TMCnet News

Simulated Convective Invigoration Processes at Trade Wind Cumulus Cold Pool Boundaries [Journal of the Atmospheric Sciences]
[August 08, 2014]

Simulated Convective Invigoration Processes at Trade Wind Cumulus Cold Pool Boundaries [Journal of the Atmospheric Sciences]


(Journal of the Atmospheric Sciences Via Acquire Media NewsEdge) ABSTRACT Observations of precipitating trade wind cumuli show convective invigoration on the downwind side of their cold pools. The authors study convection and cold pools using a nested-Weather Research and Forecasting Model simulation of 19 January 2005-a day from the Rain in Cumulus over the Ocean experiment. The temperature and water vapor mixing ratio drops in simulated cold pools fall within the envelope of observed cases, and the wind enhancement matches observations more closely. Subcloud updrafts downwind and near the cold pool boundary are statistically compared to updrafts further from cold pools. Updrafts near cold pool outflows are moister than the other updrafts and are more likely to originate from overall moister regions. Cold pool-influenced updrafts tend to exceed the other updrafts in vertical velocity and are associated with more cloud liquid water. The strength of circulation within the cold pool boundary is unable to match that because of the low-level environmental wind shear, and the lifted updrafts advect faster than the environmental wind, thereby accessing the ambient environmental moisture converged by cold pool expansion. Cases with higher rain rates correspond to larger cloud cover through the shearing off of the upper-level cloud, consistent with observations. This study suggests that it is the ability of cold pools to lift thermodynamically favorable air that is critical for secondary convection of trade wind cumuli.



1. Introduction Precipitation radar observations, both from space (Short and Nakamura 2000) and collected at the surface during the Rain in Cumulus over the Ocean (RICO) experiment (Rauber et al. 2007; Snodgrass et al. 2009; Nuijens et al. 2009), reveal rain rates exceeding 2 mm h21 for shallow precipitating cumuli in the Caribbean trade wind region. Ship-based observations have associated such precipitation with a decrease in surface specific hu- midity and equivalent potential temperature (Zuidema et al. 2012), confirming the presence of large (40-60 km in diameter) cold pools visible from space as mesoscale arcs organized around cloud-free areas (Fig. 1a). Unlike light ('1 mm day21) precipitation that generally cools and moistens the well-mixed subcloud layer while keeping the equivalent potential temperature ue constant (Nitta and Esbensen 1974; Albrecht 1993), the heavier shallow precipitation appears capable of generating convective downdrafts that originate above cloud base and descend into the subcloud layer (Zuidema et al. 2012), similar to what has been observed for deeper convection (e.g., Zipser 1969; Barnes and Garstang 1982). If the downdrafts are able to impact the surface air, a pool of evaporatively cooled air that is also drier than the surrounding surface environmental air is formed. Analysis of radar re- flectivities collected during RICO find that the leading edges of precipitating shallow cumuli propagate at speeds higher than the mean low-level wind speeds and similar to the estimated propagation speed of the asso- ciated cold pool outflow (Zuidema et al. 2012). Cold pools are therefore considered to be responsible for the observed arc-shaped organizations of precipitating shallow cumuli, with the cold pools invigorating convection at their downwind boundary and suppressing thermals in- side the stable cold pool area.

Cold pools occurring in the trade wind regions are worthy of research interest, not only for their partici- pation within the global hydrological cycle, but also because of their relationship to the global radiative budget through their influence on cloud fraction. Cold pools influence cloud fraction both by representing areas of low surface buoyancy that discourage bottom- up shallow convection and by encouraging secondary convection at the cold pool boundaries. An earlier study on feedbacks of shallow-cloud precipitation upon cloud cover neglecting cold pools suggests light precipitation should reduce cloud cover by removing liquid water (Albrecht 1993). Results from an ensemble of high- resolution simulations of precipitating cumuli, based on RICO soundings, indeed show a slight trend of de- creasing shallow cloud cover with increase of the pre- cipitation rate (vanZanten et al. 2011). RICO analyses of scanning surface-based radar data, however, find a slight positive correlation (Nuijens et al. 2009). The discrepancy can be explained by the shearing off of the upper portions of the more heavily precipitating clouds (4-5 km; Zuidema et al. 2012), compared to the modeled cloud top heights of 2-3 km in vanZanten et al. (2011). More variation in the upper levels of shallow cumulus cloud cover rather than lower levels has also been noted in a longer time series at Barbados (Nuijens et al. 2014). The deeper clouds that are associated with cold pools suggest subcloud processes at the cold pool-precipitation interface are critical for understanding how the necessary cloud depth is achieved that can allow overall cloud cover to vary.


Cold pools are also becoming an area of increasing interest to the large-scale modeling community because their parameterization is suggesting new solutions to old modeling problems. For example, cold pools as a tool for organizing convection according to the scale of the convection provides an approach for overcoming in- adequacies in entrainment assumptions (Mapes and Neale 2011). The parameterization of cold pools is also one approach for improving the representation of the diurnal cycle in precipitation within large-scale models (Rio et al. 2009). Such parameterizations can build on the increasing ability to pursue model simulations pos- sessing both the necessary resolution and domain size to represent cold pool processes (e.g., Seifert and Heus 2013). By comparing high-resolution simulations to ob- servations at similar process-level scales, as will be done here, greater confidence can be gained in the high- resolution simulations themselves.

Several mechanisms have been proposed to explain cold pool invigoration of further convection, for both shallow and deep clouds. One purely thermodynamic mechanism is highlighted within a cloud-resolving modeling study of deep convection, with early-stage subcloud evapora- tion providing the surrounding air with moisture that then enhances the convective available potential energy at the boundaries and thereby encourages convection without any other forcing (Tompkins 2001). This mech- anism for the mesoscale organization of shallow pre- cipitating cumuli is further supported in the large-eddy simulations of Seifert and Heus (2013), with moist rings generated from the evaporation of rain around the cold pool boundaries.

The density-driven outflows can also contribute dy- namically to encourage the convergence of moisture at the cold pool boundary, as mentioned in the large-eddy simulation (LES) study on trade wind cumuli mesoscale organization by Xue et al. (2008). Dynamic lifting by the denser cold pool air occurs as the cold pool air subsides and spreads into the environment at the speed of gravity current, with a propagation speed close to the surface exceeding that of the mean wind. The propagation causes convergence at the leading edge that can forcibly lift air parcels. This can strengthen already-buoyant updrafts, or transport humidity to upper levels, pre- paring for future convection. In addition, dynamic lifting at the cold pool boundary, by producing stronger and wider updrafts, can increase their available kinetic en- ergy and ability to trigger further convection. The in- fluence of such processes on deep convection has been explored as a parameterization within Rio et al. (2009), Grandpeix and Lafore (2010), and Rio et al. (2013). The processes themselves have not yet been investigated in detail for shallow convection.

The influence of the typical wind shear pattern of the trade winds on cold pool-induced updrafts should also be considered. Studies on midlatitude squall lines have noticed that the lifting is strongest when the local cir- culation associated with the vertical variation of the cold pool boundary propagation is balanced by the circulation of low-level environmental wind shear (Rotunno et al. 1988; Moncrieff and Liu 1999; Weisman and Rotunno 2004). When the environmental wind shear exceeds the vertical gradient in the cold pool propagation speed, the updraft tilts downshear of the environmental wind, and when the circulation within the cold pool boundary dominates, the updraft leans against the cold pool boundary. In the RICO cases, although the shallow cu- mulus cold pools are much weaker than the midlatitude deep convection, the influence of shear on the lifting process should remain similar.

It is a challenge to disentangle the coexisting ther- modynamic and dynamic aspects of cold pool effects, but proper attribution of causes holds the promise of improving those convective parameterizations that can be explicitly formulated to depend on cold pool processes (e.g., Bretherton et al. 2004; Mapes and Neale 2011; Rio et al. 2013). Simulation data provide four- dimensional fields that are more comprehensive for understanding the relevant processes than the two- dimensional (time-height) column data fields acquired observationally from a ship. An example can be made using the shipboard data from a motion-compensated 10-mm-wavelength Doppler lidar, a vertically pointing X-band precipitation radar, and 15-m flux tower mete- orological data across the leading edge of a raining cold pool boundary (Fig. 2). The updrafts are located on the downwind side of the major precipitation and are asso- ciated with newly developed clouds, while the down- drafts that produce cooling and drying at the surface are associated with the rain shaft of the convection. The onset of surface precipitation, shown as the yellow tri- angle, coincides with the largest drop in surface tem- perature and water vapor. The in situ observations are intriguing but, lacking Lagrangian tracking and the full three-dimensional temperature, water vapor, velocity, and precipitation fields, they do not provide enough information for distinguishing between thermodynamic and dynamic mechanisms linking cold pools and con- vection. Such intriguing but inadequate pieces of in- formation characterize field observations of spatially inhomogeneous time-varying phenomena, reinforcing the need for complementary modeling studies.

Therefore, this study analyzes realistic, high-resolution model simulation to assess the processes by which shallow cumulus cold pools invigorate further shallow precipitation, motivated by observations of trade wind cumulus cold pools during the RICO experiment (Zuidema et al. 2012). We select 19 January 2005 for the simulation, in part because the day has been previously studied (Abel and Shipway 2007; Snodgrass et al. 2009). Conditions on 19 January reflect the influence of a dis- sipated cold front (Caesar 2005), with northeasterly winds rather than the easterly trades. A frontal bound- ary can explain the observed linear cloud lines (Fig. 1b), though no obvious gradient in temperature or moisture is documented in the surface meteorological measure- ments or the ship soundings. The observed liquid-phase- only precipitation (Fig. 2) is identified as ''undisturbed'' by Nuijens et al. (2009), meaning that the fractional area covered by precipitation deviates less than three stan- dard deviations from the mean of RICO operating pe- riod (from 24 November 2004 to 25 January 2005). The influence of a dissipating cold front arriving from the north is likely a common occurrence for Caribbean trade wind cumuli during the boreal winter, so that 19 January may represent one of several typical conditions for trade wind cumulus.

The simulation uses the Weather Research and Forecasting Model (WRF) (updated version 3.2) with multiple nested domains-a modeling setup that has been previously successfully applied to study continental stratocumulus (Zhu et al. 2010). The nesting technique allows the model to explicitly resolve turbulent-scale processes in the innermost domain of 24km 3 24km at 100-m resolution, while the open lateral boundary con- ditions allow a realistic large-scale forcing imposed on the parent domain of size 972 km 3 972 km to transmit to the innermost domain. As will be shown, by incorporating a sensitivity to the realistic, spatially inhomogeneous, time-varying large-scale forcing, cold pools can develop within the relatively small inner domain of the nested- WRF simulation.

The ability of this simulation to model cold pools can be compared to the experience with large-eddy-scale simulations that typically apply doubly periodic boundary conditions, idealized initial conditions, and prescribed homogeneous large-scale forcing (e.g., Matheou et al. 2011; Seifert and Heus 2013). Such sim- ulations assume that the turbulent atmosphere at the inflow boundary is identical to that at the outflow boundary-an assumption that is more applicable to the relatively homogeneous conditions of marine stratocu- mulus or nonprecipitating shallow cumulus than to strongly precipitating shallow-cumulus conditions. When nevertheless large-eddy-scale simulations of strongly precipitating shallow cumulus are undertaken, a suffi- ciently large domain and sufficiently high resolution of the turbulence processes is required to allow such spa- tially inhomogeneous features as the asymmetric trade wind cold pools to develop. Indeed, Matheou et al. (2011) are not able to reproduce precipitating mesoscale arcs until a resolution of 20 m is achieved within a com- parable domain size of 20 km 3 20 km. Seifert and Heus (2013) also find a significant time delay for the de- velopment of a cold-pool-dominated regime within their larger domain size of 50 km 3 50 km at larger grid spacings (50 and 100 m) compared to a grid spacing of 25 m. Our 1-day hindcast simulation produces convec- tion and cold pools that can be compared to the ob- servations from this particular day, developing an understanding of the modeling strengths and weaknesses for nonhomogeneous conditions.

The specifications of this simulation, and comparisons of simulated convection and cold pools with observa- tions, are presented in section 2.Insection 3,weexamine thermodynamic and dynamic cold pool convective in- vigoration mechanisms. The findings of our analyses are summarized and discussed in section 4.

2. Characteristics of simulated convection and cold pools a. WRF simulation setup The 19 January case is simulated from 0000 UTC 19 January to 0600 UTC 20 January 2005, with only the last 24 h analyzed. National Centers for Environmental Prediction (NCEP) (Final) Operational Global Analy- ses data (FNL) at 18 resolution, available at every 6 h, supply the initial and lateral boundary conditions. The sea surface temperature is prescribed from the NCEP FNL and updates every 6 h. One parent domain (972 km 3 972 km) and four two-way nested domains centered at 188N, 61.78W are configured with nesting ratio of 1:3 (grid spacing of each outer domain is 3 times the grid spacing of its next-level nested domain). The innermost domain covers the track of the Research Vessel Seward Johnson (RVSJ), possessing a size of 24 km 3 24 km with a horizontal spacing of 100 m. The vertical domain extends from the surface to 10 hPa with in total 77 levels, of which 48 levels are below 4 km with a vertical spacing varying from 6.5 to 200 m. We refer to the 3-m first model level as the surface level of this simulation. The 3D Smagorinsky scheme (Smagorinsky 1963)treatsthe subgrid-scale turbulent mixing for domains with 100- and 300-m grid spacing, and the Mellor-Yamada-Janji^c boundary layer scheme (Janji^c 2001) is used for treating the vertical turbulent mixing for the coarse domains. The simulation uses the single-moment Thompson cloud microphysics scheme (Thompson et al. 2008) (the known bug of WRF, v3.2, Thompson scheme is fixed with the updated code), with a total cloud droplet number con- centration Nc 5 100 3 106 m23. Each of the 1441 one- minute simulation outputs from the innermost domain is treated as an individual sample for the statistical analyses.

Since the vertical resolution of NCEP FNL does not resolve the fine vertical structure of cumulus topped marine boundary layer, the thermodynamic profiles from radiosondes launched from the RVSJ every 6 h are assimilated and nudged in the outer domains. Figure 3 shows the average of the ship soundings from 19 January (19RVSJ) and the average of the profiles from the in- nermost model domain (19WRF). These are compared with the average of all soundings launched from RVSJ between 9 and 24 January (allRVSJ) and the average of the two- to six-times-daily soundings from Spanish Point, Barbuda, from 16 December 2004 through 8 January 2005 (SPNT). Spanish Point is located to the south-southwest of the ship. One of the ship radiosondes launched on 19 January penetrates a cloud that reaches above 4 km, causing the 19RVSJ water vapor mixing ratio above 3 km to be significantly higher than the simulated domain-average 19WRF values. The com- posite ship sounding of 19 January also reveals a moister atmosphere column above 1 km than the average of all soundings from the ship, indicating the additional moisture available for this day. The most significant difference is between the 19 January wind profiles and those of the 4-week-mean SPNT profile and 2-week January allRVSJ mean, with the allRVSJ wind profile more sheared than the SPNT profile. The winds are weaker on 19 January compared to allRVSJ, with the wind shear at 2.5-3 km. RICO-mean conditions at Spanish Point, Barbuda, show more easterly winds above 2 km and drier and warmer air above 2.5 km compared to the January-mean ship profile. The simu- lations described in vanZanten et al. (2011) are initiated by the mean soundings from SPNT, producing cloud tops that extend up to 2.5 km, with weak precipitation and no cold pools.

b. Comparisons to observed convection and cold pool air properties The simulated surface air temperature reveals cold pools embedded within the larger-scale defined cloud line (Fig. 4a). Three shallow precipitating convection cases along with associated surface cold pools are documented within these cloud lines by the RVSJ flux tower. The flux tower measurements are gathered ap- proximately 15 m above sea level with more detailed information regarding the instruments and data pro- cessing available in the appendix of Zuidema et al. (2012). The average and standard deviation of the 24-h flux tower measurements are compared to the 24-h simulation data from all grid points for air tempera- ture, water vapor mixing ratio, and wind speed at 13-m level, as well as surface rain rate, sensible heat flux (SHF), and latent heat flux (LHF) (Table 1). The mean simulated air temperatures and water vapor mixing ratio match the average shipboard values well. The ship mean wind speeds are slightly higher than those of the WRF domain, yet the ship mean turbulent fluxes are less, implying higher transfer coefficients are used within the model. The relatively higher variability of the simulated wind speed is thought to reflect its variability across the domain, whereas the ship mostly remained within the linear cloud line, reporting mostly north-northeasterly winds (e.g., Fig. 1b, with a wind speed of 4.7 m s21 at 1606 UTC).

The 24-h domain-averaged simulated surface rainfall rate is 2.1 mm day21, which is relatively close to the daily area-averaged rainfall rate 1.87 mm day21 derived from scanning precipitation radar reflectivities for this day (Snodgrass et al. 2009). The domain-averaged nonzero surface rain rate (RRa), is calculated for each output minute by averaging the surface rain rate of all grid points with rain rates of 0.1 mm h21 or more, and RRa is set to zero when no such rain rates are present. RRa ranges up to 10 mm h21, with 95% of the values ex- ceeding 0 and below 6 mm h21 (Fig. 5a). Rainier time periods were also cloudier, as shown in Fig. 5b. The cloud cover is defined as the fraction of grid columns that contain cloud at any level, and the mean cloud cover of all 1441 output minutes is about 0.13. This is compa- rable to the cloud cover reported by Matheou et al. (2011) using a grid spacing of 40 m. The average cloud cover of the top 10th percentile of RRa is 0.05 higher than the average of the remaining output minutes (Fig. 5b). The cloud cover increases with higher RRa for a correlation coefficient of 0.47.

The simulated surface temperature and cloud water path in Fig. 4a reveal the circular cold pools embedded within the linearly oriented cloud lines. The colder re- gions mostly correspond to drier air, as seen by the anomalous water vapor mixing ratio (Fig. 4b). The ap- parent change in surface wind can be spotted at some of these cold and dry boundaries (Fig. 4b). The simulated clouds also extend above 3 km (Fig. 6)-similar to the observations (Fig. 2). For output minutes with RRa . 0, the averaged vertical profile of cloud fraction shows three local peaks: at the cloud base (500 m), 1.6 km, and the level right below 3 km (Fig. 6a). The two upper peaks coincide with layers of slight enhanced wind shear (Figs. 3c,d). The buoyant cloud-core fraction, defined as the portion of cloudy grid cells (qc . 0.1 g kg21) that are positively buoyant relative to the domain mean in- creasingly deviate from the averaged cloud fraction up to 1.6 km (Fig. 6a) as environmental air is mixed into the cloud. The clouds that do reach up to 2 km provide the peak at 2 km in the averaged cloud and rainwater mixing ratio (Figs. 6b,d), with drier conditions aloft.

One assessment of the simulated cold pools is shown in Fig. 7 as a composite of all 1441 output minutes centered upon each output's most intense instant rain center and the same composite for only those output minutes associated with the upper 10th percentile of RRa. For each of these output minutes, we first locate the grid point containing the domain-maximum surface rain rate, then consecutively sample the nearby grid points that align with the domain-averaged surface wind vector within the same moment in time, and average these surface air properties from the selected output minutes. The negative distance on the x axis corresponds to the downwind (southwestern) side and positive dis- tance to the upwind (northeastern) side of the domain- maximum surface rain rate. Figure 7 also shows the average of the three observed cold pool passages from this day. These correspond to maximum surface rain rates of 7.4, 7.5, and 45mm h21. The ship moves in the same direction as the surface wind during the three cold pool passages, meaning that the ship samples its cold pool for longer than if it were stationary.

Figure 7 clearly shows that the simulated cold pools are, like the observations, asymmetric along the mean wind direction. The temperature change is steeper on the downwind side of cold pool, which is where the rain occurs. The wind increases ahead of the rain, enhancing the surface fluxes, and indicating the cold pool outflow. The mean rain rates in the simulation match observed values, while the average cold pool is obviously weaker than observed. The temperature depression is at best 0.5 K and the change in surface water vapor mixing ratio is negligible. The cold pools corresponding to the upper 10% of the RRa are significantly stronger than the av- erage but still weaker than the observations. To some extent, the composite may reflect the process of aver- aging, but, nevertheless, the weak simulated cold pools also reflect on the ability of the simulation to produce cold pools. The 19 January observation captures only the change in wind direction at the cold pool boundary rather than wind speed [see Fig. 12g in Zuidema et al. (2012)], which we attribute to ship location.

The decreasing of the water vapor mixing ratio qy and ue with u in simulated cold pool cases are consistent with the observed cases (Figs. 8a,b). The wind changes in the simulation are stronger for the same change in u than in the available observations, resulting in slightly larger surface fluxes (Figs. 8c,d). The simulated cold pools may be weaker than those observed in their thermodynamic properties, but the gustiness changes at the cold pool boundaries and the resulting convergence appear ade- quately captured.

c. Terminology and statistics of simulated cold pools Since the objective of this study is to examine the ability of precipitation-induced cold pools to invigorate further convection at the downwind boundary, a neces- sary first step is to articulate how arc-shaped cold pool downwind boundaries and their influences are identified within the simulation. Strict criteria are imposed to appropriately identify precipitation-driven cold pool downwind boundaries. First, the cold pools' characters as density currents are preserved through a criterion based on virtual potential temperature uy. Second, a negative anomaly of uy must be associated with the cold pool and not represent a preexisting condition. The horizontal uy anomaly is calculated at each individual levelbelowcloudbase[Duy(x, y, z) 5uy(x, y, z)2 uy(z)], and the horizontal anomalies of other scalars Dx are calculated in the same manner. For output minute t, the cold pool downwind boundary grid points must be less buoyant than the domain mean [Duy(x, y, z)jt , 0] and must have been equally or more buoyant than the domain mean within the previous minute [Duy(x, y, z)jt21 $ 0]. Third, the grid point wind speed must be higher than the speed of domain-averaged wind [U(x, y, z) . U (z)].

In addition, a connection of the cold pool downwind boundary to the precipitation and its catchment area are maintained by only selecting grid cells that already sat- isfy the conditions above and are in the proximity of significant rain. A grid point is considered near signifi- cant rain if the rectangle horizontal area of 6 km 3 6km centered on the grid point contains precipitating points and the average surface rain rate of these points are greater than 2 mm h21. The rain-rate threshold and the size of the precipitation catchment area are reasonable if subjective choices. The 6-km-spatial-scale estimate is shown adequate to encompass most of the related rain catchment area (including the top 10% of strongest rain events) (Fig. 7) but still restrictive enough to exclude other irrelevant rainy points. The 2 mm h21 catchment- area-average threshold has been identified previously as a way to distinguish cold pools that are dried by convective downdrafts from those that are moistened (Barnes and Garstang 1982). A catchment-area criterion also excludes those boundaries of cold pools that are dissipating with the decaying of convection. Based on these conditions, the cold pool downwind boundary can be automatically identified as shown in the example of Fig. 9. Overall, the combined criteria provide us with confidence that we are examining the subset of cold pool boundaries that are actively involved with promoting secondary convection. Of the 1441 output minutes, 620 contain at least one cold pool boundary satisfying this criterion-or, 43% of the total. The depth of the cold pool boundaries hcpb is estimated by the buoyancy of grid cells within the columns above the 3-m-level cold pool boundary points. For each output minute, the hcpb is the level where the number of negative Duy points within the columns becomes less than 10% of the 3-m-level cold pool boundary points.

Based on these definitions, about 91% of the cold pool boundaries have hcpb of 80 m or greater (Fig. 10), cor- responding to 565 output minutes. The negative Duy extends through the whole subcloud layer for 31% of the boundaries, indicating deeper cold pools than Seifert and Heus (2013). The length along the 3-m-level cold pool downwind boundary arc Lcpb are below 20 km for all applicable output minutes, with 144 output minutes greater than 6.8 km (equivalent to cold pool diameters of about 12.7 and 4.3 km, respectively), the latter cor- responding roughly to the scale of the highest 10% of rain catchment area in Fig. 7. The smaller Lcpb at the 80-m level relative to the surface level reflects the vertical reduction in cold pool size. Satellite images reveal arc- shaped organization of clouds spanning up to 40-60 km in equivalent diameter during RICO, and the ship- launched soundings identify some cold pools of 200 m in depth (Zuidema et al. 2012).

The difference between the wind averaged within the cold pool downwind boundary and the domain-averaged wind is defined as the expansion rate of cold pool downwind boundary C*. Its magnitude is similar to the propagation speed of a density current that the expan- sion rate depends on the magnitude of negative buoy- ancy within the cold pools and the cold pool depth (Grandpeix and Lafore 2010). For output minutes that have multiple cold pools present in the domain, C* of each output minute is the average rate for all the cold pool downwind boundaries. In the following section, a subscripted number indicates the altitude level at which the C* is estimated.

The simulated updrafts are identified as points with vertical velocity w $ 0.5 m s21. The updrafts that can potentially be influenced by the cold pool outflow by virtue of proximity need to be distinguished from the updrafts that occur thermodynamically regardless of the presence of a cold pool. This is done by delineating the area that encounters the propagating cold pool downwind boundary. A ''cold pool ambient region'' (CPAR) is defined as the area within 1 km downwind of the cold pool boundary at the 80-m level (with the cold pool boundary not included). Almost all (560 of the 565) output minutes with 80-m-level cold pool boundaries contain updrafts within 1 km of the boundary. Another distance choice of 1.5 km does not increase the sample size of updrafts within CPAR significantly, indicating the 1-km distance is capturing most of the updrafts. In contrast, a distance threshold of 0.5 km reduces the sample size significantly. Ultimately, while the choice of a 1-km distance is subjective, the choice is also reason- able. The potential reach of spatially inhomogeneous cold pool outflow to the updrafts in their vicinity is a statistical correspondence. More distant updrafts may still be influenced by cold pool density currents, and some updrafts downwind of the cold pool will already be thermodynamically buoyant, but as our results will show, a clear statistical signal of the influence of the upwind cold pool can be inferred based on a 1-km choice for the CPAR.

3. Cold pool effects The cold pool effects on organizing subsequent pre- cipitating shallow convection are examined in this sec- tion. The thermodynamic and dynamic properties of the updrafts within and outside the CPAR are compared to demonstrate the existence of cold pool effects, and the cold pool-affected updrafts are diagnosed thereafter for the mechanisms by which the convection is invigorated.

a. Cold pool effects on low-level updrafts The 80-m-level anomalies of qy, u, and ue from the domain-mean values are averaged over the updraft points within and outside (including the output minutes lacking any cold pools) the CPAR for each applicable output minute (Fig. 11). Not surprisingly, updrafts tend to be moister and warmer than the domain-mean re- gardless of proximity to the cold pool boundary. The moisture content differs markedly between the two populations. The updrafts influenced by cold pool out- flows are generally moister than for the other updrafts (Fig. 11a). When only compared to the updrafts within the same output minutes, the CPAR updrafts are moister than the other updrafts by a mean of 0.07 g kg21 and are also slightly warmer (Fig. 11b). The impact of both the enhanced moisture and warmth can be seen in the distribution of the updraft ue anomaly values from their domain mean in Fig. 11c.

In Fig. 12, the relationships of the overall moisture content of the CPAR relative to the domain mean with the difference between two populations of up- drafts and the cold pool expansion rate are examined. On average, the CPAR updrafts only cover 16% of the CPAR area; CPARs that are uniformly moister are more capable of supporting individual moister updrafts (Fig. 12a). More interesting is the finding that moister CPARs are also associated with faster-expanding cold pools (Fig. 12b), possibly because of the more efficient converging of moisture by the cold pool expansion.

Although many of the cold pool-influenced updrafts are more buoyant than the other updrafts within the same moment (Fig. 11b), enhanced buoyancy is not the main factor affecting the speed of the CPAR updrafts. This is made clear when the vertical velocities of CPAR and non-CPAR updrafts are compared to their buoy- ancy, and then each group of updrafts is divided into their buoyant and nonbuoyant (relative to the domain mean) portion (Fig. 13). Overall, the CPAR updrafts are capable of attaining higher vertical velocities than the non-CPAR updrafts (Fig. 13a). The buoyant portion non-CPAR updraft vertical velocities increase with positive buoyancy (Fig. 13b) and likewise as the stability of the nonbuoyant portion increases (Fig. 13c)-indicating compensation between the buoyant and nonbuoyant portions (Fig. 13a). In contrast, the vertical velocities of buoyant CPAR updrafts do not show an increasing trend with the positive buoyancy (Fig. 13b), and the nonbuoyant portion of CPAR updrafts are closer to neutrally buoyant than the non-CPAR updrafts (Fig. 13c); the latter contribute to the higher buoyancy of CPAR updrafts relative to the non-CPAR updrafts.

As a result of the CPAR updrafts not driven by buoyancy and the non-CPAR updrafts showing little effect of buoyancy on controlling the average updraft vertical velocities, the difference between the vertical velocity of the CPAR and non-CPAR updrafts taken from the same output minutes does not increase with higher buoyancy difference (Fig. 14a). However, we can infer from Fig. 14b that the dynamic lifting by the cold pool outflow contributes to strengthen the CPAR updrafts. The difference in vertical velocity between the CPAR and non-CPAR updrafts is enhanced with the cold pool expansion rate, as compared to the weak neg- ative correlation of the relationship shown in Fig. 14a.

The cold pool expansion strengthens the updrafts by lifting air parcels preferably with high ue. The example of a cold pool occurring between 2105 and 2123 UTC provided in Fig. 9 helps to visualize these processes. In Fig. 9, the majority of the CPAR updrafts coincide with a high u e anomaly ( D u e . 2 K) as well as a high q y anomaly (Dqy . 0.7 g kg21). A moisture patch located about 1-2 km to the downwind (southwestern) side of the cold pool boundary at 2105 UTC eventually con- verges with the cold pool boundary, becoming the locus for further updrafts.

b. Impact of ambient wind shear on the lifted updrafts The relationship between the cold pool boundary and the ambient vertical wind shear can influence the ori- entation of the lifted updrafts and thereby the ability to further propagate convection (e.g., Liu and Moncrieff 1996; Weisman and Rotunno 2004). Figure 15 shows three possible scenarios of the relationship. The mean cold pool ambient horizontal wind U, which resembles the mean horizontal wind in this simulation, increases with height up to the cloud-base level, with a difference dU between vertical level z and z 1 dz of dU 5 U(z 1 dz) 2 U(z). By the definition of C*, the mean horizontal wind within the cold pool downwind boundary equals C* 1 U, and dC* 5 C*(z 1 dz) 2 C*(z). Since the environmental wind below cloud-base level enhances with height (e.g., Fig. 3), assume dU . 0, and cold pool expansion rate decreases with height; dC* and dU are of opposite sign. The sign and magnitude of dU and (dC* 1 dU) determine the direction and strength of the circu- lation associated with the ambient vertical wind shear and the wind shear within the cold pool downwind boundary, respectively. When the two circulations compensate for each other dU ;2(dC* 1 dU) and dC* ;22dU (Fig. 15b). When the ambient wind cir- culation is stronger, dU .2(dC* 1 dU), and jdC*j , j2dUj (Fig. 15a). In addition, dU ,2(dC* 1 dU) and jdC*j . j2dUj are true when the cold pool boundary circulation dominates over that of the ambient wind (Fig. 15c). The force-lifted updrafts follow the direction of the dominant circulation or rise upright when the two circulations counter each other. This, in turn, has an impact on the air feeding farther into the cloud.

The cold pool ambient low-level vertical wind shear is estimated as the difference of mean horizontal wind relative to the 3-m level Usurf, at the 80-m level U80, 170-m level U170, and 300-m level U300, respectively. The ambient vertical wind shear in this simulation closely resembles the domain-mean vertical wind shear. The circulation that may balance the ambient vertical wind shear is associated with the difference in cold pool ex- pansion rate at the 80-m level C*80, 170-m level C*170, and 300-m level C*300 relative to the 3-m level C*surf. The magnitude of C*80,170,300 2 C*surf is smaller than the magnitude of U80,170,300 2 Usurf for nearly all output minutes, with averaged jC*80,170,300 2 C*surfj ' 0.6, 0.9, and 1 m s21 compared to an average ambient wind shear of jU80,170,300 2 Usurfj ' 1.8, 2.2, and 2.3 m s21. Thus, the simulation most closely matches the condition shown in Fig. 15a, so that the lifted updrafts should rise along the downshear of the ambient or mean wind. The scenario may differ in nature, since the simulation underestimates the cold pool strength.

Cross sections 2105-2123 UTC aligned with the domain-averaged surface wind vector illuminate how the vertical structure of the subcloud updraft relates to the cloud and rain field (Figs. 16a-d). At 2105 UTC, the cold pool boundary is still away from some updrafts extending southwestward to about 2 km in Fig. 16a (see also Fig. 9a) that support the convection without the cold pool influence. At 2111 UTC, the surface-based stable cold pool layer diminishes the updrafts, and the updrafts below 300 m along the section are too weak to detect (Figs. 16b and 9b). At 2117 and 2123 UTC, the cross sections are able to capture the rain and strong downdrafts reaching the surface and are expanding the cold pool downwind boundary (Figs. 16c,d). The hori- zontal wind anomalies of updrafts below 300 m change signs from against the mean wind direction at 2105 UTC to along the mean wind and mean wind shear direction (Figs. 16c,d,g,h). The updrafts that advance faster than the mean wind have better access to the ambient envi- ronmental moisture that converged by expanding cold pools. In addition, as evident in Fig. 16, the wind shear above the cloud base tilts the cloud into the wind, al- lowing the precipitation to fall inside the cold pool and thereby strengthen the already-existing cold pool with its evaporative cooling.

c. Cold pool effects at the cloud-base level The updrafts at the cloud-base level are more directly related to the cloud and rain development than the low- level updrafts considered so far. The 450-m level in the simulation is below and most near the average model cloud base (Fig. 6a), and is referred to as the cloud-base level. The thermodynamic and dynamic properties of the cloud-base-level updrafts are analyzed for evidence of cold pool effects. Similar to the updrafts at the 80-m level, the cloud-base-level updrafts are also divided into two groups depending on whether or not they are di- rectly above the 80-m-level CPAR.

The CPAR updrafts retain their high ue anomalies relative to the domain-mean ue from 80 to 450 m, shifting the number density peak from Dueup 5 1:5to2K(Figs. 11c and 17a). The higher Dueup of the CPAR updrafts is mainly due to their higher humidity, but the moister CPAR updrafts are also cooler than the non-CPAR up- drafts (Fig. 17b). The high-Dueup-but-cold-CPAR updrafts are thereby neutrally even slightly less buoyant relative to non-CPAR updrafts (Fig. 17c), in contrast to a slightly positive relative buoyancy at the 80-m level (Fig. 14a).

The scale of vertical velocity differences (CPAR 2 non-CPAR) at cloud-base level is 3 times that at the 80-m level (Fig. 17d versus Fig. 14b), attesting to the impact of the cold pool dynamic lifting. Although the correlation of cloud-base-level vertical velocities difference with the 80-m-level cold pool expansion rate is even weaker than the 80-m-level vertical velocity difference, attributed to the weaker correlation between the CPAR updraft vertical velocities at the cloud-base and 80-m levels (r 5 0.3) as compared to the non-CPAR updrafts (r 5 0.7). Possible interpretations include that the lifting force that depends on the cold pool strength varies over time with the precipitation fall upon the cold pool, and more mixing with environmental air takes place when air is dynamically lifted from 80 m to the cloud-base level under the influence of wind shear interaction between the cold pool boundary and the ambient winds.

The cloud-base-level updraft coverage may also be affected by the presence of cold pool boundary in the domain. A Due distribution for the CPAR cloud-base- level updrafts that resembles that for the non-CPAR updrafts (Fig. 17a) can be achieved by removing some current CPAR updrafts that contribute to the excessive number density relative to the non-CPAR updrafts for higher Due values, which include around 50% of the current CPAR updrafts. This estimation suggests that cold pool lifting approximately doubles the updraft cov- erage within CPAR at cloud-base level. However, this contribution to total updraft coverage is very small, since the CPAR updrafts only make up around 3.3% of the total cloud-base-level updrafts.

Nevertheless, the CPAR updrafts are capable of generating more cloud than the non-CPAR updrafts. The average cloud water path (CWP) associated with CPAR updrafts is 254gm22, compared to 130gm22for the non-CPAR updrafts. The CWP associated with non- CPAR updrafts depends strongly on the cloud-base- level vertical velocity (Fig. 18a), whereas the CPAR CWPup values are higher in the mean for the same vertical velocity, if also more variable. The CWPup dif- ferences (CPAR 2 non-CPAR) increase from about 2200 to 400 g m22 as the relative vertical velocity in- creases from about 20.5 to 0.6 m s21 (Fig. 18b). A sim- ilar relationship with CWP does not occur for u or humidity difference (not shown). Cold pools affect cloud growth by altering the thermodynamic properties of the updrafts to include more humid, cooler air (Fig. 17b) and by enhancing the updraft speed; Fig. 18b suggests the latter is particularly important for encouraging deeper, high-liquid water path clouds.

4. Conclusions and discussion This study attempts to identify the dominant mecha- nisms supporting the updrafts generating precipitation by shallow trade wind cumuli organized in mesoscale arcs. The mechanisms that are considered include 1) thermo- dynamically enhanced updrafts due to either environ- mental moisture from outside the cold pool or moisture generated by the evaporation of rain inside the cold pool (Tompkins 2001) and 2) forced lifting by the expanding cold pool. These mechanisms are not completely in- dependent of each other, since, for example, both mech- anisms may be implicated when a cold pool boundary moving faster than the mean wind converges and further lifts high-ue air, as sketched in Fig. 19. The individual mechanisms are examined through statistical compari- sons of the updrafts within and outside of the regions near the cold pool boundaries, as is the time evolution of a selected case study.

The analyses reveal that the updrafts in the proximity of the downwind cold pool boundary are typically moister than the other updrafts. The higher humidities generate higher ue for the cold pool-influenced updrafts. The temperatures of the CPAR updrafts are slightly warmer than those of the non-CPAR updrafts at the lower subcloud layer but are cooler at the cloud-base level. In contrast to non-CPAR updrafts that rise by their thermodynamic privilege or dynamic lifting along the larger-scale defined convergence lines (Fig. 1b), the CPAR updrafts are more subject to local dynamic pro- cesses. The updrafts near the cold pool boundaries generate more cloud water on average than the updrafts away from cold pools. The cold pool boundary collects environmental moisture as it propagates into new en- vironments, causing moisture convergence close to the boundary. The moisture excess of the updrafts within CPARs correlates well with an overall positive moisture anomaly for the CPARs. The cold pool-lifted air parcels originate from the moist air within CPAR and con- tribute to the broader ue range for the CPAR updrafts. These results speak to a preferential sampling of pre- existing environmental moisture pools that may well reflect the remnants of other convective events, but little evidence is found to support the hypothesis that updrafts at the cold pool boundary arise from air premoistened by earlier subcloud evaporation from the same convection (Tompkins 2001). Instead, these results speak more to the ability of secondary convection to thrive and continue when propagating into relatively moist environments.

The strengthening of CPAR updrafts by cold pool dynamic lifting is evident through their enhanced ver- tical velocities at the 80-m level, as compared to the non- CPAR updrafts. The dynamic lifting by the cold pool outflow slightly increases the area covered by the cloud- base-level updrafts by introducing more air parcels that may not rise otherwise. Since these parcels tend to be moist, they contribute to the higher cloud liquid water paths associated with the CPAR updrafts compared to non-CPAR updrafts.

A parameterization for cold pool dynamic lifting in- troduced by Grandpeix and Lafore (2010) and Rio et al. (2009, 2013) is based on the idea that greater expansion rates and deeper and larger cold pools favorably induce higher updraft mass flux at the cold pool downwind boundary. Compared to midlatitudes, the tropics pro- vide abundant moisture that is available for the cold pool-enhanced updrafts. The cold pool effects on en- hancing the updraft coverage and vertical velocity require consideration in GCM parameterizations of trade wind cumulus. However, such parameterizations would need to be modified to account for the moderate depth and strength of shallow cumulus cold pools, as well as focusing on parameterizing the updraft proper- ties and associated cloud water path rather than those of the cloud-base mass flux.

This study relies on a nested-WRF simulation with an innermost domain size of 24km 3 24km and a horizontal grid spacing of 100 m and vertical resolution of 48 levels below 4 km. The nesting technique provides open lateral boundary conditions that allow the reanalysis-derived large-scale forcing imposed on the parent domain of 972 km 3 972 km to be transmitted to the inner- most domain capable of resolving large-eddy-scale circulations. This simulation is able to produce cold pools, as observed for this day, although with weaker changes in surface air properties for rain rates that often exceed those observed. This ability to model cold pools can be compared with the behavior of large-eddy-scale simulations that typically apply doubly periodic boundary conditions and idealized, homogeneous initial condi- tions and forcings. Such simulations require larger do- mains at a higher resolution to be able to capture the spatially inhomogeneous, asymmetric trade wind cu- mulus cold pools (e.g., Matheou et al. 2011; Seifert and Heus 2013). Instead, this study's modeling approach demonstrates that the ability to explicitly simulate a spec- trum of scales-from the large-scale flow, to the meso- scale, and further down to the largest of the turbulent eddies-is also effective for the simulation of cold pools. This connection between the synoptic scale and the mesoscale has typically been ignored for trade wind cumulus but likely represents a common interaction for the trade wind region with the midlatitudes, of which 19 January 2005 is an example.

Wind shear is arguably overly efficient at shearing off upper-level cloud in the simulations compared with the observations (Fig. 16) and this may suggest that the up- draft strength is also weaker than in nature. Nevertheless, the increase in cloud fraction with domain-averaged rain rates is in agreement with observations. The weak sim- ulated cold pools on average could reflect issues with the spatial resolution, microphysical parameterization, and turbulent mixing. A reduced grid spacing would allow more resolution of the turbulent-scale processes and amplify the downdrafts and updrafts for forming the cold pools and invigorating convection. This in turn may also affect the relationship of the cold pool circulation with the low-level environmental wind shear. In this simulation, the low-level environmental circulation is consistently stronger than the circulation of the cold pools, and in the case study examination of updrafts lifted by a strong cold pool, the lifted updrafts point downshear, as would be expected.

The trade wind cumulus cold pools are associated with stronger winds in the lower boundary layer, which in turn will generate stronger surface fluxes [Fig. 8;see Zuidema et al. (2012) for a more complete analysis of the observed surface fluxes]. The cold pools provide a mechanism for generating the stronger updraft ve- locities that can also deepen the boundary layer. As suchthecoldpoolsmayhelpstabilizetherelativehu- midity in the trade wind regions by encouraging the drying and warming of the boundary layer through entrainment of drier, warmer air from aloft. This con- trasts with the cooler, moister cold pools that have been observed for stratocumulus (Terai and Wood 2013), wherein increased stability traps the moisture in- troduced by the latent heat fluxes off the ocean near the surface.

Possible reasons for the weak cold pools in this and other simulations despite variations in the imposed boundary conditions include numerical diffusion and grid resolution (Matheou et al. 2011), as well as the microphysical parameterization (Seifert and Heus 2013). This simulation uses the bulk microphysics scheme of Thompson et al. (2008), which employs a single- moment parameterization for rain. Although less ex- plicit than double-moment schemes, the diagnosis of the intercept parameter from the rain mixing ratio enables the Thompson scheme to generate a large number concentration of small raindrops when the rain mixing ratio is low. Because small drops evaporate more read- ily, this should favor more evaporation, leading to stronger convective downdrafts, than if the intercept parameter were a fixed value. This, in fact, is evident in a one-dimensional model intercomparison where the Thompson scheme produces the largest rain evap- oration rates of the microphysics schemes assessed for similar environmental moisture (Shipway and Hill 2012). The modeled surface rain rates are larger than observed, and the use of assimilated soundings help ensure that the subcloud environmental moisture is realistic, so that it seems unlikely that the simulated rain evaporation rates are less than occurred in na- ture. This does not rule out the importance of turbu- lent mixing in changing the properties of the downdrafts and updrafts. Further work will assess the influence of choice of microphysical parameterization more completely.

This study focuses on one day only (19 January 2005). The convection generated for this day is as- sociated with a dissipating cold front, but the asso- ciated cold pool strength falls within the envelope of values from more truly undisturbed days (Fig. 8). Further comparative modeling studies can help as- sess the representativeness of these results for days possessing cold pools within other typical trade wind conditions.

Acknowledgments. We thank the NSF Physical and Dynamic Meteorology Division and Program Man- ager Brad Smull for support through Grant AGS- 1114521 and NOAA Climate Project Office Grant NA130AR4310157.

Denotes Open Access content.

REFERENCES Abel, S. J., and B. J. Shipway, 2007: A comparison of cloud- resolving model simulations of trade wind cumulus with air- craft observations taken during RICO. Quart. J. Roy. Meteor. Soc., 133, 781-794, doi:10.1002/qj.55.

Albrecht, B. A., 1993: Effects of precipitation on the thermody- namic structure of the trade wind boundary layer. J. Geophys. Res., 98, 7327-7337, doi:10.1029/93JD00027.

Barnes, G. M., and M. Garstang, 1982: Subcloud layer energetics of precipitating convection. Mon. Wea. Rev., 110, 102-117, doi:10.1175/1520-0493(1982)110,0102:SLEOPC.2.0.CO;2.

Bretherton, C. S., J. R. McCaa, and H. Grenier, 2004: A new parameterization for shallow cumulus convection and its application to marine subtropical cloud-topped boundary layers. Part I: Description and 1D results. Mon. Wea. Rev., 132, 864-882, doi:10.1175/1520-0493(2004)132,0864: ANPFSC.2.0.CO;2.

Caesar, K.-A. L., 2005: Summary of the weather during the RICO project. UCAR Science Summary, 30 pp. [Available online at https://www.eol.ucar.edu/system/files/kathy_caesar.pdf.] Grandpeix, J.-Y., and J.-P. Lafore, 2010: A density current pa- rameterization coupled with Emanuel's convection scheme. Part I: The models. J. Atmos. Sci., 67, 881-897, doi:10.1175/ 2009JAS3044.1.

Janji^c, Z. I., 2001: Nonsingular implementation of the Mellor- Yamada Level 2.5 scheme in the NCEP Meso model. NCEP Office Note 437, 61 pp. [Available online at http:// www.emc.ncep.noaa.gov/officenotes/newernotes/on437.pdf.] Liu, C., and M. W. Moncrieff, 1996: A numerical study of the ef- fects of ambient flow and shear on density currents. Mon. Wea. Rev., 124, 2282-2303, doi:10.1175/1520-0493(1996)124,2282: ANSOTE.2.0.CO;2.

Mapes, B., and R. Neale, 2011: Parameterizing convective organization to escape the entrainment dilemma. J. Adv Model. Earth Syst., 3, M06004, doi:10.1029/2011MS000042.

Matheou, G., D. Chung, L. Nuijens, B. Stevens, and J. Teixeira, 2011: On the fidelity of large-eddy simulation of shallow pre- cipitating cumulus convection. Mon. Wea. Rev., 139, 2918- 2939, doi:10.1175/2011MWR3599.1.

Moncrieff, M. W., and C. Liu, 1999: Convection initiation by density currents: Role of convergence, shear, and dynamical organization. Mon. Wea. Rev., 127, 2455-2464, doi:10.1175/ 1520-0493(1999)127,2455:CIBDCR.2.0.CO;2.

Nitta, T., and S. Esbensen, 1974: Heat and moisture budget analyses using BOMEX data. Mon. Wea. Rev., 102, 17-28, doi:10.1175/ 1520-0493(1974)102,0017:HAMBAU.2.0.CO;2.

Nuijens, L., B. Stevens, and A. P. Siebesma, 2009: The environment of precipitating shallow cumulus convection. J. Atmos. Sci., 66, 1962-1979, doi:10.1175/2008JAS2841.1.

-, I. Serikov, L. Hirsch, K. Lonitz, and B. Stevens, 2014: The distribution and variability of low-level cloud in the North Atlantic trades. Quart. J. Roy. Meteor. Soc., doi:10.1002/ qj.2307, in press.

Rauber, R. M., and Coauthors, 2007: Rain in shallow cumulus over the ocean: The RICO campaign. Bull. Amer. Meteor. Soc., 88, 1912-1928, doi:10.1175/BAMS-88-12-1912.

Rio, C., F. Hourdin, J.-Y. Grandpeix, and J.-P. Lafore, 2009: Shifting the diurnal cycle of parameterized deep convection over land. Geophys. Res. Lett., 36, L07809, doi:10.1029/ 2008GL036779.

-, and Coauthors, 2013: Control of deep convection by sub- cloud lifting processes: The ALP closure in the LMDZ5B general circulation model. Climate Dyn., 40, 2271-2292, doi:10.1007/s00382-012-1506-x.

Rotunno,R.,J.B.Klemp,andM.L.Weisman,1988:A theory for strong, long-lived squall lines. J. Atmos. Sci., 45, 463-485, doi:10.1175/1520-0469(1988)045,0463: ATFSLL.2.0.CO;2.

Seifert, A., and T. Heus, 2013: Large-eddy simulation of organized precipitating trade wind cumulus clouds. Atmos. Chem. Phys., 13, 5631-5645, doi:10.5194/acp-13-5631-2013.

Shipway, B. J., and A. A. Hill, 2012: Diagnosis of system- atic differences between multiple parameterizations of warm rain microphysics using a kinematic framework. Quart. J. Roy. Meteor. Soc., 138, 2196-2211, doi:10.1002/ qj.1913.

Short, D. A., and K. Nakamura, 2000: TRMM radar observa- tions of shallow precipitation over the tropical oceans. J. Cli- mate, 13, 4107-4124, doi:10.1175/1520-0442(2000)013,4107: TROOSP.2.0.CO;2.

Smagorinsky, J., 1963: General circulation experiments with the primitive equations. Mon. Wea. Rev., 91, 99-164, doi:10.1175/ 1520-0493(1963)091,0099:GCEWTP.2.3.CO;2.

Snodgrass, E. R., L. Di Girolamo, and R. M. Rauber, 2009: Pre- cipitation characteristics of trade wind clouds during RICO derived from radar, satellite, and aircraft measurements. J. Appl. Meteor. Climatol., 48, 464-483, doi:10.1175/ 2008JAMC1946.1.

Terai, C. R., and R. Wood, 2013: Aircraft observations of cold pools under marine stratocumulus. Atmos. Chem. Phys., 13, 9899-9914, doi:10.5194/acp-13-9899-2013.

Thompson, G., P. R. Field, R. M. Rasmussen, and W. D. Hall, 2008: Explicit forecasts of winter precipitation using an improved bulk microphysics scheme. Part II: Implementation of a new snow parameterization. Mon. Wea. Rev., 136, 5095-5115, doi:10.1175/2008MWR2387.1.

Tompkins, A. M., 2001: Organization of tropical convection in low vertical wind shears: The role of cold pools. J. Atmos. Sci., 58, 1650-1672, doi:10.1175/1520-0469(2001)058,1650: OOTCIL.2.0.CO;2.

vanZanten, M. C., and Coauthors, 2011: Controls on precipitation and cloudiness in simulations of trade-wind cumulus as observed during RICO. J. Adv. Model. Earth Syst., 3, M06001, doi:10.1029/2011MS000056.

Weisman, M. L., and R. Rotunno, 2004: ''A theory for strong long-lived squall lines'' revisited. J. Atmos. Sci., 61, 361-382, doi:10.1175/1520-0469(2004)061,0361:ATFSLS.2.0.CO;2.

Xue, H., G. Feingold, and B. Stevens, 2008: Aerosol effects on clouds, precipitation, and the organization of shallow cu- mulus convection. J. Atmos. Sci., 65, 392-406, doi:10.1175/ 2007JAS2428.1.

Zhu, P., B. A. Albrecht, V. P. Ghate, and Z. Zhu, 2010: Multipole- scale simulations of stratocumulus clouds. J. Geophys. Res., 115, D23201, doi:10.1029/2010JD014400.

Zipser, E. J., 1969: The role of organized unsaturated convective downdrafts in the structure and rapid decay of an equatorial disturbance. J. Appl. Meteor., 8, 799-814, doi:10.1175/ 1520-0450(1969)008,0799:TROOUC.2.0.CO;2.

Zuidema, P., and Coauthors, 2012: On trade wind cumulus cold pools.J.Atmos.Sci., 69, 258-280, doi:10.1175/JAS-D-11-0143.1.

ZHUJUN LI AND PAQUITA ZUIDEMA Rosenstiel School of Marine and Atmospheric Sciences, University of Miami, Miami, Florida PING ZHU Department of Earth and Environment, Florida International University, Miami, Florida (Manuscript received 19 June 2013, in final form 27 March 2014) Corresponding author address: Zhujun Li, RSMAS/MPO, 4600 Rickenbacker Cswy., Miami, FL 33149.

E-mail: [email protected] (c) 2014 American Meteorological Society

[ Back To TMCnet.com's Homepage ]