Usama Anber 1, Shuguang Wang 2, and Adam Sobel 1,2,3



Similar documents
Cloud-resolving simulation of TOGA-COARE using parameterized largescale

Description of zero-buoyancy entraining plume model

Trimodal cloudiness and tropical stable layers in simulations of radiative convective equilibrium

Evalua&ng Downdra/ Parameteriza&ons with High Resolu&on CRM Data

MICROPHYSICS COMPLEXITY EFFECTS ON STORM EVOLUTION AND ELECTRIFICATION

Cloud-Resolving Simulations of Convection during DYNAMO

Using Cloud-Resolving Model Simulations of Deep Convection to Inform Cloud Parameterizations in Large-Scale Models

How To Model An Ac Cloud

Diurnal Cycle of Convection at the ARM SGP Site: Role of Large-Scale Forcing, Surface Fluxes, and Convective Inhibition

CRM simula+ons with parameterized large- scale dynamics using +me- dependent forcings from observa+ons

Cloud-resolving simulation of TOGA-COARE using parameterized largescale

Frank and Charles Cohen Department of Meteorology The Pennsylvania State University University Park, PA, U.S.A.

The effects of organization on convective and large-scale interactions using cloud resolving simulations with parameterized large-scale dynamics

A Review on the Uses of Cloud-(System-)Resolving Models

Limitations of Equilibrium Or: What if τ LS τ adj?

Large Eddy Simulation (LES) & Cloud Resolving Model (CRM) Françoise Guichard and Fleur Couvreux

GCMs with Implicit and Explicit cloudrain processes for simulation of extreme precipitation frequency

Mass flux fluctuation in a cloud resolving simulation with diurnal forcing

Parameterization of Cumulus Convective Cloud Systems in Mesoscale Forecast Models

Unified Cloud and Mixing Parameterizations of the Marine Boundary Layer: EDMF and PDF-based cloud approaches

Comparison of the Vertical Velocity used to Calculate the Cloud Droplet Number Concentration in a Cloud-Resolving and a Global Climate Model

Developing Continuous SCM/CRM Forcing Using NWP Products Constrained by ARM Observations

Month-Long 2D Cloud-Resolving Model Simulation and Resultant Statistics of Cloud Systems Over the ARM SGP

The formation of wider and deeper clouds through cold-pool dynamics

A new positive cloud feedback?

The impact of parametrized convection on cloud feedback.

Convective Clouds. Convective clouds 1

An economical scale-aware parameterization for representing subgrid-scale clouds and turbulence in cloud-resolving models and global models

Sensitivity studies of developing convection in a cloud-resolving model

National Center for Atmospheric Research,* Boulder, Colorado V. TATARSKII

Goal: Understand the conditions and causes of tropical cyclogenesis and cyclolysis

Lecture 3. Turbulent fluxes and TKE budgets (Garratt, Ch 2)

Stochastic variability of mass flux in a cloud resolving simulation

Can latent heat release have a negative effect on polar low intensity?

Chapter 7 Stability and Cloud Development. Atmospheric Stability

Fundamentals of Climate Change (PCC 587): Water Vapor

SPOOKIE: The Selected Process On/Off Klima Intercomparison Experiment

Simulation of low clouds from the CAM and the regional WRF with multiple nested resolutions

Turbulent mixing in clouds latent heat and cloud microphysics effects

Assessing the performance of a prognostic and a diagnostic cloud scheme using single column model simulations of TWP ICE

Improving Representation of Turbulence and Clouds In Coarse-Grid CRMs

Stability and Cloud Development. Stability in the atmosphere AT350. Why did this cloud form, whereas the sky was clear 4 hours ago?

Tropical Cloud Population

Convective Systems over the South China Sea: Cloud-Resolving Model Simulations

1D shallow convective case studies and comparisons with LES

What the Heck are Low-Cloud Feedbacks? Takanobu Yamaguchi Rachel R. McCrary Anna B. Harper

How To Write A New Vertical Diffusion Package

Impact of microphysics on cloud-system resolving model simulations of deep convection and SpCAM

High-Resolution Simulation of Shallow-to-Deep Convection Transition over Land

A Mass-Flux Scheme View of a High-Resolution Simulation of a Transition from Shallow to Deep Cumulus Convection

Understanding and Improving CRM and GCM Simulations of Cloud Systems with ARM Observations. Final Report

Simulations of Clouds and Sensitivity Study by Wearther Research and Forecast Model for Atmospheric Radiation Measurement Case 4

Evaluating Clouds in Long-Term Cloud-Resolving Model Simulations with Observational Data

This chapter discusses: 1. Definitions and causes of stable and unstable atmospheric air. 2. Processes that cause instability and cloud development

A Real Case Study of Using Cloud Analysis in Grid-point Statistical Interpolation Analysis and Advanced Research WRF Forecast System

Long-Term Cloud-Resolving Model Simulations of Cloud Properties and Radiative Effects of Cloud Distributions

Physics of the Atmosphere I

Atmospheric Stability & Cloud Development

Storms Short Study Guide

ATMS 310 Jet Streams

Effects of Tropospheric Wind Shear on the Spectrum of Convectively Generated Gravity Waves

7B.2 MODELING THE EFFECT OF VERTICAL WIND SHEAR ON TROPICAL CYCLONE SIZE AND STRUCTURE

Comment on An ensemble cumulus convection parameterisation. with explicit cloud treatment by T. M. Wagner and H.-F. Graf. Robert S.

Convective Vertical Velocities in GFDL AM3, Cloud Resolving Models, and Radar Retrievals

Harvard wet deposition scheme for GMI

CHARACTERISTICS OF CLOUDS AND THE NEAR CLOUD ENVIRONMENT IN A SIMULATION OF TROPICAL CONVECTION

Weak pressure gradient approximation and its analytical solutions

A limited area model (LAM) intercomparison study of a TWP-ICE active monsoon mesoscale convective event

Atmospheric Processes

DETAILED STORM SIMULATIONS BY A NUMERICAL CLOUD MODEL WITH ELECTRIFICATION AND LIGHTNING PARAMETERIZATIONS

RADIATION IN THE TROPICAL ATMOSPHERE and the SAHEL SURFACE HEAT BALANCE. Peter J. Lamb. Cooperative Institute for Mesoscale Meteorological Studies

Observed Cloud Cover Trends and Global Climate Change. Joel Norris Scripps Institution of Oceanography

J4.1 CENTRAL NORTH CAROLINA TORNADOES FROM THE 16 APRIL 2011 OUTBREAK. Matthew Parker* North Carolina State University, Raleigh, North Carolina

Titelmasterformat durch Klicken. bearbeiten

Chapter 6: Cloud Development and Forms

Indian Ocean and Monsoon

Theory of moist convection in statistical equilibrium

Including thermal effects in CFD simulations

IMPACT OF SAINT LOUIS UNIVERSITY-AMERENUE QUANTUM WEATHER PROJECT MESONET DATA ON WRF-ARW FORECASTS

Surface Rainfall Cold Cloud Fractional Coverage Relationship in TOGA COARE: A Function of Vertical Wind Shear

The horizontal diffusion issue in CRM simulations of moist convection

Improving Low-Cloud Simulation with an Upgraded Multiscale Modeling Framework

Investigating mechanisms of cloud feedback differences.

Research Objective 4: Develop improved parameterizations of boundary-layer clouds and turbulence for use in MMFs and GCRMs

The ARM-GCSS Intercomparison Study of Single-Column Models and Cloud System Models

Chapter Overview. Seasons. Earth s Seasons. Distribution of Solar Energy. Solar Energy on Earth. CHAPTER 6 Air-Sea Interaction

Comment on "Observational and model evidence for positive low-level cloud feedback"

Formation & Classification

Convection in water (an almost-incompressible fluid)

ASR CRM Intercomparison Study on Deep Convective Clouds and Aerosol Impacts

Ecosystem-land-surface-BL-cloud coupling as climate changes

Relationship between the Subtropical Anticyclone and Diabatic Heating

Reply to No evidence for iris

39th International Physics Olympiad - Hanoi - Vietnam Theoretical Problem No. 3

Various Implementations of a Statistical Cloud Scheme in COSMO model

STRATEGY & Parametrized Convection

Group Session 1-3 Rain and Cloud Observations

Lecture 4: Pressure and Wind

Project Title: Quantifying Uncertainties of High-Resolution WRF Modeling on Downslope Wind Forecasts in the Las Vegas Valley

Aspects of the parametrization of organized convection: Contrasting cloud-resolving model and single-column model realizations

Transcription:

Response of Atmospheric Convection to Vertical Wind Shear: Cloud Resolving Simulations with Parameterized Large-Scale Circulation. Part I: Specified Radiative Cooling. Usama Anber 1, Shuguang Wang 2, and Adam Sobel 1,2,3 1 Department of Earth and Environmental Sciences, Columbia University. 2 Department of Applied Physics and Applied Mathematics, Columbia University. 3 Lamont Doherty Earth Observatory, Earth Institute of Columbia University

1 2 3 4 5 6 7 8 9 ABSTRACT It is well known that vertical wind shear can organize deep convective systems and greatly extend their lifetimes. We know much less about the influence of shear on the bulk properties of tropical convection in statistical equilibrium. To address the latter question, the authors present a series of cloud-resolving simulations on a doubly periodic domain with parameterized large scale dynamics. The horizontal mean horizontal wind is relaxed strongly in these simulations towards a simple unidirectional linear vertical shear profile in the troposphere. The strength and depth of the shear layer are varied as control parameters. Surface enthalpy fluxes are prescribed. 10 11 12 13 14 15 16 17 18 The results fall in two distinct regimes. For weak wind shear, time-averaged rainfall decreases with shear and convection remains disorganized. For larger wind shear, rainfall increases with shear, as convection becomes organized into linear mesoscale systems. This non-monotonic dependence of rainfall on shear is observed when the imposed surface fluxes are moderate. For larger surface fluxes, convection in the unsheared basic state is already strongly organized, but increasing wind shear still leads to increasing rainfall. In addition to surface rainfall, the impacts of shear on the parameterized large-scale vertical velocity, convective mass fluxes, cloud fraction, and momentum transport are also discussed. 19 20 21 22 23 24 1

25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 1. Introduction Environmental wind shear can organize atmospheric moist convection into structures that contain mesoscale motions with a range of horizontal scales, from a few kilometers to thousands of kilometers. Severe convective storms, supercells, squall lines, mesoscale convective systems, tropical cloud clusters, trade cumulus, tropical cyclones, and the Madden Julian Oscillation are all influenced by wind shear, in various ways (e.g., Cotton and Anthes 1989; Houze 1993). Understanding the role of the vertical wind shear in the formation of thunderstorms, squall lines and other mesoscale features has been a long-standing problem, studied since the availability of upper air soundings, (e.g. Newton 1950). Simulations of squall lines with numerical models (e.g. Hane 1973; Moncrieff 1981; Thorpe et al. 1982; Rotunno et al. 1988; Weisman et al. 1988; Liu and Moncrieff 2001; Robe and Emanuel 2001; Weisman et al. 2004) show that environmental wind shear is crucial to their organization. Rotunno et al. (1988) and Weisman et al. (1988), widely known as RKW and WKR, respectively, emphasized that the cold pool-shear interaction may greatly prolong the lifetimes of squall lines and enhance their intensities. 40 41 42 43 44 45 46 47 In an environment without shear, the cold pool generally brings low entropy air from the mid-troposphere down to near the surface, where it is detrimental to further convection. On the other hand, the cold pool can propagate horizontally as a density current, generating circulations which lift environmental boundary layer air to its level of free convection, and triggering new convective cells. When the environment is sheared, the circulation associated with the shear may balance the circulation associated with the cold pool on the downshear side, promoting deeper lifting. The case where the cold pool is roughly in balance with the shear has been called the optimal state, meaning a state in 2

48 49 which the system maintains an upright updraft and repeatedly generates new cells on its leading edge (Xu et al. 1996; Xue et al.1997; Xue 2000; Robe and Emanuel 2001). 50 51 52 53 54 Despite its importance, the effect of vertical wind shear has not been included explicitly in parameterization schemes for large-scale models. This may cause biases in quantities that directly regulate energy balance, such as cloud fraction, optical depth, and radiative fluxes. These quantities are related to convective organization, which is modulated by shear (e.g., Liu and Moncrieff 2001). 55 56 57 58 59 60 The aim of this study is not to simulate squall lines or other specific mesoscale storm types per se, but rather to investigate the influence of vertical shear on mean precipitation and convective organization in the tropics in statistical equilibrium. We investigate this in cloud-resolving simulations of a small tropical region in which the interaction of that region with the surrounding environment is represented through a simple parameterization of the large-scale circulation. 61 62 63 64 65 66 67 68 69 When the large-scale circulation is parameterized, the model itself can determine the occurrence and intensity of deep convection. This is as opposed to more traditional methods, in which otherwise similar numerical experiments are performed with specified large-scale vertical motion, strongly constraining the bulk properties of convection a priori (e.g., Mapes 2004). The large-scale parameterization here is the weak temperature gradient (WTG) method (e.g., Sobel and Bretherton 2000; Raymond and Zeng 2005; Wang and Sobel 2011, 2012, 2013). In this method, horizontal mean temperature anomalies in the free tropical troposphere, relative to a prescribed target profile, are strongly relaxed towards zero. This represents the effects of compensating vertical 3

70 71 72 73 motion, communicated to large distances by gravity waves (e.g., Bretherton and Smolarkiewicz 1989) so as to maintain weak horizontal pressure and temperature gradients. The parameterized adiabatic cooling by the large-scale vertical velocity approximately balances the diabatic heating explicitly simulated by the CRM. 74 75 76 77 78 79 80 81 82 83 84 85 Parameterizations of large-scale dynamics have been used for a range of idealized calculations. Among these, Sessions et al. (2010) studied the response of precipitation to surface horizontal wind speed with fixed sea surface temperature (SST) using a 2-D cloud resolving model (CRM) under WTG. They demonstrated the existence of multiple equilibria corresponding to precipitating and non-precipitating states for the same boundary conditions but different initial conditions, corroborating the findings of Sobel et al. (2007) in a single column model with parameterized convection. Wang and Sobel (2011) showed the equilibrated precipitation as a function of SST using both a 2-D and 3- D CRM. Both of these studies show monotonic increases in surface precipitation rate (in the precipitating state, where it exists) with either increasing surface wind speed at fixed SST or vice versa. Here we extend these studies to examine the response of convection to vertical shear of the horizontal wind. 86 87 88 89 90 91 This paper is organized as follows. We describe the model setup and experiment in section 2. In section 3 we show how convective organization, mean precipitation, and other thermodynamic and dynamic quantities vary with the magnitude of shear when the depth of the shear layer is fixed and approximately equal to the depth of the troposphere. In section 4 we show the effect of variations in the depth of the shear layer. We conclude in section 5. 4

92 93 94 2. Experiment Design and Model Setup 2.1 Experiment design 95 96 97 98 99 100 101 102 103 104 105 106 107 108 The model used here is the Weather Research and Forecast (WRF) model version 3.3, in three spatial dimensions, with doubly periodic lateral boundary conditions. The experiments are conducted with Coriolis parameter f = 0. The domain size is 196 196 22 km 3, the horizontal resolution is 2 km, and 50 vertical levels are used, with 10 levels in the lowest 1 km. The microphysics scheme is the Purdue-Lin bulk scheme (Lin et al., 1983; Rutledge and Hobbs, 1984; Chen and Sun, 2002). This scheme has six species: water vapor, cloud water, cloud ice, rain, snow, and graupel. The 2-D Smagorinsky first order closure scheme is used to parameterize the horizontal transports by sub-grid eddies. The surface fluxes of moisture and heat are parameterized following Monin-Obukhov similarity theory. The Yonsei University (YSU) first order closure scheme is used to parameterize boundary layer turbulence and vertical subgrid scale eddy diffusion (Hong and Pan 1996; Noh et al. 2003; Hong et al. 2006). In this scheme nonlocal counter gradient transport (Troen and Mahrt 1986) is represented, and the local Richardson number, temperature, and wind speed determine the depth of the boundary layer. 109 110 111 112 Radiative cooling is set to a constant rate of 1.5K/day in the troposphere. The stratospheric temperature is relaxed towards 200 K over 5 days, and the radiative cooling in between is matched smoothly depending on temperature as in Pauluis and Garner (2006): 113 5

114 K day for T > K 1 1.5 207.5 Q = rad. cooling K T elsewhere 200 5 day (1) 115 116 117 118 119 120 121 122 123 124 The vertical wind shear is maintained by a term in the horizontal momentum equation which relaxes the horizontal mean zonal wind to a prescribed profile, U(z), with a relaxation time scale of 1 hour. Figure 1a shows U(z) for a constant shear layer depth of 12 km and varying shear magnitude, specified by varying the wind speed at the top of the shear layer from 10 m/s to 40 m/s in increments of 10 m/s while maintaining U=0 at the surface. Hereafter we refer to these experiments as U0, U10, U20 U30 and U40. Figure 1b, shows profiles in which we vary the depth of the shear layer while fixing the wind speed at the top of the layer at 20 m/s. In all cases, we relax the horizontal mean meridional wind to zero. 125 We prescribe surface fluxes of sensible heat (SH) and latent heat (LH). Three sets 126 of values are used for the total heat flux and its individual components: 160 W/m 2 (low), 127 128 129 130 131 132 133 partitioned as SH=15 W/m 2, LH=145 W/m 2 ; 206 W/m 2 (moderate), partitioned as SH=22 W/m 2, LH=184 W/m 2 ; and 280 W/m 2 (high), partitioned as SH=30 W/m 2, LH=250 W/m 2. These represent a range of deviations from the vertically integrated radiative cooling, which is 145 W/m 2. We consider only positive deviations from the radiative cooling, as we are most interested in equilibrium cases featuring significant convection. The parameterized circulations in our model tend to result in positive gross moist stability, so that the mean precipitation rate increases with the excess of surface fluxes 6

134 over radiative cooling (e.g., Wang and Sobel, 2011, 2012). 135 136 137 138 139 140 141 142 We prescribe surface fluxes, rather than just SST (as in previous work) because for different shear profiles, convective momentum transport induces significant changes in surface wind speeds, even with a strong relaxation imposed on the surface winds. This causes surface fluxes to vary significantly with shear if the fluxes are calculated interactively. Varying fluxes by themselves will change the occurrence and properties of deep convection, apart from any effects of shear on convective organization. We wish to completely isolate the direct effect of shear, without considering this modulation of surface fluxes. 143 144 145 146 147 148 149 2.2 WTG We use the WTG method to parameterize the large scale circulation, as in Wang and Sobel (2011), whose methods in turn are closely related to those of Sobel and Bretherton (2000) and Raymond and Zeng (2005). Specifically, we add a term representing largescale vertical advection of potential temperature to the thermodynamic equation. This term is taken to relax the horizontal mean potential temperature in the troposphere to a prescribed profile: 150 151 152 θ t +... = θ θ RCE τ (2), 153 154 155 Where θ is potential temperature,θ is the mean potential temperature of the CRM domain (the overbar indicates the CRM horizontal domain average), θ is the target RCE potential temperature, and τ is the Newtonian relaxation time scale, taken to be 3 hours in 7

156 157 158 159 160 161 162 163 our simulations. As τ approaches zero (a limit one may not be able to reach due to numerical issues), this becomes a strict implementation of WTG and the horizontal mean free troposphere temperature must equal θ. In general, τ is interpreted as the time RCE scale over which gravity waves propagate out of the domain, thus reducing the horizontal pressure and temperature gradients (Bretherton and Smolarkiewicz 1989). Finite τ allows the temperature to vary in response to convective and radiative heating. The large scale vertical circulation implied by this relaxation constraint is W wtg, the WTG vertical velocity, 164 θ W WTG η ( ρg µ) = θ θ RCE τ (3), 165 166 Where ρ is the density, η is the mass based vertical coordinate of WRF, and ρ g µ is T pd pd part of the coordinate transformation from z to η. η is defined as η =, where p d µ 167 is the dry pressure, T p d is a constant dry pressure at the model top, and µ is the dry 168 169 170 171 172 173 174 175 176 column mass. Within the boundary layer, following Sobel and Bretherton (2000) we do not apply (1), but instead obtain W WTG by linear interpolation from surface to the PBL top. Unlike in previous studies with our implementation of WTG, the PBL top is not fixed, but is diagnosed in the boundary layer parameterization scheme (discussed below). The scheme determines a PBL top at each grid point, and we use the maximum value in the computational domain at each time step as the PBL top for the computation of W wtg. Transport of moisture by the large-scale vertical motion introduces an effective source or sink of moisture to the column. The moisture equation is updated at each time step by adding the following terms associated with the WTG vertical velocity: 8

177 q t +... = W q WTG ( ρg µ) η (4), 178 179 180 Where q is the moisture mixing ratio. The right hand side of equation (4) is the advection by the large scale vertical velocity W WTG. We assume that the moisture field is horizontally uniform on large scales, thus neglecting horizontal advection. 181 182 183 184 185 186 187 188 189 190 191 3. Response to a shear layer of fixed depth a. Convective Organization and Precipitation Figure 2 shows randomly chosen snapshots of hourly surface rain for different values of shear strength and surface fluxes. In the case of low surface fluxes (Figure 2a) convection in the unsheared flow is random in appearance and has a popcorn structure. For small shear (cases U10 and U20) convection starts to organize but does not maintain a particular pattern. In some snapshots it has lines normal to the shear direction, while in others it loses this structure and appears to form small clusters. For strong shear, lines of intense precipitation form parallel to the shear direction. 192 193 194 195 196 197 198 In the case of moderate surface fluxes (Figure 2b) the unsheared flow s convection is random and distributed uniformly across the domain, but less so than in the case of low surface fluxes, as there are arcs and semi-circular patterns. Under weak shear (case U10) there are organized convective clusters in part of the domain. As the shear increases further, lines of intense precipitation (in brown shading) are trailed by lighter rain (blue shading) which propagate downshear (eastward). The organization in all cases is three-dimensional. 9

199 200 201 202 For high surface fluxes (Figure 2c), convection is organized in linear forms in the absence of vertical wind shear, loses linear structure with moderate shear (U10), and transform into aggregated states in which intense precipitation only occurs in a small part of the domain. 203 204 205 206 207 208 209 210 To further quantify the impact of the shear on precipitation, Figure 3 shows time series of the domain mean daily rain rate for the different shear strengths. Statistical equilibrium is reached in the first few days. The magnitude of the temporal variability is minimized in the unsheared case, and increases with increasing surface fluxes, while it is maximized in the cases U30 and U40 for low surface fluxes (Figure 3. a), and in U20 for higher surface fluxes (Figure 3. b and c). In many of the simulations the variability is quasi-periodic, with periods on the order of days. We are interested here primarily in the time-averaged statistics and have not analyzed these oscillations in any detail. 211 212 213 214 215 216 217 218 219 To define a quantitative measure of convective organization, we first define a blob as a contiguous region of reflectivity greater than 15 dbz in the vertical layer 0-2 km (Holder et al. 2008). Figure 4a shows a normalized probability density function (PDF) of the total (as the sum of all times for which we have data) number density of blobs, and Figure 4b shows the normalized PDF of the total area of blobs in the domain for the moderate surface fluxes case. The number of blobs decreases as convection clusters into aggregated structures with stronger shear. The areal coverage of convection increases with stronger shear as the tail of the PDF spreads towards larger areas. The other cases of surface fluxes are qualitatively similar and not shown. 220 10

221 222 223 224 225 226 227 228 229 230 231 232 233 234 c. Mean Precipitation, Thermodynamic Budget and Large-Scale Circulation 1) Mean Precipitation A key question we wish to address here is the dependence of the mean precipitation in statistical equilibrium as a function of the shear. Figure 5 shows the domain and time mean precipitation as a function of shear for three cases of surface fluxes. The most striking feature is that for low and moderate surface fluxes, the mean precipitation is a non-monotonic function of the shear (Figure 5a and 5b). For low surface fluxes, small shear brings the precipitation below the unsheared case, achieving its minimum at U20. For strong shear (U30 and U40), however, precipitation increases not only relative to the unsheared case but also above the surface fluxes, which indicates moisture import by the large-scale circulation. For moderate surface fluxes, the structure of the non-monotonicity differs from that in the low surface fluxes case. The minimum precipitation now shifts to U10, above which the behavior is monotonic. Again, strong shear is required to bring precipitation above that in the unsheared case. 235 236 237 238 239 240 241 242 243 For high surface fluxes, the behavior is monotonic, and small shear suffices to bring precipitation above that in the unsheared case. The difference between the minimum and maximum precipitation in the high surface flux case exceeds that in the lower surface flux cases, which is also (as one would expect) manifested in parameterized large scale vertical velocity W wtg as we will show in Figures 7c and 7f. There is no obvious relationship, however, between mean precipitation and organization. While small shear can organize convection from a completely random state, the mean of that more organized convection need not to be larger than that in the unsheared random state (Figures 2a and 2b). 11

244 245 246 247 2) Moist Static Energy (MSE) budgets The moist static energy (MSE) budget can be a useful diagnostic for precipitating convection as MSE is approximately conserved in adiabatic processes. It is the sum of thermal, potential and latent heat terms: 248 h = c p T + gz + L v q (5) 249 250 Where c p,, L v, and q are the heat capacity of dry air at constant pressure, latent heat of condensation, and water vapor mixing ratio, respectively. 251 252 253 254 255 256 257 Figure 6 shows the time and domain mean profiles of temperature, water vapor mixing ratio, and moist static energy, all expressed as perturbations from the same quantities in the unsheared experiment. There is a weak sensitivity to the shear in the temperature profile for low surface fluxes, becoming larger as surface fluxes increase. Recall that the temperature is continually being strongly relaxed back towards the target profile. The temperature perturbation is positive in the mid to upper troposphere and negative in the boundary layer in the strong shear case. 258 259 260 261 The moisture profile is more sensitive to the shear than temperature, but varies little with increasing surface fluxes. Above the boundary layer the strong shear case is moister than the unsheared case by about 0.2 g/kg and drier in the boundary layer by about 0.5 g/kg. The weak shear case is drier than the unsheared case at all levels. 262 263 264 The moist static energy variations with shear are dominated by the moisture term, are very similar for all three cases of surface fluxes, and depend non-monotonically on the shear. The non-monotonicity remains the same for all cases of surface fluxes, unlike 12

265 the large-scale vertical velocity and mean precipitation. 266 267 268 269 270 271 272 3) Large-Scale Circulation Figure 7 shows the vertical profiles of large-scale vertical velocity W WTG (a)-(c), and its maximum value with respect to the vertical (d)-(f) for the three cases of surface fluxes. For low surface flux, W WTG is very weak and varies little with shear. As the surface fluxes increase, W WTG increases, but the effect of the shear is much more significant in the highest surface fluxes case, where W WTG increases by 4 cm/s from the unsheared case to the case with the strongest shear. 273 274 275 276 4) Normalized Gross Moist Stability In order to diagnose the mean precipitation from the point of view of the moist static energy budget, we use the diagnostic equation for precipitation as in, e.g., Sobel (2007) or Wang and Sobel (2011): 277 1 P= ( E+ H+< QR > ) < QR > H (6) M 278 Where <. >= p T p0 ρdz is the mass weighted vertical integral from the bottom to the top 279 280 281 of the domain. P, E, H, and Q R are precipitation, latent heat flux, sensible heat flux, and radiative cooling rate. We have also utilized the large-scale vertical velocity and the moist static energy introduced in the above two subsections to define 282 M = < W WTG h z > < W WTG s > as the normalized gross moist stability, which represents z 283 284 the export of moist static energy by the large-scale circulation per unit of dry static energy export. Here s is the dry static energy (sum of the thermal and potential energy 13

285 only, without the latent heat term). 286 287 288 289 290 291 The above equation is not predictive because we use the model output to compute the terms; in particular, we do not have a theory for M. Nonetheless, it gives a quantitative, if diagnostic, understanding of the role played by large-scale dynamics in controlling precipitation. We must first verify that our moist static energy budget closes well enough that the precipitation calculated from equation (6) agrees well with the model output precipitation. This is the case, as shown in blue in Figure 5. 292 293 294 295 296 As shown in Figure 5, the variations of M with respect to shear (themselves shown in fig. 8) explain those in the mean precipitation. Since both surface fluxes and radiative cooling are constant in each plot in Fig. 5, variations in M are the only factor controlling the variations in precipitation computed by equation (6), which is quite close to the actual simulated precipitation. 297 298 299 300 301 302 303 304 305 Maximum M means maximum export of moist static energy by the large-scale circulation per unit of dry static energy export. Since radiative cooling and surface fluxes are prescribed, the moist static energy export is also prescribed, and variations in M must correspond to variations in dry static energy export. Since dry static energy export changes can only be balanced by changes in the rate of condensation (radiative cooling, again, being fixed) and, hence, in the precipitation rate. For the case of low surface fluxes, M exceeds 1 for the cases U10 and U20; this is because of the low level subsidence and upper level ascent in the W WTG profile, with midlevel inflow bringing in relatively low moist static energy air. 306 14

307 308 309 310 311 312 313 314 c. Convective Cloud Properties 1) Cloud Fraction The vertical profile of cloud fraction for the three values of surface fluxes is shown in Figure 9. The cloud is defined as the set of grid points where the mixing ratio of cloud hydrometeors (water and ice) is greater than 0.005 g/kg. Generally, there are three peaks: a shallow peak at 1km, a midlevel congestus peak at 4km, and a peak associated with deep convection at 9-12 km. When the cloud fraction in the deep convective peak becomes large, we interpret it as stratiform cloud. 315 316 317 318 In all three cases, the shear has almost no effect on the shallow convection. For low surface fluxes (Figure 9a) the effect of strong shear is seen mainly in the midlevels, where cloud fraction increases from 30% in the unsheared flow to 60%. A modest stratiform layer forms as cloudiness increases from 10% to 30%. 319 320 321 322 323 324 325 326 327 328 329 For moderate surface fluxes (Figure 9b) the strong shear, while causing an increase of more than 40% in the midlevel congestus over the unsheared case, also leads to 100% cloud cover in the layer 9-12 km. It is also interesting to notice a transition from a single peak in the high clouds at 9 km in the unsheared and weak shear cases to a cloudy layer 5-12 km deep in the strong shear. In Figure 9c near-total deep cloudiness is determined by the strong, persistent convection associated with the high surface fluxes. The shear has no effect on the cloud fraction at upper levels, since it is 100% even with no shear. The effect of shear on mid-level cloudiness is clear, however, as cloud fraction increases over 35% from the unsheared to the strongest shear flow. The cloud cover in the midlevel increases monotonically with the shear, but the deep cloudiness does not (Figure 9a and 9b), behaving instead more like mean precipitation. We discuss this 15

330 further in section 3.c.3. 331 332 333 334 335 336 337 338 339 2) Convective Mass Flux Figure 10 shows the convective mass fluxes in both updrafts and downdrafts. Updrafts are defined here as including all grid points with liquid water content greater than 0.005 g/kg and vertical velocity greater than 1 m/s, normalized by the total number of grid points. The downdrafts are defined including all grid points with vertical velocity less than -1 m/s. The updraft mass fluxes are about twice as large as the downdraft mass fluxes. Both have two peaks, in the lower and upper troposphere, both of which strengthen by about a factor of 3 from lowest to highest surface fluxes. 340 341 342 343 344 345 346 347 Despite different convective organization for different shear profiles, the vertical profiles of the updrafts and downdrafts have similar shapes. For the downdrafts, the lower tropospheric peak exceeds the upper tropospheric one for all three cases of surface fluxes, and are similar for different shear strengths. The largest difference occurs in the case of highest surface fluxes, in which the downdraft in the lower peak for the smallest shear is stronger than for the strongest shear by about 20%. The effect of the shear on the upper tropospheric peak, however, is more prominent. The strongest shear causes the downdraft mass flux to strengthen by 50%. 348 349 350 351 The updraft mass fluxes in the low surface fluxes case are greater in the lower troposphere than in the upper troposphere, due to the dominant role of the shallow convection, while in the moderate surface fluxes case both peaks are almost comparable, and in the high surface fluxes case the upper peak is larger due to deep clouds. Although 16

352 353 354 355 with high surface fluxes there is an increase of 25% in the upper tropospheric updraft from the weakest to strongest shear, the effect of the shear on the updraft is smaller than in the downdraft and has negligible impact on the lower tropospheric peak, and, unlike the downdraft, is non-monotonic, similarly to cloud fraction. 356 357 358 359 360 361 362 3) Momentum Fluxes Another collective effect of the cloud population and its organization on the largescale flow results from vertical fluxes of horizontal momentum, including both convective momentum transport and gravity wave momentum transport (Lane and Moncrieff 2010, Shaw and Lane 2013). We evaluate these momentum fluxes as a function of the shear strength. Figure 11 shows the vertical profiles of the time and 363 domain mean zonal momentum fluxes ρuw ' ' where u and w are the zonal and vertical 364 365 velocity perturbations from the horizontal average, respectively, and the overbar is the horizontal mean. 366 367 368 369 370 371 372 373 The magnitude of the momentum transport is a monotonic function of the shear. It is negligible for all unsheared cases, as expected, and negative (down gradient) and peaks close to the top of the shear layer for the sheared cases. The negative momentum transport represents a downward transport of horizontal momentum within and immediately above the shear layer, which further accelerates the horizontal wind in the lower troposphere and decelerates it in the upper troposphere and the stratosphere. The amplitude of the momentum transport also increases with surface fluxes. Interestingly, for the highest surface fluxes the momentum flux in the U10 case (Figure 11c), is up 17

374 375 gradient in the lower troposphere and peaks at about 4 km, while it is down gradient in the upper troposphere. 376 377 378 379 380 381 382 383 4. Response to Different Shear Depths In this section we investigate the effect of varying the depth of the shear layer. We repeat the above calculations but with shear depths of 1500 m, 3000 m, 4500 m, 6000 m, and 9000 m. In each case, the wind speed is held fixed at 20 m/s at the top of the layer (Figure 1. b). Surface latent and sensible heat fluxes are held fixed at the intermediate values from the previous section, totaling 206 W/m 2. 384 385 386 387 388 389 390 391 392 393 394 395 Figure 12 shows time and domain mean precipitation as a function of the shear depth. The shallowest shear layer, with depth 1.5 km, produces the least precipitation, about 30% less than the unsheared case, despite a greater degree of convective organization compared to the unsheared flow (not shown). Doubling the shear layer depth almost doubles the precipitation. The mid-depth shear layers, with tops at 3-4.5 km produce the greatest precipitation; in this sense, these intermediate shear layer depths are optimal. Although the convection is in statistical equilibrium, this is in agreement with studies of transient storms of Weisman et al. (2004). While it was not the case that the optimal state produces the greatest surface precipitation in their original papers and later work, Bryan et al. (2006) demonstrated that optimal state is indeed associated with the greatest precipitation in several mesoscale models. Bryan et al. (2006) further attributed this discrepancy to numerical artifacts in the model used by Weisman et al. (2004). 18

396 397 Increasing the shear depth above 4.5 km reduces the precipitation again, with the values for 6 km and 9 km depths being smaller than those for the intermediate depths. 398 399 400 401 402 403 404 405 The vertical profiles of the large-scale vertical velocity as a function of the shear layer depth are shown in Figure 13. As we expect, the amplitude differences are largely consistent with those in the mean precipitation. The peak value, for instance, is smallest for the shallow shear case, and largest for the mid-level shear. The cases of 3 km and 4.5 km are almost identical, while there is a small increase in precipitation for 4.5 km over that at 3 km. This is due to the more abundant moisture in the lower to mid troposphere for the case of 4.5 km, as shown in Figure 14. Dominated by moisture, moist static energy has a maximum for the 4.5 km case. 406 407 408 409 410 411 412 Figure 15 shows the mass fluxes in updrafts and downdrafts for the different shear depths. Although the maximum mean precipitation occurs for the shear depth of 4.5 km, the maximum updraft mass flux occurs at 3km shear depth. This demonstrates that convective mass flux need not vary in the same way as the precipitation rate. The response of the downdrafts is much weaker than that of the updrafts. In fact, the downdraft mass flux for the shallow shear is almost identical to that for the deeper shear layers with depths 6km and 9km. 413 414 415 416 417 Finally, Figure 16 shows the domain averaged time mean zonal momentum fluxes ρuw ' ' for different shear depths. The momentum flux is down gradient for all shear depths, as in Fig. 12, and achieves the minimum at the top of the shear layer in each profile. The downward momentum transport is greatest when the shear is shallow: 1.5 km and 3 km deep. This differs from the variations in the mean precipitation. As with other 19

418 419 measures of organization, momentum flux apparently does not have a direct relationship to precipitation. 420 5. Summary and Discussion 421 422 423 424 425 426 427 428 429 We have quantified the effect of vertical wind shear on atmospheric convection by conducting a series of 3-D CRM simulations with large-scale dynamics parameterized according to the weak temperature gradient approximation. We varied the shear while holding surface heat fluxes fixed. We repeated the calculations for three different values of surface fluxes. As surface fluxes are increased, the strength of the simulated convection increases, as measured by domain-averaged precipitation. Precipitation increases more rapidly than the surface latent heat flux, because of parameterized largescale moisture convergence. The three cases thus correspond to weak, moderate and strong convection. 430 431 432 433 434 435 436 The response of mean precipitation to shear in statistical equilibrium for the lower two surface flux cases is non-monotonic in the magnitude of the shear, with a minimum precipitation at an intermediate shear value. The shear value at which the minimum precipitation occurs is somewhat different for low and moderate surface fluxes. Only for strong shear does the precipitation exceed that in the unsheared flow. In the case of the highest surface fluxes used, the response to shear is monotonic and a small amount of shear is enough to cause the precipitation to exceed that of the unsheared flow. 437 438 The dependences of large-scale vertical velocity and the moist static energy on shear are both similar to the dependence of precipitation on shear. The dependence on 20

439 440 441 shear of the normalized gross moist stability, which combines these two quantities in a way relevant to the column-integrated moist static energy budget, is found to explain the variations in mean precipitation with shear well. 442 443 444 445 446 447 448 Despite the relative smallness of the changes in the mean precipitation with shear, shear has a strong effect on convective organization. Weak shear is sufficient to organize convection even under weak surface fluxes when the large-scale vertical motion is negligible. Stronger shear can organize convection into convective clusters and squall line-like structures with lines of intense precipitation trailed by lighter rain. As the surface fluxes and domain-averaged precipitation increase, however, convection is found to be organized even in the absence of vertical wind shear. 449 450 451 452 The wind shear also has a significant effect on cloud cover, which increases with the shear (especially high clouds); convective mass flux (in particular the downdraft mass flux); and momentum fluxes, as more momentum is transported downward when the shear increases. 453 454 455 456 457 458 When the total change in velocity over the shear layer is fixed but the layer depth is varied, mid-level shear depths are found to be optimal for producing maximum mean precipitation, maximum large-scale vertical velocity and maximum column moist static energy. Convective organization, for this shear depth, is very similar to that for deeper shear layers, but with more intense lines of precipitation. Confining the shear in a shallow layer dries out the atmospheric column, leaving precipitation at a minimum. 459 460 We have neglected the complications of cloud-radiation feedback by fixing the radiative cooling rate in a simple relaxation scheme. In reality, longwave radiative 21

461 462 463 cooling can be suppressed by the abundant upper tropospheric ice cloud in the deep tropics, which by itself is controlled by the shear. This effect will be investigated in a separate, future study. 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 22

479 480 481 482 483 484 485 486 487 488 References Bretherton, C. S., P. N. Blossey, M. Khairoutdinov, 2005: An Energy-Balance Analysis of Deep Convective Self-Aggregation above Uniform SST. J. Atmos. Sci., 62, 4273 4292. Bryan, G. H., J. C. Knievel, and M. D. Parker, 2006: A multimodel assessment of RKW theory s relevance to squall-line characteristics. Mon. Wea. Rev., 134, 2772 2792. Chen, S. H., and W. Y. Sun 2002, A one dimensional time dependent cloud model, J. Meteorol. Soc. Jpn., 80, 99 118. 489 490 Cotton, W. R., and R. A. Anthes, 1989: Storm and Cloud Dynamics. International Geophysical Series, Vol. 44, Academic Press. 491 492 Hane, C. E., 1973: The squall line thunderstorm: Numerical experimentation. J. Atmos. Sci., 30, 1672 1690. 493 494 495 Held, I. M., R. S. Hemler, V. Ramaswamy, 1993: Radiative-Convective Equilibrium with Explicit Two-Dimensional Moist Convection. J. Atmos. Sci., 50, 3909 3927. 496 497 498 499 500 Holder, C. T., S. E. Yuter, A. H. Sobel, A. R. Aiyyer, 2008: The Mesoscale Characteristics of Tropical Oceanic Precipitation during Kelvin and Mixed Rossby Gravity Wave Events. Mon. Wea. Rev., 136, 3446 3464. Hong, S. Y., and H. L. Pan,1996, Nonlocal boundary layer vertical diffusion in a 23

501 Medium Range Forecast model, Mon. Weather Rev., 124, 2322 2339. 502 503 Hong, S.Y., Y. Noh, and J. Dudhia, 2006: A new vertical diffusion pack- age with an explicit treatment of entrainment processes, Mon. Weather Rev., 134, 2318 2341. 504 505 Houze, R. A., Jr., 1993: Cloud Dynamics. International Geophysical Series, Vol. 53, Academic Press, 573 pp. 506 507 508 509 Lane, T. P., M. W. Moncrieff, 2010: Characterization of Momentum Transport Associated with Organized Moist Convection and Gravity Waves. J. Atmos. Sci., 67, 3208 3225. 510 511 Lin, Y.L., R. D. Farley, and H. D. Orville, 1983, Bulk parameterization of the snow field in a cloud model, J. Clim. Appl. Meteorol., 22, 1065 1092. 512 513 Liu, C., M. W. Moncrieff, 2001: Cumulus Ensembles in Shear: Implications for Parameterization. J. Atmos. Sci., 58, 2832 2842. 514 515 516 Mapes, B. E. (2004), Sensitivities of cumulus ensemble rainfall in a cloud resolving model with parameterized large scale dynamics, J. Atmos. Sci., 61, 2308 2317. 517 518 Moncrieff, M. W., 1981: A theory of organized steady convection and its transport properties. Quart. J. Roy. Meteor. Soc., 107, 29 50. 519 520 Muller, C. J., I. M. Held, 2012: Detailed Investigation of the Self-Aggregation of Convection in Cloud-Resolving Simulations. J. Atmos. Sci., 69, 2551 2565. 521 522 Noh, Y., W. G. Cheon, S. Y. Hong, and S. Raasch, 2003, Improvement of the K profile 24

523 524 model for the planetary boundary layer based on large eddy simulation data, Boundary Layer Meteorol., 107, 401 427. 525 526 Pauluis, O., and S. Garner, 2006, Sensitivity of radiative convective equilibrium simulations to horizontal resolution, J. Atmos. Sci., 63, 1910 1923. 527 528 529 Raymond, D. J., and X. Zeng, 2005, Modelling tropical atmospheric convection in the context of the weak temperature gradient approximation, Q. J. R. Meteorol. Soc., 131, 1301 1320. 530 531 Robe, F. R., and K. A. Emanuel, 2001: The effect of vertical wind shear on radiative convective equilibrium states. J. Atmos. Sci., 58, 1427 1445. 532 533 534 Rotunno, R., J. B. Klemp, and M. L. Weisman, 1988: A theory for strong, long-lived squall lines. J. Atmos. Sci., 45, 463 485. 535 536 537 538 539 Rutledge, S. A., and P. V. Hobbs 1984, The mesoscale and microscale structure and organization of clouds and precipitation in midlatitude cyclones. XII: A diagnostic modeling study of precipitation development in narrow cloud frontal rainbands, J. Atmos. Sci., 41, 2949 2972. 540 541 Sessions, S., D. J. Raymond, and A. H. Sobel, 2010, Multiple equilibria in a cloud resolving model, J. Geophys. Res., 115, D12110. 542 543 544 Shaw, T. A., and T. P. Lane, 2013: Toward an understanding of vertical momentum transports in cloud system resolving model simulations of multiscale tropical convection. J. Atmos. Sci., in press. 25

545 546 547 548 Sobel, A. H. 2007, Simple models of ensemble averaged precipitation and surface wind, given the SST, in The Global Circulation of the Atmosphere, edited by T. Schneider and A. H. Sobel, pp. 219 251, Princeton Univ. Press, Princeton, N. J. 549 550 Sobel, A. H., and C. S. Bretherton, 2000, Modeling tropical precipitation in a single column, J. Clim., 13, 4378 4392. 551 552 553 Thorpe, A. J., M. J. Miller, and M. W. Moncrieff, 1982: Two-dimensional convection in non-constant shear: A model of mid-latitude squall lines. Quart. J. Roy. Meteor. Soc., 108, 739 762. 554 555 556 Troen, I., and L. Mahrt 1986, A simple model of the atmospheric boundary layer: Sensitivity to surface evaporation, Boundary Layer Meteorol., 37, 129 148. 557 558 559 Wang, S., and A. H. Sobel 2011, Response of convection to relative sea surface temperature: Cloud resolving simulations in two and three dimensions, J. Geophys. Res., 116, D11119. 560 561 Wang, S. and A. H. Sobel, 2012: Impact of imposed drying on deep convection in a cloud-resolving model. J. Geophys. Res., 117, D02112, doi:10.1029/2011jd016847 562 563 564 Wang, S., A. H. Sobel, and Z. Kuang, 2013: Cloud-resolving simulation of TOGA- COARE using parameterized large-scale dynamics. J. Geophys. Res., 118, doi:10.1002/jgrd.50510. 26

565 566 Weisman, M. L., J. B. Klemp, and R. Rotunno, 1988: Structure and evolution of numerically simulated squall lines. J. Atmos. Sci., 45, 1990 2013. 567 568 569 570 Xu, Q., M. Xue, and K. K. Droegemeier, 1996: Numerical simulation of density currents in sheared environments within a vertically confined channel. J. Atmos. Sci., 53, 770 786. 571 572 573 Xue, M., 2000: Density currents in two-layer shear flows. Quart. J. Roy. Meteor. Soc., 126, 1301 1320. 574 575 576 Xue, M., Q. Xu, and K. K. Droegemeier, 1997: A theoretical and numerical study of density currents in nonconstant shear flows. J. Atmos. Sci., 54, 1998 2019. 577 578 579 580 Yanai, M., S. K. Esbensen, and J.-H. Chu, 1973: Determination of the bulk properties of tropical cloud clusters from large-scale heat and moisture budgets. J. Atmos. Sci., 30, 611 627. 581 582 583 27

584 585 586 List of Figures FIG. 1. Wind profiles used in the simulations. (a) fixed depth of 12 km, (b) varying depths. 587 588 589 FIG. 2. Random snapshots of hourly surface precipitation for a period of 35 hours. Each row corresponds to a different shear case. From top to bottom: U0, U10, U20, U30, and U40. (a) for low (b) moderate, and (c) high surface fluxes. 590 591 FIG. 3. Time series of daily precipitation. (a) low, (b) moderate, and (c) high surface fluxes. Colors indicate the value of the shear (m/s). 592 593 FIG. 4. Normalized probability density functions (PDFs) of (a) the total number density of blobs (number every 4x10 4 km 2 ), and (b) the size of blobs in the domain (km 2 ). 594 595 596 FIG. 5. Time and domain mean model output precipitation (red), derived precipitation (blue), and surface fluxes (black) as a function of the shear. (a) low, (b) moderate, and (c) high surface fluxes. 597 598 FIG. 6. Temperature, water vapor mixing ratio, and moist static energy, all expressed as differences from the unsheared case. (a) low, (b) moderate, and (c) high surface fluxes. 599 600 FIG. 7. (a)-(c) vertical profile of the large-scale vertical velocity W WTG. (d)-(f) W WTG Maximum. (a) and (d) low, (b) and (e) moderate, and (c) and (f) high surface fluxes. 601 602 FIG. 8. Normalized gross moist stability M as a function of the shear. (a) low, (b) moderate, and (c) high surface fluxes. 603 28

604 FIG. 9. Cloud fraction. (a) low, (b) moderate, and (c) high surface fluxes. 605 606 FIG. 10. Convective mass flux of updraft and downdraft. (a) low, (b) moderate, (c) high surface fluxes. 607 FIG. 11. Momentum fluxes. (a) low, (b) moderate, and (c) high surface fluxes. 608 609 FIG. 12. Time and domain mean model output precipitation (red), derived precipitation (blue), and surface fluxes (black) as a function of the shear layer depth. 610 611 FIG. 13. Vertical profile of the large-scale vertical velocity W WTG for different shear depths. 612 613 FIG. 14. Temperature, water vapor mixing ratio, and moist static energy, all expressed as differences from the unsheared case, for different shear layer depths. 614 FIG. 15. Convective mass flux of updraft and downdraft for different shear depths. 615 FIG. 16. Momentum fluxes for different shear depths. 616 617 618 29

619 620 621 622 FIG. 1. Wind profiles used in the simulations. (a) fixed depth of 12 Km, (b) varying depths. 623 624 625 626 627 628 629 630 631 632 30

633 634 635 636 637 FIG. 2. Random snapshots of hourly surface precipitation for a period of 35 hours. Each row corresponds to a different shear case. From top to bottom: U0, U10, U20, U30, and U40. (a) for low (b) moderate, and (c) high surface fluxes. 638 31

639 640 641 642 FIG. 3. Time series of daily precipitation. (a) low, (b) moderate, and (c) high surface fluxes. Colors indicate the value of the shear (m/s). 643 32

644 645 646 FIG. 4. Normalized probability density functions (PDFs) of (a) the total number density of blobs (number every 4x10 4 km 2 ), and (b) the size of blobs in the domain (km 2 ). 647 648 649 650 651 33

652 653 FIG. 5. Time and domain mean model output precipitation (red), derived precipitation 34

654 655 (blue), and surface fluxes (black) as a function of the shear. (a) low, (b) moderate, and (c) high surface fluxes. 656 657 658 659 660 661 662 663 664 665 35

666 667 668 FIG. 6. Temperature, water vapor mixing ratio, and moist static energy, all expressed as differences from the unsheared case. (a) low, (b) moderate, and (c) high surface fluxes. 669 36

670 671 672 FIG. 7. (a)-(c) vertical profile of the large-scale vertical velocity W WTG. (d)-(f) W WTG Maximum. (a) and (d) low, (b) and (e) moderate, and (c) and (f) high surface fluxes. 673 37

674 675 FIG. 8. Normalized gross moist stability M as a function of the shear. (a) low, (b) 38

676 moderate, and (c) high surface fluxes. 677 39

678 679 FIG. 9. Cloud fraction. (a) low, (b) moderate, and (c) high surface fluxes. 40

680 681 FIG. 10. Convective mass flux of updrafts and downdrafts. (a) low, (b) moderate, (c) high 41

682 surface fluxes. 683 42

684 685 FIG. 11. Momentum fluxes. (a) low, (b) moderate, and (c) high surface fluxes. 43

14 13 12 Derived Precip, Model Precip Surface Fluxes mm/d 11 10 9 8 7 686 6 Unsheared 1500 3000 4500 6000 9000 Shear Depth 687 688 FIG. 12. Time and domain mean model output precipitation (red), derived precipitation (blue), and surface fluxes (black) as a function of the shear layer depth. 689 690 691 44

692 693 694 FIG. 13. Vertical profile of the large-scale vertical velocity W WTG for different shear depths. 695 45

696 697 FIG. 14. Temperature, water vapor mixing ratio, and moist static energy, all expressed as 46

698 differences from the unsheared case, for different shear layer depths. 699 700 701 702 703 704 705 706 707 708 47

709 710 711 FIG. 15. Convective mass fluxes of updrafts and downdrafts for different shear layer depths. 712 713 714 715 716 717 48

718 719 FIG. 16. Momentum fluxes for different shear layer depths. 49