1. Introduction
Tropical islands are hot spots for cumulus convection because of their strong perturbation of conditionally unstable airflow (e.g., Qian 2008; Smith et al. 2009). Under weak winds, the stronger daytime heating over land than over the surrounding sea destabilizes the island flow and gives rise to thermally direct circulations that promote vertical motion. These thermal circulations often collapse into sea breeze–land breeze fronts, with potent updrafts along and ahead of their frontal surfaces (e.g., Kingsmill 1995; Carbone et al. 2000; Fovell 2005). As the wind strengthens and/or the island narrows, the island heat anomaly is carried into the wake and organizes into quasi-linear convergence zones. The associated updrafts may initiate cloud trails extending far downwind (e.g., Yang et al. 2008; Kirshbaum and Fairman 2015).
Convection may also be initiated mechanically as impinging moist flow is forced to ascend or divert around mountainous island terrain. Over the Hawaiian islands of Kauai and Oahu, trade winds are often strong enough to surmount the 1000–1500-m-terrain crests, leading to persistent heavy precipitation over the windward slopes (e.g., Schroeder 1977; Ramage and Schroeder 1999). As the island height increases (e.g., over the island of Hawaii) or the cross-barrier wind decreases, the impinging airflow eventually becomes blocked, with convergence zones shifting upwind and/or downwind of the island (e.g., Smolarkiewicz et al. 1988; Yang and Chen 2008). Mechanical and thermal forcings often interact, particularly in blocked flows over taller and larger islands like Hawaii and Taiwan, to produce rich diurnal variations in clouds and precipitation (e.g., Yeh and Chen 1998; Yang and Chen 2008).
Although less studied than Hawaii, the Lesser Antilles islands in the Caribbean Sea also regularly initiate cumulus convection. These islands are situated in the subtropical Atlantic trade wind belt with similar heights and sizes as Kauai. The island of Dominica, with a peak height of ~1.5 km (see Fig. 1), has received recent attention because of its dramatic orographic enhancement of trade wind precipitation and its pronounced dynamical wakes (Kirshbaum and Smith 2009; Smith et al. 2009, 2012; Minder et al. 2013). During the Dominica Experiment (DOMEX) in April–May 2011, surface rain gauges, meteorological stations, operational radars, and the Wyoming King Air (WKA) research aircraft observed Dominican airflow and convection (Smith et al. 2012, hereafter S12). As expected, many events featured strong trade winds, large orographic precipitation enhancements, and turbulent island wakes. However, four other events featured unexpectedly weak trade winds and thermally driven convection, with strikingly different dynamical and microphysical signatures from the strong-wind events.
While thermally forced convection over flat islands and mechanically forced convection over mountainous islands have each received attention separately, few studies have examined thermally forced convection over mountainous islands. This subject is important because a large fraction of tropical islands (e.g., the Lesser Antilles, Hawaii, and the Maritime Continent) are mountainous, and the mountains may play a pivotal role in convection initiation. In particular, elevated heating enhances local baroclinicity, which may enhance flow convergence and convection initiation (e.g., Kirshbaum and Wang 2014). On the other hand, high terrain can block the inland penetration of sea breezes, thus inhibiting island convection (Qian et al. 2012; Barthlott and Kirshbaum 2013). While recent studies have attempted to quantify elevated-heating effects over idealized terrain using linear models and heat-engine theory (e.g., Tian and Parker 2003; Crook and Tucker 2005; Kirshbaum and Wang 2014), the applicability of such simple models to turbulent flows over realistic topography remains uncertain.
Herein we exploit DOMEX observations and cloud-resolving numerical simulations to investigate thermally forced convection over Dominica. We study the “golden” event of 18 April 2011, which was characterized by the most vigorous and quasi-steady cumulus convection of the four weak-wind events. Section 2 provides an observational overview of this event, illustrating the diurnal evolution and finescale structure of the island convection. To complement the observations, sections 3 and 4 provide cloud-resolving numerical simulations of the event. Section 3 observationally verifies a set of “control” simulations at different grid resolutions, and section 4 presents sensitivity experiments that evaluate the impacts of specific environmental and topographic parameters on the island circulations and convection. Section 5 uses heat-engine theory to quantify and interpret the subcloud thermal circulations, and section 6 presents the conclusions.
2. Observations
The 18 April 2011 convection event over Dominica, observed by the seventh research flight (RF07) of DOMEX, was identified by S12 as a prototypical case of thermally forced island convection under weak background winds. In a preliminary analysis of the event, S12 contrasted WKA observations of RF07 against a prototypical strong-wind event (research flight 13, or RF13) on 27 April 2011. Whereas the cross-barrier winds in the lowest kilometer were less than 2 m s−1 in RF07, they exceeded 12 m s−1 in RF13. S12 identified several key differences between the two events, including (i) a transition from divergent (RF07) to convergent (RF13) flight-level (1.8 km) winds over the island terrain; (ii) a sharp increase in the cloud aerosol and droplet-number concentrations, and a corresponding decrease in the droplet diameter, in RF07; and (iii) much heavier precipitation in RF13. These striking differences have motivated more detailed studies of each event: the RF13 case was studied by Minder et al. (2013) and the RF07 case is studied herein. In contrast to S12, we focus primarily on the island thermal circulations, rather than the cloud microphysics, of RF07.
a. Data sources
A full description of the DOMEX observational facilities is provided in S12. Here we briefly review the subset of DOMEX observations used in this study, which are summarized in Fig. 1. Operational surface meteorological data were obtained from Météo France stations on Martinique (Aimé Césaire Airport, station identifier TFFF) and Guadeloupe (Pointe-à-Pitre Airport, station identifier TFFR). An additional “mountain weather” station on Dominica was installed for DOMEX on an east-facing ridge near Freshwater Lake (FWL) at a height of 860 m above mean sea level (MSL), with data at 2-min resolution. Plan-position-indicator scans were provided from an operational Météo France radar on Martinique with a nominal frequency of 5 min (a radar was also present on Guadeloupe but it was not used). The radar elevation angle was 0.5° everywhere except over and past terrain obstacles, where the lowest unblocked elevation above ground was used. In addition, raw visible satellite data from the National Oceanic and Atmospheric Administration (NOAA) Geostationary Operational Environmental Satellite-13 (GOES-13), with a pixel spacing of 1 km and a nominal time resolution of 30 min, were obtained from NOAA’s Comprehensive Large Array-Data Stewardship System (CLASS).
DOMEX featured 21 WKA research flights over April–May 2011. The nominal WKA flight pattern consisted of eight horizontal flight legs plus an upstream aircraft sounding from 150 to 4000 m (S12). Three ocean legs on different sides of the island were flown at two different altitudes (low, or “L,” at ~300 m and high, or “H,” at ~1200 m) to sample upstream and downstream airflow at different levels, and two over-island legs were flown at ~1800 m. On 18 April 2011, however, the flight track was altered to repeatedly sample the island convection in a racetrack pattern over the high terrain. Ten legs were flown, including leg 1 well upstream of Dominica (both low and high), the leg 2 dogleg just off the east coast (both low and high), and three legs 3 and 4 directly over the island (Fig. 1).
In situ aircraft data include positional measurements using an inertial navigation station (INS) corrected with differential GPS (using a base station on Martinique), three-dimensional winds using a gust probe with INS, static temperature using a reverse-flow thermometer, static pressure using a Rosemount 1501 sensor with INS, water-vapor mixing ratio using a LICOR-6262 infrared gas analyzer, cloud-droplet size distributions using a forward-scattering cloud-droplet probe (CDP), and rain size distributions using two-dimensional optical array precipitation probe (2D-P). Although the in situ data were available at a sampling frequency of 25 Hz, we use a lower sampling rate (1 Hz) for ease of analysis. With a WKA airspeed of around 90 m s−1, this gives a sampling distance of around 90 m, which is sufficient for comparison with our numerical simulations. We also incorporate reflectivity and radial velocity profiles from the 95-GHz Wyoming Cloud Radar (WCR; http://www.atmos.uwyo.edu/wcr/) aboard the WKA.
b. Overview of RF07
The WKA sounding, recorded as the WKA looped around the southern half of leg 1 over 1138–1153 local solar time (LST) while descending from ~4000 to ~150 m, is shown in Fig. 2. To deepen the sounding into the stratosphere, we patched the operational 1200 UTC Guadeloupe (TFFR) sounding to the top of the aircraft sounding at 3792 m (642 hPa). Despite the reduced resolution of the TFFR sounding, the two profiles connect smoothly with no obvious jumps in the temperature, dewpoint, or wind profiles.
The merged sounding contains a shallow subcloud layer (from 0 to 600 m, or 1007 to 940 hPa) and a dry, conditionally unstable cloud layer (from 800 to 3500 m, or 917 to 660 hPa) separated by a sharp transition layer. Subcloud winds are weak and northerly while cloud-layer winds are also weak and westerly. The cloud layer is overlaid by a very dry and nearly moist-neutral layer extending to ~12 km. The sounding has a mean-layer (0–500 m) convective available potential energy (CAPE) and convective inhibition (CIN) of 119 and 87 J kg−1, respectively, suggesting weak instability and strong inhibition. The lifting condensation level (LCL), level of free convection (LFC), and level of neutral buoyancy (LNB) are 810 m (916 hPa), 1177 m (877 hPa), and 8535 m (347 hPa), respectively.
Time series of surface observations from the FWL mountain weather station reveal that the incoming shortwave radiation (
The sea level pressure difference between FWL and the base of the ocean sounding (ΔSLP) indicates the diurnal evolution of the thermally induced pressure anomalies that drove the island-scale circulation (Fig. 3e). Care is required in this computation because the FWL time series contains diurnal and semidiurnal tides while the aircraft sounding is instantaneous. To create an ocean SLP time series that is directly comparable to the FWL time series, we follow a similar procedure to Minder et al. (2013): we average the SLP from two coastal surface stations (the Guadeloupe and Martinique airports) and isolate the tidal variations by applying a low-pass Butterworth filter with a cutoff frequency of four cycles per day. We then fit the filtered time series to the instantaneous ocean SLP measurement to create an ocean SLP time series, which is subtracted from the FWL SLP time series to give ΔSLP.
As expected from hydrostatic considerations, the ΔSLP thus produced follows the opposite evolution as
The diurnal cloud development is depicted by the satellite-derived effective albedo at four times (0815, 1115, 1415, and 1715 LST) in Figs. 4a–d. The effective albedo is converted from raw satellite counts using online calibration routines for GOES imagers. The island clouds were largely nonexistent in the early morning and formed over the higher terrain at ~1000 LST (Figs. 4a,b). They became more widespread in the afternoon then gradually dissipated (Figs. 4c,d). The cloud fraction (
Figure 5 shows in situ WKA horizontal velocities along legs 1L, 2L, 3, and 4, overlaid upon the averaged cloud albedo over the 1200–1500 LST flight period. Only the third passes of legs 3 and 4 are shown. Over leg 1 the subcloud winds transitioned from north-northeasterly in the north to northerly in the south, suggesting flow deflection around the island. In the absence of thermal forcing, such deflection is favored when the nondimensional mountain height (
WCR reflectivity and radial velocity profiles along legs 3 and 4 (again the third pass of each leg) indicate that the leg-4 clouds were shallow with mainly ascending hydrometeors while the leg-3 hydrometeors were deeper and precipitating (Fig. 6). Consistent with S12’s conceptual diagram (their Fig. 15a), the cloud initiation was focused over the western ridge flank because of advection of the island thermal anomaly by the north-northeasterly subcloud winds. The clouds matured as they traversed the ridge (under the westerly cloud-layer winds) and then decayed over eastern Dominica.
The leg-3 rainshafts produced radar echoes of 10–20 dBZ near flight level (
3. Control simulations
a. Setup
The Advanced Research Weather Research and Forecasting (WRF) Model version 3.5 is used to perform quasi-idealized simulations. The simulations use a third-order Runge–Kutta time-integration scheme and fifth (third)-order advection in the horizontal (vertical). The simulations are nonhydrostatic and apply the Coriolis force to perturbations from the initial state (which is assumed to be geostrophically balanced) using an f-plane approximation at a latitude of 15°N. A single domain is used with dimensions of
Other parameterizations include shortwave (Dudhia scheme) and longwave [Rapid Radiative Transfer Model (RRTM)] radiation. Sloping-terrain effects and topographic shading are considered in the shortwave scheme. Surface-layer physics are parameterized based on Monin–Obukhov similarity theory, and the atmospheric boundary layer is either represented explicitly (for
Cumulus convection is represented explicitly and cloud microphysics are parameterized using the Thompson scheme, which is double moment for rain and ice and single moment for cloud liquid water, snow, and graupel. At the largest grid spacing (
The simulations use a horizontally homogeneous, single-sounding initialization based on the sounding in Fig. 2 and are integrated for 12 h (from 0500 to 1700 LST) to capture half the diurnal cycle. Because the ocean has a weak diurnal cycle, we assume a quasi-steady oceanic trade wind flow throughout the day. To enforce this steady state we add tendencies to the model variables at each time step to offset the two key processes causing model drift: low-level moistening due to ocean evaporation and clear-air radiative cooling. Because few cumuli develop over the ocean, no additional tendencies are required to offset their effects. To quantify these forcings, we performed a separate simulation with
In the course of our numerical experimentation a complication arose in simulations with Δ = 125 m: large-amplitude sound waves developed aloft directly over the island. To alleviate this problem we changed the WRF time off-centering parameter (or epssm) from its default value of 0.1 to 0.2, which eliminated the sound waves without noticeably changing the flow dynamics and convection of interest. However, by the time this problem was diagnosed and corrected we had already conducted most of our sensitivity simulations of section 4. Thus, the sensitivity simulations (which use Δ = 250 m and do not exhibit noticeable sound-wave propagation) use a different value of epssm (0.1) than the control simulations herein (0.2).
b. Description
The four “control” simulations use the above configuration with Δ progressively halved from 1000 m (CTL-1000), to 500 m (CTL-500), to 250 m (CTL-250), to 125 m (CTL-125), the last being the finest resolution permitted by our computing resources. These grid spacings span the range from modern limited-area explicit-convection forecasts (e.g., Lean et al. 2008) to large-eddy simulations of shallow convection (e.g., Kirshbaum and Grant 2012). Because the smallest Δ (125 m) is just sufficient to resolve deep convection (Bryan et al. 2003), it is still too coarse to resolve the shallow convection during RF07. Nonetheless, as will be demonstrated shortly, the CTL-250 and CTL-125 cases verify similarly against observations, suggesting diminishing marginal returns of higher resolution.
Before proceeding to the model verification, we illustrate the diurnal cycle for the CTL-250 case, which is revealed by snapshots of liquid water path (LWP) and first-model-level wind vectors at 0800, 1100, 1400, and 1700 LST in Figs. 7a–d. A katabatic land breeze propagates offshore in the morning hours, causing flow convergence over the surrounding ocean at 0800 LST. By 1100 LST strong thermally forced onshore flow develops over the high terrain, with intense convergence over the ridge axis giving rise to a band of densely packed shallow cumuli. Although the convergence and convection continue into the afternoon, both become less sharply focused over the ridge axis by 1400 LST, as cumulus detrainment produces shallow stratiform clouds in the upper cloud layer that drift eastward. As the insolation weakens so does the thermal forcing and horizontal convergence, causing the cumuli to largely dissipate by 1700 LST. Although LWP and albedo are not directly comparable, this simulated diurnal cloud evolution is qualitatively consistent with the RF07 satellite observations in Fig. 4.
Figure 8 compares surface wind vectors and LWP at 1400 LST for the four simulations. The mesoscale flow patterns are similar except for enhanced small-scale wind variability over the more rugged terrain at smaller Δ. Compared to the in situ wind observations at 300 m MSL along legs 1 and 2 (Fig. 5), the simulated surface winds are more northerly. This difference is owing to two effects: (i) the surface winds undergo more frictional backing than the flight-level winds at 300 m and (ii) the simulation was initialized by the aircraft sounding, which was flown over the southern half of leg 1 where the island-deflected winds were more northerly (see Fig. 5). As the grid resolution increases the simulated cumuli become more turbulent and irregular, with the area of nonzero LWP increasing but the patches of large LWP (2 kg m−2 and higher) shrinking. The stratiform clouds east of the ridge develop only in simulations with
A vertical cross section of liquid-water mixing ratio (
c. Verification
Because of various numerical idealizations such as the highly simplified land surface scheme, poorly constrained skin temperatures, and steady-state background flow, the quasi-idealized simulations should not be expected to exactly reproduce the observations. However, some model verification is useful to establish whether the simulated flows are broadly realistic. We begin with a comparison of simulated and observed FWL surface data, in which
Similarly, the diurnal RH variations are muted in the CTL-1000 case yet highly accurate for the other three cases (Fig. 3c). However, the impressive performance of the latter group is partially owing to a cancellation of errors. Both
Table 1 compares observed and simulated values of the above quantities, averaged over all points sampled in the three legs. While the simulations reproduce conditionally averaged quantities such as
Comparison of in situ WKA aircraft observations with corresponding quantities from the control simulations. All quantities are defined in the text. The entries contain the corresponding observed/simulated values at the grid resolutions (Δ) shown in the top row for legs 3 and 4. A value of “NA,” or not applicable, denotes zero samples of the quantity along the flight leg.
Despite the underestimated
The simulations with
Part of the above discrepancy between observed and simulated rainfall likely stems from radar precipitation underestimation. The TFFF radar was insensitive to echoes weaker than 12 dBZ, implying that the lighter rain over the ridge was undetectable. Moreover, the radar grid resolution (nominally 1 km) was coarser than most of the island showers, which reduced their effective reflectivity and caused some of them to go undetected. For a characteristic rain shower width of 500 m, averaging over a grid cell implies a 6-dBZ reduction in reflectivity and a 69% reduction in radar-derived rain rate. Finally, beam blocking past the southern island peaks (see Fig. 1) may also have shielded the TFFF radar from some of the heavier rainfall cores over the central island.
The remaining rainfall discrepancies likely stem from deficiencies in the numerical simulations. Given that the cloud macrophysics (e.g., b, w,
Overall, the control simulations are far from perfect, with underestimated cloud-layer mass fluxes and substantially overestimated precipitation. Nonetheless, the simulated thermal circulations and cloud macrophysics are sufficiently realistic (for
4. Sensitivity tests
To better understand the processes regulating thermally forced island convection, we perform a series of “sensitivity” simulations that isolate the impacts of various environmental, topographic, and cloud-related parameters. The reference case for these experiments is the CTL-250 simulation with epssm = 0.1, which for convenience (and to distinguish it from the original CTL-250 simulation with epssm = 0.2) will be henceforth denoted as the reference (REF) simulation. We choose
a. Topographic forcing
To quantify the impacts of island topography on the simulated convection, we perform two simulations that are identical to REF except that in one the terrain height is capped at 2 m (THERM) and in the other the surface heat fluxes are shut off over the domain (MECH). The former corresponds to pure thermal forcing while the latter corresponds to pure mechanical forcing. As shown in the LWP and surface wind comparison at 1400 LST in Figs. 12a–c, both REF and THERM produce strong low-level inflow and cumulus convection while MECH exhibits upstream blocking, a turbulent wake, and cloud-free conditions. This comparison supports the hypothesis that thermal (rather than mechanical) forcing drove the cumulus convection in RF07. The field of island cumuli in THERM is similar to REF except that the patches of large LWP are more focused over the western side of the island, coinciding with a narrow zone of sea-breeze convergence.
The island-averaged mean-layer (0–500 m above ground level, or AGL) equivalent potential temperature (
To characterize the bulk island-scale circulation we compare the island horizontal influx (
Interestingly, the maximum
b. Boundary layer winds
In conditionally unstable flow over mountainous islands, the convection-initiation regime transitions from thermal to mechanical as the ambient cross-barrier winds increase. For typical trade wind flows over Dominica, this transition occurs at a cross-barrier wind speed of ~5 m s−1 (Nugent et al. 2014). Of the four thermally forced cases observed in DOMEX, the range of low-level wind speeds was small (1.6–3.6 m s−1) but the range of WKA-observed
We thus perform three simulations that are all identical to the REF case except for minor modifications to the 0–2-km wind profile: the HALFWIND and DBLWIND cases respectively halve and double the wind speeds over 0–1 km with no change in direction, and the CBWIND case rotates the 0–1-km winds to a more cross-barrier orientation (60° clockwise from due north) with no change in wind speed. The modified winds are linearly relaxed back to the REF profile over 1–2 km. Because the modified winds in HALFWIND and DBLWIND change the oceanic latent heat fluxes, we perform separate 2D ocean-only simulations (as described in section 3) to recompute
Although the above changes to the ambient winds are small (less than 2 m s−1 in magnitude), they significantly impact the island thermal circulation and cumulus convection. While the island flow pattern is broadly similar among the REF, HALFWIND, and DBLWIND cases, the CBWIND case forms a distinct island wake with lee vortices and reversed (westerly) flow in the center (Figs. 12d–f). Qualitative changes in the cloud fields include a more prominent stratiform outflow layer in HALFWIND, a southward cloud-field displacement in DBLWIND, and a narrower cloud field centered over the ridge axis in CBWIND.
Relative to the REF case, the island mean-layer
The differences in low-level
c. Cloud feedbacks
The island clouds feed back both positively (through latent heat release) and negatively (through shading and evaporative cold pools) on the thermal circulations giving rise to them. As an obvious example, Fig. 3 suggests a strong negative feedback on the thermal forcing at FWL once cumuli develop and locally block solar insolation. To quantify such effects we perform four additional simulations that are identical to REF except for changes to the microphysics and radiation schemes: NORN shuts off the autoconversion from clouds to rain in the microphysics scheme (eliminating precipitation), NOCOD shuts off the cloud effects on optical depth in the shortwave radiation scheme (eliminating cloud shading), NORNCOD eliminates both above effects, and NOMP shuts off the microphysics scheme (eliminating clouds entirely).
Relative to the REF case, the convection appears to be suppressed to the east of the ridge in the NORN and NORNCOD cases and intensified in the NOCOD case (Figs. 12g–i). The latter arises from a stronger island thermal anomaly, and a correspondingly stronger thermal circulation, in the absence of cloud shading. The former likely stems from the elimination of subcloud cold pools, which tend to initiate secondary cumuli away from the ridge crest. Because the thermal circulation and convection in the rain-free NORN and NORNCOD cases are stronger and more focused directly over the ridge, they create more organized and stronger compensating subsidence over the ridge flanks, which effectively suppresses convection there. As for the NOMP case, its flow appears broadly similar to the REF case despite the absence of clouds (Fig. 12j).
The partial to complete elimination of negative cloud feedbacks in the NORN, NOCOD, and NORNCOD cases causes
Cloud shading, precipitation, and latent-heat release all have similar-magnitude impacts on
5. Heat-engine theory
Recent studies have used thermodynamic heat-engine theory to quantify the strength of mountain thermal circulations in highly idealized simulations (Tian and Parker 2003; Kirshbaum 2013; Kirshbaum and Wang 2014). To evaluate whether such theory holds for the more realistic simulations considered in this study, we apply it to a subset of our experiments: the REF and THERM cases from section 4a, the latter of which exhibits slightly larger
To diagnose the variables in (2) from the simulations, we create 40 parallel transects crossing the central portion of the ridge axis, spaced at 1-km intervals with an along-transect grid spacing of Δ (Fig. 17). For reference, we define the along- and cross-transect coordinates as
Figure 18 shows time series of
6. Conclusions
We have performed an observational and numerical investigation of a thermally forced cumulus convection event over the mountainous Caribbean island of Dominica during the Dominica Experiment (DOMEX) field campaign. Surface meteorological data from a mountain weather station at Freshwater Lake (FWL) and visible satellite observations revealed a clear diurnal cycle in island thermal forcing and cumulus convection, with strong insolation and a rapid temperature increase (and pressure decrease) in the morning hours. Shallow, lightly precipitating cumuli formed by midday, which gradually weakened in the afternoon. Wyoming King Air (WKA) in situ observations from flight legs on both sides of the ridge indicated flow divergence at flight level (1.8 km), reflecting outflow from the island thermal circulation and cumulus convection. The clouds tended to initiate just downwind of the ridge axis (on its southwestern side) and deepen as they drifted eastward across the ridge.
To better understand the mechanisms and sensitivities of the island convection, we performed cloud-resolving numerical simulations with the WRF Model. Although these quasi-idealized simulations could not represent the full complexity of the observed flows, the higher-resolution runs (with horizontal grid spacings
Sensitivity tests varying the island heating and terrain revealed that, as expected, the cumulus convection was driven by island thermal forcing. Perhaps counterintuitively, however, flattening the terrain led to stronger island-scale mass convergence
Simulations were also conducted to isolate and quantify the feedbacks of cloud shading, rainfall, and latent-heat release on the island-scale horizontal inflow and cumulus convection. While cloud shading and rainfall each reduced
To evaluate the skill of thermodynamic heat-engine theory at predicting the island thermal circulations, we applied the theory to the above comparison between mountainous and flat islands, except that the simulations were run in “dry” mode without cloud microphysics. Consistent with the moist simulations, the island thermal circulation strengthened slightly when the terrain was flattened. The heat-engine theory captured this sensitivity and offered a simple explanation for it: the elevated terrain in the mountain case reduced the mixed-layer depth and thus lowered the circulation’s thermodynamic efficiency. This finding contrasts with recent studies of continental mountains, where taller terrain tended to enhance the circulation strength (Tian and Parker 2003; Kirshbaum and Wang 2014). We hypothesize that this disagreement stems from differences in effective thermal forcing between islands and continental mountains. Whereas islands experience strong thermal forcing regardless of terrain height, continental mountains are the main source of regional baroclinicity and, as a result, more responsible for driving thermal circulations. Further investigation is required to quantify the impacts of orography on island thermal circulations under a broad range of environmental conditions.
Acknowledgments
We thank Ron Smith for sharing the DOMEX dataset, Justin Minder for the tidal pressure correction and Météo France radar analysis, Jeff French and Samuel Haimov for help deciphering the WKA data, and Greg Thompson for insight into the cloud microphysics. Comments from two anonymous reviewers and Christian Barthlott helped to improve the manuscript. Funding was provided by the Natural Science and Engineering Research Council Discovery Grant NSERC/RGPIN 418372-12, Fonds de Recherche Nature et Techonologies (FRQNT) Grant FQRNT NC-171838. Numerical simulations were performed on the Guillimin supercomputer at McGill University, under the auspices of Calcul Québec and Compute Canada.
APPENDIX
Calculating and
REFERENCES
Barthlott, C., and D. J. Kirshbaum, 2013: Sensitivity of deep convection to terrain forcing over Mediterranean islands. Quart. J. Roy. Meteor. Soc., 139, 1762–1779, doi:10.1002/qj.2089.
Bjerknes, J., 1938: Satured-adiabatic ascent of air through dry-adiabatically descending environment. Quart. J. Roy. Meteor. Soc., 64, 325–330.
Bryan, G. H., J. C. Wyngaard, and J. M. Fritsch, 2003: Resolution requirements for the simulation of deep moist convection. Mon. Wea. Rev., 131, 2394–2416, doi:10.1175/1520-0493(2003)131<2394:RRFTSO>2.0.CO;2.
Carbone, R. E., J. W. Wilson, T. D. Keenan, and J. M. Hacker, 2000: Tropical island convection in the absence of significant topography. Part I: Life cycle of diurnally forced convection. Mon. Wea. Rev., 128, 3459–3480, doi:10.1175/1520-0493(2000)128<3459:TICITA>2.0.CO;2.
Crook, N. A., and D. F. Tucker, 2005: Flow over heated terrain. Part I: Linear theory and idealized numerical simulations. Mon. Wea. Rev., 133, 2552–2564, doi:10.1175/MWR2964.1.
Crosman, E. T., and J. D. Horel, 2010: Sea and lake breezes: A review of numerical studies. Bound.-Layer Meteor., 137, 1–29, doi:10.1007/s10546-010-9517-9.
Fovell, R. G., 2005: Convective initiation ahead of the sea-breeze front. Mon. Wea. Rev., 133, 264–278, doi:10.1175/MWR-2852.1.
Galewsky, J., 2008: Orographic clouds in terrain-blocked flows: An idealized modeling study. J. Atmos. Sci., 65, 3460–3478, doi:10.1175/2008JAS2435.1.
Geerts, B., Q. Miao, and J. C. Demko, 2008: Pressure perturbations and upslope flow over a heated, isolated mountain. Mon. Wea. Rev., 136, 4272–4288, doi:10.1175/2008MWR2546.1.
Jiang, Q., 2003: Moist dynamics and orographic precipitation. Tellus, 55A, 301–316, doi:10.1034/j.1600-0870.2003.00025.x.
Kingsmill, D. E., 1995: Convection initiation associated with a sea-breeze front, a gust front, and their collision. Mon. Wea. Rev., 123, 2913–2933, doi:10.1175/1520-0493(1995)123<2913:CIAWAS>2.0.CO;2.
Kirshbaum, D. J., 2013: On thermally forced circulations over heated terrain. J. Atmos. Sci., 70, 1690–1709, doi:10.1175/JAS-D-12-0199.1.
Kirshbaum, D. J., and R. B. Smith, 2009: Orographic precipitation in the tropics: Large-eddy simulations and theory. J. Atmos. Sci., 66, 2559–2578, doi:10.1175/2009JAS2990.1.
Kirshbaum, D. J., and A. L. M. Grant, 2012: Invigoration of cumulus cloud fields by mesoscale ascent. Quart. J. Roy. Meteor. Soc., 138, 2136–2150, doi:10.1002/qj.1954.
Kirshbaum, D. J., and C.-C. Wang, 2014: Boundary layer updrafts driven by airflow over heated terrain. J. Atmos. Sci., 71, 1425–1442, doi:10.1175/JAS-D-13-0287.1.
Kirshbaum, D. J., and J. G. Fairman Jr., 2015: Cloud trails past the Lesser Antilles. Mon. Wea. Rev., 143, 995–1017, doi:10.1175/MWR-D-14-00254.1.
Lean, H. W., P. A. Clark, M. Dixon, N. M. Roberts, A. Fitch, R. Forbes, and C. Halliwell, 2008: Characteristics of high-resolution versions of the Met Office Unified Model for forecasting convection over the United Kingdom. Mon. Wea. Rev., 136, 3408–3424, doi:10.1175/2008MWR2332.1.
Minder, J., R. B. Smith, and A. D. Nugent, 2013: The dynamics of ascent-forced orographic convection in the tropics: Results from Dominica. J. Atmos. Sci., 70, 4067–4088, doi:10.1175/JAS-D-13-016.1.
Nugent, A. D., R. B. Smith, and J. R. Minder, 2014: Wind speed control of tropical orographic convection. J. Atmos. Sci., 71, 2695–2712, doi:10.1175/JAS-D-13-0399.1.
Qian, J.-H., 2008: Why precipitation is mostly concentrated over islands in the Maritime Continent. J. Atmos. Sci., 65, 1428–1441, doi:10.1175/2007JAS2422.1.
Qian, T., C. C. Epifanio, and F. Zhang, 2012: Topographic effects on the tropical land and sea breeze. J. Atmos. Sci., 69, 130–149, doi:10.1175/JAS-D-11-011.1.
Ramage, C. S., and T. A. Schroeder, 1999: Trade wind rainfall atop Mount Waialeale, Kauai. Mon. Wea. Rev., 127, 2217–2226, doi:10.1175/1520-0493(1999)127<2217:TWRAMW>2.0.CO;2.
Reisner, J. M., and P. K. Smolarkiewicz, 1994: Thermally forced low Froude number flow past three-dimensional obstacles. J. Atmos. Sci., 51, 117–133, doi:10.1175/1520-0469(1994)051<0117:TFLFNF>2.0.CO;2.
Rogers, R. R., and M. K. Yau, 1989: A Short Course in Cloud Physics. Butterworth-Heineman, 290 pp.
Schroeder, T. A., 1977: Meteorological analysis of an Oahu flood. Mon. Wea. Rev., 105, 458–468, doi:10.1175/1520-0493(1977)105<0458:MAOAOF>2.0.CO;2.
Smith, R. B., P. Schafer, D. J. Kirshbaum, and E. Regina, 2009: Orographic precipitation in the tropics: Experiments in Dominica. J. Atmos. Sci., 66, 1698–1716, doi:10.1175/2008JAS2920.1.
Smith, R. B., and Coauthors, 2012: Orographic precipitation in the tropics: The Dominica Experiment. Bull. Amer. Meteor. Soc., 93, 1567–1579, doi:10.1175/BAMS-D-11-00194.1.
Smolarkiewicz, P. K., and R. Rotunno, 1989: Low Froude number flow past three-dimensional obstacles. Part I: Baroclinically generated lee vortices. J. Atmos. Sci., 46, 1154–1164, doi:10.1175/1520-0469(1989)046<1154:LFNFPT>2.0.CO;2.
Smolarkiewicz, P. K., R. M. Rasmussen, and T. L. Clark, 1988: On the dynamics of Hawaiian cloud bands: Island forcing. J. Atmos. Sci.,45, 1872–1905, doi:10.1175/1520-0469(1988)045<1872:OTDOHC>2.0.CO;2.
Tian, W. S., and D. J. Parker, 2003: A modeling study and scaling analysis of orographic effects on boundary layer shallow convection. J. Atmos. Sci., 60, 1981–1991, doi:10.1175/1520-0469(2003)060<1981:AMSASA>2.0.CO;2.
Wang, C.-C., 2014: Thermally-driven circulation and convection over a mountainous tropical island. M.S. thesis, Dept. of Atmospheric and Oceanic Sciences, McGill University, 90 pp.
Yang, Y., and Y.-L. Chen, 2008: Effects of terrain heights and sizes on island-scale circulations and rainfall for the island of Hawaii during HaRP. Mon. Wea. Rev., 136, 120–146, doi:10.1175/2007MWR1984.1.
Yang, Y., S.-P. Xie, and J. Hafner, 2008: The thermal wake of Kauai Island: Satellite observations and numerical simulations. J. Climate, 21, 4568–4586, doi:10.1175/2008JCLI1895.1.
Yeh, H.-C., and Y.-L. Chen, 1998: Characteristics of rainfall distributions over Taiwan during the Taiwan Area Mesoscale Experiment (TAMEX). J. Appl. Meteor., 37, 1457–1469, doi:10.1175/1520-0450(1998)037<1457:CORDOT>2.0.CO;2.