0.22.
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
Various terms in the zonal momentum budget for (a) Q0 = −0.05 and (b) Q0 = 0.22 at t = 0. All quantities are averaged between y = 1.45 and 1.55.
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
b. Budgets of zonal momentum and moist static energy
In addition to its effect on the wave morphology, the magnitude of the meridional moisture gradient has a strong influence on the dominant zonal momentum balance. Figure 6 shows the zonal moment budget around y = 1.5. For small negative Q0 = −0.05, all four terms in Eq. (1)—tendency, Coriolis force, pressure gradient, and friction—have similar magnitudes. For large Q0 (0.22), the familiar geostrophic balance is struck between the Coriolis and pressure gradient forces, while the tendency and friction terms are small. This indicates that the use of the quasi-steady state assumption for u (Gill 1980) in several MJO theories (Sobel and Maloney 2012, 2013; Adames and Kim 2016) is appropriate.
Projection of different terms in the moist static energy (MSE) budget terms to the MSE tendency for (a) Q0 = −0.05 and (b) Q0 = 0.22. The six terms are Wadv, vertical advection; Uadv, zonal advection; E, surface fluxes; CRF, cloud-radiative feedback; Vadv, meridional advection; and damping in Eq. (28). The sum of these six terms equals 1 due to normalization:
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
Figure 8 shows the budgets for growth rate and frequency for Q0 = −0.05 and 0.22 for zonal wave k = 1, normalized by Im(ω) and −Re(ω), respectively. The dominant contribution of CRF to the growth rate is borne out for both Q0 cases (Figs. 8a,c), while vertical advection acts as the leading sink in the growth rate budget. The large contribution by the two processes is due to their phasing relative to the MSE, moisture, and precipitation anomalies; CRF is in phase with precipitation as it is expressed as proportionality to precipitation, while vertical advection is out of phase with precipitation. For frequency and propagation (Figs. 8a,c), zonal advection and surface fluxes oppose each other, while the former dominates the contribution to the frequency and propagation budgets. Meridional advection becomes important for Q0 = 0.22, only secondary to CRF, the leading contributor.
Budget of (left) growth rate and (right) frequency derived from the MSE equation for (a),(b) Q0 = −0.05 and (c),(d) Q0 = 0.22. The six terms are Wadv, vertical advection; Uadv, zonal advection; E, surface fluxes; CRF, cloud-radiative feedback; Vadv, meridional advection; and damping in Eq. (28). The sum of these six terms equals 1 due to normalization.
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
c. Sensitivity to parameters
, Γ, and α
The parameter space has eight dimensions, a number sufficiently large as to make it impossible to fully explore the space. Here, we only examine parameter sensitivity within limited ranges for
Figure 9 shows the sensitivity of the n = 1 and υ = 0 modes using the same control wave parameters as in Fig. 4 to the parameter that combines WISHE and the zonal moisture gradient,
Sensitivity of wave parameters to the parameter representing the net effect of zonal moisture advection and WISHE,
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
Figure 10 shows the sensitivity of the solutions to effective gross moist stability Γe by changing Γ only. For Γe < −0.05, the most unstable mode is short waves instead of planetary wavenumbers (e.g., 0.5 or 1). For Q0 = −0.05, growth rate of the n = 1 and υ = 0 solutions generally decreases with positive Γe (Fig. 10a), while it is relatively flat for Q0 = 0.22. Phase speed increases with Γe (Fig. 10b). The meridional scale (Fig. 10c) increases significantly with positive Γe, and decreases with negative Γe. The term
Sensitivity to effective gross moist stability Γe for zonal wave k = 1: (a) growth rate, (b) zonal phase speed, (c) meridional scale, and (d) local meridional wavenumber ly at y = 1.
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
Figure 11 shows the sensitivity to the convective rate constant α from 0.2 to 0.75, corresponding to a time scale varying from 1.7 days to 0.5 days. The growth rate (Fig. 11a) decreases with α while the phase speed increases with α for both the n = 1 and υ = 0 solutions when Q0 = 0, but both vary little with α when Q0 = 0.22. The meridional decay scale increases with α for both Q0 cases, while
Sensitivity to convective time scale α for zonal wave k = 1: (a) growth rate, (b) zonal phase speed, (c) meridional scale, and (d) local meridional wavenumber ly at y = 1. No solutions exist for larger values of α than those shown.
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
5. Matching between the υ = 0 mode and f-plane moisture wave
As discussed above, the algebraic structures of the n = 1 mode and their associated spatial patterns have two components, one dominant near the equator and the other in the subtropics. This suggests that some understanding might be gained by considering local approximations that represent these two components separately, and then considering under what conditions these might combine to form a single mode. We consider the υ = 0 mode near the equator; for higher latitudes, say |y| > 1, we consider the modes that would exist on an f plane. The spatial patterns of the υ = 0 and the f-plane wave modes are shown in the appendix (Figs. A3 and A4). We will argue that the structure of the n = 1 mode can be thought of as a consequence of phase speed matching between these two more local modes.
Using
We again consider the two cases Q0 = −0.05 and Q0 = 0.22. We do not expect the meridional wavenumber of the f-plane waves to match those of the beta plane solution (due to complete lack of constraint imposed by spherical geometry). Instead, we ask under what conditions the f-plane and υ = 0 modes have the same phase speeds, and in particular where the phase speed of the f-plane mode is relatively insensitive to the reference latitude used to choose the value of f so that the matching is robust to this somewhat arbitrary choice, as follows. If coupling between the υ = 0 and the f-plane moisture wave is a useful characterization of the dynamics of the n = 1 mode, the latter should have larger l for Q0 = −0.05 and smaller l for Q0 = 0.22. The relative magnitudes of these two would be consistent with those derived from the local wavenumber using Eq. (16). Figure 12 shows the zonal phase speed for k = 1 as a function of meridional wavenumber l for the two Q0 values with three different f values: 1, 1.5, and 2. Other parameters are kept the same as Fig. 2. In both Q0 cases, the υ = 0 solution has the same phase speed: 8.3 m s−1. For Q0 = −0.05, the three f waves approximately agree on the zonal phase speed for l ∼ 13 with a zonal phase speed ∼ 7.5 m s−1. For Q0 = 0.22, l ∼ 0.5–3 with a zonal phase speed ∼ 12–13 m s−1. The ratio between the values of l corresponding to these two values of Q0 for this f-plane wave,
(a) Phase speed of the zonal wave-1 f-plane moisture wave for Q0 = −0.05. (b) As in (a), but for Q0 = 0.22. Three f values are used: 1 (blue), 1.5 (green), and 2 (red). Open circles indicate stable solutions; closed symbols are unstable solutions. The horizontal bar in (a) indicates the phase speed 8.2 m s−1 of the n = 1 mode at k = 1 for Q0 = −0.05, and 10.8 m s−1 in (b).
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
The υ = 0 mode is a moisture mode (as a direct analog to the Kelvin mode) whose instability results from the net effect of WISHE and zonal moisture advection, while the f-plane moisture mode propagates eastward due to those effects as well as to meridional moisture advection (as in Sobel et al. 2001), although the latter becomes small as the meridional moisture gradient does (as is relevant in northern summer). The two share similar eastward phase speeds, hence phase speed matching between the equatorial υ = 0 mode and the subtropical wave solution can occur. This matching is best achieved at planetary wavelengths (k = 0.5 or 1), and the combined n = 1 mode on the β plane is also most unstable at that scale.
It is useful to think about how the phase speed matching between the υ = 0 moisture mode and the f-plane moisture wave differs from the conventional Kelvin–Rossby paradigm for the MJO. The free dry Rossby wave propagates westward while the free dry Kelvin wave propagates eastward, so the MJO cannot be thought of as resulting from an interaction between these free waves, but only their quasi-stationary forced counterparts (e.g., Gill 1980). Here, the matching of interest is between true free (moist) modes, and this allows us to explain the relationship of the MJO to the BSISO, something to our knowledge previously not well explained in terms of Rossby and Kelvin waves.
6. Conclusions
We have presented a moist shallow water model, representing a first baroclinic vertical mode structure as is common in theoretical studies of tropical atmosphere dynamics, linearized about a resting atmosphere on an equatorial beta plane. Idealized representations of moist physics are similar to those in many past studies. We assume a basic state with a positive zonal moisture gradient and westerly low-level wind, consistent with the observed mean state over the Indian Ocean (Sobel and Maloney 2013; Ahmed 2021). The moisture distribution is assumed to vary quadratically with latitude, a form that simplifies the analysis while remaining qualitatively consistent with observations.
The behavior of this simple model is interpreted as relevant to the dynamics of the MJO and BSISO over the Indian Ocean, where both modes most frequently develop. Normal mode solutions are obtained through linear stability analysis. The natural meridional structures of the modes are fundamentally different from those of the dry system; for some parameters, the gravest meridional mode with nonzero meridional velocity produces meridional oscillations whose wavelength decreases with latitude. Our analysis focuses on this gravest (n = 1) mode, as well as the mode with zero meridional velocity (broadly analogous to the Kelvin mode).
It is shown that, under the longwave approximation, the υ = 0 and n = 1 eastward-propagating modes share the same fundamental instability mechanisms. It is essential to both the instability and eastward propagation of the MJO- and BSISO-like modes that the moistening effect of low-level easterly anomalies due to horizontal advection exceeds their drying effect due to reduced surface fluxes (since the wind anomalies are superimposed on a westerly background state). The magnitude of the growth rate is controlled largely by the combination of these two effects, as well as by cloud-radiative instability.
The effect of the meridional moisture gradient is explored systematically, and we argue that this gradient determines whether the unstable intraseasonal modes are MJO-like or BSISO-like. It is found (Fig. 3) that, with nearly zero or small positive meridional moisture gradient (typical of the Northern Hemisphere in northern summer), the horizontal structure exhibits a northwest–southeast tilt in the Northern Hemisphere (and the opposite tilt in the Southern Hemisphere), consistent with the observed BSISO. As the meridional moisture gradient increases, the horizontal tilt decreases and the pattern transitions to a more swallowtail-like wave pattern. When the meridional moisture gradient reaches values typical of northern winter, the MJO-like structure emerges, with cyclonic gyres straddling the equatorial precipitation maximum in the subtropics of both hemispheres. We conclude that the meridional moisture gradient shapes the horizontal structures, leading to the transformation from the BSISO-like tilted horizontal structure to the MJO-like neutral wave structure, as the gradient increases from small positive values in northern summer to large negative values in northern winter. On the other hand, the meridional moisture gradient does not significantly change the time scale or growth rate of the most unstable mode.
We further argue that the n = 1 mode can be thought of as resulting from phase speed matching between two simpler and more local modes: the υ = 0 mode, which is trapped near the equator, and the waves found on an f plane representing the subtropics. The meridional moisture gradient Qy does not affect the υ = 0 mode, since that gradient enters the problem only through meridional advection. The meridional wavenumbers of the f-plane mode whose phase speeds best match those of the υ = 0 mode vary with Qy in a manner broadly consistent with the increasing meridional scale of the n = 1 mode with Qy.
While the idealized MJO here is recognizable in its similarity to the observed one, the idealized BSISO is perhaps less obviously so. We summarize its characteristics that compare favorably with observations as follows: 1) It has planetary scale. 2) There is poleward propagation in both hemisphere; see Figs. 3 and 6 of Waliser et al. (2009). 3) There is a northwest–southeast tilt of horizontal structure in Northern Hemisphere. 4) The MSE budget (Figs. 7 and 8) is dominated by zonal advection and surface fluxes, consistent with observational analysis in Fig. 4 of Jiang et al. (2018). 5) It exhibits seasonality: it is active in the boreal summer season with negative or small positive meridional moisture gradient. 6) The idealized BSISO initiates and grows in the Indian Ocean. One discrepancy between the theory and observation is that the present theory has neglected hemispheric asymmetry: the idealized BSISO is symmetric about the equator, while the observed BSISO is highly antisymmetric and mainly active in the Northern Hemisphere. To address this issue, we have performed linear stability analysis using a moisture profile that maximizes at off-equatorial latitude (e.g., 20°N). Our preliminary analysis suggests that while this indeed adds some realism, the fundamental dynamics of the moisture mode instability and propagation remain unchanged compared to the idealized symmetric BSISO mode. The results will be reported elsewhere.
The present study uses estimated parameter values relevant to the Indian Ocean basin to understand the initialization dynamics of the ISOs there. No attempt is made to explore the dynamics over the Pacific Ocean, where the MJO’s convection usually weakens while its circulation continues to propagate eastward in a manner often more consistent with a free Kelvin wave (e.g., Sobel and Kim 2012; Powell 2017). Our analysis fits into a conceptual picture in which the MJO and BSISO, as moisture modes per se, exist primarily over the Indian Ocean, and change form over the Maritime Continent and western North Pacific. In the latter case, the mean surface wind changes to easterly in the eastern Pacific Ocean, and the zonal moisture decreases to the east; both are opposite to their respective values in the Indian Ocean. The degree of cancellation between the two warrant careful analysis. The present study assumes constant zonal moisture gradient. While this assumption does not satisfy periodic boundary conditions, we consider it as a useful starting point, given the distinct behavior of the MJO and BSISO over the Indian Ocean and the much greater difficulty of conducting an analysis with a basic state gradient that varies in longitude. A more sophisticated analysis that explicitly accounts for zonal variation of the basic state and the transition from the Indian Ocean to the western Pacific is left for future work.
The limited range of physical processes considered here is not intended to imply that excluded processes are irrelevant. Boundary layer frictional convergence (Wang and Rui 1990) and shallow circulation (Wu 2003), for example, may also play important roles, perhaps by introducing more complex vertical structures and thus influencing the gross moist stability. Some aspects of the large-scale basic flow, neglected here except for its implicit presence in the WISHE term, might be especially important. In particular, the interaction of convection with the vertical wind shear associated with the Asian summer monsoon may be important to the behavior of the real BSISO, as argued in previous studies (e.g., Jiang et al. 2004; Bellon and Sobel 2008a).The trailing stratiform heating and dipolar vertical velocity in climate model simulations (Wang et al. 2017) and cloud-permitting simulations (Wang et al. 2015), together with the leading shallow circulation to the east, may alter eastward propagation (Wang and Li 2020) and the instability. Both effects may be considered by introducing the second baroclinic mode (Mapes 2000; Khouider and Majda 2006; Kuang 2008; Andersen and Kuang 2008) in the theory. Nonetheless, we find it compelling that the model used here appears to explain the different structures of the MJO and BSISO in terms of the observed meridional moisture gradient without including those processes. We view this as a reason to consider the MJO and BSISO as different manifestations of the same underlying dynamics, in the most basic sense, and the meridional moisture gradient as the critical parameter that controls the change in structure from one to the other with the seasons.
Acknowledgments.
We thank three anonymous reviewers for their insightful comments. We are thankful to K. A. Emanuel for sharing his moist shallow water model solver, which inspired the present study. This research was supported by Office of Naval Research Grant N00014-16-1-3073 and National Science Foundation Grant AGS-1543932, and by the Monsoon Mission Project under India’s Ministry of Earth Sciences. SW acknowledges support from NSFC 41875066.
Data availability statement.
The ERA5 reanalysis dataset and the DYNAMO sounding dataset are publicly available. Matlab code is available at Github (https://github.com/wangsg2526/mjobsiso), or upon request to the corresponding author.
APPENDIX
Formulation and Simplifications
a. Formulation of the model
We choose H0 ∼ 5 km. The value of Λw may be analytically specified [e.g.,
Surface sensible heat flux,
The factor
Note that Γq is negative. Defining normalized gross moist stability as Γ = (Γs + Γq)/Γs, an estimate based on the time mean of the DYNAMO sounding array data (Johnson et al. 2015) gives Γ = 0.18. Other estimates in the past suggest that Γ ranges from small negative values to ∼0.2.
Changing the variables as
This leads to the nondimensional moist shallow water equation set at the beginning of section 2 (after dropping the subscript 1 and subscript *). A similar form can also be derived from the primitive equation set in Neelin and Zeng (2000), or a simplifed Boussinesq equation set in Fuchs and Raymond (2005, 2017). It is also broadly similar to the bulk model in Emanuel (1987, 2020), or the 1.5-layer model in Liu and Wang (2016, 2017) and Wang and Chen (2017).
b. Approximations of the equation set
The leading-order balance in the moisture equation is between convergence (second term on the lhs) and rain (first term on the lhs). Further neglecting the right-hand side leads to conventional moisture convergence closure. Evaluation of these approximations would be straightforward with the same solver and parameters.
c. Rain and column-integrated moisture
We derive the linear relationship between precipitation and first moisture mode q1 using the DYNAMO northern sounding dataset. First, empirical orthogonal analysis is performed on 6-hourly moisture anomalies. The leading empirical orthogonal function (EOF; Fig. A1a) explains more than 66% of total variances. Second, the column-integrated first EOF of moisture anomalies is regressed to surface precipitation. Regression coefficient gives α, which varies from 1 to 1.3 days with different rain rate estimate (Fig. A1b). This estimate agrees with FR17, who used a 1-day time scale.
(a) Λq as derived from the leading EOF of the moisture anomalies from the DYNAMO northern sounding array. (b) Column-integrated first EOF anomaly 〈q1〉 and precipitation P anomalies. Both quantities are daily averaged.
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
Spectrum of the solution for the values of Q0 used in Fig. 3. Open circles indicate stable solutions; closed symbols are unstable solutions. The sizes of the symbols indicate growth rate.
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
d. The υ = 0 mode
Using S and rearranging it gives
Figure A3 shows the spatial structure of the υ = 0 mode using the following parameters: α = 0.35, αφ = 0.01, αu = 0.03,
The spatial structure of υ = 0 mode, showing ϕ (shading), convergence (black contour), and wind vectors. Parameters are as in Fig. 2a.
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
e. The f-plane moisture wave
Figure A4 shows the spatial structure of the f-plane moisture wave using the following parameters: α = 0.35, αφ = 0.01, αu = 0.03,
The spatial structure of f plane wave for k = 1, l = 15: ϕ (shading), convergence (black contour; 1.1, solid, −1.1, dashed), and wind vectors. Parameters are as in Fig. 2a.
Citation: Journal of Climate 35, 4; 10.1175/JCLI-D-21-0361.1
REFERENCES
Abramowitz, M., and I. A. Stegun, 1964: Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. Dover, 1046 pp.
Adames, Á. F., and D. Kim, 2016: The MJO as a dispersive, convectively coupled moisture wave: Theory and observations. J. Atmos. Sci., 73, 913–941, https://doi.org/10.1175/JAS-D-15-0170.1.
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.
Ahmed, F., 2021: The MJO on the equatorial beta-plane: An eastward propagating Rossby wave induced by meridional moisture advection. J. Atmos. Sci., 78, 3115–3135, https://doi.org/10.1175/JAS-D-21-0071.1.
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.
Andersen, J. A., and Z. Kuang, 2008: A toy model of the instability in the equatorially trapped convectively coupled waves on the equatorial beta plane. J. Atmos. Sci., 65, 3736–3757, https://doi.org/10.1175/2008JAS2776.1.
Andersen, J. A., and Z. Kuang, 2012: Moist static energy budget of MJO-like disturbances in the atmosphere of a zonally symmetric aquaplanet. J. Climate, 25, 2782–2804, https://doi.org/10.1175/JCLI-D-11-00168.1.
Bellon, G., and A. H. Sobel, 2008a: Poleward-propagating intraseasonal monsoon disturbances in an intermediate-complexity axisymmetric model. J. Atmos. Sci., 65, 470–489, https://doi.org/10.1175/2007JAS2339.1.
Bellon, G., and A. H. Sobel, 2008b: Instability of the axisymmetric monsoon flow and intraseasonal oscillation. J. Geophys. Res., 113, D07108, https://doi.org/10.1029/2007JD009291.
Drbohlav, H.-K. L., and B. Wang, 2005: Mechanism of the northward-propagating intraseasonal oscillation: Insights from a zonally symmetric model. J. Climate, 18, 952–972, https://doi.org/10.1175/JCLI3306.1.
Emanuel, K. A., 1987: An air–sea interaction model of intraseasonal oscillations in the tropics. J. Atmos. Sci., 44, 2324–2340, https://doi.org/10.1175/1520-0469(1987)044<2324:AASIMO>2.0.CO;2.
Emanuel, K. A., 2020: Slow modes of the equatorial waveguide. J. Atmos. Sci., 77, 1575–1582, https://doi.org/10.1175/JAS-D-19-0281.1.
Fuchs, Ž., and D. J. Raymond, 2002: Large-scale modes of a nonrotating atmosphere with water vapor and cloud–radiation feedbacks. J. Atmos. Sci., 59, 1669–1679, https://doi.org/10.1175/1520-0469(2002)059<1669:LSMOAN>2.0.CO;2.
Fuchs, Ž., and D. J. Raymond, 2005: Large-scale modes in a rotating atmosphere with radiative-convective instability and WISHE. J. Atmos. Sci., 62, 4084–4094, https://doi.org/10.1175/JAS3582.1.
Fuchs, Ž., and D. J. Raymond, 2017: A simple model of intraseasonal oscillations. J. Adv. Model. Earth Syst., 9, 1195–1211, https://doi.org/10.1002/2017MS000963.
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.
Hersbach, H., and Coauthors, 2020: The ERA5 global reanalysis. Quart. J. Roy. Meteor. Soc., 146, 1999–2049, https://doi.org/10.1002/qj.3803.
Jiang, X., T. Li, and B. Wang, 2004: Structures and mechanisms of the northward propagating boreal summer intraseasonal oscillation. J. Climate, 17, 1022–1039, https://doi.org/10.1175/1520-0442(2004)017<1022:SAMOTN>2.0.CO;2.
Jiang, X., Á. F. Adames, M. Zhao, D. Waliser, and E. Maloney, 2018: A unified moisture mode framework for seasonality of the Madden–Julian oscillation. J. Climate, 31, 4215–4224, https://doi.org/10.1175/JCLI-D-17-0671.1.
Johnson, R. H., P. E. Ciesielski, J. H. Ruppert, and M. Katsumata, 2015: Sounding-based thermodynamic budgets for DYNAMO. J. Atmos. Sci., 72, 598–622, https://doi.org/10.1175/JAS-D-14-0202.1.
Khairoutdinov, M. F., and K. Emanuel, 2018: Intraseasonal variability in a cloud-permitting near-global equatorial aquaplanet model. J. Atmos. Sci., 75, 4337–4355, https://doi.org/10.1175/JAS-D-18-0152.1.
Khouider, B., and A. J. Majda, 2006: A simple multicloud parameterization for convectively coupled tropical waves. Part I: Linear analysis. J. Atmos. Sci., 63, 1308–1323, https://doi.org/10.1175/JAS3677.1.
Kikuchi, K., 2021: The boreal summer intraseasonal oscillation (BSISO): A review. J. Meteor. Soc. Japan, 99, 933–972, https://doi.org/10.2151/jmsj.2021-045.
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.
Krishnamurti, T. N., and D. Subrahmanyam, 1982: The 30–50 day mode at 850 mb during MONEX. J. Atmos. Sci., 39, 2088–2095, https://doi.org/10.1175/1520-0469(1982)039<2088:TDMAMD>2.0.CO;2.
Kuang, Z., 2008: A moisture-stratiform instability for convectively coupled waves. J. Atmos. Sci., 65, 834–854, https://doi.org/10.1175/2007JAS2444.1.
Lawrence, D. M., and P. J. Webster, 2002: The boreal summer intraseasonal oscillation: Relationship between northward and eastward movement of convection. J. Atmos. Sci., 59, 1593–1606, https://doi.org/10.1175/1520-0469(2002)059<1593:TBSIOR>2.0.CO;2.
Li, T., L. Wang, M. Peng, B. Wang, C. Zhang, W. Lau, and H.-C. Kuo, 2018: A paper on the tropical intraseasonal oscillation published in 1963 in a Chinese journal. Bull. Amer. Meteor. Soc., 99, 1765–1779, https://doi.org/10.1175/BAMS-D-17-0216.1.
Liu, F., and B. Wang, 2016: Role of horizontal advection of seasonal-mean moisture in the Madden–Julian oscillation: A theoretical model analysis. J. Climate, 29, 6277–6293, https://doi.org/10.1175/JCLI-D-16-0078.1.
Liu, F., and B. Wang, 2017: Roles of the moisture and wave feedbacks in shaping the Madden–Julian oscillation. J. Climate, 30, 10 275–10 291, https://doi.org/10.1175/JCLI-D-17-0003.1.
Madden, R. A., and P. R. Julian, 1971: Detection of a 40–50 day oscillation in the zonal wind in the tropical Pacific. J. Atmos. Sci., 28, 702–708, https://doi.org/10.1175/1520-0469(1971)028<0702:DOADOI>2.0.CO;2.
Majda, A. J., and S. N. Stechmann, 2009: The skeleton of tropical intraseasonal oscillations. Proc. Natl. Acad. Sci. USA, 106, 8417–8422, https://doi.org/10.1073/pnas.0903367106.
Mapes, B. E., 2000: Convective inhibition, subgrid-scale triggering energy, and stratiform instability in a toy tropical wave model. J. Atmos. Sci., 57, 1515–1535, https://doi.org/10.1175/1520-0469(2000)057<1515:CISSTE>2.0.CO;2.
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., 2008: Primary and successive events in the Madden–Julian Oscillation. Quart. J. Roy. Meteor. Soc., 134, 439–453, https://doi.org/10.1002/qj.224.
Nanjundiah, R. S., J. Srinivasan, and S. Gadgil, 1992: Intraseasonal variation of the Indian summer monsoon. J. Meteor. Soc. Japan, 70, 529–550, https://doi.org/10.2151/jmsj1965.70.1B_529.
Neelin, J. D., and J.-Y. Yu, 1994: Modes of tropical variability under convective adjustment and the Madden–Julian oscillation. Part I: Analytical theory. J. Atmos. Sci., 51, 1876–1894, https://doi.org/10.1175/1520-0469(1994)051<1876:MOTVUC>2.0.CO;2.
Neelin, J. D., and N. Zeng, 2000: A quasi-equilibrium tropical circulation model—Formulation. J. Atmos. Sci., 57, 1741–1766, https://doi.org/10.1175/1520-0469(2000)057<1741:AQETCM>2.0.CO;2.
Powell, S. W., 2017: Successive MJO propagation in MERRA-2 reanalysis. Geophys. Res. Lett., 44, 5178–5186, https://doi.org/10.1002/2017GL073399.
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.
Sentić, S., Ž. Fuchs-Stone, and D. J. Raymond, 2020: The Madden–Julian oscillation and mean easterly winds. J. Geophys. Res. Atmos., 125, e2019JD030869, https://doi.org/10.1029/2019JD030869.
Shi, X., D. Kim, A. F. Adames, and J. Sukhatme, 2018: WISHE-moisture mode in an aquaplanet simulation. J. Adv. Model. Earth Syst., 10, 2393–2407, https://doi.org/10.1029/2018MS001441.
Sikka, D. R., and S. Gadgil, 1980: On the maximum cloud zone and the ITCZ over Indian longitudes during the southwest monsoon. Mon. Wea. Rev., 108, 1840–1853, https://doi.org/10.1175/1520-0493(1980)108<1840:OTMCZA>2.0.CO;2.
Sobel, A. H., and D. Kim, 2012: The MJO–Kelvin wave transition. Geophys. Res. Lett., 39, 2012GL053380, https://doi.org/10.1029/2012GL053380.
Sobel, A. H., and E. Maloney, 2012: An idealized semi-empirical framework for modeling the Madden–Julian oscillation. J. Atmos. Sci., 69, 1691–1705, https://doi.org/10.1175/JAS-D-11-0118.1.
Sobel, A. H., and E. Maloney, 2013: Moisture modes and the eastward propagation of the MJO. J. Atmos. Sci., 70, 187–192, https://doi.org/10.1175/JAS-D-12-0189.1.
Sobel, A. H., J. Nilsson, and L. M. Polvani, 2001: The weak temperature gradient approximation and balanced tropical moisture waves. J. Atmos. Sci., 58, 3650–3665, https://doi.org/10.1175/1520-0469(2001)058<3650:TWTGAA>2.0.CO;2.
Stechmann, S. N., and S. Hottovy, 2020: Asymptotic models for tropical intraseasonal oscillations and geostrophic balance. J. Climate, 33, 4715–4737, https://doi.org/10.1175/JCLI-D-19-0420.1.
Sugiyama, M., 2009: The moisture mode in the quasi-equilibrium tropical circulation model. Part I: Analysis based on the weak temperature gradient approximation. J. Atmos. Sci., 66, 1507–1523, https://doi.org/10.1175/2008JAS2690.1.
Waliser, D., and Coauthors, 2009: MJO simulation diagnostics. J. Climate, 22, 3006–3030, https://doi.org/10.1175/2008JCLI2731.1.
Wang, B., and H. Rui, 1990: Dynamics of the coupled moist Kelvin–Rossby wave on an equatorial β plane. J. Atmos. Sci., 47, 397–413, https://doi.org/10.1175/1520-0469(1990)047<0397:DOTCMK>2.0.CO;2.
Wang, B., and G. S. Chen, 2017: A general theoretical framework for understanding essential dynamics of Madden–Julian oscillation. Climate Dyn., 49, 2309–2328, https://doi.org/10.1007/s00382-016-3448-1.
Wang, L., and T. Li, 2020: Effect of vertical moist static energy advection on MJO eastward propagation: Sensitivity to analysis domain. Climate Dyn., 54, 2029–2039, https://doi.org/10.1007/s00382-019-05101-8.
Wang, L., T. Li, E. Maloney, and B. Wang, 2017: Fundamental causes of propagating and nonpropagating MJOS in MJOTF/GASS models. J. Climate, 30, 3743–3769, https://doi.org/10.1175/JCLI-D-16-0765.1.
Wang, S., A. H. Sobel, F. Zhang, Y. Q. Sun, Y. Yue, and L. Zhou, 2015: Regional simulation of the October and November MJO events observed during the CINDY/DYNAMO field campaign at gray zone resolution. J. Climate, 28, 2097–2119, https://doi.org/10.1175/JCLI-D-14-00294.1.
Wang, S., A. H. Sobel, and J. Nie, 2016: Modeling the MJO in a cloud-resolving model with parameterized large-scale dynamics: Vertical structure, radiation, and horizontal advection of dry air. J. Adv. Model. Earth Syst., 8, 121–139, https://doi.org/10.1002/2015MS000529.
Wang, S., D. Ma, A. H. Sobel, and M. K. Tippett, 2018: Propagation characteristics of BSISO indices. Geophys. Res. Lett., 45, 9934–9943, https://doi.org/10.1029/2018GL078321.
Wang, S., A. H. Sobel, C.-Y. Lee, D. Ma, S. Chen, M. Curcic, and J. Pullen, 2021: Propagating mechanisms of the 2016 summer BSISO event: Air–sea coupling, vorticity, and moisture. J. Geophys. Res. Atmos., 126, e2020JD033284, https://doi.org/10.1029/2020JD033284.
Webster, P. J., 1983: Mechanisms of low-frequency variability: Surface hydrological effects. J. Atmos. Sci., 40, 2110–2124, https://doi.org/10.1175/1520-0469(1983)040<2110:MOMLFV>2.0.CO;2.
Wu, Z., 2003: A shallow CISK, deep equilibrium mechanism for the interaction between large-scale convection and large-scale circulations in the tropics. J. Atmos. Sci., 60, 377–392, https://doi.org/10.1175/1520-0469(2003)060<0377:ASCDEM>2.0.CO;2.
Wu, Z., E. S. Sarachik, and D. S. Battisti, 2001: Thermally driven tropical circulations under Rayleigh friction and Newtonian cooling: Analytic solutions. J. Atmos. Sci., 58, 724–741, https://doi.org/10.1175/1520-0469(2001)058<0724:TDTCUR>2.0.CO;2.
Xie, Y.-B., S.-J. Chen, I.-L. Zhang, and Y.-L. Hung, 1963: A preliminarily statistic and synoptic study about the basic currents over southeastern Asia and the initiation of typhoon (in Chinese). Acta Meteor. Sin., 33, 206–217.
Yano, J.-I., and K. Emanuel, 1991: An improved model of the equatorial troposphere and its coupling with the stratosphere. J. Atmos. Sci., 48, 377–389, https://doi.org/10.1175/1520-0469(1991)048<0377:AIMOTE>2.0.CO;2.
Yasunari, T., 1979: Cloudiness fluctuations associated with the Northern Hemisphere summer monsoon. J. Meteor. Soc. Japan, 57, 227–242, https://doi.org/10.2151/jmsj1965.57.3_227.
Zhang, C., 2005: Madden–Julian Oscillation. Rev. Geophys., 43, RG2003,https://doi.org/10.1029/2004RG000158.
Zhang, C., and J. Ling, 2012: Potential vorticity of the Madden–Julian oscillation. J. Atmos. Sci., 69, 65–78, https://doi.org/10.1175/JAS-D-11-081.1.
Zhang, C., Á. F. Adames, B. Khouider, B. Wang, and D. Yang, 2020: Four theories of the Madden–Julian oscillation. Rev. Geophys., 58, e2019RG000685, https://doi.org/10.1029/2019RG000685.