1. Introduction
Diurnal processes determine much of the variability of tropical weather. For rainfall, the amplitude of the mean diurnal harmonic exceeds 70% of the seasonal mean in many coastal tropical regions (Minobe and Takebayashi 2015). In locations like the Maritime Continent and the Bight of Panama, coastal processes themselves likely induce between 40% and 60% of rainfall (Bergemann et al. 2015). Satellite observations indicate that the peak of the rainfall diurnal harmonic occurs progressively later with distance from the coastline, in some locations propagating up to 1000 km offshore (e.g., Yang and Slingo 2001). Analogous behavior has been observed for the diurnal cycle of winds (Gille et al. 2005), leading to many hypotheses on the interconnectedness of these processes (e.g., Mori et al. 2004; Robinson et al. 2008; Vincent and Lane 2016; Kilpatrick et al. 2017).
One theory, originating with Mapes et al. (2003), is that gravity waves forced along or near coastlines propagate both offshore and inland, destabilizing the atmosphere and promoting convection as they go. These waves may be forced by topography or convective heating near coastlines, or by the diurnally oscillating temperature gradient associated with the coastline itself. In the tropics, gravity waves forced by coastal temperature gradients can be interpreted as the linear component of the land–sea breeze, manifesting as onshore surface winds in the afternoon and evening (the sea breeze), and offshore winds in the morning (the land breeze).
Rotunno (1983) demonstrated using a simple, two-dimensional linear theory that for latitudes between +30° and −30°, oscillating coastal temperature gradients force gravity wave rays, i.e., packets of gravity waves which propagate far offshore. The rays resemble the upper half of “Saint Andrew’s cross,” the signature of internal gravity waves generated by oscillatory forcings (Mowbray and Rarity 1967). The differing behavior of the land–sea breeze poleward and equatorward of 30° was subsequently corroborated by nonlinear numerical simulations (Yan and Anthes 1987).
Rotunno’s original theory employed the Boussinesq equations, linearized about a hydrostatic, stationary atmosphere, with a free-slip lower boundary. Subsequent studies modified his approach to incorporate momentum diffusivity (Niino 1987), friction (Dalu and Pielke 1989; Du and Rotunno 2015; Das Gupta et al. 2015), shore-parallel thermal wind shear (Drobinski et al. 2011), nonlinear coastlines (Li and Chao 2016; Jiang 2012a), constant background winds (Qian et al. 2009; Du and Rotunno 2018), and background winds that shear linearly with height (Du et al. 2019). Of particular relevance to the present study is the theory of Jiang (2012b), who divided the vertical domain into at most 3 distinct layers, allowing discontinuities in the Brunt–Väisälä frequency N, potential temperature, and background winds between each layer, but requiring they be constant within each layer. To solve the resulting equations, Jiang modified the heating function introduced by Rotunno to constrain heating to lie within a single layer.
In the present study, we consider some new extensions of Rotunno’s theory to situations involving nonconstant stability, i.e., a nonconstant Brunt–Väisälä frequency N. We first consider the case of a step change in N, i.e., N = N1 below some height H1 and N = N2 above H1, where N1 and N2 are distinct positive numbers. This analysis is similar to that of Lindzen and Tung (1976), who considered the anelastic equations, and calculated the reflection and refraction coefficients associated with a stability discontinuity for a single gravity wave forced below the stability change. Here we consider the simpler Boussinesq equations, but the more complex case of a forced spectrum of waves, and the effect of forcing above the stability change. This analysis is also similar to that of Jiang (2012b) discussed above, although here we consider the simpler problem of resting horizontal background winds, but the more complex situation of forcing above and below the stability change, and the possibility of a stability decrease with height.
Second, we consider the case where N = N1 below some height H1, N = N2 above another height H2, with H2 > H1, and with N transitioning linearly between N1 and N2 between H1 and H2. This setup admits an analytic solution that generalizes the piecewise-constant solution. To the best of our knowledge, a solution to this problem is yet to appear in the published literature, with the closest solution being for the three-layer model considered by Jiang (2012b), in which waves are forced only in the lowest vertical layer, and N is constant within each layer. We will refer to our two solutions as the “piecewise-constant” and “transition-layer” solutions, respectively. We present illustrative examples of both solutions with a low-level stability change emblematic of that between the boundary layer and free troposphere, as this setup is easiest to interpret and compare with previous work.
We then generalize the theory by considering an alternative heating function which emulates the upper-level heating associated with a diurnally recurring convective line. This extension is straightforward, as our core mathematical results are independent of the spatial structure of the heating function. We use this extended theory to interpret soundings acquired during the Australian leg of the Years of the Maritime Continent (YMC) campaign (Yoneyama and Zhang 2020), focusing on the stability change between the troposphere and stratosphere.
The remainder of this paper is organized as follows. In sections 2 and 3 we present the piecewise-constant and transition-layer solutions, respectively, and provide illustrative example solutions using the original low-level surface heating function of Rotunno (1983). In section 4 we introduce the YMC sounding data and the alternative upper-level convective diurnal heating function, and interpret the YMC data in light of the extended theory. Section 5 provides discussion and conclusions.
2. Piecewise-constant solution
The heating function given by Eq. (6) at 1200 LST, i.e.,
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
The solution for ψ2 proceeds almost identically to that for ψ1. As in the solutions of Rotunno (1983) and Qian et al. (2009), ψ is then recovered from
-
Nondimensionalize and combine the governing equations into a single partial differential equation (PDE) for the streamfunction ψ.
-
Perform a Fourier transform with respect to x and consider the e±it modes separately.
-
Solve the resulting PDE analytically using Green’s method.
-
Recover ψ by performing the inverse Fourier transform numerically using the trapezoid rule.
We now consider an illustrative example solution, presented in dimensional coordinates to aid intuition. We state the governing parameters in dimensional coordinates, with the corresponding nondimensional numbers provided in Table 1. We set the coastal width parameter L = 50 km, the heating depth H = 1 km, Brunt–Väisälä frequencies N1 = 0.01 s−1 and N2 = 0.03 s−1, giving
The parameters used for the various example solutions presented in this study, noting f is the Coriolis parameter and ω the angular frequency of Earth. See text for the definitions of each parameter and nondimensional number. Nonapplicable parameters are marked with N/A. The surface and convective heating functions are given by Eqs. (6) and (55), and depicted in Figs. 1 and 12, respectively.
Figure 2 depicts the example solution’s coastline perpendicular horizontal velocity
Coastline perpendicular vertical cross sections of the piecewise-constant solution for the coastline perpendicular horizontal winds in dimensional coordinates
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
Figures 3a and 3b decompose the
As in Fig. 2, but for (a) the horizontal velocities
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
Figures 4a and 4b illustrate the vertical velocity
As in Fig. 2, but for (a) the vertical velocity in dimensional coordinates
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
The reflection, refraction, and ducting coefficients defined above apply to the amplitudes of the streamfunctions and vertical velocities associated with individual waves. For the horizontal velocities, the refraction coefficient acquires an additional factor of
3. Transition-layer solution
Equation (23) is again solved using Green’s method. The Green’s function G can be obtained by considering the z′ ∈ D1, z′ ∈ DTL, and z′ ∈ D2 cases separately, as in the piecewise-constant N case. The resulting expressions are analogous to those for the piecewise-constant solution, but more complex to state, and so are relegated to appendix B. Equation (32) can then be decomposed
Figure 5 provides an illustrative example solution for the coastline perpendicular horizontal velocity
As in Fig. 2, but for the transition-layer solution for the coastline-perpendicular horizontal winds
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
Figures 6a–c decompose the
As in Fig. 5, but for (a) the horizontal velocities
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
Figures 7a and 7b illustrate the vertical velocity
As in Fig. 4, but for the transition-layer solutions for (a)
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
Plots illustrating how (a) the reflection coefficient r1, (b) the refraction coefficient r2, and (c),(d) the ducting coefficient r3, change as the transition layer deepens, for various values of the nondimensional vertical wavenumber m, noting
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
The short limit expression for the ducting coefficient r3 depends on both m and
To provide physical intuition for this result, consider again the problem of an unforced atmosphere with no lower boundary, a step change in stability at
Consider now the analogous transition-layer case, where stability increases by the same amount overall, but the change occurs linearly between
For a transition layer of finite, nonzero thickness, r1 and r2 depend on m, and the reflection and refraction coefficients of the rays will lie somewhere between the long and short limit values for the component waves comprising the rays. As the thickness of the transition layer increases, inequalities (49) and (50) imply a greater proportion of the ray’s wavenumber spectrum behaves according to the long limit, rather than the short limit. Thus, the reflection and refraction coefficients for the rays will also approach the long limits 0 and
4. Years of the Maritime Continent soundings
The original motive for formulating the transition-layer problem was to interpret a set of radiosonde observations obtained from the second Australian leg of the YMC field campaign during November and December 2019 (Yoneyama and Zhang 2020; Protat and McRobert 2020b; Protat et al. 2022). During this campaign, the R/V Investigator, Australia’s scientific research vessel, sailed through the waters near Darwin, the capital of Australia’s Northern Territory, as depicted in Fig. 9. Between 20 November and 2 December 2019, the Investigator traveled from the waters northwest of Darwin, past Croker Island, and back again, launching eight radiosondes per day as it did so.
Maps depicting, (a) Australia’s Northern Territory, its capital Darwin, and surrounding landmarks, and (b) the region within the red dashed box in (a), showing the location of the R/V Investigator between the 20 Nov and the 2 Dec 2019.
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
Figure 10 illustrates the vertical profile of the Brunt–Väisälä frequency N obtained from the soundings taken between 20 November and 2 December 2019. The light gray line depicts the profile of N at 0000 LST 20 November 2019, and the dark gray line the average over this day. The thicker black line provides the average over the entire 20 November–2 December 2019 time period, with a 200 m running mean applied to reduce the small-scale noise. Note the average profile of N is ≈0.01 s−1 below ≈15 km and ≈0.025 s−1 above ≈19 km, transitioning roughly linearly between these two values between these two heights. This structure is not an artifact of the averaging: it is also present, albeit in a noisier form, in the gray lines. Furthermore, an N transition layer is consistent with previous work demonstrating that under various tropopause definitions utilizing diverse variables, the tropopause resembles a transition layer, rather than a discontinuity (Pan et al. 2004; Schmidt et al. 2006; Fueglistaler et al. 2009; Feng et al. 2012).
The vertical profile of the Brunt–Väisälä frequency N from the soundings obtained during the portion of the R/V Investigator voyage depicted in Fig. 9b. The light gray line depicts the Brunt–Väisälä frequency at 0000 LST 20 Nov 2019, the dark gray line depicts the average over 20 Nov 2019, while the thicker black line depicts the average over the entire 20 Nov–2 Dec 2019 period, with a 200 m running mean applied to this average to further smooth small-scale noise.
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
Figure 11a provides a Hovmöller diagram of the meridional winds from the soundings between 0000 LST 20 November and 0000 LST 2 December 2019, with winds linearly interpolated in time onto a regular hourly time step. White regions indicate missing values, where the radiosonde’s balloon burst before ascending above 25 km. Note there is significant vertical shear in the meridional winds, particularly between ≈8 and 17 km. Recall that we assumed background winds were zero in the theory presented in sections 2 and 3, as sheared background winds significantly complicate the analytic approach (Du et al. 2019): our theory should therefore be applied cautiously to this YMC dataset. Nevertheless, note the weak stratospheric meridional background winds above ≈17 km in Fig. 11a, with clear downward-propagating diurnal perturbations.
Hovmöller diagrams of (a) the meridional winds, (b) the meridional wind perturbations against a 24 h centered-running-mean background wind, and (c) the average perturbations at each time of day, from the soundings taken on board the R/V Investigator between 20 Nov and 2 Dec 2019. The location of the R/V Investigator during this period is depicted in Fig. 9b.
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
Figure 11b provides the corresponding Hovmöller diagram of the wind perturbations against a 24-h centered running mean background wind, with missing values interpolated linearly when calculating the background wind, but not the perturbations. Diurnal perturbations are now evident throughout both the stratosphere and troposphere. Sometimes the tropospheric perturbations also propagate downward, but at a faster vertical phase speed than those in the stratosphere, while at other times they are essentially stationary. Figure 11c provides a Hovmöller diagram of the composite diurnal cycle, obtained by averaging the perturbations depicted in Fig. 11b at each hour of the day: the cycle is repeated once for clarity. A downward-propagating signal is again evident in the stratosphere, with the signal more stationary below 15 km, and showing a degree of discontinuity at ≈4 km. (Note that in Figs. 17b and 17c, the magnitude of the winds above 17 km are of comparable or greater magnitude than those below 17 km.)
Suppose we interpret the perturbations in Figs. 11b and 11c as representing gravity wave rays forced at the diurnal frequency. Such waves would likely have multiple sources, including the immediate Australian coastline, remote coastlines, and nearby and remote diurnal cycles of convection. The basic solution of Rotunno (1983), and Figs. 2 and 5, suggest that wave rays forced at the surface along a coastline would not reach the stratosphere until thousands of kilometers away from that coastline. Over such distances, the rays would likely disperse through the effect of background winds, friction, and violations of the f-plane assumption.
Suppose then that the waves are forced nearby, but at upper levels. Indeed, Hankinson et al. (2014) applied ray-tracing methods to stratospheric gravity waves observed in radiosonde observations over Darwin, concluding the waves originated from convection over Indonesia, the Philippines, and New Guinea. In particular, Hankinson et al. (2014) attributed the lower-stratospheric waves with vertical wavelengths of 2–4 km, similar magnitudes to those of the stratospheric signals in Fig. 11, to New Guinea convection. Furthermore, in a high-resolution simulation, Vincent and Lane (2016, Fig. 15) documented wave rays forced along the northern New Guinea coastline, with additional rays forced at upper levels around 12 km altitude, roughly in phase with those forced near the surface. These upper-level rays were likely forced by the convective and stratiform heating associated with the convective line that occurs daily over the New Guinea mountain range during austral summer.
The heating function given by Eq. (55) at 1200 LST, i.e.,
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
Figure 13 presents an example piecewise-constant solution using the revised heating function, with H1 = 17 km, L = 100 km, D = 4 km, and H = 12 km (see Table 1). It is somewhat unclear how Q0 should be chosen: as a first guess we take Q0 = 6 × 10−6 m s−3, half that of the surface forcing considered in sections 2 and 3, with this choice motivated by the potential temperature tendencies and perturbations in the simulation and observational results of Vincent and Lane (2016, 2018). From Fig. 10 we take N1 = 0.01 s−1 and N2 = 0.025 s−1, so that
The piecewise-constant solution for the convective-line-perpendicular horizontal winds
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
A Saint Andrew’s cross pattern is evident in Fig. 13, with reflection and refraction of the upper rays at H1. The reflected rays, which are narrow, superpose on the lower rays forced at 12 km. Vertical phase speeds associated with the upper rays are negative as before, but positive for the lower rays, as required by the radiation condition. Also, the horizontal winds are now out of phase either side of
Figure 14 decomposes the 1500 LST horizontal winds depicted in Fig. 13b into those arising from F1 and F2, respectively, i.e., forcing below and above the stability change at H1 = 17 km. As in section 2, the overall response is mostly determined by F1, recalling that the heating is concentrated at H = 12 km.
As in Fig. 13, but for (a) the horizontal velocities
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
Figure 15 provides the corresponding transition-layer solution, with H1 = 15 km and H2 = 19 km (see Table 1). As before, the rays reflect less than in the piecewise-constant solution, with more energy escaping into the stratosphere. Figure 16 decomposes the 1500 LST
As in Fig. 13, but for the transition-layer solution, with H1 = 15 km and H2 = 19 km, as depicted by the horizontal dotted lines.
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
As in Fig. 15, but for the horizontal velocities
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
To qualitatively compare the solutions above with the YMC soundings, suppose we interpret
Figure 17a provides a Hovmöller diagram of the
Hovmöller diagrams of the convective-line-perpendicular horizontal winds
Citation: Journal of the Atmospheric Sciences 80, 10; 10.1175/JAS-D-23-0074.1
Timings and magnitudes of the upper-level signals are comparable to those in Fig. 11, although the upper-level response ends below 20 km in Fig. 17, but extends to at least 25 km in Fig. 11. One explanation for this difference is the lack of background winds in the theory, with background winds in the convective-line-perpendicular direction acting to partially disperse the rays, resulting in more oscillations in the vertical (Qian et al. 2009).
A major difference between Figs. 11 and 17 is in the response below 12 km. In Fig. 17 the perturbations below 12 km exhibit positive vertical phase speeds as discussed above, whereas in Fig. 11 phase speeds are negative, or the signal is stationary. One explanation for this difference is that, in reality, additional surface or low-level convective diurnal heating is present, generating rays with negative phase speeds which superpose with the downward pointing rays with positive phase speeds generated by the upper-level forcing. This superpositioning could result in a signal that is stationary, or downward propagating. The apparent continuity of the upper and lower perturbations in Figs. 11b and 11c would then be essentially coincidental. These ideas could be critically tested using ray-tracing methods applied to more complex numerical model data (e.g., Hecht et al. 2004; Alexander et al. 2004; Vincent et al. 2004; Hankinson et al. 2014).
To generalize the above ideas, note that the line-parallel winds
5. Discussion and conclusions
In this study we extended the tropical linear sea-breeze theory of Rotunno (1983) to situations involving nonconstant Brunt–Väisälä frequency N. We first presented an illustrative example of a low-level stability change, emblematic of that between the boundary layer and troposphere. Because solutions are formulated using the Green’s function G, generalizing to other heating functions Q is straightforward. We therefore also considered an alternative heating function representing diurnally recurrent convective or stratiform heating aloft, and considered how forced waves behave as they passed through the tropopause.
In both cases wave reflection, refraction, and ducting behavior depends not only on the height and magnitude of the stability change, but on the thickness of the transition layer, at least in the idealized circumstances of the theory. This behavior follows from the structure of the Green’s function G, implying broad generality of this result. It may therefore be instructive to consider how the solutions presented here behave across diverse climatic conditions, and to compare this behavior with more realistic numerical simulations, or observations. For example, increases in tropopause height are anticipated under warming (e.g., Hu and Vallis 2019), and this will affect the tropopause reflection, refraction, and ducting behavior in the theory presented here. Furthermore, the thickness of the tropopause transition layer varies with latitude and season, and there is an observed thickening trend over the last four decades (Schmidt et al. 2006; Feng et al. 2012), with idealized models indicating the structure of the tropopause transition layer is sensitive to tropospheric warming, stratospheric cooling, and the vertical ozone profile (Lin et al. 2017; Dacie et al. 2019). The solutions presented in this study therefore suggest tropopause wave reflection, refraction, and ducting will vary significantly across season and latitude, and potentially also with climate. These hypotheses could be investigated with more realistic modeling experiments.
The solutions also provide test cases for assessing the numerical methods used in more complex models. In particular, if the vertical discretization of a finite-difference-based time stepping numerical method is too coarse over a stability change, we might expect simulated wave rays to reflect and refract as in the piecewise-constant solution, rather than the appropriate transition-layer solution, with finite-difference-based schemes therefore exaggerating the total energy reflected. In the online supplement we provide a preliminary investigation of these ideas.
Simplifying assumptions were made to derive the linear equations considered in this study, as in the related studies described in section 1. Most significant is perhaps linearity itself, which eliminates density current dynamics, an important aspect of the land–sea breeze at low levels near the coast (Qian et al. 2012). Nonlinear and nonhydrostatic effects may also be significant in certain contexts (e.g., Pandya et al. 1993, 2000). Furthermore, viscosity may increasingly attenuate wave rays with horizontal distance from their source (Yan and Anthes 1987; Dalu and Pielke 1989), so that if the stability change occurs a large vertical distance from the forcing, reflection and refraction become a moot point.
Another potential limitation is the Boussinesq approximation, which is usually thought to be valid only over vertical scales much less than the density scale height Hρ, where Hρ ≈ 8 km in the troposphere. However, Du et al. (2019) compared a similar set of analytic solutions, involving a nonconstant background wind
The most physically important process missing from the theory presented here is probably background winds, as these not only alter the basic structure of the wave rays (Qian et al. 2009), but changes in background winds with height also induce attenuation, reflection, and refraction (Du et al. 2019; Lane 2021) of their own. A straightforward extension of our piecewise-constant solution would be to ignore the Coriolis term, but include constant background winds above and below the stability change, with a discontinuity permitted at the stability change. Analogously, the transition-layer solution could be extended by ignoring Coriolis, but allowing the background winds to shear linearly between the constant values above and below the transition layer. Equation (41) would then become a form of Whittaker’s equation, with Whittaker function solutions, which may permit a similar analysis to that of the present paper.
To summarize, in this study we began by extending the linear sea-breeze theory developed by Rotunno (1983) to include vertical changes in the Brunt–Väisälä frequency N. In section 2 we presented the solution in the case where N = N1 below some height H1, and N = N2 above H1, where N1 and N2 are constant. The behavior of this solution is determined by the nondimensional parameters
In the piecewise-constant solution, gravity wave rays emanate from the origin, as in the base solution of Rotunno (1983). These rays are generated by heating below H1. The individual waves comprising the rays reflect and refract, with the amplitudes of the reflected and refracted waves governed by the coefficients
When stability decreases with height, or H1 is comparable or smaller than the vertical scale of the heating H, waves forced by heating above H1 also play a significant role. Below the forcing, the rigid lower boundary implies vertically nonpropagating waves are present above and below H1, with the ratio of their amplitudes given by the ducting coefficient
Overall, the reflection, refraction, and ducting behavior in the transition-layer solution is significantly different from the piecewise-constant solution. The coefficients r1, r2, and r3 now depend on m, but for each m the amplitudes of the reflected wave is lower, and the amplitude of the refracted wave greater, than in the corresponding piecewise-constant solution. The reflection, refraction, and ducting coefficients approach their piecewise-constant expressions when H2 → H1. Furthermore, the reflection, refraction, and ducting coefficients approach the limits 0,
The core reflection, refraction, and ducting behaviors described above all follow from the structure of the Green’s function G, not the spatial structure of the heating function Q, and it is therefore straightforward to consider other functions Q. In section 4 we considered an alternative heating function which emulates the upper-level convective or stratiform heating associated with a convective line. In this case H is the height at which heating is concentrated, and the solution depends on an additional nondimensional parameter
Acknowledgments.
Funding for this study was provided for Ewan Short by the Australian Research Council’s Centre of Excellence for Climate Extremes (CE170100023), and by the Australian Bureau of Meteorology. Thanks are due to Alain Protat, chief scientist during the Australian leg of the YMC campaign, the Australian Marine National Facility, the R/V Investigator crew, and the YMC science team for the soundings used in this study.
Data availability statement.
The code written to generate and plot the solutions, and animated versions of key figures, are freely available online (Short 2023). The YMC sounding data are also freely available online (Protat and McRobert 2020a).
APPENDIX A
Parabolic Cylinder Functions
Plots of A(z) and θ(z) are provided in the online supplement, and while difficult to prove formally, it is apparent that these functions do not oscillate for z ∈ DTL. Furthermore, in this form the functions Da(z) and Db(z) bear a close resemblance to the first-order Wentzel, Kramers, and Brillouin (WKB) approximate solutions to Eq. (41), and thus, the functions Da(z)eit and Db(z)e−it can be interpreted as modulated upward- and downward-propagating waves, respectively; further details are provided in the online supplement. In the form of Eq. (A10), algebra involving Da(z) and Db(z) is considerably simpler, and in their Bessel function forms, A and θ can be rapidly calculated to machine precision, simplifying numerical implementation.
APPENDIX B
Transition-Layer Expressions
As discussed in section 3, the transition-layer solution can be derived by separately solving for G(z, z′) for the subcases z′ ∈ D1, z′ ∈ DTL, and z′ ∈ D2. Here we present the expressions for G associated with
It is straightforward to show each of the above expressions for G approach their piecewise-constant forms as
Note that Bessel functions
One final limiting behavior remains to be checked. Suppose that instead of n(z) varying linearly over the transition layer between 1 and
REFERENCES
Abramowitz, M., and I. A. Stegun, 1972: Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. Dover, 1076 pp.
Alexander, M. J., P. T. May, and J. H. Beres, 2004: Gravity waves generated by convection in the Darwin area during the Darwin Area Wave Experiment. J. Geophys. Res., 109, D20S04, https://doi.org/10.1029/2004JD004729.
Bergemann, M., C. Jakob, and T. P. Lane, 2015: Global detection and analysis of coastline-associated rainfall using an objective pattern recognition technique. J. Climate, 28, 7225–7236, https://doi.org/10.1175/JCLI-D-15-0098.1.
Dacie, S., and Coauthors, 2019: A 1D RCE study of factors affecting the tropical tropopause layer and surface climate. J. Climate, 32, 6769–6782, https://doi.org/10.1175/JCLI-D-18-0778.1.
Dalu, G. A., and R. A. Pielke, 1989: An analytical study of the sea breeze. J. Atmos. Sci., 46, 1815–1825, https://doi.org/10.1175/1520-0469(1989)046<1815:AASOTS>2.0.CO;2.
Das Gupta, A. J., A. S. V. Murthy, and R. S. Nanjundiah, 2015: A study of singularities in the inviscid linear theory of sea breezes. Int. J. Adv. Eng. Sci. Appl. Math., 7, 33–37, https://doi.org/10.1007/s12572-015-0129-y.
Drobinski, P., R. Rotunno, and T. Dubos, 2011: Linear theory of the sea breeze in a thermal wind. Quart. J. Roy. Meteor. Soc., 137, 1602–1609, https://doi.org/10.1002/qj.847.
Du, Y., and R. Rotunno, 2015: Thermally driven diurnally periodic wind signals off the east coast of China. J. Atmos. Sci., 72, 2806–2821, https://doi.org/10.1175/JAS-D-14-0339.1.
Du, Y., and R. Rotunno, 2018: Diurnal cycle of rainfall and winds near the south coast of China. J. Atmos. Sci., 75, 2065–2082, https://doi.org/10.1175/JAS-D-17-0397.1.
Du, Y., R. Rotunno, and F. Zhang, 2019: Impact of vertical wind shear on gravity wave propagation in the land–sea-breeze circulation at the equator. J. Atmos. Sci., 76, 3247–3265, https://doi.org/10.1175/JAS-D-19-0069.1.
Feng, S., Y. Fu, and Q. Xiao, 2012: Trends in the global tropopause thickness revealed by radiosondes. Geophys. Res. Lett., 39, L20706, https://doi.org/10.1029/2012GL053460.
Fueglistaler, S., A. E. Dessler, T. J. Dunkerton, I. Folkins, Q. Fu, and P. W. Mote, 2009: Tropical tropopause layer. Rev. Geophys., 47, RG1004, https://doi.org/10.1029/2008RG000267.
Gille, S. T., S. G. L. Smith, and N. M. Statom, 2005: Global observations of the land breeze. Geophys. Res. Lett., 32, L05605, https://doi.org/10.1029/2004GL022139.
Hankinson, M. C. N., M. J. Reeder, and T. P. Lane, 2014: Gravity waves generated by convection during TWP-ICE: I. Inertia-gravity waves. J. Geophys. Res. Atmos., 119, 5269–5282, https://doi.org/10.1002/2013JD020724.
Hecht, J. H., S. Kovalam, P. T. May, G. Mills, R. A. Vincent, R. L. Walterscheid, and J. Woithe, 2004: Airglow imager observations of atmospheric gravity waves at Alice Springs and Adelaide, Australia during the Darwin Area Wave Experiment (DAWEX). J. Geophys. Res., 109, D20S05, https://doi.org/10.1029/2004JD004697.
Hu, S., and G. K. Vallis, 2019: Meridional structure and future changes of tropopause height and temperature. Quart. J. Roy. Meteor. Soc., 145, 2698–2717, https://doi.org/10.1002/qj.3587.
Jiang, Q., 2012a: A linear theory of three-dimensional land–sea breezes. J. Atmos. Sci., 69, 1890–1909, https://doi.org/10.1175/JAS-D-11-0137.1.
Jiang, Q., 2012b: On offshore propagating diurnal waves. J. Atmos. Sci., 69, 1562–1581, https://doi.org/10.1175/JAS-D-11-0220.1.
Kilpatrick, T., S.-P. Xie, and T. Nasuno, 2017: Diurnal convection-wind coupling in the Bay of Bengal. J. Geophys. Res. Atmos., 122, 9705–9720, https://doi.org/10.1002/2017JD027271.
Lane, T. P., 2021: Does lower-stratospheric shear influence the mesoscale organization of convection? Geophys. Res. Lett., 48, e2020GL091025, https://doi.org/10.1029/2020GL091025.
Li, Y., and J. Chao, 2016: An analytical solution for three-dimensional sea–land breeze. J. Atmos. Sci., 73, 41–54, https://doi.org/10.1175/JAS-D-14-0329.1.
Lin, P., D. Paynter, Y. Ming, and V. Ramaswamy, 2017: Changes of the tropical tropopause layer under global warming. J. Climate, 30, 1245–1258, https://doi.org/10.1175/JCLI-D-16-0457.1.
Lindzen, R. S., and K.-K. Tung, 1976: Banded convective activity and ducted gravity waves. Mon. Wea. Rev., 104, 1602–1617, https://doi.org/10.1175/1520-0493(1976)104<1602:BCAADG>2.0.CO;2.
Mapes, B. E., T. T. Warner, and M. Xu, 2003: Diurnal patterns of rainfall in northwestern South America. Part III: Diurnal gravity waves and nocturnal convection offshore. Mon. Wea. Rev., 131, 830–844, https://doi.org/10.1175/1520-0493(2003)131<0830:DPORIN>2.0.CO;2.
Minobe, S., and S. Takebayashi, 2015: Diurnal precipitation and high cloud frequency variability over the Gulf Stream and over the Kuroshio. Climate Dyn., 44, 2079–2095, https://doi.org/10.1007/s00382-014-2245-y.
Mori, S., and Coauthors, 2004: Diurnal land–sea rainfall peak migration over Sumatera Island, Indonesian Maritime Continent, observed by TRMM satellite and intensive rawinsonde soundings. Mon. Wea. Rev., 132, 2021–2039, https://doi.org/10.1175/1520-0493(2004)132<2021:DLRPMO>2.0.CO;2.
Mowbray, D. E., and B. S. H. Rarity, 1967: A theoretical and experimental investigation of the phase configuration of internal waves of small amplitude in a density stratified liquid. J. Fluid Mech., 28, 1–16, https://doi.org/10.1017/S0022112067001867.
Niino, H., 1987: The linear theory of land and sea breeze circulation. J. Meteor. Soc. Japan, 65, 901–921, https://doi.org/10.2151/jmsj1965.65.6_901.
Ogura, Y., and N. A. Phillips, 1962: Scale analysis of deep and shallow convection in the atmosphere. J. Atmos. Sci., 19, 173–179, https://doi.org/10.1175/1520-0469(1962)019<0173:SAODAS>2.0.CO;2.
Pan, L. L., W. J. Randel, B. L. Gary, M. J. Mahoney, and E. J. Hintsa, 2004: Definitions and sharpness of the extratropical tropopause: A trace gas perspective. J. Geophys. Res., 109, D23103, https://doi.org/10.1029/2004JD004982.
Pandya, R., D. Durran, and C. Bretherton, 1993: Comments on “Thermally forced gravity waves in an atmosphere at rest.” J. Atmos. Sci., 50, 4097–4101, https://doi.org/10.1175/1520-0469(1993)050<4097:COFGWI>2.0.CO;2.
Pandya, R., D. Durran, and M. L. Weisman, 2000: The influence of convective thermal forcing on the three-dimensional circulation around squall lines. J. Atmos. Sci., 57, 29–45, https://doi.org/10.1175/1520-0469(2000)057<0029:TIOCTF>2.0.CO;2.
Protat, A., and I. McRobert, 2020a: RV Investigator IN2019_v06 BASTA radar and radiosonde data. Commonwealth Scientific and Industrial Research Organisation, accessed 23 February 2023, https://doi.org/10.25919/5ef5a70ae2be9.
Protat, A., and I. McRobert, 2020b: Three-dimensional wind profiles using a stabilized shipborne cloud radar in wind profiler mode. Atmos. Meas. Tech., 13, 3609–3620, https://doi.org/10.5194/amt-13-3609-2020.
Protat, A., V. Louf, J. Soderholm, J. Brook, and W. Ponsonby, 2022: Three-way calibration checks using ground-based, ship-based, and spaceborne radars. Atmos. Meas. Tech., 15, 915–926, https://doi.org/10.5194/amt-15-915-2022.
Qian, T., C. C. Epifanio, and F. Zhang, 2009: Linear theory calculations for the sea breeze in a background wind: The equatorial case. J. Atmos. Sci., 66, 1749–1763, https://doi.org/10.1175/2008JAS2851.1.
Qian, T., C. C. Epifanio, and F. Zhang, 2012: Topographic effects on the tropical land and sea breeze. J. Atmos. Sci., 69, 130–149, https://doi.org/10.1175/JAS-D-11-011.1.
Robinson, F. J., S. C. Sherwood, and Y. Li, 2008: Resonant response of deep convection to surface hot spots. J. Atmos. Sci., 65, 276–286, https://doi.org/10.1175/2007JAS2398.1.
Rotunno, R., 1983: On the linear theory of the land and sea breeze. J. Atmos. Sci., 40, 1999–2009, https://doi.org/10.1175/1520-0469(1983)040<1999:OTLTOT>2.0.CO;2.
Schmidt, T., G. Beyerle, S. Heise, J. Wickert, and M. Rothacher, 2006: A climatology of multiple tropopauses derived from GPS radio occultations with CHAMP and SAC-C. Geophys. Res. Lett., 33, L04808, https://doi.org/10.1029/2005GL024600.
Short, E., 2023: rotunno83. GitHub, accessed 17 March 2023, https://github.com/eshort0401/rotunno83.
Sutherland, B. R., 2010: Internal Gravity Waves. Cambridge University Press, 377 pp., https://doi.org/10.1017/CBO9780511780318.
Vincent, C. L., and T. P. Lane, 2016: Evolution of the diurnal precipitation cycle with the passage of a Madden–Julian oscillation event through the Maritime Continent. Mon. Wea. Rev., 144, 1983–2005, https://doi.org/10.1175/MWR-D-15-0326.1.
Vincent, C. L., and T. P. Lane, 2018: Mesoscale variation in diabatic heating around Sumatra, and its modulation with the Madden–Julian oscillation. Mon. Wea. Rev., 146, 2599–2614, https://doi.org/10.1175/MWR-D-17-0392.1.
Vincent, R. A., A. MacKinnon, I. M. Reid, and M. J. Alexander, 2004: VHF profiler observations of winds and waves in the troposphere during the Darwin Area Wave Experiment (DAWEX). J. Geophys. Res., 109, D20S02, https://doi.org/10.1029/2004JD004714.
Whittaker, E. T., and G. N. Watson, 1996: A Course of Modern Analysis. 4th ed. Cambridge Mathematical Library, Cambridge University Press, 608 pp., https://doi.org/10.1017/CBO9780511608759.
Wünsche, A., 2003: Generalized Hermite polynomials associated with functions of parabolic cylinder. Appl. Math. Comput., 141, 197–213, https://doi.org/10.1016/S0096-3003(02)00333-8.
Yan, H., and R. A. Anthes, 1987: The effect of latitude on the sea breeze. Mon. Wea. Rev., 115, 936–956, https://doi.org/10.1175/1520-0493(1987)115<0936:TEOLOT>2.0.CO;2.
Yang, G.-Y., and J. Slingo, 2001: The diurnal cycle in the tropics. Mon. Wea. Rev., 129, 784–801, https://doi.org/10.1175/1520-0493(2001)129<0784:TDCITT>2.0.CO;2.
Yoneyama, K., and C. Zhang, 2020: Years of the Maritime Continent. Geophys. Res. Lett., 47, e2020GL087182, https://doi.org/10.1029/2020GL087182.