Influence of Latitude and Moisture Effects on the Barotropic Instability of an Idealized ITCZ

Eric Bembenek aDepartment of Atmospheric and Oceanics Sciences, McGill University, Montreal, Quebec, Canada

Search for other papers by Eric Bembenek in
Current site
Google Scholar
PubMed
Close
,
Timothy M. Merlis aDepartment of Atmospheric and Oceanics Sciences, McGill University, Montreal, Quebec, Canada

Search for other papers by Timothy M. Merlis in
Current site
Google Scholar
PubMed
Close
, and
David N. Straub aDepartment of Atmospheric and Oceanics Sciences, McGill University, Montreal, Quebec, Canada

Search for other papers by David N. Straub in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

A large fraction of tropical cyclones (TCs) are generated near the intertropical convergence zone (ITCZ), and barotropic instability of the related wind shear has been shown to be an important generation mechanism. The latitudinal position of the ITCZ shifts seasonally and may shift poleward in response to global warming. Aquaplanet GCM simulations have shown TC-generation frequency to vary with position of the ITCZ. These results, and that moisture plays an essential role in the dynamics, motivate the present study on the growth rates of barotropic instability in ITCZ-like zonal wind profiles. Base-state zonal wind profiles are generated by applying a prescribed forcing (representing zonally averaged latent heat release in the ITCZ) to a shallow-water model. Shifting the latitudinal position of the forcing alters these profiles, with a poleward shift leading to enhanced barotropic instability. Next, an examination of how latent release impacts the barotropic breakdown of these profiles is considered. To do this, moisture is explicitly represented using a tracer variable. Upon supersaturation, precipitation occurs and the related latent heat release is parameterized as a mass transfer out of the dynamically active layer. Whether moisture serves to enhance or reduce barotropic growth rates is found to depend on how saturation humidity is represented. In particular, taking it to be constant or a function of the layer thickness (related to temperature) leads to a reduction, whereas taking it to be a specified function of latitude leads to an enhancement. Simple arguments are given to support the idea that moisture effects should lead to a reduction in the moist shallow-water model and that a poleward shift of the ITCZ should lead to an enhancement of barotropic instability.

© 2021 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Eric Bembenek, eric.bembenek@mail.mcgill.ca

Abstract

A large fraction of tropical cyclones (TCs) are generated near the intertropical convergence zone (ITCZ), and barotropic instability of the related wind shear has been shown to be an important generation mechanism. The latitudinal position of the ITCZ shifts seasonally and may shift poleward in response to global warming. Aquaplanet GCM simulations have shown TC-generation frequency to vary with position of the ITCZ. These results, and that moisture plays an essential role in the dynamics, motivate the present study on the growth rates of barotropic instability in ITCZ-like zonal wind profiles. Base-state zonal wind profiles are generated by applying a prescribed forcing (representing zonally averaged latent heat release in the ITCZ) to a shallow-water model. Shifting the latitudinal position of the forcing alters these profiles, with a poleward shift leading to enhanced barotropic instability. Next, an examination of how latent release impacts the barotropic breakdown of these profiles is considered. To do this, moisture is explicitly represented using a tracer variable. Upon supersaturation, precipitation occurs and the related latent heat release is parameterized as a mass transfer out of the dynamically active layer. Whether moisture serves to enhance or reduce barotropic growth rates is found to depend on how saturation humidity is represented. In particular, taking it to be constant or a function of the layer thickness (related to temperature) leads to a reduction, whereas taking it to be a specified function of latitude leads to an enhancement. Simple arguments are given to support the idea that moisture effects should lead to a reduction in the moist shallow-water model and that a poleward shift of the ITCZ should lead to an enhancement of barotropic instability.

© 2021 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Eric Bembenek, eric.bembenek@mail.mcgill.ca

1. Introduction

The intertropical convergence zone (ITCZ) is a region of near-surface convergence of horizontal winds, typically located north of the equator in the Pacific ocean. It is characterized by ascending vertical velocities and, as a result, high climatological rainfall. It is also an important region for the generation of tropical cyclones (TCs), which can result from a process known as “ITCZ breakdown.” This breakdown can occur from a combination of barotropic instability of the ITCZ itself and in association with westward-propagating disturbances known as easterly waves. The breakdown can be viewed as a transition from linearly growing waves via barotropic instability to finite-amplitude vortices that can subsequently intensify into TCs. A classic example is described by Agee (1972), who suggested that Tropical Storm Anna formed as a result of ITCZ instability. Wang and Magnusdottir (2006) found that 65 TCs in the central and eastern Pacific were produced between 1999 and 2003 with an equal number of TCs formed due to easterly waves and barotropic instability. In another observational study, Berry and Reeder (2013) showed that, between 1979 and 2010, more than 50% of all TCs globally formed within 600 km of the ITCZ. There are, of course, important mechanisms other than barotropic instability in TC production (e.g., Montgomery et al. 2006; Emanuel 2003; Wing et al. 2016). Our focus here, however, is on barotropic instability.

The latitudinal location of ITCZ varies seasonally with the annual average being slightly north of the equator (Waliser and Gautier 1993; Berry and Reeder 2013; Marshall et al. 2014; Wodzicki and Rapp 2016). The impact of latitudinal location on TC frequency was discussed in Merlis et al. (2013) and Ballinger et al. (2015) in the context of aquaplanet global climate models at TC-permitting resolution. In these studies, sensible and latent heat fluxes associated with an anomalously warm zonal band of sea surface temperature (SST) led to ascending vertical motion and convergent surface winds analogous to those of Earth’s ITCZ. When the ITCZ was shifted poleward, the frequency of simulated TCs increased (see also Hsieh et al. 2020). In aquaplanet simulations with an imposed SST generating an ITCZ, Hsieh et al. (2020) proposed that increases in the large-scale vorticity leads to an increase in the probability of TC formation. However, the underlying mechanism that causes increases in large-scale vorticity as the ITCZ shifts poleward in these simulations is unclear. In this study, we present an analysis of how barotropic instability—the first phase of an important TC development process—depends on the latitudinal position of an imposed heat source representing the zonal-mean latent heat release in the ITCZ region.

To examine the influence of latitudinal position on the barotropic instability associated with ITCZ breakdown, we choose the simplest model that captures this phenomenon. In particular, we use a single-layer shallow-water model similar to that of Nieto Ferreira and Schubert (1997). They used a prescribed heat source to produce a zonal-wind profile similar to the low-level winds of the ITCZ. This profile was found to be barotropically unstable and simulations showed the development of vortices analogous to TCs. More recently, Asaadi et al. (2016) used a shallow-water model with prescribed heating to study effects of parameterized latent heat release associated with easterly waves and the ITCZ has also been studied in Matsuno–Gill-type models (e.g., Adam 2018). Results similar to those of Nieto Ferreira and Schubert (1997) have also been reproduced in more comprehensive models. For example, Yokota et al. (2012, 2015) used a nonhydrostatic mesoscale model with a prescribed warm belt of SST that produced an ITCZ-like wind profile. An energy budget analysis showed the initial breakdown of the ITCZ to be through low-level barotropic instability (similar to the observational results in Cao et al. 2012). This produced vortices that later intensified through latent heat release. Wang and Magnusdottir (2005) used a primitive equation model with prescribed heating that generated ITCZ-like winds at the beginning stages of a simulation. They studied the resulting ITCZ breakdown with and without a prescribed background flow and found that barotropic instability increased if the background flow reinforced the ITCZ wind shear produced by the heat source (e.g., Dritschel 1989). In short, the ITCZ may undergo breakdown via dry barotropic instability, and latent heat release can play a role in near-neutral conditions or in the subsequent finite-amplitude growth.

This study seeks to answer two questions:

  1. How does the latitudinal position of the ITCZ affect the barotropic instability of the associated horizontal flow?

  2. How does flow-dependent moisture and its associated latent heat release impact the ITCZ breakdown in our idealized framework?

To address the first question, we force our shallow-water model using a prescribed latitudinally dependent heat source whose center latitude we vary. This heat source is an idealized representation of the zonal-mean latent heat release associated with the ITCZ, similar to that used by Nieto Ferreira and Schubert (1997). We run a suite of spinup simulations with the heat source located at different latitudes in order to generate a set of balanced base-state wind profiles. These base-state profiles depend on latitude only. As such, zonal dependencies in the base state—which are important in the context of monsoon depressions (Diaz and Boos 2019)—are not represented. The barotropic instability of our base-state profiles is then assessed in a series of dry, unforced, initial value simulations, from which we assess growth rates and determine how these vary with the latitudinal position of the model ITCZ. We find that, for an equivalent amount of heating, a poleward shift of the ITCZ leads to an increased growth rate. A simple mechanism is presented to explain this result; essentially, heating at higher latitudes results in a stronger wind shear.

To address the second question, we include humidity as a tracer variable that precipitates and releases latent heat upon supersaturation (Bouchut et al. 2009). [This model is similar to Gill (1982).] Here, we use the barotropically unstable profiles from question 1 and assess how barotropic growth rates vary as a function of latent heat release strength. For most of our simulations, a spatially constant saturation humidity is assumed. However, we also examine the sensitivity of growth rates to different spatially varying prescriptions of saturation humidity, as it has been noted that gradients in saturation humidity can have a strong influence on tropical dynamics (e.g., Sobel et al. 2001; Sukhatme 2014; Suhas and Sukhatme 2020).

Using a similar model with constant saturation humidity, Lambaerts et al. (2011) found that precipitation led to a decrease in the growth rate of barotropically unstable jet. Here, we find that whether an increased latent heat release strength leads to stronger or weaker growth rates depends on how saturation humidity is specified. If it is either taken to be constant or to be a function of temperature (i.e., proportional to the interface height field), a decrease results. Instead, taking saturation humidity to have a fixed meridional profile mimicking the base-state temperature field leads to faster growth. We present mechanisms that explain these results.

Section 2 presents the model, the energy budget of our system, and the numerical model and parameter values used. Section 3 describes the results of shifting the latitudinal position of ITCZ and the subsequent dry barotropic instability generated by those base states. Section 4 describes the flow-dependent moisture effects on the linear instability phase of the barotropic instability and interprets the results using the simulations’ energy budget. Section 5 presents our conclusions.

2. Methods

a. Shallow-water model

We represent the lower troposphere with a one-layer shallow-water model on a β plane with heating, represented as a mass sink, in the thickness equation. The equations of motion are given by
ut+uu+f×u=gh,
DhDt+hu=F,
where u = (u, υ) is the horizontal velocity, g is the gravitational acceleration, and f = (f0 + βy)k is the Coriolis parameter. The thickness of the fluid column is h = H + η, where H is the mean-layer depth with η being deviations from the mean. That is, ⟨h⟩ = H and ⟨η⟩ = 0, where ⟨⋅⟩ denotes the domain average. The mean-layer depth H is chosen to represent a typical equivalent depth of the tropical atmosphere (e.g., Kiladis et al. 2009).1
To shift the latitudinal position of the ITCZ (which is produced following the method described below), we vary f0 and β according to
f0=2Ωsinϕc,
β=2ΩRecosϕc,
where Ω is Earth’s rotation rate and Re is Earth’s radius. Different base-state profiles are produced by varying the central latitude of heating ϕc.

The forcing term F in (2) represents heating. Positive F corresponds to a thinning of the layer. Our model can be understood as the lower layer of a two-layer model where the upper layer is quiescent and has a rigid lid. Heating is then represented as a mass transfer from the lower, cooler layer into the upper, warmer layer.

We begin by producing ITCZ-like base states for different ϕc in section 3a. For this, a set of spinup simulations taking the forcing term in (2) to be a prescribed mass sink are carried out. That is, F = Q, where Q can be thought of as a zonally averaged representation of convection associated with the ITCZ (Nieto Ferreira and Schubert 1997). This provides a convenient way to spin up our base states. The heat source is defined as the product of a latitudinally dependent top-hat function Q˜(y) and a ramping function r(t). Specifically, Q(y,t)=Q˜(y)r(t), where
Q˜(y)=Q02[tanh(y+252km25km)tanh(y252km25km)]
and
r(t)=12[tanh(t4days1day)tanh(t9days1day)].
The latitudinally and time-varying components of Q are shown in the top panels of Fig. 1. The ramping function r(t) slowly increases in magnitude over several days and then decays to zero smoothly. This is done to limit the amount of gravity waves that are produced during this process. Following Nieto Ferreira and Schubert (1997), Q0 = 55.2 m day−1 and the width of the heating is 504 km. This choice of heat source is similar in effect to Merlis et al. (2013) and Ballinger et al. (2015), where a typical half-width of the GCM-simulated zonal-mean precipitation is ≈500 km. As in Nieto Ferreira and Schubert (1997), we use Q˜(y) to generate ITCZ-like horizontal winds: the mass sink is balanced by mass convergence of the layer’s horizontal winds, as is found in the near-surface flow of the ITCZ region. We note that theories of the tropical general circulation of the atmosphere typically do not take the latent heating as external to the flow (e.g., Held and Hou 1980; Neelin and Held 1987; Emanuel et al. 1994; Merlis 2015) because of its intimate link to the vertical velocity (Sobel and Bretherton 2000). Here, our aim is singly to spin up idealized ITCZs for a subsequent instability analysis done in section 3b and this is an expedient approach.
Fig. 1.
Fig. 1.

(top left) Heating function Q˜(y) and (top right) ramping function r(t). (bottom left) Zonal-mean vorticity ζ¯ and (bottom right) free surface η¯ of the spun-up base state. Meridional position y is Re(ϕϕc), where Re is Earth’s radius. Vertical dotted lines indicate position of the equator.

Citation: Journal of the Atmospheric Sciences 78, 9; 10.1175/JAS-D-20-0346.1

In section 4, we consider moisture effects by taking F to be proportional to precipitation P. To do this, we follow Lambaerts et al. (2011) and include a tracer representing a layer-averaged specific humidity q (g kg−1):
DqDt+qu=P.
Precipitation occurs when the specific humidity supersaturates and is given by
P=qqsτH(qqs),
where qs represents the saturation specific humidity, τ represents the time scale of precipitation, and H() is the Heaviside function. Saturation specific humidity qs takes several forms in our simulations. In particular, we consider the case of a spatially and temporally constant qs (as in Lambaerts et al. 2011), a fixed meridional profile qs = qs(y) (similar to Suhas et al. 2017), and qs = qs(η) following the Clausius–Clapeyron relation (as in Vallis and Penn 2020). The latent heat release associated with precipitation implies a mass sink in (2), F = αP, where α corresponds to the strength of latent heating. Precipitation thus couples the thickness and q equations [(2) and (5), respectively] in these simulations. Our focus is on how variations in α impact growth rates for the ITCZ profiles generated in section 3a. A similar study, but for forced-dissipative simulations in a two-layer quasigeostrophic β plane was carried out by Lapeyre and Held (2004). Their main result was to show a transition from jet-dominated to vortex-dominated turbulence as the strength of latent heat release increased. In another two-layer shallow-water version of the problem, Bembenek et al. (2020) varied qs and found a similar transition in part of parameter space.

In summary, in section 3a, we set F = Q in spinup simulations to generate ITCZ-like wind profiles at different latitudes. In section 3b, we set F = 0 and run initial value problem simulations of these ITCZ profiles to calculate the dry growth rate. Section 4 then takes F = αP to investigate how precipitation [determined by Eqs. (5) and (6)] impacts the growth. This is done by varying α, which relates precipitation to a mass sink in (2).

b. Energy budget

One can define an moist static energy for our model as m = hαq or m = ηαq. This is both materially conserved (see also Gill 1982; Bouchut et al. 2009) and conserved in an area-integrated sense. In words, conservation of m simply states that the mass removed from our active shallow-water layer is proportional to the moisture removed by precipitation. Note that m is independently conserved: it does not represent a reservoir with which potential and kinetic energy can be exchanged. Because of this, and because we will compare moist and dry simulations, it is more convenient to use the dry energy norm. Precipitation then appears as a forcing term for potential energy, analogous to that for any mass source or sink driving a shallow-water system.

We will also decompose energy into zonal and eddy (or perturbation) quantities, defined by A(x,y,t)=A¯(y,t)+A(x,y,t). Although energy for the shallow-water equations is cubic, it is more convenient to use a quadratic approximation. That is, we approximate energy as
E=huu2+gη22Huu2+gη22,
where ⟨⋅⟩ denotes a domain average. Comparing time series of this quadratic approximation to energy with its conserved (cubic) analog shows errors of less than 0.5% (not shown), suggesting that this approximation is not leading to significant errors in our analysis. The zonal and eddy energy are similarly approximated as
Ez(t)=Hu¯u¯2+gη¯22,
Ee(t)=Huu2+gη22.
Taking time derivatives of (8) and (9) and using the zonal and eddy momentum and η equations gives the energy budget for the zonal and eddy components:
dEzdt=Cz+Pz+residual
and
dEedt=Cz+Pe+residual.
Since we have approximated the kinetic energy in Eq. (7), these energy budgets have a small residual associated with them. For completeness, a derivation of the energy budget terms can be found in appendix A. The precipitation forcing terms are
Pz=gαη¯P¯,
Pe=gαηP¯.
These act on the zonal and eddy budgets, respectively. Whether precipitation acts to energize or damp the flow is determined by the sign of the correlation between η and P. The conversion between zonal and eddy components of energy is given by
Cz=Huυ¯u¯y,
which is common to both budgets. If Cz>0, the conversion is from zonal kinetic energy to eddy kinetic energy, driving barotropic instability.

Note that this budget is similar to that of Lambaerts et al. (2011), but we make a further decomposition into zonal and eddy components. Additionally, in the model used in Lambaerts et al. (2011), the free surface η represents the bottom of the atmosphere, so their precipitation forcing terms should be multiplied by −1 relative to ours.

c. Numerical model and parameter values

To solve the system in (1)(5), we use a finite-difference model on an Arakawa C grid with leapfrog time stepping. A bi-Laplacian dissipation (with a coefficient of Ah ≈ 9 × 1011 m4 s−1) in the momentum and humidity equations removes energy at the smallest scales. The domain is a periodic channel with walls at the north and south boundaries, where free-slip boundary conditions are applied.

The physical parameters of the problem are stated in Table 1. The domain size is 13 440 km in x (≈120° in longitude) and 3360 km in y (≈30° in latitude) with 512 × 128 grid points. This implies that Δx = Δy ≈ 26 km. The gravitational wave speed is cd=gH47ms1. As a result, the time step is limited to Δt=0.25Δx/cd140s.

Table 1.

Parameter values used in dry and constant qs simulations.

Table 1.

There are several parameters associated with moisture: τ, qs, and α. We fix τ = 3600 s ≈ 25Δt to be the fastest time scale in the problem. For the case of constant saturation humidity, we take qs = 1 g kg−1. The parameter α is related to the moist gravity wave speed by cm=g(Hαqs) for constant qs (Bouchut et al. 2009). Because of this, α cannot be chosen to be arbitrarily large as this would lead to exponentially growing moist gravity waves. We thus require α < αmax, where αmaxH/qs = 222 m kg g−1. Specifically, we consider values of α corresponding to 25%, 50%, 75%, and 90% of αmax. Note that the dynamics are controlled by the product αqs rather than by α or qs per se. For example, simulations using qs = 10 g kg−1 and αmax = 22.2 m kg g−1 gave similar results to simulations using qs = 1 g kg−1 and αmax = 222 m kg g−1 (not shown). This suggests that increasing saturation humidity, as in climate change simulations, is equivalent to increasing the strength of latent heat release in this model. Therefore, we vary α in section 4. We will also explore the sensitivity of the system to different functional forms of qs In particular, we set qs to be a fixed meridional profile and qs to be an exponential function of η, following the Clausius–Clapeyron equation in section 4b.

3. Dependence of dry dynamics on central forcing latitude ϕc

a. Base-state profiles

Base-state profiles are spun up by forcing the thickness Eq. (2) using F=Q=Q˜(y)r(t) [see (3) and (4)] for ϕc = 6°, 8°, 10°, 12°, and 14°. Variations in ϕc imply a variation in f0 and β. These spinup simulations are run for 20 days. As expected, given that r(t) → 0 near the end of the simulations, this produces a set of base-state profiles for which υ¯(y)0. Our reference base states are then defined by u¯(y) and η¯(y) at the end of these simulations.

Profiles of zonal-mean vorticity ζ¯=u¯y and η¯ for different choices of ϕc are shown in the bottom panels of Fig. 1. Vorticity increases with ϕc and has profiles similar to those found in Nieto Ferreira and Schubert (1997) and qualitatively match the lower-level zonal winds in three-dimensional models (Wang and Magnusdottir 2005; Yokota et al. 2012, 2015). In particular, the maximum zonal-wind shear of our ϕc = 10° case is ≈10 m s−1 and this is similar to those studies. Profiles of η¯ show a qualitative change with ϕc. That is, a dip in η¯ centered on y = 0 is all but absent at ϕc = 6° and becomes increasingly pronounced at higher latitudes. We next examine how these differences impact the barotropic instability properties of these base states.

b. Instability

To determine the changes in instability, we run simulations initialized using the above meridional profiles for ζ¯ and η¯ in Fig. 1 as base states to which we add white noise. Figure 2 shows snapshot of the vorticity at t = 15 days for three cases of ϕc. The instability is not apparent in the case where ϕc = 6° and is only beginning to emerge at this time in the case where ϕc = 10°. For ϕc = 14°, on the other hand, the flow is much more developed, with nonlinear effects such as vortex roll up already apparent. For all ϕc tested, the instability shows a wavenumber 8 pattern corresponding to a length scale of around 1700 km, which is close to the equatorial Rossby radius, cd/β1400km.

Fig. 2.
Fig. 2.

Vorticity normalized by 10−5 s−1 for ϕc = (top) 6°, (middle) 10°, and (bottom) 14°. All snapshots are taken at t = 15 days.

Citation: Journal of the Atmospheric Sciences 78, 9; 10.1175/JAS-D-20-0346.1

That the flow is more developed for larger ϕc indicates that the growth rate of the instability σ increases with increasing ϕc. This is confirmed in Fig. 3, which shows time series of eddy energy, (9), for each ϕc. Evident in all simulations is a period of exponential growth, and estimation of the growth rates (stated in parentheses in the legend of Fig. 3) show the expected increase with ϕc. The largest growth rate found is approximately 0.92 day−1, which is similar to Yokota et al. (2015), who found a rate of 0.896 day−1. Most of our simulations have a σ comparable to the 2-day e-folding time found in Nieto Ferreira and Schubert (1997).

Fig. 3.
Fig. 3.

Time series of eddy energy Ee as a function of latitude ϕc. Least squares estimate of the growth rate σ indicated in parentheses in the legend (with units of days−1). To quantify σ, we calculate a least squares fit to the slope of these curves for the periods where ln(Ee) grows linearly, indicated by the vertical ticks in each time series.

Citation: Journal of the Atmospheric Sciences 78, 9; 10.1175/JAS-D-20-0346.1

That our more poleward base-state profiles showed faster barotropic growth rates is a simple consequence of these profiles also showing stronger shear. Here, we use potential vorticity to rationalize why a more poleward heat source should be expected to produce stronger shear in the base-state wind profile. Shallow-water potential vorticity, PV = (f + ζ)/h, evolves according to
DDtln(PV)=QhQH,
where ηH has been assumed. Ignoring the advection term in the material derivative, we integrate this expression in time, and since Q=Q˜(y)r(t), it follows that perturbations relative to the rest state qualitatively follow
PV(y,t)=PV(y,0)exp(0tQHdt).
Subtracting PV(y,0) = f/H from both sides and approximating ff0, we can estimate that the base-state ζ¯ and η¯ profiles obey
ζ¯f0η¯Hf0[exp(0tQdt)1].
In words, this implies that for an equivalent forcing Q, the perturbation PV will be larger when the center latitude of the forcing profile is at higher latitudes (so that |f0| is larger). For forcing profiles with horizontal length scales small compared to the deformation radius (as is the case for the latent heat release in a tropical convergence zone), the partition between relative and stretching vorticity heavily favors relative vorticity. For example, Fig. 1 shows that the magnitude of ζ¯ is approximately 60 times that of f0η¯/H. Given this, the above analysis implies that, all else being equal, shifting the heat source poleward implies a more strongly sheared base state. By extension, this yields larger barotropic growth rates.

The growth rate is determined by the shear. This, in turn, depends on the forcing strength, forcing duration, and on ϕc. Because β varies weakly with ϕc in low latitudes, the ϕc dependence can be thought of as a dependence on f0. Thus, for example, a spinup at ϕc = 6° with r(t) [see Eq. (4)] redefined so that forcing is applied for nine instead of five days produces a vorticity profile similar to our ϕc = 14° (not shown). By contrast, the stretching vorticity profiles remain qualitatively different. Since the dynamics is dominated by relative vorticity, differences in the stretching profiles do not appreciably impact the growth rates. In other words, our result that a poleward shift in the ITCZ implies faster growth rates assumes the position, but not the strength of the heating to be varying with latitude. It is a combination of the two that determines the base-state shear, and by extension, the growth rate.

Vertical wind shear can break up the vertical coherence of developing vortices, potentially inhibiting growth. To explore the robustness of our barotropic results to this process, we ran similar simulations in a two-layer shallow-water model with a prescribed heat source for ϕc ∈ [6°, 14°] configured to produce comparable barotropic shear in our one-layer model. Results showed that the ITCZ breakdown process remained dominated by barotropic instability (not shown). Similar results were found by Wang and Magnusdottir (2005), who studied three-dimensional ITCZ breakdown in a primitive equation model with prescribed heating. Specifically, they found that ITCZ breakdown was dominated by barotropic instability.

This barotropic framework allows us to propose a possible interpretation for the increased frequency of TCs in Merlis et al. (2013) and Ballinger et al. (2015) as a result of an increase in PV forcing as the heating moves to higher latitudes. In those GCM simulations, the latent heating in the ITCZ is internally determined, unlike the prescribed heating here. However, the tropical-mean precipitation in GCMs is constrained by energy budgets (O’Gorman et al. 2012) that are insensitive to ITCZ latitude. In this context, our use of a prescribed heating is reasonable. We can similarly interpret some of the results in Chavas and Reed (2019), who examined aquaplanet simulations with uniform thermal forcing. They showed that increased rotation rates led to an increased frequency of TCs. This is consistent with the point we emphasize above whereby the diabatic PV source term is proportional to the rotation rate. In their case, the forcing was due to storms rather than to an imposed meridional profile. Nevertheless, a similar logic applies: for the same magnitude of latent heat release, faster rotation implies stronger PV forcing. Finally, in Earthlike aquaplanet simulations, Hsieh et al. (2020) suggest that the increase in TC formation results from an increased probability of TC seeds forming when f0 gets larger and this is consistent with the enhanced diabatic PV generation mechanism that we described.

4. Moisture effects on barotropic instability

a. Sensitivity to latent heat release strength α

Here, we consider how flow-dependent moisture affects the barotropic instability of our model ITCZ. The thickness equation, (2), is forced by F = αP where P, defined by Eq. (6), is determined by the evolution of the humidity equation, (5). We first treat the case of constant qs and later explore how the dynamics are impacted if qs is a fixed meridional profile and if qs has a temperature dependency. Latent heat release strength is varied in a manner similar to Lapeyre and Held (2004). In our system, latent heat release strength is controlled by α. As described in section 2c, α < αmaxH/qs is required in order to ensure that moist gravity wave speeds remain real. We consider simulations with α/αmax = 0.25, 0.5, 0.75, and 0.9. We will consider these values for each ϕc profile shown in Fig. 1. In other words, for each of the ITCZ base states from section 3a, we repeat the growth rate analysis of section 3b, this time allowing for precipitation effects.

Snapshots of vorticity are similar to those in the dry simulations (not shown), suggesting that precipitation has only a modest effect on the dynamics. To quantify this effect, we calculate growth rates, σ, and compare them with results from the previous section. Results are shown in Fig. 4. Overall, the reduction compared to dry simulations is modest (10%). In other words, precipitation is seen to have a weak de-energizing effect on the instability. For fixed ϕc, the growth rate shows a larger fractional decrease for larger α (i.e., for stronger precipitation forcing). This effect is more pronounced at larger values of ϕc but remains modest overall. The wavenumber of the most unstable mode remains the same as the dry case (a wavenumber-8 pattern, ≈1700 km) and is insensitive to α.

Fig. 4.
Fig. 4.

Percentage change in moist growth rate σ with respect to the dry (α = 0) simulation for each ϕc. See Fig. 3 for dry growth rates.

Citation: Journal of the Atmospheric Sciences 78, 9; 10.1175/JAS-D-20-0346.1

Although the effect is modest, that precipitation has a damping effect on barotropic instability merits some discussion. Results in the literature have been mixed. In the context of easterly waves, Thorncroft and Hoskins (1994) found a modest increase in the growth rate in a moist simulation using a three-dimensional primitive equation model. Kwon (1989), in contrast, found negligible differences in growth rates between a dry and moist case of easterly waves using a quasigeostrophic model. Differences in the base-state flow and parameterization of latent heat release are both likely important in comparing these layered or three-dimensional model results to those in our one-layer model. Finally, using a model similar to our own, Lambaerts et al. (2011) studied the barotropic instability of a Bickley jet in simulations taking qs to be constant. Unlike us, they found that moisture led to an increase in the growth rate. One difference in our analysis relative to theirs is the definition of zonal and eddy components. Specifically, in their analysis, time-varying zonal averages were considered part of the eddy field. Repeating our analysis using the same base states and methodology as theirs, however, we continue to find that precipitation leads to a decrease in the growth rate. It is thus unclear why Lambaerts et al. (2011) find a growth rate increase in their simulations. In section 4c, we present a physical mechanism that explains how precipitation leads to an increase or a decrease in the growth rate, depending on how saturation humidity is parameterized.

We next analyze the energy budget associated with the zonal and eddy components [(10) and (11), respectively] of the flow. Figure 5 shows terms associated with precipitation, Pz and Pe [(12) and (13), respectively] as well as the zonal-to-eddy conversion term Cz. During the initial instability, Pe<0 and Pz>0. That is, precipitation generates zonal potential energy and reduces eddy potential energy. After the linear phase of the instability, an opposite-signed effect is evident, although this is weaker by comparison. The amplitudes of these tendencies increase markedly with α. This contrasts the zonal-to-eddy conversion term Cz, (14), which is essentially independent of α. As expected for barotropic instability, this conversion term is positive. Overall, then, precipitation adds zonal PE and removes eddy PE. The zonal PE added, however, is small compared to that present initially (not shown). Related to this, changes in both the base-state shear and the eddy conversion term do not change appreciably in response to this forcing. That Pe is a net sink, however, implies a reduction in barotropic growth rates.

Fig. 5.
Fig. 5.

Energy budget terms for simulations with flow-dependent precipitation (α > 0) and the corresponding dry case (α = 0). (top) Precipitation forcing on eddies Pe, (13); (middle) precipitation forcing on the zonal component Pz, (12); and (bottom) conversion from zonal to eddy kinetic energy Cz, (14), for the ϕc = 10° case.

Citation: Journal of the Atmospheric Sciences 78, 9; 10.1175/JAS-D-20-0346.1

b. Nonconstant saturation humidity

So far, our experiments have considered qs to be spatially and temporally constant. More realistically, saturation humidity follows the Clausius–Clapeyron relation which suggests an exponential dependence on temperature (proportional to −η in our model). In the tropics, this implies that saturation humidity peaks near the equator and decays meridionally. We model this in two ways. First, following Suhas and Sukhatme (2020), we set saturation humidity to a fixed meridional profile, qs = qs(y). Specifically, we take qs to depend on the base-state η¯ at t = 0. Second, following Vallis and Penn (2020), we set saturation humidity to evolve with the flow, qs = qs(η). Intuitively, during the initial instability, one would expect similar results between qs = qs(y) and qs = qs(η) since ηη¯ which does not vary significantly. However, we will show that the formulation of qs has a significant impact on how precipitation affects the growth rate.

We choose two formulations of saturation humidity,
qs=qs(y)=q0exp[γη¯(y,t=0)H],
qs=qs(η)=q0exp(γηH),
where γ = Lυ/RυT0 ~ 20 is representative of Earth’s tropics (Lυ = 2.4 × 106 J kg−1, Rυ = 462 J kg−1 K−1, and T0 = 300 K) (Vallis and Penn 2020) and q0 = 1 g kg−1. We tested the sensitivity of the growth to variation in γ with ϕc = 10° and α = 100 m kg g−1. Profiles of qs(y) are shown in the left panel of Fig. 6 for different values of γ. Saturation humidity peaks near the center of the domain and decays away from y = 0, similar to Suhas and Sukhatme (2020). As in the constant qs case, simulations are initialized with humidity near saturation throughout the domain so as not to bias where barotropic instability triggers precipitation. Larger γ increases humidity and leads to a stronger gradient in qs. The right panel of Fig. 6 shows the zonally averaged qs, denoted qs¯, for the simulation with qs = qs(η) and γ = 20 at different times. The solid curves denotes qs¯ during the initial instability (t ≤ 16 days) and shows a profile that is remarkably fixed in time. At later times (shown in dotted and dashed lines), qs¯ has more pronounced variations as is to be expected when eddies have a stronger influence on the flow. Given that qs¯ is essentially fixed in time over the course of the instability, one would expect the impact of precipitation in these simulations to be similar that in simulations using qs = qs(y). As seen from Fig. 7, however, this is not the case. Specifically, assuming qs = qs(y) leads to an increase in σ, whereas assuming qs = qs(η) leads to a decrease.
Fig. 6.
Fig. 6.

(left) Profiles of saturation humidity, qs = qs(y), as a function of γ. (right) Zonally averaged qs for simulations with qs = qs(η) and γ = 20 at t = 0, 10, 15, 20, and 30 days. The solid lines indicate qs during linear instability (for t ≤ 16 days) and the dotted lines indicates qs during nonlinear phase of simulation.

Citation: Journal of the Atmospheric Sciences 78, 9; 10.1175/JAS-D-20-0346.1

Fig. 7.
Fig. 7.

Percentage change in growth rate σ as a function of γ for simulations with qs = qs(y) (black) and qs = qs(η) (gray). The nondimensional parameter γ controls the magnitude of the derivative of qs, see Eqs. (17) and (18). The simulation with γ = 0 corresponds to the constant qs case with α = 100 m kg g−1 and ϕc = 10°.

Citation: Journal of the Atmospheric Sciences 78, 9; 10.1175/JAS-D-20-0346.1

To understand why this occurs, recall that the sign of Pe determines whether precipitation increases or decreases the growth rate. Recall also that whether Pe acts as energy source or sink depends on the sign of correlations between P and η′. For constant qs, precipitation occurs where η′ > 0 (see top panel of Fig. 8) and this leads to a reduction in eddy energy (see top panel of Fig. 5). When qs = qs(y), precipitation shifts to the northeastern and southwestern flanks of negative η′ (see middle panel of Fig. 8), thus implying an energy source (and an increased growth rate). Finally, when qs = qs(η), precipitation reverts to a pattern similar to that seen for constant qs (see bottom panel of Fig. 8), so that once again, a decrease the growth rate results. In the next section, we propose mechanisms by which the location of precipitation is determined for each of these cases.

Fig. 8.
Fig. 8.

Instantaneous eddy free surface η′ for (top) constant qs, (middle) qs = qs(y), and (bottom) qs = qs(η), with black contours indicating precipitation and dashed (solid) white contours indicate convergence (divergence). All simulations have α = 100 m kg g−1 and ϕc = 10° and fields are shown on day 15. The cases with qs = qs(y) and qs(η) have γ = 20.

Citation: Journal of the Atmospheric Sciences 78, 9; 10.1175/JAS-D-20-0346.1

c. Mechanisms leading to precipitation

To understand how Pe affects growth, it suffices to understand where precipitation occurs with respect to η′. To this end, we consider the linearized eddy moisture equation for each of the three qs formulations considered. Since simulations are initialized close to saturation, we substitute q = qs + q′ into the humidity Eq. (5) for each form of qs and neglect eddy correlation terms. This gives
qs= constqt=u¯qxqsuP,
qs=qs(y)qt=u¯qxqsuυdqsdyP,
qs=qs(η)qt=u¯qx+(qs+dqsdηH)u+(1+αdqsdη)P.
Equation (21) can be further simplified using qs = q0 exp(−γη/H):
qt=u¯qxqs(1+γ)u(1+αγqsH)P.
A derivation of (21) and (22) can be found in appendix B.

For constant qs, we establish a negative correlation between horizontal divergence and precipitation, i.e., precipitation in our model must be preceded by a convergent horizontal flow (at least during the initial instability). Since eddy humidity evolves according to (19), it follows that a shallow-water fluid column can only become supersaturated following a period in which it is in a convergent part of the flow. That is, (19) implies that −∇ ⋅ u′ > 0 is required for q′ to increase. It then follows that precipitation can only occur in regions of convergence. A similar argument follows for qs = qs(η). Even though Eqs. (21) and (22) are more complex, convergence remains the only mechanism that increases q′ in a parcel. As a result, precipitation is driven by convergent modes in simulations where qs is constant or qs = qs(η). This mechanism is apparent in the top and bottom panels of Fig. 8 where η′ > 0 coincides with precipitation (black contours) and convergence (white dashed contours).

For qs = qs(y), on the other hand, there is an additional mechanism by which humidity increases. In particular, the meridional advection of humidity, −υdqs/dy, in Eq. (20) increases humidity and leads to precipitation. This mechanism is apparent in the center panel of Fig. 8 which shows that precipitation shifts to the flanks of η′ < 0. That the location of precipitation is shifted indicates that rotational modes instead of convergent are driving precipitation as has been suggested for the Madden–Julian oscillation (Pritchard and Bretherton 2014) and moist equatorial Rossby waves (Wheeler et al. 2000).

Next, we establish a relationship between convergence and η′. Since convergence drives precipitation for constant qs and qs = qs(η) this mechanism allows us to conclude that Pe<0 is expected in our model when these formulations of qs are used. In addition to being a prerequisite for precipitation, convergence also implies a thickening of fluid columns. Convergence in the eddy velocity field then implies an increase in eddy thickness η′. For example, in the linear growth phase of barotropic instability η′ evolves according to
ηt=u¯ηxHuαP,
where nonlinear advection is neglected. Assuming a Fourier mode solution for η=η^eσteikx, gives ηt=ση. A positive growth rate (σ) implies that ηt is positively proportional to η′. But since ηt>0 implies convergence, it follows trivially that regions of convergence correspond to regions of positive η′. Put together, these arguments imply that regions of precipitation occur where η′ > 0, implying Pe to be a damping term in the energy equation, (11).

For the constant qs and qs = qs(η) cases, precipitation is associated with convergence. In the growth phase of the instability, ηt (and thus η′ itself) are also associated with convergence. As such, the mass sink due to precipitation occurs in regions of elevated η′, and a sink of potential energy results. Note that this reasoning also applies to the simulations of Lambaerts et al. (2011), despite having their free surface at the bottom of the fluid column. More generally, convergence implies a thickening fluid column and precipitation, the confluence of which leads to a destruction of potential energy and hence a reduction in the growth rate. For the qs(y) case, advection of the base state qs(y) profile provides an additional term in the moisture equation, cf. (20). Since qs decreases from the center of the ITCZ, meridional flow away from the center of the jet serves to increase η′. As such, precipitation tends to occur in regions of meridional flow away from the jet center. That is, precipitation straddles regions of positive and negative η′, but with a slight bias toward negative regions, so that a net sink results.

5. Conclusions

In this study, we aimed to answer two questions:

  1. How does the latitudinal position of the ITCZ affect the barotropic instability of the associated horizontal flow?

  2. How does flow-dependent moisture and the associated eddy latent heat release impact the ITCZ breakdown in our idealized framework?

To answer these questions, we used a shallow-water model with a prescribed heat source similar to that of Nieto Ferreira and Schubert (1997) and a parameterization for latent heat release that makes use of a humidity tracer as in Bouchut et al. (2009) and Lambaerts et al. (2011).

The prescribed heat source Q represents a zonal-mean idealization of the convection associated with the ITCZ and was used to generate the ITCZ-like winds presented in section 3a. Shifting the center latitude of the heating poleward (while keeping the temporal and spatial structure of Q fixed) led to increased shear at higher latitudes. It may be, of course, that in a more realistic setting, the heat source itself would also vary with its latitudinal position. Our choice of fixed Q was motivated by Merlis et al. (2013) and Ballinger et al. (2015), whose GCM simulations showed similar shifts in ITCZ position without substantial changes to the overall heating. We also performed simulations (not shown) using a stronger heat source at lower latitudes and found that the effect of increasing Q is similar to that of shifting the center latitude poleward. A caveat is that a stronger Q at lower latitude could be made to produce a shear profile similar to that for a weaker Q at higher latitude with the η¯ profiles being different. Because the dynamics are dominated by relative, rather than stretching vorticity, however, the dry dynamics is essentially unaffected, though the modified η¯ profiles might affect qs.

That heating at higher latitudes should result in enhanced shear was explained by a simple integration of the shallow-water PV equation. This showed that an equivalent heat source produced a stronger PV response at higher latitudes and a stronger PV gradient implied a stronger shear. We also ran simulations in a two-layer shallow-water model to assess the relative importance of barotropic and baroclinic instability. For the range of latitudes considered, ϕc ∈ [6°, 14°], barotropic instability remained dominant. This is consistent with Wang and Magnusdottir (2005), who showed that their prescribed heating led to barotropic instability in a three-dimensional primitive equation model.

Then, for each of our base-state profiles, we ran a series of simulations to assess the effect of precipitation on barotropic growth rate σ. This was done by varying a parameter related to the strength of latent heat release. First, we considered qs to be constant, and found that precipitation led to a decrease in σ. This results from a correlation between precipitation and column thickness, both of which increase in response to horizontal convergence. Because of this, precipitation is systematically found in regions of larger η′ and an energy sink results. Other prescriptions of qs were also considered. For example, taking qs = qs(y) (similar to Suhas and Sukhatme 2020), we found precipitation to lead to an increase in the growth rate. In these simulations, meridional advection shifts precipitation to the flank of warm eddies (where η′ < 0), leading to an increase in available potential energy. Finally, if saturation humidity depends on η (or temperature, following the Clausius–Clapeyron relation), precipitation leads to a reduction in the growth rate. As in the constant qs case, precipitation is driven by convergence which is collocated with cold eddies. In summary, if qs is constant or qs = qs(η), precipitation is driven by convergent modes and leads to a growth rate decrease. If qs = qs(y), precipitation can be driven by rotational modes and can lead to a growth rate increase.

To summarize, we found that dry barotropic instability increases for ITCZs generated by a fixed amount of heating when the heat source was placed further from the equator. This was due to the increase in potential vorticity generation with latitude that generated stronger horizontal wind shear. Latent heating during the growth of the instability modestly reduces growth rate if qs is constant or a function of η. On the other hand, if qs is a fixed function of y (related to the initial temperature profile), then precipitation led to an increase in the growth rate. This suggests that the choice of saturation humidity prescription can have a significant impact on how precipitation affects the flow. Our results used a simple one-layer shallow-water model and focused on initial value problems and the linear growth phase of instability. Bridging the gap between this sort of simple analysis and one allowing both for more complex dynamics and that considers the more climate-relevant case of statistical equilibria with the ITCZ migrating, undergoing breakdowns, and subsequent reformation present interesting avenues for future research.

Acknowledgments

This research was supported by the Natural Sciences and Engineering Research Council of Canada. We appreciate three perceptive reviewers, who prompted the systematic investigation of the form of the saturation specific humidity.

APPENDIX A

Energy Budget Derivation

The energy in the shallow-water system is given by
E=huu2+gh22,
which is conserved absent heating and small-scale dissipation. The first term is the kinetic energy and the second term is the total potential energy. Using h = H + η, the total potential energy can be written as ⟨gh2/2⟩ = ⟨gH2/2⟩ + ⟨2/2⟩, where terms of the form ⟨gηH⟩ = 0 since ⟨η⟩ = 0. Since H is constant in time, the dynamically interesting component of the total potential energy is the available potential energy, ⟨2/2⟩.
The cubic dependence of the kinetic energy on the flow variables makes a zonal-eddy decomposition ambiguous. To rectify this, it is common to approximate the kinetic energy as
huu2Huu2,
since h = H + η with ηH. A time series of the true total energy and approximate total energy in the system shows a less than 1% variation with respect to the initial energy (not shown). This suggests that the approximation does not lead to substantial errors.
Precipitation’s effect on the energy budget results from it impact on the η equation. In particular, under the true total energy definition, precipitation appears as
dEdt=uu2αPgαηP.
The first term results from the h dependence in the kinetic energy. Under the approximation hH, the first term vanishes, so that the approximated total energy budget reads
dEdt=gαηP.
To derive Eqs. (10) and (11), we decompose the η equation into its zonal and eddy components as
η¯t=(uh¯)αP¯,
ηt=(uhuh¯)αP.
To construct the available potential energy budgets for zonal (APE¯) and eddy (APE′) components, we multiply Eq. (A1) by gη¯ and Eq. (A2) by ′ and average the resulting equations over the domain:
dAPE¯dt=gη¯(uh¯)gαη¯P¯,
dAPEdt=gη(uhuh¯)gαηP.
The first terms in both budgets cancel (approximately) when combined with the kinetic energy budget. This shows the derivation of the precipitation effect on the energy budget, Eqs. (12) and (13).
The conversion term from zonal to eddy kinetic energy appears with opposite signs in the zonal and eddy momentum equations. Applying a zonal averaging of the momentum equation, one can show
u¯t=yuυ¯,
which can be multiplied by Hu¯ and integrated by parts to give the zonal kinetic energy (KE¯) budget:
dKE¯dt=Huυ¯u¯y.

Finally, combining the KE¯ and APE¯ gives the zonal energy budget in Eq. (8). With similar steps, one can derive the eddy energy budget in Eq. (11).

APPENDIX B

Eddy Moisture Budget Derivation

Here, we derive the moisture eddy budget for a saturation humidity that depends on η, i.e., Eq. (21). We linearize the humidity equation, Eq. (5), about the state q = qs(η) + q′(x, y, t). Using the chain rule, we have
qt+u¯qx+dqsdη[ηt+(uη)]=qsuP.
Using the η equation and approximating hH, we replace the term in the square brackets and rearrange to find Eq. (21):
qt+u¯qx=(qs+dqsdηH)u+(1+αdqsdη)P.
Finally, qs(η)=γ/Hqs because qs = q0exp(−γη/H), and Eq. (22) results:
qt+u¯qx=qs(1+γ)u(1+αγqsH)P.

REFERENCES

  • Adam, O., 2018: Zonally varying ITCZs in a Matsuno-Gill-type model with an idealized Bjerknes feedback. J. Adv. Model. Earth Syst., 10, 13041318, https://doi.org/10.1029/2017MS001183.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Agee, E. M., 1972: Note on ITCZ wave disturbances and formation of Tropical Storm Anna. Mon. Wea. Rev., 100, 733737, https://doi.org/10.1175/1520-0493(1972)100<0733:NOIWDA>2.3.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Asaadi, A., G. Brunet, and M. K. Yau, 2016: On the dynamics of the formation of the Kelvin cat’s-eye in tropical cyclogenesis. Part II: Numerical simulation. J. Atmos. Sci., 73, 23392359, https://doi.org/10.1175/JAS-D-15-0237.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ballinger, A. P., T. M. Merlis, I. M. Held, and M. Zhao, 2015: The sensitivity of tropical cyclone activity to off-equatorial thermal forcing in aquaplanet simulations. J. Atmos. Sci., 72, 22862302, https://doi.org/10.1175/JAS-D-14-0284.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Bembenek, E., D. N. Straub, and T. M. Merlis, 2020: Effects of moisture in a two-layer model of the midlatitude jet stream. J. Atmos. Sci., 77, 131147, https://doi.org/10.1175/JAS-D-19-0021.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Berry, G., and M. J. Reeder, 2013: Objective identification of the intertropical convergence zone: Climatology and trends from the ERA-Interim. J. Climate, 21, 18941909, https://doi.org/10.1175/JCLI-D-13-00339.1.

    • Search Google Scholar
    • Export Citation
  • Bouchut, F., J. Lambaerts, G. Lapeyre, and V. Zeitlin, 2009: Fronts and nonlinear waves in a simplified shallow-water model of the atmosphere with moisture and convection. Phys. Fluids, 21, 116604, https://doi.org/10.1063/1.3265970.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Cao, X., P. Huang, G. H. Chen, and W. Chen, 2012: Modulation of western North Pacific tropical cyclone genesis by intraseasonal oscillation of the ITCZ: A statistical analysis. Adv. Atmos. Sci., 29, 744754, https://doi.org/10.1007/s00376-012-1121-0.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Chavas, D. R., and K. A. Reed, 2019: Dynamical aquaplanet experiments with uniform thermal forcing: System dynamics and implications for tropical cyclone genesis and size. J. Atmos. Sci., 76, 22572274, https://doi.org/10.1175/JAS-D-19-0001.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Diaz, M., and W. R. Boos, 2019: Barotropic growth of monsoon depressions. Quart. J. Roy. Meteor. Soc., 145, 824844, https://doi.org/10.1002/qj.3467.

  • Dritschel, D., 1989: On the stabilization of a two-dimensional vortex strip by adverse shear. J. Fluid Mech., 206, 193221, https://doi.org/10.1017/S0022112089002284.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Emanuel, K., 2003: Tropical cyclones. Annu. Rev. Earth Planet. Sci., 31, 75104, https://doi.org/10.1146/annurev.earth.31.100901.141259.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Emanuel, K., J. D. Neelin, and C. S. Bretherton, 1994: On large-scale circulations in convecting atmospheres. Quart. J. Roy. Meteor. Soc., 120, 11111143, https://doi.org/10.1002/qj.49712051902.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Gill, A. E., 1982: Studies of moisture effects in simple atmospheric models: The stable case. Geophys. Astrophys. Fluid Dyn., 19, 119152, https://doi.org/10.1080/03091928208208950.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Held, I. M., and A. Y. Hou, 1980: Nonlinear axially symmetric circulations in a nearly inviscid atmosphere. J. Atmos. Sci., 37, 515533, https://doi.org/10.1175/1520-0469(1980)037<0515:NASCIA>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hsieh, T.-L., G. A. Vecchi, W. Yang, I. M. Held, and S. T. Garner, 2020: Large-scale control on the frequency of tropical cyclones and seeds: A consistent relationship across a hierarchy of global atmospheric models. Climate Dyn., 55, 31773196, https://doi.org/10.1007/s00382-020-05446-5.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • 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.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Kwon, H. J., 1989: A reexamination of the genesis of African waves. J. Atmos. Sci., 46, 36213631, https://doi.org/10.1175/1520-0469(1989)046<3621:AROTGO>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lambaerts, J., G. Lapeyre, and V. Zeitlin, 2011: Moist versus dry barotropic instability in a shallow-water model of the atmosphere with moist convection. J. Atmos. Sci., 68, 12341252, https://doi.org/10.1175/2011JAS3540.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lapeyre, G., and I. M. Held, 2004: The role of moisture in the dynamics and energetics of turbulent baroclinic eddies. J. Atmos. Sci., 61, 16931710, https://doi.org/10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lindzen, R. S., and S. Nigam, 1987: On the role of sea surface temperature gradients in forcing low-level winds and convergence in the tropics. J. Atmos. Sci., 44, 24182436, https://doi.org/10.1175/1520-0469(1987)044<2418:OTROSS>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Marshall, J., A. Donohoe, D. Ferreira, and D. McGee, 2014: The ocean’s role in setting the mean position of the inter-tropical convergence zone. Climate Dyn., 42, 19671979, https://doi.org/10.1007/s00382-013-1767-z.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Merlis, T. M., 2015: Direct weakening of tropical circulations from masked CO2. Proc. Natl. Acad. Sci. USA, 112, 13 16713 171, https://doi.org/10.1073/pnas.1508268112.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Merlis, T. M., M. Zhao, and I. M. Held, 2013: The sensitivity of hurricane frequency to ITCZ changes and radiatively forced warming in aquaplanet simulations. Geophys. Res. Lett., 40, 41094114, https://doi.org/10.1002/grl.50680.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Montgomery, M., M. Nicholls, T. Cram, and A. Saunders, 2006: A vortical hot tower route to tropical cyclogenesis. J. Atmos. Sci., 63, 355386, https://doi.org/10.1175/JAS3604.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Neelin, J. D., and I. M. Held, 1987: Modeling tropical convergence based on the moist static energy budget. Mon. Wea. Rev., 115, 312, https://doi.org/10.1175/1520-0493(1987)115<0003:MTCBOT>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Nieto Ferreira, R. N., and W. H. Schubert, 1997: Barotropic aspects of ITCZ breakdown. J. Atmos. Sci., 54, 261285, https://doi.org/10.1175/1520-0469(1997)054<0261:BAOIB>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • O’Gorman, P. A., R. P. Allan, M. P. Byrne, and M. Previdi, 2012: Energetic constraints on precipitation under climate change. Surv. Geophys., 33, 585608, https://doi.org/10.1007/s10712-011-9159-6.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Pritchard, M. S., and C. S. Bretherton, 2014: Causal evidence that rotational moisture advection is critical to the superparameterized Madden–Julian oscillation. J. Atmos. Sci., 71, 800815, https://doi.org/10.1175/JAS-D-13-0119.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Sobel, A. H., and C. Bretherton, 2000: Modeling tropical precipitation in a single column. J. Climate, 13, 43784392, https://doi.org/10.1175/1520-0442(2000)013<4378:MTPIAS>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Sobel, A. H., J. Nilsson, and L. M. Polvani, 2001: The weak temperature gradient approximation and balanced tropical moisture waves. J. Atmos. Sci., 58, 36503665, https://doi.org/10.1175/1520-0469(2001)058<3650:TWTGAA>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Suhas, D. L., and J. Sukhatme, 2020: Moist shallow-water response to tropical forcing: Initial-value problems. Quart. J. Roy. Meteor. Soc., 146, 36953714, https://doi.org/10.1002/qj.3867.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Suhas, D. L., J. Sukhatme, and J. M. Monteiro, 2017: Tropical vorticity forcing and superrotation in the spherical shallow-water equations. Quart. J. Roy. Meteor. Soc., 143, 957965, https://doi.org/10.1002/qj.2979.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Sukhatme, J., 2014: Low-frequency modes in an equatorial shallow-water model with moisture gradients. Quart. J. Roy. Meteor. Soc., 140, 18381846, https://doi.org/10.1002/qj.2264.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Thorncroft, C. D., and B. J. Hoskins, 1994: An idealized study of African easterly waves. I: A linear view. Quart. J. Roy. Meteor. Soc., 120, 953982, https://doi.org/10.1002/qj.49712051809.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Vallis, G. K., and J. Penn, 2020: Convective organization and eastward propagating equatorial disturbances in a simple excitable system. Quart. J. Roy. Meteor. Soc., 146, 22972314, https://doi.org/10.1002/qj.3792.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Waliser, D. E., and C. Gautier, 1993: A satellite-derived climatology of the ITCZ. J. Climate, 6, 21622174, https://doi.org/10.1175/1520-0442(1993)006<2162:ASDCOT>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wang, C.-C., and G. Magnusdottir, 2005: ITCZ breakdown in three-dimensional flows. J. Atmos. Sci., 62, 14971512, https://doi.org/10.1175/JAS3409.1.

  • Wang, C.-C., and G. Magnusdottir, 2006: The ITCZ in the central and eastern Pacific on synoptic time scales. Mon. Wea. Rev., 134, 14051421, https://doi.org/10.1175/MWR3130.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wheeler, M., G. N. Kiladis, and P. J. Webster, 2000: Large-scale dynamical fields associated with convectively coupled equatorial waves. J. Atmos. Sci., 57, 613640, https://doi.org/10.1175/1520-0469(2000)057<0613:LSDFAW>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wing, A. A., S. J. Camargo, and A. H. Sobel, 2016: Role of radiative–convective feedbacks in spontaneous tropical cyclogenesis in idealized numerical simulations. J. Atmos. Sci., 73, 26332642, https://doi.org/10.1175/JAS-D-15-0380.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wodzicki, K. R., and A. D. Rapp, 2016: Long-term characterization of the Pacific ITCZ using TRMM, GPCP, and ERA-Interim. J. Geophys. Res. Atmos., 121, 31533170, https://doi.org/10.1002/2015JD024458.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Yokota, S., H. Niino, and W. Yanase, 2012: Tropical cyclogenesis due to breakdown of intertropical convergence zone: An idealized numerical experiment. SOLA, 8, 103106, https://doi.org/10.2151/sola.2012-026.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Yokota, S., H. Niino, and W. Yanase, 2015: Tropical cyclogenesis due to breakdown of ITCZ: Idealized numerical experiments and a case study of the event in July 1988. J. Atmos. Sci., 72, 36633684, https://doi.org/10.1175/JAS-D-14-0328.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
1

Note that in Lambaerts et al. (2011), η is at the bottom of the atmosphere, while our model setup has η at the top of the lower layer of the atmosphere (e.g., Lindzen and Nigam 1987).

Save
  • Adam, O., 2018: Zonally varying ITCZs in a Matsuno-Gill-type model with an idealized Bjerknes feedback. J. Adv. Model. Earth Syst., 10, 13041318, https://doi.org/10.1029/2017MS001183.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Agee, E. M., 1972: Note on ITCZ wave disturbances and formation of Tropical Storm Anna. Mon. Wea. Rev., 100, 733737, https://doi.org/10.1175/1520-0493(1972)100<0733:NOIWDA>2.3.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Asaadi, A., G. Brunet, and M. K. Yau, 2016: On the dynamics of the formation of the Kelvin cat’s-eye in tropical cyclogenesis. Part II: Numerical simulation. J. Atmos. Sci., 73, 23392359, https://doi.org/10.1175/JAS-D-15-0237.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ballinger, A. P., T. M. Merlis, I. M. Held, and M. Zhao, 2015: The sensitivity of tropical cyclone activity to off-equatorial thermal forcing in aquaplanet simulations. J. Atmos. Sci., 72, 22862302, https://doi.org/10.1175/JAS-D-14-0284.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Bembenek, E., D. N. Straub, and T. M. Merlis, 2020: Effects of moisture in a two-layer model of the midlatitude jet stream. J. Atmos. Sci., 77, 131147, https://doi.org/10.1175/JAS-D-19-0021.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Berry, G., and M. J. Reeder, 2013: Objective identification of the intertropical convergence zone: Climatology and trends from the ERA-Interim. J. Climate, 21, 18941909, https://doi.org/10.1175/JCLI-D-13-00339.1.

    • Search Google Scholar
    • Export Citation
  • Bouchut, F., J. Lambaerts, G. Lapeyre, and V. Zeitlin, 2009: Fronts and nonlinear waves in a simplified shallow-water model of the atmosphere with moisture and convection. Phys. Fluids, 21, 116604, https://doi.org/10.1063/1.3265970.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Cao, X., P. Huang, G. H. Chen, and W. Chen, 2012: Modulation of western North Pacific tropical cyclone genesis by intraseasonal oscillation of the ITCZ: A statistical analysis. Adv. Atmos. Sci., 29, 744754, https://doi.org/10.1007/s00376-012-1121-0.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Chavas, D. R., and K. A. Reed, 2019: Dynamical aquaplanet experiments with uniform thermal forcing: System dynamics and implications for tropical cyclone genesis and size. J. Atmos. Sci., 76, 22572274, https://doi.org/10.1175/JAS-D-19-0001.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Diaz, M., and W. R. Boos, 2019: Barotropic growth of monsoon depressions. Quart. J. Roy. Meteor. Soc., 145, 824844, https://doi.org/10.1002/qj.3467.

  • Dritschel, D., 1989: On the stabilization of a two-dimensional vortex strip by adverse shear. J. Fluid Mech., 206, 193221, https://doi.org/10.1017/S0022112089002284.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Emanuel, K., 2003: Tropical cyclones. Annu. Rev. Earth Planet. Sci., 31, 75104, https://doi.org/10.1146/annurev.earth.31.100901.141259.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Emanuel, K., J. D. Neelin, and C. S. Bretherton, 1994: On large-scale circulations in convecting atmospheres. Quart. J. Roy. Meteor. Soc., 120, 11111143, https://doi.org/10.1002/qj.49712051902.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Gill, A. E., 1982: Studies of moisture effects in simple atmospheric models: The stable case. Geophys. Astrophys. Fluid Dyn., 19, 119152, https://doi.org/10.1080/03091928208208950.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Held, I. M., and A. Y. Hou, 1980: Nonlinear axially symmetric circulations in a nearly inviscid atmosphere. J. Atmos. Sci., 37, 515533, https://doi.org/10.1175/1520-0469(1980)037<0515:NASCIA>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hsieh, T.-L., G. A. Vecchi, W. Yang, I. M. Held, and S. T. Garner, 2020: Large-scale control on the frequency of tropical cyclones and seeds: A consistent relationship across a hierarchy of global atmospheric models. Climate Dyn., 55, 31773196, https://doi.org/10.1007/s00382-020-05446-5.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • 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.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Kwon, H. J., 1989: A reexamination of the genesis of African waves. J. Atmos. Sci., 46, 36213631, https://doi.org/10.1175/1520-0469(1989)046<3621:AROTGO>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lambaerts, J., G. Lapeyre, and V. Zeitlin, 2011: Moist versus dry barotropic instability in a shallow-water model of the atmosphere with moist convection. J. Atmos. Sci., 68, 12341252, https://doi.org/10.1175/2011JAS3540.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lapeyre, G., and I. M. Held, 2004: The role of moisture in the dynamics and energetics of turbulent baroclinic eddies. J. Atmos. Sci., 61, 16931710, https://doi.org/10.1175/1520-0469(2004)061<1693:TROMIT>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lindzen, R. S., and S. Nigam, 1987: On the role of sea surface temperature gradients in forcing low-level winds and convergence in the tropics. J. Atmos. Sci., 44, 24182436, https://doi.org/10.1175/1520-0469(1987)044<2418:OTROSS>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Marshall, J., A. Donohoe, D. Ferreira, and D. McGee, 2014: The ocean’s role in setting the mean position of the inter-tropical convergence zone. Climate Dyn., 42, 19671979, https://doi.org/10.1007/s00382-013-1767-z.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Merlis, T. M., 2015: Direct weakening of tropical circulations from masked CO2. Proc. Natl. Acad. Sci. USA, 112, 13 16713 171, https://doi.org/10.1073/pnas.1508268112.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Merlis, T. M., M. Zhao, and I. M. Held, 2013: The sensitivity of hurricane frequency to ITCZ changes and radiatively forced warming in aquaplanet simulations. Geophys. Res. Lett., 40, 41094114, https://doi.org/10.1002/grl.50680.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Montgomery, M., M. Nicholls, T. Cram, and A. Saunders, 2006: A vortical hot tower route to tropical cyclogenesis. J. Atmos. Sci., 63, 355386, https://doi.org/10.1175/JAS3604.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Neelin, J. D., and I. M. Held, 1987: Modeling tropical convergence based on the moist static energy budget. Mon. Wea. Rev., 115, 312, https://doi.org/10.1175/1520-0493(1987)115<0003:MTCBOT>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Nieto Ferreira, R. N., and W. H. Schubert, 1997: Barotropic aspects of ITCZ breakdown. J. Atmos. Sci., 54, 261285, https://doi.org/10.1175/1520-0469(1997)054<0261:BAOIB>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • O’Gorman, P. A., R. P. Allan, M. P. Byrne, and M. Previdi, 2012: Energetic constraints on precipitation under climate change. Surv. Geophys., 33, 585608, https://doi.org/10.1007/s10712-011-9159-6.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Pritchard, M. S., and C. S. Bretherton, 2014: Causal evidence that rotational moisture advection is critical to the superparameterized Madden–Julian oscillation. J. Atmos. Sci., 71, 800815, https://doi.org/10.1175/JAS-D-13-0119.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Sobel, A. H., and C. Bretherton, 2000: Modeling tropical precipitation in a single column. J. Climate, 13, 43784392, https://doi.org/10.1175/1520-0442(2000)013<4378:MTPIAS>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Sobel, A. H., J. Nilsson, and L. M. Polvani, 2001: The weak temperature gradient approximation and balanced tropical moisture waves. J. Atmos. Sci., 58, 36503665, https://doi.org/10.1175/1520-0469(2001)058<3650:TWTGAA>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Suhas, D. L., and J. Sukhatme, 2020: Moist shallow-water response to tropical forcing: Initial-value problems. Quart. J. Roy. Meteor. Soc., 146, 36953714, https://doi.org/10.1002/qj.3867.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Suhas, D. L., J. Sukhatme, and J. M. Monteiro, 2017: Tropical vorticity forcing and superrotation in the spherical shallow-water equations. Quart. J. Roy. Meteor. Soc., 143, 957965, https://doi.org/10.1002/qj.2979.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Sukhatme, J., 2014: Low-frequency modes in an equatorial shallow-water model with moisture gradients. Quart. J. Roy. Meteor. Soc., 140, 18381846, https://doi.org/10.1002/qj.2264.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Thorncroft, C. D., and B. J. Hoskins, 1994: An idealized study of African easterly waves. I: A linear view. Quart. J. Roy. Meteor. Soc., 120, 953982, https://doi.org/10.1002/qj.49712051809.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Vallis, G. K., and J. Penn, 2020: Convective organization and eastward propagating equatorial disturbances in a simple excitable system. Quart. J. Roy. Meteor. Soc., 146, 22972314, https://doi.org/10.1002/qj.3792.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Waliser, D. E., and C. Gautier, 1993: A satellite-derived climatology of the ITCZ. J. Climate, 6, 21622174, https://doi.org/10.1175/1520-0442(1993)006<2162:ASDCOT>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wang, C.-C., and G. Magnusdottir, 2005: ITCZ breakdown in three-dimensional flows. J. Atmos. Sci., 62, 14971512, https://doi.org/10.1175/JAS3409.1.

  • Wang, C.-C., and G. Magnusdottir, 2006: The ITCZ in the central and eastern Pacific on synoptic time scales. Mon. Wea. Rev., 134, 14051421, https://doi.org/10.1175/MWR3130.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wheeler, M., G. N. Kiladis, and P. J. Webster, 2000: Large-scale dynamical fields associated with convectively coupled equatorial waves. J. Atmos. Sci., 57, 613640, https://doi.org/10.1175/1520-0469(2000)057<0613:LSDFAW>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wing, A. A., S. J. Camargo, and A. H. Sobel, 2016: Role of radiative–convective feedbacks in spontaneous tropical cyclogenesis in idealized numerical simulations. J. Atmos. Sci., 73, 26332642, https://doi.org/10.1175/JAS-D-15-0380.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wodzicki, K. R., and A. D. Rapp, 2016: Long-term characterization of the Pacific ITCZ using TRMM, GPCP, and ERA-Interim. J. Geophys. Res. Atmos., 121, 31533170, https://doi.org/10.1002/2015JD024458.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Yokota, S., H. Niino, and W. Yanase, 2012: Tropical cyclogenesis due to breakdown of intertropical convergence zone: An idealized numerical experiment. SOLA, 8, 103106, https://doi.org/10.2151/sola.2012-026.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Yokota, S., H. Niino, and W. Yanase, 2015: Tropical cyclogenesis due to breakdown of ITCZ: Idealized numerical experiments and a case study of the event in July 1988. J. Atmos. Sci., 72, 36633684, https://doi.org/10.1175/JAS-D-14-0328.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Fig. 1.

    (top left) Heating function Q˜(y) and (top right) ramping function r(t). (bottom left) Zonal-mean vorticity ζ¯ and (bottom right) free surface η¯ of the spun-up base state. Meridional position y is Re(ϕϕc), where Re is Earth’s radius. Vertical dotted lines indicate position of the equator.

  • Fig. 2.

    Vorticity normalized by 10−5 s−1 for ϕc = (top) 6°, (middle) 10°, and (bottom) 14°. All snapshots are taken at t = 15 days.

  • Fig. 3.

    Time series of eddy energy Ee as a function of latitude ϕc. Least squares estimate of the growth rate σ indicated in parentheses in the legend (with units of days−1). To quantify σ, we calculate a least squares fit to the slope of these curves for the periods where ln(Ee) grows linearly, indicated by the vertical ticks in each time series.

  • Fig. 4.

    Percentage change in moist growth rate σ with respect to the dry (α = 0) simulation for each ϕc. See Fig. 3 for dry growth rates.

  • Fig. 5.

    Energy budget terms for simulations with flow-dependent precipitation (α > 0) and the corresponding dry case (α = 0). (top) Precipitation forcing on eddies Pe, (13); (middle) precipitation forcing on the zonal component Pz, (12); and (bottom) conversion from zonal to eddy kinetic energy Cz, (14), for the ϕc = 10° case.

  • Fig. 6.

    (left) Profiles of saturation humidity, qs = qs(y), as a function of γ. (right) Zonally averaged qs for simulations with qs = qs(η) and γ = 20 at t = 0, 10, 15, 20, and 30 days. The solid lines indicate qs during linear instability (for t ≤ 16 days) and the dotted lines indicates qs during nonlinear phase of simulation.

  • Fig. 7.

    Percentage change in growth rate σ as a function of γ for simulations with qs = qs(y) (black) and qs = qs(η) (gray). The nondimensional parameter γ controls the magnitude of the derivative of qs, see Eqs. (17) and (18). The simulation with γ = 0 corresponds to the constant qs case with α = 100 m kg g−1 and ϕc = 10°.

  • Fig. 8.

    Instantaneous eddy free surface η′ for (top) constant qs, (middle) qs = qs(y), and (bottom) qs = qs(η), with black contours indicating precipitation and dashed (solid) white contours indicate convergence (divergence). All simulations have α = 100 m kg g−1 and ϕc = 10° and fields are shown on day 15. The cases with qs = qs(y) and qs(η) have γ = 20.

All Time Past Year Past 30 Days
Abstract Views 430 0 0
Full Text Views 1800 1152 394
PDF Downloads 554 103 20