1. Introduction
Synoptic-scale variability in the tropics is governed by a variety of disturbances that couple to convection. Among the least understood of these disturbances are cyclonic features that occur during the active months of the Indian monsoon. These disturbances are known as monsoon low pressure systems, or synoptic-scale monsoonal disturbances (SMDs). These systems exhibit a horizontal scale of ~2000 km (Godbole 1977; Krishnamurti et al. 1975; Sikka 1977; Lau and Lau 1990) and circulation anomalies that are strongest in the lower troposphere (Cohen and Boos 2016; Hunt et al. 2016; Adames and Ming 2018). SMDs also exhibit significant precipitation anomalies, which are a maximum to the west or southwest of the center of low pressure (Mooley 1973; Yoon and Chen 2005; Hunt et al. 2016). SMDs predominantly develop over the Bay of Bengal, propagate northwestward toward the Indian subcontinent with a phase speed of ~5 m s−1 (Mooley 1973; Godbole 1977; Krishnamurti et al. 1975, 1976; Sikka 1977; Hunt and Turner 2017). As they propagate northwestward, they produce a large fraction of the total monsoonal rainfall received by India (Stano et al. 2002; Ding and Sikka 2006; Yoon and Chen 2005; Yoon and Huang 2012). The influence of SMDs on total monsoon precipitation indicates that understanding the mechanisms by which they propagate and grow is of central importance to our understanding of the South Asian monsoon.
It has long been thought that SMDs are destabilized by a variant of baroclinic instability (Charney and Stern 1962; Bretherton 1966) that is modified by diabatic heating, referred to “moist baroclinic instability” (Salvekar et al. 1986; Krishnakumar et al. 1992; Aravequia et al. 1995; de Vries et al. 2010; Krishnamurti et al. 2013). Similarly to the original dry baroclinic instability model, this moist model includes counterpropagating waves that can phase lock and induce mutual growth. For this phase locking to occur, the anomalies in potential vorticity (PV) must tilt against the climatological-mean vertical wind shear (de Vries et al. 2010; Cohen and Boos 2016). However, it was recently shown by Cohen and Boos (2016) that the structure of SMDs, as described by modern reanalysis products, is inconsistent with the criteria necessary for moist baroclinic instability to occur. For example, they found that SMDs exhibit an upright vertical structure, which is inconsistent with the tilting necessary for baroclinic instability to occur. Furthermore, Boos et al. (2015) found that the quasigeostrophic (QG) approximation, which is often employed in studies of monsoon low pressure systems (Subrahmanyam et al. 1981; Sanders 1984), does not sufficiently describe the dynamics of SMDs. These results suggest that other processes may be responsible for the growth of these systems.
In a recent study, Adames and Ming (2018) investigated the potential role that moist processes may have in SMDs simulated in GFDL’s atmospheric general circulation model (AM4.0). They found a strong coupling between anomalous precipitation and column moisture. By jointly analyzing the moisture and moist static energy budgets, they found that the moisture anomalies propagate because of horizontal advection of mean dry static energy (DSE) by the anomalous winds. Horizontal advection of DSE induces a moisture tendency by forcing ascent along the sloping isentropes of the monsoon region. The anomalous precipitation is shifted ~5° to the east of the region of maximum ascent, where the moisture tendency is a maximum. This result suggests that the traditional QG assumption of forced ascent being tightly coupled to the precipitation field (Mak 1982; Sanders 1984) is incomplete, and moist thermodynamics may need to be included to fully explain the interaction between convection and the circulation in SMDs.
In this study, we propose a linear framework for SMDs. It incorporates a prognostic equation for column moisture, akin to the “moisture mode” framework commonly used to study the Madden–Julian oscillation (MJO; Fuchs and Raymond 2005, 2017; Raymond and Fuchs 2009; Sobel and Maloney 2012, 2013; Adames and Kim 2016). By incorporating a prognostic equation for column moisture, a wave instability arises that may describe the growth of SMDs. This instability involves a coupling between lower-tropospheric vorticity and column moisture. Warm-air and moisture advection by the anomalous winds moisten the free troposphere ahead of the cyclone, creating an environment favorable for deep convection. The subsequent convection intensifies the SMD through vortex stretching. The framework involves the use of equations reminiscent to those of the QG approximation but applied to the outer tropics ~16°N and with the addition of a prognostic moisture equation, which couples to precipitation through a linearized version of empirical relationship described by Bretherton et al. (2004).
While the focus of this study is on the growth and propagation of SMDs, we will show that it might be applicable to other modes of synoptic-scale tropical variability. Results from previous studies suggest that the same mechanism may operate in easterly waves both in the eastern Pacific and over Africa (Lau and Lau 1990, 1992; Kiladis et al. 2006; Serra et al. 2008; Berry and Thorncroft 2012; Janiga and Thorncroft 2013; Rydbeck and Maloney 2015) and monsoon depressions across the globe (Hurley and Boos 2015).
This paper is structured as follows. The next section offers a simple theoretical framework that can describe the structure and growth of SMDs. The dispersion relation of an idealized SMD is described in section 3. The role of precipitation in the growth of SMDs is discussed in section 4. Section 5 offers a simple equation that describes the zonal scale of SMDs. A discussion of the implications of the framework presented here is offered in section 6. A few concluding remarks are given in section 7.
2. A linear model for SMDs with prognostic moisture
In this section, we present a linear theoretical framework that can explain the growth and propagation of SMDs. We begin by presenting a basic set of equations linearized with respect to a background state with the following properties:
Zonally uniform mean state that is characterized by a weak zonal flow
A climatological-mean precipitation of ~10 mm day−1
The mean humidity and temperature increase with latitude












The terms on the right-hand sides (rhs) of Eqs. (7a) and (7b) are the Coriolis acceleration (f0 + βy) and the pressure gradient force, respectively. The rhs terms in Eq. (7c) are the meridional advection of mean temperature by the anomalous winds, vertical DSE advection, and heating by latent heat release, respectively. Finally, the equation for column moisture, which is converted to a precipitation equation via Eq. (6), contains three terms on the rhs: horizontal advection of mean moisture by the anomalous winds, vertical moisture advection, and the loss of moisture through precipitation. The terms in Eqs. (7c) and (7d) have been rearranged so that their similarities are more readily apparent. Note that advection of the terms in Eqs. (7a)–(7d) by the mean flow are neglected. This is because of the weak flow required to maintain simplicity in the analysis sought here. While strong zonal winds would be consistent with the observed monsoonal mean state, they would imply a strong mean vertical wind shear that would require inclusion of terms such as vertical advection of zonal momentum, which significantly complicate our analysis. Equations (7a)–(7d) are similar to those of Sukhatme (2014) and Monteiro and Sukhatme (2016), who also looked at the effects of moisture in a shallow-water system.


Schematic describing the structure and propagation of (a) a dry Rossby wave and (b) an idealized SMD in the gross PV framework. In the dry PV framework, meridional advection of mean PV by the anomalous flow induces westward propagation of the Rossby wave. Isentropic (QG) ascent, denoted by a red arrow, occurs in phase with the northerly anomalies while descent occurs with southerly wind anomalies. The gross PV framework in (b) is a generization of (a) that includes meridional moisture gradient and the impact of moist processes in the PV budget. In it, horizontal moisture advection and QG ascent moisten the atmosphere. This moistening causes deep convection to increase (blue arrow), which results in a vorticity tendency from vortex stretching. The propagation of the wave is then from a combination of dry PV advection and vortex stretching from moist processes.
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1
Equations (18a)–(18c) describe the relative importance of dry and moist processes to the propagation of an idealized SMD. This relationship is demonstrated in Fig. 2 for fixed values of βd and βd, but it also applies to
Contributions of βm and βd to βG as a function of the NGMS (
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1
In the following section, it will be shown that unstable Rossby-like modes arise from the equations described in this section. This implies that the gross PV is not a materially conserved quantity. Nonetheless, it will be shown that it is a useful quantity for understanding the dynamics of SMDs.
3. Dispersion relation
In this section, we will seek linear wave solutions to the equations presented in the previous section. To do this, several additional approximations are required. Meridional variations in
a. Full solution
Real component of the dispersion relation obtained from Eq. (20) for the (a) first and (b) second mode and the growth rate for the (c) first and (d) second mode. For all panels, τc = 12 h, βm = 1 × 10−11 m−1 s−1, and βd = 2.4 × 10−11 m−1 s−1. Dotted lines correspond to constant phase speeds of −15, −10, −5, and −2.5 m s−1, with decreasing values away from the origin. The location of maximum growth in (c) is denoted by an asterisk.
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1
b. Approximate solution

Figure 4 shows the approximate zonal phase speed and group velocity of the approximate solution for the largest meridional scales and values of βm ranging from 0 to 1 × 10−11 m−1 s−1. This corresponds to a temperature gradient that monotonically increases from 0 to 0.5 K (1000 km)−1 and a moisture gradient that increases from 0 to 8 mm (1000 km)−1. The wave exhibits westward phase propagation and a group velocity that is westward at the largest scales and eastward for smaller scales, and it is ~0 near zonal wavenumber 10, which is when
(a) Approximate phase speed obtained from Eq. (22a) for l = 0, Ld = 1300 km,
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1



(a) As in Fig. 3c, but for a value of τc of 1 h. (b) Approximate growth rate as obtained from Eq. (24). For both panels, βm = 1 × 10−11 m−1 s−1 and βd = 2.4 × 10−11 m−1 s−1. Dotted lines correspond to constant phase speeds of −15, −10, −5, −2.5, and −1 m s−1, with decreasing values away from the origin. The location of maximum growth is denoted by an asterisk.
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1
Figure 6a shows the growth rate for various values of βm. When βm = 0, the wave is weakly damped from zonal wavenumbers 0–30 and approximately neutral thereafter. For a small value of βm of 2 × 10−12 m−1 s−1, the wave is unstable for zonal wavenumbers 15 and larger, and no clear scale selection is seen. For larger values of βm, growth is largest at the synoptic scales, between zonal wavenumbers 10 and 25.
(a) As in Fig. 4, but showing the approximate growth rate as defined in Eq. (24). (b) Contributions to the growth rate by the contribution from
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1
4. Precipitation and its relation to wave propagation and growth
The propagation and growth of the wave solution in the previous subsection can be elucidated by examining the relationship between
Horizontal structure of an idealized SMD of zonal wavenumber 15 and a meridional structure that decays exponentially with the square of the distance from the reference latitude (16°N). The meridional decay is expressed through the formula exp[−(y − y0)2
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1

A schematic describing the phasing between P′ and ψ′, based on Eq. (27), is shown in Fig. 8. The interpretation is as follows. As the vortex propagates westward, the phasing between P′ and ψ′ depends on the duration of P′ and the speed in which the circulation anomalies propagate. The P′ is highly sensitive to 〈q′〉 when τc is small (~1 h) and responds quickly to moistening. Thus, small values of τc yield a P′ response that is closely in phase with moistening, namely, horizontal advection of moist PV. When τc is long, the sensitivity is lower, and precipitation increases slowly, resulting in a shift toward the vortex center. The magnitude of the shifting also depends on the wave frequency since faster propagation will result in the wave propagating toward the region of precipitation more quickly. The relation to the NGMS can be understood as follows. When the NGMS is larger, moisture is removed quickly in a precipitating column, limiting the duration of precipitation. When the NGMS is small, moisture is removed slowly, allowing for longer duration of precipitation. Longer-lasting precipitation results in a larger shift toward the center of low pressure.
Schematic describing the phase relationship between the circulation and precipitation for the wave solution in Eq. (20). For synoptic-scale waves, the phase angle can be broken down into four quadrants, the location which is determined by the direction of the total zonal phase speed
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1
The phasing between P′ and ψ′ also depends on cp and cpm, which determine the growth rate [Eq. (24)]. As shown in Fig. 7a, the wave grows when cp and cpm are in the same direction. Figure 7c shows the case in which cp is westward but cpm is eastward. Anomalous southerly flow now induces moistening, and precipitation is a maximum to the east of the moistening. This results in a P′ that is slightly in phase with the anticyclonic anomalies. Because precipitation generates a cyclonic tendency, it follows that it will damp the anticyclonic anomalies.
From inspection of Eq. (25), we can define two different processes that contribute to precipitation, corresponding to the terms in parentheses in the numerator of Eq. (25). The first is the advection of moist PV by the anomalous winds, which we will refer to as the advective contribution
Figure 9 shows the P anomalies from Fig. 7 decomposed into
As in Fig. 7a, but with (a)
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1
The growing and damped modes can also be understood in terms of
As in Fig. 7, but showing (a),(c)
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1
5. A qualitative relation for the SMD zonal scale









6. Synthesis and discussion
In this study, we presented a linear framework that may explain the propagation and growth of SMDs. Based on results from recent observational and modeling studies (Hunt et al. 2016; Adames and Ming 2018), we propose that column moisture plays a central role in SMDs. By including a prognostic moisture equation and coupling it to precipitation through a simplified Betts–Miller scheme, we derive a dispersion that describes the following features of SMDs, including
propagation mechanism,
instability and growth, and
synoptic scale (wavelengths of 2000–3000 km).
The instability occurs only if the wave’s dry and moist phase speeds are in the same direction. The dry phase speed includes dry processes such as advection of planetary vorticity by the anomalous winds and vortex stretching from dry isentropic ascent. The moist phase speed only includes propagation that is induced by vortex stretching from deep convection. This stretching is a result of isentropic ascent and horizontal moisture advection moistening the free troposphere. These processes produce an instability by producing an environment conducive to deep convection. This subsequent convection occurs in phase with the cyclonic anomalies, intensifying them through vortex stretching.
a. Comparison to baroclinic instability
Previous studies have sought to describe SMDs as a result of moist baroclinic instability (Charney and Stern 1962; Bretherton 1966; de Vries et al. 2010), which relies on the existence of counterpropagating waves for their growth (Sanders 1984; Snyder and Lindzen 1991). In this framework, observed SMDs would exhibit structures that tilt eastward with height. This tilt was not found in observations (Cohen and Boos 2016), suggesting that SMDs are not the result of moist baroclinic instability. Unlike existing models of moist baroclinic instability [de Vries et al. 2010; see review by Cohen and Boos (2016)], the framework presented here only involves dry and moist waves in the lower troposphere propagating in the same direction. The middle and upper troposphere do not play a role in the proposed instability. Furthermore, the vertical structure of the idealized SMDs exhibit no tilt, which is more consistent with observations (Cohen and Boos 2016).
Most studies that involve the use of baroclinic instability diagnose convection from the QG omega equation. In the QG framework, ascent occurs in regions of warm-air advection, P′ ∝ υ′βT. Using this parameterization, where moisture is not prognostic, yields solutions similar to those shown in section 2, but they are neutrally stable.
In the framework presented here, convection is parameterized in terms of column water vapor. By doing this, isentropic ascent is not just associated with convection but to a positive moisture tendency, as seen in Eq. (15b) and Fig. 10. Such a relationship was found by Adames and Ming (2018), suggesting that moisture may play a critical role in the growth of SMDs.
b. Comparison to the balanced moisture waves of Sobel et al. (2001)
While the framework presented here differs significantly from models of moist baroclinic instability, it resembles the balanced moisture waves described by Sobel et al. [2001; see their Eqs. (42) and (45)]. The main difference between the solutions shown here and theirs is the presence of a temperature gradient in addition to a meridional moisture gradient and that temperature tendencies are not neglected. This implies that the low-frequency background is not in weak temperature gradient balance. This deviation arises from the land–sea contrast over the South Asian monsoon region and the finite value of the Rossby radius of deformation. We can obtain solutions similar to those of Sobel et al. (2001) by applying the weak temperature gradient (WTG) approximation to Eq. (20). Thus, the balanced moisture waves of Sobel et al. (2001) are a special case of the SMDs described here. Furthermore, the gross PV equation [Eq. (18)] is consistent with the discussion of Sobel (2002), who likened the moisture field to the potential vorticity field in the deep tropics. The idealized SMD solutions are also similar to the moist Rossby waves described by Monteiro and Sukhatme (2016), who looked at the effects of moisture on midlatitude Rossby waves. Our results differ from theirs only because we investigate the case when both temperature and moisture increase with latitude, as observed in the South Asian monsoon.
c. Comparison with moisture mode theory of the MJO
It is worth comparing the moist waves described here to the linear moisture mode theory of the MJO (Fuchs and Raymond 2005, 2007, 2017; Sobel and Maloney 2012, 2013; Adames and Kim 2016). In the moisture mode framework, the growth rate is negatively related to the NGMS and inversely proportional to τc. In the approximate solutions shown in section 3 [Eq. (24)], the relationships between NGMS, τc, and growth are inverse to those in MJO moisture mode theory. This difference is due to the way that moisture interacts with the anomalies in both frameworks. In idealized SMDs, propagation and growth are partly due to the impact of moisture–convection feedbacks on the vorticity (wind) field. In the MJO, propagation is due to the modulation of moisture by the wind anomalies.
This result implies that the existence of a prognostic moisture equation can cause instability through different mechanisms. In the MJO, the process has been referred to as moisture mode instability or simply moisture instability. The instability for the idealized SMDs may be interpreted as a different type of moisture instability but related to the impact moisture has on PV. Thus, it may be more appropriate to refer to this instability as “moisture vortex” instability. That the interaction between moisture and the large-scale flow can lead to more than one instability is interesting, and their manifestation in observations is worth exploring in the future.
d. Radiative heating



e. Relevance to other tropical depression disturbances
The dispersion relation in Eq. (20) may shed some insight onto other synoptic-scale tropical disturbances. For example, Kiladis et al. (2006) found that divergence of the Q vectors (which is related to βT) leads convection by 1/8 cycle. Furthermore, Rydbeck and Maloney (2015) and Rydbeck et al. (2017) analyzed easterly waves occurring over the tropical northeastern Pacific, while Hsieh and Cook (2005, 2007) analyzed easterly waves occurring over Africa. The former found that when the waves are developing, meridional moisture advection and vortex stretching play a key role in the growth and propagation of these waves. The latter found that the position of the ITCZ and latent heat release from convection was more important for African easterly development than for the strength of the African easterly jet. Furthermore, Cornforth et al. (2009) found that including moist processes in easterly waves caused their amplitude to triple when compared to dry easterly waves. Supporting these results, a regional modeling study by Berry and Thorncroft (2012) found that African easterly waves weaken steadily when moist convection is turned off in their model (their Fig. 11). They conclude that easterly waves exhibit hybrid adiabatic and diabatic structures. These results suggest that the dynamics of easterly waves over the eastern Pacific and Africa may be described in terms of a gross potential vorticity, and the waves may grow through moisture vortex instability. It is possible that this instability is also relevant in other regions where disturbances that resemble SMDs form (see Hurley and Boos 2015).
Several studies have suggested that interactions between PV and water vapor play an important role in tropical cyclogenesis (Raymond et al. 2007, 2011). Furthermore, It was found by Bracken and Bosart (2000) that QG ascent may play a role in developing cyclones. By using a convection scheme similar to Betts–Miller, Brammer and Thorncroft (2015) found that developing cyclones exhibit higher positive horizontal moisture advection ahead of the disturbance than nondeveloping ones. Collectively, these results suggest that the framework presented here could be generalized and made applicable to tropical cyclone development.
f. Caveats
There are several caveats to the framework presented here. The mean monsoonal state is characterized by zonal westerlies on the order of 5–10 m s−1 and a stronger temperature gradient of 2–4 K (106 m)−1, which yields a substantial vertical wind shear. We chose a small temperature gradient that results in a negligible wind shear in order to keep the analysis simple. This results in meridional moisture advection playing a larger role than temperature advection in the idealized SMDs presented here. This is inconsistent with the results of Adames and Ming (2018), who found that temperature advection is the largest contributor to SMD propagation. Because of the weak flow assumed in this study, advection of moisture and vorticity by the background flow do not play a role in the idealized SMDs analyzed here. However, observed mean winds in the tropics have been shown to play a significant role in the propagation of SMDs and other tropical waves (Yanai and Nitta 1967; Lau and Lau 1992; Rydbeck and Maloney 2015). Another caveat is assuming a constant value of τc and the NGMS. Large variations in mean moisture precipitation and temperature across the monsoon region can lead to large spatial variations in both the NGMS and τc.
Furthermore, the vertical structure of the wave is treated as being directly related to the profile of vertical motion. As a result, the idealized SMDs have a first baroclinic vertical structure. Observed SMDs, however, exhibit a vertical structure where the maximum winds are strongest near the surface. Wang and Xie (1996) interpreted this structure as a combination of barotropic and baroclinic modes in the presence of easterly shear. A more realistic SMD structure could be obtained by having a more realistic mean state with strong easterly shear. While we hypothesize that the instability proposed here is still valid for a more realistic mean state, the simplifications done here to more clearly elucidate our proposed moisture vortex instability is a caveat of this study.
7. Concluding remarks
The linear framework in this study suggests an instability mechanism for tropical lows in a troposphere characterized by high humidity and horizontal moisture and temperature gradients. The proposed instability implies that SMDs may grow without barotropic or baroclinic instabilies. It could also complement these instabilities in regions where they have been shown to be relevant, such as in African easterly waves (Yanai and Nitta 1967; Hall et al. 2006; Hsieh and Cook 2007). It would be interesting to see if such a mechanism can be clearly identified in observations.
The gross PV (qG) equation presented could be a useful tool in understanding the propagation and growth of tropical disturbances. By just inspecting the magnitude and phase of horizontal advection of qG with respect to the qG center, it can be determined whether the disturbance is growing and its propagation speed and direction. A generalized version of Eq. (18a) may be obtained that could be used in diagnostic studies of tropical waves. It would be interesting to see if a generalized form of the gross PV equation can explain the meridional propagation of monsoon depressions by incorporating interactions between moisture and nonlinear horizontal vorticity advection. It was found by Boos et al. (2015) that this “beta drift” plays a central role in the northwestward propagation of monsoon low pressure systems.
The linear framework presented here presents an intriguing interpretation of SMDs, which can be tested in model simulations. For example, simulations in which moisture is treated diagnostically may produce weaker disturbances than those that treat moisture prognostically, as suggested by the growth rate in Eq. (24). These and other studies are interesting directions for future study.
Acknowledgments
This work was supported by the National Oceanic and Atmospheric Administration (NOAA) Grant NA15OAR4310099. We thank Eric Maloney for conversations that motivated some of the discussion presented here. We would also like to thank Kuniaki Inoue, Nadir Jeevanjee, Isaac Held, George Kiladis, David Raymond, and an anonymous reviewer for comments that greatly helped improve the contents of the manuscript.
APPENDIX
Derivation and Scaling of the Linear Equations
a. Scaling of the basic equations
For a wind field U on the order of 1 m s−1, a planetary vorticity of f0 = 4 × 10−5 s−1, and a horizontal-scale L on the order of 106 m, the mean zonal flow is, to first order, in geostrophic balance. It can be shown that this scaling is consistent with a Rossby number on the order of ~10−1 [Ro = U/(f0L)]. It follows that the thermal wind UT is on the order of 1 m s−1 for a horizontal temperature change on the order of 10−1 K (106 m)−1.





The moisture equation [Eq. (A1f)] is scaled as follows. Because vertical variations in moisture are much larger than horizontal variations, two different scales are used for them (δpq and δLq). In the tropics, where water vapor concentrations are high, vertical variations in mean latent heat are comparable to those in mean dry static energy, such that





b. Vertical truncation of the equations

Basis functions of (a) vertical velocity Ω, (b) horizontal winds and geopotential Λ, and (c) temperature a. The Ω has been normalized by dividing by
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1


c. Mean state









Vertical profiles of (a) mean dry static energy
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1









d. Propagation and growth for various parameter values
In section 3, we obtained linear wave solutions that could describe monsoon low pressure systems. The solutions were shown for specific parameter values. In this section, we analyze how changing these parameter values affects these solutions. Figure A3a shows the real component of the wave frequency, as obtained from Eq. (20) for l = 0 and various values of βq. Frequency is a maximum near zonal wavenumber 6, and frequency increases as βq increases. Figure A3b shows the wave frequency for different values of τc. It is clear that larger values of τc result in weaker propagation. This is due to a longer delay in convective onset, which results in precipitation contributing less to the vorticity tendency, as shown in section 4. When variation in NGMS are considered (Fig. A3c), we see a shift of the maximum frequency toward smaller k as the NGMS increases. Between zonal wavenumbers 10 and 40, we observe a decrease in frequency as NGMS increases, likely from a decreasing contribution of moist processes to propagation. Increasing the Rossby radius of deformation, shown in Fig. A3d, yields faster propagation across all wavenumbers.
(a) Real component of the dispersion relation in Eq. (20) for l = 0 and various values of βq. (b) As in (a), but as a function of zonal wavenumber and τc. (c) Frequency as a function of k and NGMS. (d) Frequency as a function of k and Ld. Values of other parameters are shown above each panel.
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1
Figure A4a shows the growth rate of the dispersion shown in Eq. (20) for l = 0 and various values of βq. Stronger growth is observed for larger values of βq, and the maximum growth shifts to smaller k as βq increases. Damping at the largest scales weakens as βq increases. The growth rate for l = 0 but for various values of τc is shown in Fig. A4b. The strongest growth occurs for time scales of 4–6 h and near zonal wavenumber 15. For larger values of τc, growth is weaker and is shifted toward smaller scales, with the strongest growth occurring near zonal wavenumber 20 for τc values of 1 day. For variations in the NGMS (
As in Fig. A3, but showing the growth rate. (c),(d) Dots show the zonal scale estimated from Eq. (28), which roughly corresponds to the maximum growth rate in the approximate solution (Eq. 24).
Citation: Journal of the Atmospheric Sciences 75, 6; 10.1175/JAS-D-17-0310.1
REFERENCES
Adames, Á. F., 2017: Precipitation budget of the Madden–Julian oscillation. J. Atmos. Sci., 74, 1799–1817, https://doi.org/10.1175/JAS-D-16-0242.1.
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., and Y. Ming, 2018: Moisture and moist static energy budgets of South Asian monsoon low pressure systems in GFDL AM4.0. J. Atmos. Sci., https://doi.org/10.1175/JAS-D-17-0309.1, in press.
Aravequia, J. A., V. B. Rao, and J. P. Bonatti, 1995: The role of moist baroclinic instability in the growth and structure of monsoon depressions. J. Atmos. Sci., 52, 4393–4409, https://doi.org/10.1175/1520-0469(1995)052<4393:TROMBI>2.0.CO;2.
Berry, G. J., and C. D. Thorncroft, 2012: African easterly wave dynamics in a mesoscale numerical model: The upscale role of convection. J. Atmos. Sci., 69, 1267–1283, https://doi.org/10.1175/JAS-D-11-099.1.
Betts, A. K., 1986: A new convective adjustment scheme. Part I: Observational and theoretical basis. Quart. J. Roy. Meteor. Soc., 112, 677–691, https://doi.org/10.1002/qj.49711247307.
Betts, A. K., and M. J. Miller, 1986: A new convective adjustment scheme. Part II: Single column tests using GATE wave, BOMEX, ATEX and Arctic air-mass data sets. Quart. J. Roy. Meteor. Soc., 112, 693–709, https://doi.org/10.1002/qj.49711247308.
Boos, W. R., J. V. Hurley, and V. S. Murthy, 2015: Adiabatic westward drift of Indian monsoon depressions. Quart. J. Roy. Meteor. Soc., 141, 1035–1048, https://doi.org/10.1002/qj.2454.
Bracken, W. E., and L. F. Bosart, 2000: The role of synoptic-scale flow during tropical cyclogenesis over the North Atlantic Ocean. Mon. Wea. Rev., 128, 353–376, https://doi.org/10.1175/1520-0493(2000)128<0353:TROSSF>2.0.CO;2.
Brammer, A., and C. D. Thorncroft, 2015: Variability and evolution of African easterly wave structures and their relationship with tropical cyclogenesis over the eastern Atlantic. Mon. Wea. Rev., 143, 4975–4995, https://doi.org/10.1175/MWR-D-15-0106.1.
Bretherton, C. S., M. E. Peters, and L. E. Back, 2004: Relationships between water vapor path and precipitation over the tropical oceans. J. Climate, 17, 1517–1528, https://doi.org/10.1175/1520-0442(2004)017<1517:RBWVPA>2.0.CO;2.
Bretherton, F. P., 1966: Baroclinic instability and the short wavelength cut-off in terms of potential vorticity. Quart. J. Roy. Meteor. Soc., 92, 335–345, https://doi.org/10.1002/qj.49709239303.
Charney, J. G., and M. E. Stern, 1962: On the stability of internal baroclinic jets in a rotating atmosphere. J. Atmos. Sci., 19, 159–172, https://doi.org/10.1175/1520-0469(1962)019<0159:OTSOIB>2.0.CO;2.
Cohen, N. Y., and W. R. Boos, 2016: Perspectives on moist baroclinic instability: Implications for the growth of monsoon depressions. J. Atmos. Sci., 73, 1767–1788, https://doi.org/10.1175/JAS-D-15-0254.1.
Cornforth, R. J., B. J. Hoskins, and C. D. Thorncroft, 2009: The impact of moist processes on the African easterly jet–African easterly wave system. Quart. J. Roy. Meteor. Soc., 135, 894–913, https://doi.org/10.1002/qj.414.
de Vries, H., J. Methven, T. H. A. Frame, and B. J. Hoskins, 2010: Baroclinic waves with parameterized effects of moisture interpreted using Rossby wave components. J. Atmos. Sci., 67, 2766–2784, https://doi.org/10.1175/2010JAS3410.1.
Ding, Y., and D. R. Sikka, 2006: Synoptic systems and weather. The Asian Monsoon, Springer, 131–201, https://doi.org/10.1007/3-540-37722-0_4.
Frierson, D. M. W., 2007: The dynamics of idealized convection schemes and their effect on the zonally averaged tropical circulation. J. Atmos. Sci., 64, 1959–1976, https://doi.org/10.1175/JAS3935.1.
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, 2007: A simple, vertically resolved model of tropical disturbances with a humidity closure. Tellus, 59A, 344–354, https://doi.org/10.1111/j.1600-0870.2007.00230.x.
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.
Godbole, R. V., 1977: The composite structure of the monsoon depression. Tellus, 29, 25–40, https://doi.org/10.3402/tellusa.v29i1.11327.
Haertel, P. T., G. N. Kiladis, A. Denno, and T. M. Rickenbach, 2008: Vertical-mode decompositions of 2-day waves and the Madden–Julian oscillation. J. Atmos. Sci., 65, 813–833, https://doi.org/10.1175/2007JAS2314.1.
Hall, N. M. J., G. N. Kiladis, and C. D. Thorncroft, 2006: Three-dimensional structure and dynamics of African easterly waves. Part II: Dynamical modes. J. Atmos. Sci., 63, 2231–2245, https://doi.org/10.1175/JAS3742.1.
Hsieh, J.-S., and K. H. Cook, 2005: Generation of African easterly wave disturbances: Relationship to the African easterly jet. Mon. Wea. Rev., 133, 1311–1327, https://doi.org/10.1175/MWR2916.1.
Hsieh, J.-S., and K. H. Cook, 2007: A study of the energetics of African easterly waves using a regional climate model. J. Atmos. Sci., 64, 421–440, https://doi.org/10.1175/JAS3851.1.
Hunt, K. M. R., and A. G. Turner, 2017: The effect of horizontal resolution on Indian monsoon depressions in the Met Office NWP model. Quart. J. Roy. Meteor. Soc., 143, 1756–1771, https://doi.org/10.1002/qj.3030.
Hunt, K. M. R., A. G. Turner, P. M. Inness, D. E. Parker, and R. C. Levine, 2016: On the structure and dynamics of Indian monsoon depressions. Mon. Wea. Rev., 144, 3391–3416, https://doi.org/10.1175/MWR-D-15-0138.1.
Hurley, J. V., and W. R. Boos, 2015: A global climatology of monsoon low-pressure systems. Quart. J. Roy. Meteor. Soc., 141, 1049–1064, https://doi.org/10.1002/qj.2447.
Janiga, M. A., and C. D. Thorncroft, 2013: Regional differences in the kinematic and thermodynamic structure of African easterly waves. Quart. J. Roy. Meteor. Soc., 139, 1598–1614, https://doi.org/10.1002/qj.2047.
Jiang, X., M. Zhao, E. D. Maloney, and D. E. Waliser, 2016: Convective moisture adjustment time scale as a key factor in regulating model amplitude of the Madden-Julian oscillation. Geophys. Res. Lett., 43, 10 412–10 419, https://doi.org/10.1002/2016GL070898.
Kiladis, G. N., C. D. Thorncroft, and N. M. J. Hall, 2006: Three-dimensional structure and dynamics of African easterly waves. Part I: Observations. J. Atmos. Sci., 63, 2212–2230, https://doi.org/10.1175/JAS3741.1.
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; Corrigendum, 49, RG3004, https://doi.org/10.1029/2008RG000266.
Kim, D., M.-S. Ahn, I.-S. Kang, and A. D. Del Genio, 2015: Role of longwave cloud–radiation feedback in the simulation of the Madden–Julian oscillation. J. Climate, 28, 6979–6994, https://doi.org/10.1175/JCLI-D-14-00767.1.
Krishnakumar, V., R. N. Keshavamurty, and S. V. Kasture, 1992: Moist baroclinic instability and the growth of monsoon depressions—Linear and nonlinear studies. Proc. Indian Acad. Sci., Earth Planet. Sci., 101, 123–152.
Krishnamurti, T.-N., M. Kanamitsu, R. Godbole, C.-B. Chang, F. Carr, and J. H. Chow, 1975: Study of a monsoon depression (I): Synoptic structure. J. Meteor. Soc. Japan, 53, 227–239, https://doi.org/10.2151/jmsj1965.53.4_227.
Krishnamurti, T.-N., M. Kanamitsu, R. Godbole, C.-B. Chang, F. Carr, and J. H. Chow, 1976: Study of a monsoon depression (II): Dynamical structure. J. Meteor. Soc. Japan, 54, 208–226, https://doi.org/10.2151/jmsj1965.54.4_208.
Krishnamurti, T. N., A. Martin, R. Krishnamurti, A. Simon, A. Thomas, and V. Kumar, 2013: Impacts of enhanced CCN on the organization of convection and recent reduced counts of monsoon depressions. Climate Dyn., 41, 117–134, https://doi.org/10.1007/s00382-012-1638-z.
Lau, K.-H., and N.-C. Lau, 1990: Observed structure and propagation characteristics of tropical summertime synoptic scale disturbances. Mon. Wea. Rev., 118, 1888–1913, https://doi.org/10.1175/1520-0493(1990)118<1888:OSAPCO>2.0.CO;2.
Lau, K.-H., and N.-C. Lau, 1992: The energetics and propagation dynamics of tropical summertime synoptic-scale disturbances. Mon. Wea. Rev., 120, 2523–2539, https://doi.org/10.1175/1520-0493(1992)120<2523:TEAPDO>2.0.CO;2.
Mak, M., 1982: On moist quasi-geostrophic baroclinic instability. J. Atmos. Sci., 39, 2028–2037, https://doi.org/10.1175/1520-0469(1982)039<2028:OMQGBI>2.0.CO;2.
Monteiro, J. M., and J. Sukhatme, 2016: Quasi-geostrophic dynamics in the presence of moisture gradients. Quart. J. Roy. Meteor. Soc., 142, 187–195, https://doi.org/10.1002/qj.2644.
Mooley, D. A., 1973: Some aspects of Indian monsoon depressions and the associated rainfall. Mon. Wea. Rev., 101, 271–280, https://doi.org/10.1175/1520-0493(1973)101<0271:SAOIMD>2.3.CO;2.
Neelin, J. D., and I. M. Held, 1987: Modeling tropical convergence based on the moist static energy budget. Mon. Wea. Rev., 115, 3–12, https://doi.org/10.1175/1520-0493(1987)115<0003:MTCBOT>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.
Raymond, D. J., and Ž. Fuchs, 2009: Moisture modes and the Madden–Julian oscillation. J. Climate, 22, 3031–3046, https://doi.org/10.1175/2008JCLI2739.1.
Raymond, D. J., S. L. Sessions, and Ž. Fuchs, 2007: A theory for the spinup of tropical depressions. Quart. J. Roy. Meteor. Soc., 133, 1743–1754, https://doi.org/10.1002/qj.125.
Raymond, D. J., S. L. Sessions, A. H. Sobel, and Ž. Fuchs, 2009: The mechanics of gross moist stability. J. Adv. Model. Earth Syst., 1 (3), https://doi.org/10.3894/JAMES.2009.1.9.
Raymond, D. J., S. L. Sessions, and C. López Carrillo, 2011: Thermodynamics of tropical cyclogenesis in the northwest Pacific. J. Geophys. Res., 116, D18101, https://doi.org/10.1029/2011JD015624.
Raymond, D. J., Ž. Fuchs, S. Gjorgjievska, and S. Sessions, 2015: Balanced dynamics and convection in the tropical troposphere. J. Adv. Model. Earth Syst., 7, 1093–1116, https://doi.org/10.1002/2015MS000467.
Rushley, S. S., D. Kim, C. S. Bretherton, and M.-S. Ahn, 2018: Reexamining the nonlinear moisture-precipitation relationship over the tropical oceans. Geophys. Res. Lett., 45, 1133–1140, https://doi.org/10.1002/2017GL076296.
Rydbeck, A. V., and E. D. Maloney, 2015: On the convective coupling and moisture organization of east Pacific easterly waves. J. Atmos. Sci., 72, 3850–3870, https://doi.org/10.1175/JAS-D-15-0056.1.
Rydbeck, A. V., E. D. Maloney, and G. J. Alaka Jr., 2017: In situ initiation of east Pacific easterly waves in a regional model. J. Atmos. Sci., 74, 333–351, https://doi.org/10.1175/JAS-D-16-0124.1.
Salvekar, P. S., L. George, and S. K. Mishra, 1986: Low level wind shear and baroclinic growth of monsoon depression scale waves. Meteor. Atmos. Phys., 35, 10–18, https://doi.org/10.1007/BF01029520.
Sanders, F., 1984: Quasi-geostrophic diagnosis of the monsoon depression of 5–8 July 1979. J. Atmos. Sci., 41, 538–552, https://doi.org/10.1175/1520-0469(1984)041<0538:QGDOTM>2.0.CO;2.
Sardeshmukh, P. D., and B. J. Hoskins, 1988: The generation of global rotational flow by steady idealized tropical divergence. J. Atmos. Sci., 45, 1228–1251, https://doi.org/10.1175/1520-0469(1988)045<1228:TGOGRF>2.0.CO;2.
Serra, Y. L., G. N. Kiladis, and M. F. Cronin, 2008: Horizontal and vertical structure of easterly waves in the Pacific ITCZ. J. Atmos. Sci., 65, 1266–1284, https://doi.org/10.1175/2007JAS2341.1.
Sikka, D. R., 1977: Some aspects of the life history, structure and movement of monsoon depressions. Pure Appl. Geophys. (PAGEOPH), 115, 1501–1529, https://doi.org/10.1007/BF00874421.
Smith, L. M., and S. N. Stechmann, 2017: Precipitating quasigeostrophic equations and potential vorticity inversion with phase changes. J. Atmos. Sci., 74, 3285–3303, https://doi.org/10.1175/JAS-D-17-0023.1.
Snyder, C., and R. S. Lindzen, 1991: Quasi-geostrophic wave-CISK in an unbounded baroclinic shear. J. Atmos. Sci., 48, 76–86, https://doi.org/10.1175/1520-0469(1991)048<0076:QGWCIA>2.0.CO;2.
Sobel, A. H., 2002: Water vapor as an active scalar in tropical atmospheric dynamics. Chaos, 12, 451–459, https://doi.org/10.1063/1.1480795.
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.
Stano, G., T. N. Krishnamurti, T. S. V. V. Kumar, and A. Chakraborty, 2002: Hydrometeor structure of a composite monsoon depression using the TRMM radar. Tellus, 54A, 370–381, https://doi.org/10.1034/j.1600-0870.2002.01330.x.
Subrahmanyam, D., M. K. Tandon, L. George, and S. K. Mishra, 1981: Role of barotropic mechanism in the development of a monsoon depression: A MONEX study. Pure Appl. Geophys. (PAGEOPH), 119, 901–912, https://doi.org/10.1007/BF00878958.
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.
Sukhatme, J., 2014: Low-frequency modes in an equatorial shallow-water model with moisture gradients. Quart. J. Roy. Meteor. Soc., 140, 1838–1846, https://doi.org/10.1002/qj.2264.
Vallis, G. K., 2017: Atmospheric and Oceanic Fluid Dynamics. Cambridge University Press, 964 pp.
Wallace, J. M., and P. V. Hobbs, 2006: Atmospheric Science: An Introductory Survey. 2nd ed. International Geophysics Series, Vol. 92. Academic Press, 504 pp.
Wang, B., and X. Xie, 1996: Low-frequency equatorial waves in vertically sheared zonal flow. Part I: Stable waves. J. Atmos. Sci., 53, 449–467, https://doi.org/10.1175/1520-0469(1996)053<0449:LFEWIV>2.0.CO;2.
Yanai, M., and T. Nitta, 1967: Computation of vertical motion and vorticity budget in a Caribbean easterly wave. J. Meteor. Soc. Japan, 45, 444–466, https://doi.org/10.2151/jmsj1965.45.6_444.
Yanai, M., S. Esbensen, and J. Chu, 1973: Determination of bulk properties of tropical cloud clusters from large-scale heat and moisture budgets. J. Atmos. Sci., 30, 611–627, https://doi.org/10.1175/1520-0469(1973)030<0611:DOBPOT>2.0.CO;2.
Yano, J.-I., and M. Bonazzola, 2009: Scale analysis for large-scale tropical atmospheric dynamics. J. Atmos. Sci., 66, 159–172, https://doi.org/10.1175/2008JAS2687.1.
Yoon, J.-H., and T.-C. Chen, 2005: Water vapor budget of the Indian monsoon depression. Tellus, 57A, 770–782, https://doi.org/10.3402/tellusa.v57i5.14737.
Yoon, J.-H., and W.-R. Huang, 2012: Indian monsoon depression: Climatology and variability. Modern Climatology, InTech, 45–72.