• Aucun résultat trouvé

Impact of melting snow on the valley flow field and precipitation phase transition

N/A
N/A
Protected

Academic year: 2021

Partager "Impact of melting snow on the valley flow field and precipitation phase transition"

Copied!
44
0
0

Texte intégral

(1)

Impact of melting snow on the valley flow field and

precipitation phase transition

Julie M. Th´eriaulta,∗, Jason A. Milbrandtb, Jonathan Doylea, Justin R. Minderc, Gregory Thompsond, Noemi Sarkadie, Istvan Geresdie

aUniversit´e du Qu´ebec `a Montr´eal, Montr´eal, Qu´ebec, CANADA

bAtmospheric Numerical Prediction Research, Environment Canada, Dorval, Canada

cUniversity at Albany, Albany, New York, US

dNational Center for Atmospheric Research, Boulder, US

eUniversity of P´ecs, Faculty of Science, Szentagothai Research Center, P´ecs, Hungary

Abstract

The prediction of precipitation phase and intensity in complex terrain is challenging when the surface temperature is near 0◦C. In calm weather con-ditions, melting snow often leads to a 0◦C-isothermal layer. The tempera-ture feedback from melting snow generates cold dense air moving downslope, hence altering the dynamics of the storm. A correlation has been commonly observed between the direction of the valley flow and the precipitation phase transition in complex terrain. This study examines the impact of tempera-ture feedback from melting snow on the direction of the valley flow when the temperature is near 0◦C. Semi-idealized two-dimensional simulations using the Weather Research and Forecasting model were conducted for a case of moderate precipitation in the Pacific Coast Ranges. The results demonstrate that the temperature feedbacks caused by melting snow affects the direc-tion of the flow in valleys. Several microphysics schemes (1-moment bulk,

Corresponding author

(2)

2-moment bulk, and bin), which parameterize snow in different ways, all pro-duced a valley flow reversal but at different rates. Experiments examining sensitivity to the initial prescribed snow mixing ratio aloft were conducted to study the threshold precipitation at which this change in the direction of the valley flow field can occur. All prescribed snow fields produced a change in the valley wind velocity but with different timings. Finally, the evolution of the rain-snow boundary with the different snowfields was also studied and compared with the evolution of the wind speed near the surface. It was found that the change in the direction of the valley flow occurs after the 0◦C isotherm reaches the base of the mountain. Overall this study showed the importance to account for the latent heat exchange from melting snow. This weak temperature feedback can impact, in some specific weather conditions, the valley flow field in mountainous area.

Keywords: precipitation, complex terrain, rain-snow boundary, dynamic meteorology, microphysics

1. Introduction 1

Precipitation is one of the most important weather elements affecting our 2

society. Its occurrence represents a crucial part of the global water cycle 3

and it is a fundamental aspect of storms. The precipitation phase (i.e., 4

rain versus snow) has a major impact on the water resources in the spring 5

snowmelt season and plays an important role in determining flood hazard 6

(e.g. Barnett et al., 2005; Elsner et al., 2010; White et al., 2002). 7

Formation and phase changes of precipitation are associated with diabatic 8

heating and cooling of the environmental air due to latent heat exchanges. 9

(3)

Cooling due to melting snow can alter the temperature profile, which can in 10

turn induce mesoscale circulations and influence the evolution of the storm. 11

Lin and Stewart (1986) showed that melting-induced mesoscale circulation 12

could extend as far as 50 km horizontally. Furthermore, in still weather 13

conditions, the melting of snow often produces a deep isothermal layer of 14

0◦C (Findeisen, 1940), which also leads to a change in precipitation from 15

rain to snow. 16

These thermodynamic and dynamical feedbacks have been studied over 17

complex terrain. Steiner et al. (2003) demonstrated through radar measure-18

ments that a change from up-valley to down-valley flow and a precipitation 19

phase transition occur simultaneously. In particular, they observed that the 20

top of the radar bright band correlated with the shear level where the flow 21

direction changed. On the other hand, Z¨angl (2007) conducted numerical 22

simulations of the same event and concluded that the melting process only 23

has a small contribution to the change of the wind flow in the valley. 24

Similar radar patterns to those discussed in Steiner et al. (2003) were 25

observed in other regions of the world. For instance, in the St-Lawrence 26

River Valley, Quebec, Canada during The 1998 Ice Storm (Henson et al., 27

2011) as well as in the Whistler Area, British Columbia, Canada (Fig. 1) 28

during the Vancouver 2010 Winter Olympics. In particular, a correlation 29

between the change in precipitation phase, valley flow field, and a rapid 30

decrease in surface temperature was observed on 13-14 February 2010 in the 31

Whistler Area (Th´eriault et al., 2012) during the SNOW-V10 field project 32

(Isaac et al., 2014). It was hypothesized that the cooling from the melting 33

snow was associated with the change in direction of the valley flow field. 34

(4)

The characterization of the rain-snow boundary in mountainous terrain has 35

been addressed in several studies including Medina et al. (2005), Minder 36

et al. (2011), Z¨angl (2007), Minder and Kingsmill (2013). In particular, 37

Minder et al. (2011) performed numerical simulations to study the mesoscale 38

features of the rain-snow boundary along mountainside. It was demonstrated 39

that diabatic cooling by melting precipitation, adiabatic cooling from vertical 40

motion, and microphysical timescales associated with melting all influenced 41

the location of the rain-snow boundary along the mountainside, causing it 42

to descend over a mountain windward slope. Their study also showed the 43

predicted magnitude of the rain-snow boundary’s descent varies substantially 44

depending on microphysical parameterization. 45

The sensitivity to microphysical assumptions related to snow on the di-46

abatic cooling effects and the resulting precipitation phase changes was ex-47

amined in Milbrandt et al. (2014) in a simple one-dimensional framework 48

for the 13-14 February 2010 case in the Whistler area. The snow quantity 49

aloft corresponding to radar observations was prescribed with an observed 50

temperature and humidity profile with melting and cooling rates simulated 51

with a bulk microphysics scheme. It was shown that the cooling rate due 52

to melting, and hence the resulting timing of the phase transition at the 53

surface, can be quite sensitive aspects of the representation of snow in the 54

model. This includes the assumed fall speed parameters, the number of prog-55

nostic moments, and constraints on the size distribution such as the lower 56

limit of the slope parameter and the assumptions of the melting processes in 57

schemes. 58

Given the difficulty of predicting the precipitation phase and intensity 59

(5)

when the temperature is near 0 ◦C, this study aims to better understand 60

the impact of temperature feedbacks from melting snow on the direction 61

of the valley flow field and on the precipitation phase. Semi-idealized two-62

dimensional simulations of the 13-14 February 2010 Whistler case were con-63

ducted using a mesoscale model in a systematic manner. First, the link be-64

tween the temperature feedback from melting snow and the direction of the 65

valley flow field is verified. A sensitivity experiment with various microphysi-66

cal parameterization approach was also conducted. Second, the sensitivity to 67

different precipitation rates, through prescribing different initial snowfields, 68

is studied to investigate the threshold precipitation rates required to produce 69

a change in the valley flow direction. Third, the evolution of the rain-snow 70

line is compared to the rate of change of the valley flow for the different 71

initial precipitation rates. 72

The paper is organized as follows. Section 2 describes the model config-73

uration and experimental design. Section 3 summarizes the results from the 74

control simulation along with the effects of suppressing diabatic cooling due 75

to melting snow. The factor impacting the timing of the valley flow field are 76

presented in section 4. The concluding remarks are given in Section 5. 77

2. Experimental design 78

2.1. Case overview 79

To test the impact of the temperature feedbacks from melting on the 80

valley flow field and the precipitation phase transition, semi-idealized nu-81

merical simulations were performed based on the well-documented case of 13 82

February 2010 (Th´eriault et al., 2012). On this day, an intense warm-frontal 83

(6)

system slowly approached the British Columbia coastline as it elongated in 84

a north-south orientation. This weather system was associated with heavy 85

snow and a transition of precipitation from rain to snow along the moun-86

tainsides throughout most of the day. One particular geographic area and a 87

multi-hour time period of this storm are used for this semi-idealized study. 88

This study focuses on the Callaghan Valley (VOD) located west of the 89

base of Whistler Mountain (VOT) and south-west from the rawinsonde sta-90

tion (VOC) (Fig. 1) during the SNOW-V10 project (Isaac et al, 2014). A 91

rapid decrease of the surface air temperature was observed in the Callaghan 92

Valley (Fig. 2a) from 2230 UTC 13 February 2013 to 0000 UTC 14 February 93

2010. Soon after 0000 UTC, surface temperature reached 0◦C and remained 94

constant until 0600 UTC 14 February 2010. Figure 2b also shows that pre-95

cipitation started in the valley (at VOT) when the air temperature began to 96

drop. The fact that air temperature remained constant at 0◦C for several 97

hours strongly suggests that the temperature feedbacks from melting snow 98

was a dominant forcing during that time period. Furthermore, the radar 99

Doppler velocity (Fig. 3) in that region suggested a strong correlation be-100

tween the rapid cooling of the surface air and the change in the direction 101

of the valley flow. In particular, it showed a transition from up-valley flow 102

prior to the onset of the rapid decrease in surface air temperature and to 103

down-valley flow when the surface temperature reached 0◦C (Fig. 3). The 104

wind blows down valley flow the depth of the initial melting layer near the 105

surface approximately 150 min after the onset of precipitation. The radar 106

data used is discussed in (Joe et al., 2014). 107

(7)

2.2. Model setup 108

2.2.1. Model description 109

All of the simulations in this study were performed using the Weather 110

Research and Forecasting (WRF) model, version 3.3.1 (Skamarock et al., 111

2008). The set-up involves a semi-idealized two-dimensional configuration, 112

with modified orography corresponding to the Callaghan Valley area. The 113

initial conditions and inflow lateral boundary conditions were based on ob-114

servations of the case. The control run and the sensitivity tests with dif-115

ferent prescribed precipitation rates (described below) were done using a 116

two-moment bulk microphysics scheme as described in Milbrandt and Yau 117

(2005b) (hereafter referred to as MY2). Given the sensitivity to the param-118

eterization of snow shown in Milbrandt et al. (2014), it should be noted that 119

the original version of the scheme has been used in this study. 120

The two-dimensional transect of interest is a vertical cross section oriented 121

north-south passing through the radar location (VVO) and the Callaghan 122

Valley (VOD) (Fig. 1b). The orography around the VOD station was cen-123

tered in the domain; with 30 km of simplified orography, 60 km of flat ter-124

rain upwind (south of VOD), and 140 km downwind. The orography field is 125

smoothed six times with a 1-2-6-2-1 filter and then interpolated to a 250 m 126

grid spacing. The smoothing is necessary to remove numerical noise. For nu-127

merical stability at the time of model set-up, the elevation of the downwind 128

side of the mountain was fabricated to continue the mountain topography. 129

This prevents having an abrupt variation of the orography at the base of the 130

mountain. The result is a near-bell shaped mountain (Fig. 4a) similar to 131

that used in the default WRF idealized two-dimensional case. 132

(8)

The domain was chosen to be sufficiently large to minimize reflections 133

from the lateral boundaries. Hence, the domain was set to 200 km with 134

a 250 m grid spacing. The inflow and outflow boundaries were both set 135

to open. To have a maximum number of vertical levels within the melting 136

layer, 72 vertical grids have been used where the grid spacing varied from 137

20 m to 750 m for the z <9 km and 750 m for z >9 km. The top of the 138

atmosphere is at 22 km. To minimize reflections from the lateral boundary, 139

a 10 km damping layer was used at the top of the model to minimize the 140

reflection of gravity waves from the upper boundary (Klemp et al., 2008). 141

The simulation was integrated with time steps of 1 s for a total of 8 h. 142

The Coriolis force and surface fluxes have been neglected in the simulations. 143

All clouds and precipitation are represented by the microphysics scheme, 144

which varies according to the experiment. No subgrid-scale condensation or 145

convective schemes were used. 146

Note that the two-dimensional nature of our runs neglects the effects of 147

the valley geometry on the thermodynamic and dynamic evolution of the 148

valley atmosphere. For instance, in valleys with sloping side-walls a volume 149

effect occurs that causes the valley to cool more rapidly by melting than a 150

plain or a valley with vertical walls (e.g. Steinacker, 1983; Unterstrasser and 151

Z¨angl, 2006). Thus by neglecting three-dimensional valley geometry we are 152

likely underestimating the cooling rate of the valley air. 153

2.2.2. Initial conditions 154

The vertical temperature profile measured from VOC at 0000 UTC and 155

0600 UTC are shown in Fig. 5. At 0000 UTC, a shallow melting layer was 156

present near the surface and 6 h later, that melting was replaced by a near 157

(9)

0◦C-isothermal layer (Fig. 5b). Note that the wind direction also changed 158

with time and elevation. For example, the flow changed from southerly wind 159

to northerly wind between 0000 UTC 14 February 2010 and 0600 UTC 14 160

February 2010 at lower levels but stayed from the south at higher levels. 161

The meteorological fields were initialized using the Whistler (VOC) sound-162

ing at 0000 UTC 14 February 2010 (Fig. 4 b-d). We assumed that the 163

meteorological conditions were similar in the Callaghan Valley at the onset 164

of precipitation. Note that the Callaghan Valley is located south-west of the 165

Whistler sounding station (VOC) and that VOT observed the rapid cooling 166

of air temperature 2 h later than in Callaghan valley. The observed meteoro-167

logical fields were smoothed to prevent numerical instabilities. A comparison 168

of the real and smoothed temperature, relative humidity and horizontal wind 169

speed vertical fields are shown in Figs. 4b, c and d, respectively. To pre-170

vent snow sublimation aloft, the relative humidity has been increased to 95 171

% where the snow field is initialized (section 2.3). Finally, the north-south 172

component of the wind speed was used. 173

For the control run, the MY2, which predicts the mixing ratio and to-174

tal number concentration of 6 hydrometeor categories: clouds droplets, rain, 175

pristine ice crystals, snow, graupel and hail, was used. From the precipita-176

tion sensor located at VOT, the precipitation rate was around 3 mm h−1 but 177

the quantitative precipitation forecast (QPF) suggests that more precipita-178

tion occurred in the Callaghan Valley (VOD) (Th´eriault et al., 2012). We 179

based our assumptions on QPF because no precipitation sensor was installed 180

at VOD. Therefore, it was assumed that snow is falling continuously from 181

above the melting layer to yield a surface precipitation rate of approximately 182

(10)

5 mm h−1. The snow field was initialized with a mixing ratio, qs = 1.25 g

183

kg−1, and total number concentration, Ns = 9860 m−3, at the model level

184

corresponding to an elevation of 2.3 km, based on observed radar reflectiv-185

ity and temperature, assuming the relation between the snow intercept size 186

distribution parameter and temperature from Thompson et al. (2004) (see 187

Milbrandt et al. (2014) for details). 188

2.3. Sensitivity experiments 189

First, to show the impact of the temperature feedback from melting snow 190

on the direction of the valley flow field, the control run was run while neglect-191

ing the temperature tendency due to melting snow. This sensitivity experi-192

ment has been repeated with another microphysical scheme, the Thompson 193

et al. (2008) referred to as THOMP. Furthermore, we also repeated the con-194

trol simulation with a bin microphysics scheme (e.g. Geresdi, 1998; Geresdi 195

et al., 2014) referred as GERBIN. However the experiment neglecting the 196

temperature feedback from melting was not performed due to the complex-197

ity of the scheme. 198

Second, the initial precipitation rates were varied to determine the sen-199

sitivity on the time at which the flow reversal/stagnation is reached. In 200

weather forecasting in British Columbia, a rule of thumb is commonly used 201

(Trevor Smith, personal communication Environment Canada, 2010) that a 202

precipitation rate of 3 mm h−1 or more may lead to a valley flow reversal 203

in complex terrain. The investigation of the precipitation rate thresholds 204

associated with the change of direction of the valley flow is conducted for 205

many initial snow mixing ratio, qs = 2.5, 1.875, 1.25, 0.625, 0.3125 g kg−1,

206

corresponding approximately to resulting surface precipitation rates of 10, 207

(11)

7.5, 5, 2.5 and 1.25 mm h−1, respectively. The initial wind speed will also 208

affect the change in the direction of the valley flow field but only the initial 209

precipitation rate has been studied. For each initial snow mixing ratio (or 210

precipitation rate), the timing of the flow reversal at different stages was 211

investigated by comparing the location of the 0◦C-line on the mountainside 212

as well as the location associated with mixed precipitation types (50% snow 213

and 50% rain). 214

3. Impact of melting snow on the valley flow field 215

3.1. Control simulation 216

To ensure that the model reproduced acceptable atmospheric conditions 217

we have studied the weather conditions at 60 min. This time was chosen 218

because the 0◦C-isotherm is located approximately between VVO and VOD 219

(Fig. 6). The top panel shows the vertical cross-section of the wind, temper-220

ature, and snow fields and the bottom panel the surface precipitation rates 221

along the domain cross-section. First, the maximum simulated wind speed 222

is in excess of 20 m s−1 above the barrier between 2 and 4.5 km. At this 223

time, the wind at VVO remained in the up-valley direction but had started 224

weakening with respect to the initial conditions. Second, the 0◦-isotherm 225

reached the ground approximately 5 km north of VVO on the mountainside. 226

It has lowered by 400 m within 60 min in the simulations. Third, the snow 227

field is initiated upstream of the barrier at altitude between 2.3-6 km with a 228

mixing ratio value decreasing with height. The mixing ratio was chosen to 229

yield a precipitation rate of 5 mm h−1 on the upstream side of the mountain. 230

Snow is advected up to 50 km downstream of the barrier by the southerly 231

(12)

wind. Note that the rain and graupel fields aloft are not shown here because 232

a specific attention is paid to the main atmospheric and precipitation fields 233

occurring over the domain. 234

The surface precipitation rates along the cross-section show a mixture of 235

rain and snow on the upstream side of the barrier (Fig. 6b). The precipita-236

tion changed to mainly snow with some graupel 5 km north of VVO, which 237

corresponds to the location of where the 0 ◦C-isotherm reached the ground. 238

Note that some graupel are produced by accretion with cloud droplets that 239

formed due to ascending air along the mountainside. Graupel are also formed 240

downwind, which is possibly caused by the updraft associated with gravity 241

waves. 242

Now that we have an overall view of the weather conditions along the 243

domain, the remaining analysis will focus on the horizontal distance in the 244

vicinity of VVO and VOD (-10 km <y <15 km) and up to an altitude of 3 245

km above sea level. 246

3.2. Effects of melting on the valley flow field 247

To assess the impact of melting snow on the direction of the valley flow 248

field, the temperature tendency due to the melting of snow was disabled in 249

the microphysics scheme and the simulation was re-run and compared with 250

the control run. A similar experiment was also conducted with the control 251

setup but with the temperature tendency due to evaporation and sublimation 252

suppressed. First, the impact of the processes are minor and were only 253

present at the beginning (first 20 min) of the simulation until the atmosphere 254

conditions are saturated. This was excepted since the atmosphere was so near 255

to saturation with respect to liquid water. 256

(13)

The impact of the cooling associated with the melting of snow is clearly 257

shown in Fig. 7. This illustrates the horizontal wind speed at two times 258

of interest, which are 120 and 210 min, during the simulation. These times 259

correspond to when the valley flow begins to change direction and when the 260

down valley flow is distributed throughout the depth of the melting layer, 261

respectively. The results are similar to the radar radial velocity fields at 262

times corresponding to the onset of the change of the valley flow direction. 263

For example, precipitation started at approximately 2230 UTC 13 February 264

2010 and the flow began to change direction 60 min later. Therefore, it took 265

90 min for the flow to change direction and fill the initial depth of the melting 266

layer. The simulations suggest there is also 90 min time lapse between the 267

start of the reversal and the moment when the valley flow field has entirely 268

changed direction. This timing is comparable to the observed radar-inferred 269

winds. Figure 7c and d show simulation results obtained when the diabatic 270

cooling due to melting snow is turned off. The direction and strength of the 271

valley flow remains constant throughout the simulation time. These results 272

suggest that the temperature feedback associated with melting snow has an 273

impact on the valley flow direction or stagnation. 274

Melting also affects the small-scale structure of the valley airflow. As 275

melting begins, the associated cooling is not uniform with height. This lo-276

calized cooling destabilizes the atmospheric profile (Findeisen, 1940). As a 277

result, shallow convection temporarily occurs within and below the melting 278

layer in our simulation. These can be seen in Fig. 8, which is discussed in the 279

next section. Such melting-induced convection has been simulated previously 280

using more idealized settings (e.g. Szyrmer and Zawadzki, 1999; Unterstrasser 281

(14)

and Z¨angl, 2006). While this has a notable but temporary effect on the valley 282

flow structure, it appears that convective overturning has little overall effect 283

on the rate at which the valley cools. This is consistent with the fact that 284

the two-dimensional simulation results are similar, in terms of temperature 285

changes and precipitation phase transition, to the one-dimensional results of 286

Milbrandt et al. (2014), where convection is absent. 287

3.3. Mechanisms of the valley flow stagnation and reversal 288

The above analysis demonstrates the connection between melting and the 289

valley flow, but what are the dynamical mechanisms whereby melting causes 290

the flow to stagnate and reverse? In our simulations the effects of surface 291

friction and Coriolis have been neglected. Thus, in a Lagrangian framework, 292

horizontal pressure gradients are the only term capable of decelerating the 293

horizontal momentum of an air parcel (baring substantial internal dissipa-294

tion). 295

To analyze horizontal pressure gradients and their causes, Fig. 8 shows 296

perturbation fields from the control experiment at t =120 and 240 min. Tem-297

perature and horizontal velocity are plotted as anomalies with respect to the 298

initial profile. Pressure is plotted as an anomaly with respect to the p(z) 299

profile on the upwind boundary at the given time (to focus on instantaneous 300

horizontal pressure gradients). All panels also show contours of equivalent 301

potential temperature. These provide approximate streamlines where the 302

flow is reversible moist-adiabatic. The vertical gradient also provides an ap-303

proximate measure of the moist static stability. By t = 120 min, melting has 304

nearly cooled all the air at the base of the mountain to 0◦C, but at locations 305

further upwind there is still a substantial layer of above-freezing air with the 306

(15)

remnants of melting-induced convection. Within the melting-cooled air is 307

a zone of decelerated flow, centered at x = 60 km (Fig. 8a). A secondary 308

zone of deceleration is found further up the mountain slope. A high-pressure 309

anomaly of up to 0.5 Pa is found over the mountain slopes extending upwind 310

(Fig. 8c). As flow along streamlines crosses isobars, air parcels increase their 311

pressure and decelerate (Fig. 8d). Near z = 1 km streamlines rise over the 312

layer of cooled and decelerated air just upwind of the base of the mountain. 313

At t = 240 min the flow configuration is similar, except the low-level high 314

pressure anomaly and region of decelerated flow have expanded upwind past 315

x = 40 km (Fig. 8g-h). This is coincident with a deepening of the layer of 316

cooled air in the same region (Fig. 8e). Streamlines begin to rise far upwind 317

of the mountain to surmount this layer. 318

When the temperature effect of melting is suppressed in the simulations, 319

low-level cooling upwind is eliminated. A localized cool anomaly is still found 320

above the windward slope (Fig. 9). A high-pressure anomaly is still found 321

over the terrain, but it is confined to the upper portion of the windward 322

slope. Flow deceleration is found near 1.5 km but stagnation and reversal 323

do not occur. Near z = 1 km there is a localized acceleration of the flow 324

associated with the lifting of an upwind near surface jet of faster flow. 325

What is the source of the horizontal variations in the low-level pressure 326

anomaly that act to decelerate the flow and the differences between the sim-327

ulations? One plausible hypothesis is that horizontal variations in snowfall 328

rate above the melting layer lead to horizontal variations in melting-induced 329

cooling that produce (by hydrostatic balance) horizontal pressure variations. 330

Such horizontal variations in precipitation rate could be produced by oro-331

(16)

graphic enhancement. However, Fig. 6 shows that the total precipitation 332

rate is essentially uniform across the windward slope. Thus, while melting of 333

orographically enhanced snowfall may sometimes enhance flow deceleration, 334

such a mechanism does not explain the results of our experiment. 335

An alternative way to produce horizontal pressure gradients and decel-336

erate the flow is through the lifting of stratified air. In a stratified atmo-337

sphere, local lifting over the windward slope of a mountain produces cold-338

temperature and high-density anomalies. If the atmosphere is hydrostatically 339

balanced, these density anomalies directly result in high-pressure anomalies 340

beneath them that can cause the low-level flow to decelerate, stagnate, or 341

reverse (e.g. Smith, 1988, 1989). For uniform upwind conditions, simple ter-342

rain geometry, and in the absence of latent heating and Coriolis forcing the 343

occurrence of low-level flow stagnation is controlled by the horizontal aspect 344

ratio of the terrain and the non-dimensional mountain height: H = N h/U , 345

where N is the Brunt-Vaisalla frequency, U is the horizontal wind speed, and 346

h is the height of the terrain (e.g. Pierrehumbert and Wyman, 1985; Smith, 347

1989; Smolarkiewicz and Rotunno, 1990). This parameter measures the non-348

linearity of the flow and stagnation tends to occur above some threshold 349

value of H that depends on the terrain shape. In scenarios such as the one 350

we consider here, condensation of water vapor releases latent heat and mod-351

ifies the flow dynamics, reducing windward pressure perturbations and flow 352

deceleration. For flows of near-saturated air, the effects of condensation can 353

be roughly accounted for in the above theory by replacing H with a moist 354

non-dimensional mountain height Hm = Nmh/U (e.g. Jiang, 2003) where

355

Nm is a moist version of the Brunt-Vaisalla frequency (Durran and Klemp,

(17)

1982). 357

The interaction of the above mechanism with melting can be understood 358

by considering the evolution of the lowest 1 km of the atmospheric profile 359

upwind of the terrain (Fig. 10). Initially the low-level flow is unsaturated. 360

The profile is stable with respect to dry lifting, as evidenced by the profile 361

of θ (Fig. 10 d). Although Nm2 is initially slightly negative below z = 1 362

km (Fig. 10 f), since the flow is unsaturated, the profile is actually moist 363

stable, evidenced by the constant θe in the lowest 0.4 km. As the simulation

364

progresses, snow falls into the lowest layers and melts (Fig. 10 b). This 365

gradually cools the air below 1 km to 0◦C (Fig. 10 a). The cooling brings 366

this layer to saturation and excess water vapor is condensed out (Fig. 10 c). 367

This isothermal layer increases the dry stratification as represented by θ and 368

the moist stratification as represented by θe and Nm (Fig. 10 d-f). When the

369

temperature effect from melting is suppressed, only very modest amounts 370

of moistening and cooling occur in the lowest few hundred meters due to 371

sublimation (Fig. 10 g-i). These changes only cause slight modifications to 372

the upwind stratification (Fig. 10 j-l). 373

To explore changes in the dynamical regime associated with the observed 374

changes in the upwind profile, values of Hm are calculated before and after

375

melting. These are computed by averaging the horizontal winds and Nm of

376

the upwind profile from the surface to 2 km, and setting h=1.2 km. Before 377

melting Hm = 1.9. Although this is above the typical threshold of about

378

H=0.85 for flow stagnation for a two-dimensional ridge (e.g. Huppert and 379

Miles, 1969), our results are not directly comparable with previous studies 380

due to the non-uniform vertical profile. More importantly, the excess Nm

(18)

produced by melting increases Hmto 3.3, indicating that melting has changed

382

the upwind profile in such a way that flow stagnation is much more favored. 383

In the no-temperature effect from melting snow experiment sublimation Hm

384

only increases to 2.2, a much more modest change. 385

When the temperature effect of melting is suppressed, the lack of a sub-386

stantial pressure perturbation upwind of about x = 65 km is due to the 387

minimal lifting upwind of this region and lack of stratification below z = 1 388

km (Fig. 9 c and f). In this experiment the lifting-induced pressure anomaly 389

causes deceleration that is focused above 1.2 km (Fig. 9 g and h). When 390

melting is included the air below z =1 km becomes much more strongly 391

stratified (Fig. 8). Lifting of this melting-cooled air near the base of the 392

mountain produces a pressure perturbation, which produces much more sub-393

stantial flow deceleration (Fig. 8 c-d and g-h). As air upwind lifts over the 394

decelerated air, pressure anomalies are produced upwind of the foot of the 395

terrain (Fig. 8 c and g). As found in previous studies (e.g. Pierrehumbert 396

and Wyman, 1985) the region of decelerated flow propagates far upwind of 397

the mountain. In this and other two-dimensional simulations without Corio-398

lis forcing the upwind propagation continues indefinitely (Pierrehumbert and 399

Wyman, 1985). Eventually, the windward deceleration proceeds to the point 400

where the windward flow stagnates and reverses (Fig. 8 h). 401

In summary, melting modifies the windward flow dynamics primarily via 402

its effects on the upwind atmospheric profile. By stratifying the low-level 403

air, melting moves the flow into a dynamical regime with high Hm where

404

low-level pressure perturbations produced by lifting are able to decelerate 405

the flow to the point of stagnation and reversal. 406

(19)

4. Factors impacting the timing of the valley flow field 407

4.1. Cloud and precipitation microphysics parameterization approach 408

To examine the sensitivity to different parameterizations of snow, the 409

same experiments were run using two other microphysics schemes. They are 410

the THOMP bulk scheme, which has some notable differences in treatment 411

of snow compared to MY2, and the bin-resolving scheme (GERBIN). 412

Each microphysics scheme used in this study (MY2, THOMP and GERBIN) 413

is constructed differently but considers the same microphysical processes for 414

melting snow. These processes are (1) equilibrium between melting rate of 415

snow, condensational heating and diffusion of sensible heat by conduction; 416

(2) mass conversion from accretion and collection of cloud droplets and rain 417

drops in the melting layer. However, the behavior of each scheme in this 418

intercomparison is different because of specific basic assumptions such as ini-419

tial size distribution, number of moments predicted and characteristics of 420

categories. For example, the amount of mass converted into rain depends on 421

the atmospheric conditions (wet bulb temperature) but this equation is inte-422

grated over an analytic size distribution (for the bulk microphysics scheme). 423

Therefore, even if the atmospheric conditions are the same, the amount of 424

mass melted into rain depends on the parameters of the size distribution (N0

425

and λ) assumed. In the case of a bin microphysics scheme, the evolution of 426

the precipitation characteristics is highly detailed and considered to be more 427

realistic because there is discrete number of mass bins of snow. This allows 428

for a more accurate representation of the melting rate throughout the size 429

distribution of melting snow (Geresdi et al., 2014). Furthermore, the treat-430

ment of melted water is different between bulk and bin scheme approaches. 431

(20)

For instance, in the bin schemes no shedding occur as opposed to bulk scheme 432

where the melted water immediately sheds off the snowflakes. This difference 433

affects the number and the mean size of rain drops produced by the melting 434

process. 435

As shown in Figure 11, all microphysics schemes produced valley flow 436

stagnation in the control case. On the other hand, when the temperature 437

tendency from melting snow was shut off, none of the schemes produced a 438

change in the valley flow direction. We were able to disable the temperature 439

feedbacks from melting snow in the bulk scheme only due to the complexity 440

of the parameterization in a bin-resolving approach. Wind speed starts to de-441

crease 20 min after temperature starts to decrease, finally reaching 0 m s−1at 442

180 min, 210 min and 200 min for the MY2, THOMP and GERBIN, respec-443

tively. The wind speed and temperature remain nearly constant throughout 444

the time evolution when the cooling rate from melting snow is turned off 445

(MY2 and THOMP only). For a precipitation rate of 5 mm h−1, the wind 446

speed decreases to 0 m s−1 2.5 to 3 h after the surface temperature started 447

to decrease depending on the microphysics scheme used. 448

In terms of the surface temperature evolution, as expected, all micro-449

physics parameterizations show a cooling when the temperature feedback 450

from melting is considered. The temperature starts to decrease at the sur-451

face after 20 min in the simulation. This initial decrease in temperature 452

early in the simulations is likely due to evaporation and sublimation, as the 453

atmosphere is not yet saturated with respect to liquid water. The cooling is 454

mainly due to the temperature feedback from the melting of snow because it 455

eventually reaches a constant value of 0◦C at 100 min for the MY2 scheme 456

(21)

and 130 min for the THOMP and 110 for the GERBIN schemes. The timing 457

of the cooling rates is faster for the MY2 scheme because snow is assumed to 458

fall faster than in THOMP and GERBIN schemes. However, when snow falls 459

in the melting layer, its terminal velocity is doubled when falling at temper-460

ature above 0 ◦C in THOMP whereas it remains the same in MY2. In that 461

case, the residence time of snow in the melting layer is reduced which de-462

creases the melting rate and therefore the temperature feedback from melting 463

snow. Furthermore, the difference between the bin-resolving approach and 464

the bulk are the assumptions associated with the ice category transferred to 465

liquid water category. While the bulk schemes continuously produce water 466

drops by shedding snow and graupel, the bin approach forms water drops at 467

low concentration and of relatively larger size throughout the melting process 468

(Geresdi et al., 2014). In GERBIN the melted snowflakes are transferred to 469

the water drop category if the fraction of the melted water is larger than 0.85. 470

Note that when the temperature is near 0◦C even the smaller snowflakes do 471

not melt completely, which is not the case in the bulk approach. 472

This impact on the temperature evolution directly affects the types of 473

precipitation reaching the surface. As expected, when the temperature feed-474

back from the melting snow is neglected, the temperature and wind speed do 475

not vary significantly so a mixture of rain and snow is produced at VVO with 476

5 times more wet snow compared to rain. Only the MY2 scheme produced 477

graupel at the surface. This could be due to the decrease in the strength of 478

the vertical motion over the barrier, and in turn in the amount of available 479

cloud water to enable snow conversion to graupel. Also, the treatment of 480

snow conversion to graupel are different in THOMP and MY2 as MY2 tends 481

(22)

to generally produce more graupels than THOMP. For the case where full 482

microphysics is considered, a transition from rain to snow is produced at the 483

base of the mountain in addition to a trace of graupel between 60 to 150 min. 484

Note that the rate of transition from snow to rain is similar. For instance, 485

the time at which the precipitation type at the surface is half rain and snow 486

occurs between 68 and 80 min for all schemes. The differences are caused 487

by the different parameterization of snow and by the different treatment of 488

shedding in the three approaches tested. 489

4.2. Prescribed snow field aloft 490

The time evolution of the direction of the valley flow field and the timing 491

of the 0◦C line traveling along the mountainside have been studied by varying 492

the initial precipitation rate prescribed. This has been conducted only with 493

the MY2 microphysics scheme. These sensitivity experiments allow us to 494

address the following questions: (1) Is there a minimum precipitation rates 495

that would cause reversal or stagnation of the valley flow field? (2) What 496

is the relative timing of the 0◦C-line descending the mountainside and the 497

complete reversal of the valley flow? 498

The change in the valley flow direction with full microphysics has been 499

studied for 6 prescribed precipitation rates. The times at which the valley 500

flow field starts to change direction and when the valley flow has completely 501

changed direction (<1 km ASL) have been calculated and are shown in Fig. 502

12a. The time at which the valley flow starts changing direction is defined by 503

the time at which the wind speed becomes negative below the 0◦C-isotherm 504

and between VVO and VOD. On the other hand, the timing associated with 505

completely reversed valley flow (<1 km ASL) was obtained based on the 506

(23)

maximum number of model grid points between VVO and VOD associated 507

with a negative wind speed. 508

First, the onset of the change in the wind direction occurs near the top 509

of the melting layer where cooling due to melting snow is present. As snow 510

continues to fall towards the surface, the change in the flow direction starts 511

to propagate downward. Second, all prescribed precipitation rates lead to a 512

change in the direction of the valley flow field but at different timings (Fig. 513

12a). Finally, the onset of direction change varies by 40 min from 1.25 mm 514

h−1 to 5 mm h−1 and this event occurs much faster for precipitation rates 515

>5 mm h−1. In particular, a precipitation of 10 mm h−1 has to be sustained 516

for nearly 2 h to start changing the direction of the valley wind whereas it 517

would take 4 h for a precipitation rate of 1.25 mm h−1. 518

Because of all the other processes influencing the temperature evolution 519

along the mountainside such as adiabatic cooling/warming, the complete 520

reversal of the flow occurred before the 0◦C-line reached the base of the 521

mountain. It takes nearly 360 min for the valley flow to completely change 522

direction at a precipitation rate of 1.25 mm h−1 and 180 min at 10 mm 523

h−1. Finally, the elapsed time between the onset of direction change and the 524

completion of the reversal increases with increasing precipitation rate. For 525

example, this time span is 65 min at 10 mm h−1 and 110 min at 1.25 mm 526

h−1. 527

Figure 12b plots the time when the 0◦C-line and the 50/50 rain/snow line 528

have, respectively, descended the mountainside under different precipitation 529

rates. As the precipitation rate increases, the time for the 0◦C-line to descend 530

the mountainside varied from 480 min for a 1.25 mm h−1 to 90 min for 10 531

(24)

mm h−1. The time for the 50/50 rain/snow line is shorter because that line is 532

located at lower elevation than the 0◦C-line. As expected, the time needed for 533

these lines to reach the base of the mountain is faster for higher precipitation 534

rate (10 mm h−1) than for lower ones (1.25 mm h−1). This is due to the fact 535

that low precipitation rate is associated with less mass melting into snow 536

and cooling off the top of the melting layer. Notice that the timing of change 537

in the flow direction and the timing of rain-snow boundary descending the 538

mountainside vary similarly. 539

The time associated with the change in the valley flow direction through-540

out the depth of the valley generally occurs after the 0◦C reached the base 541

of the mountain (Fig. 13). For example, the valley flow change direction 542

∼100 min after the 0◦C-line has reached the base of the mountain (VVO)

543

for a precipitation rate of 5 mm h−1. In this case, the onset of reversal and 544

the arrival of the 0 ◦C-line at the base of the mountain occur simultaneously. 545

On the other hand, the flow starts to reverse just when the 0◦C-line is still 546

going down the valley for a precipitation rate of 1.25 mm h−1. For the lowest 547

precipitation rate, the cooling rate is slower hence decreasing the lag time 548

between the cooling from melting snow and the change in the valley flow 549

direction. 550

5. Summary and conclusion 551

This study investigated the impact of the temperature feedback from 552

melting snow on the low-level flow field and precipitation phase transition 553

in complex terrain. In particular, experiments examining the temperature 554

feedbacks on the direction of the valley flow and the evolution of the rain-snow 555

(25)

line have been conducted. To address these issues, numerical simulations 556

using a semi-idealized setup have been used. The WRF model was initialized 557

with the vertical temperature, humidity and wind speed measured at VOC. 558

Snow was allowed to fall continuously from aloft at a constant rate on the 559

upstream side of the mountain, with no synoptic forcing. 560

A series of sensitivity experiments were conducted and the results showed 561

that the simulations with all of the microphysics schemes tested reproduced 562

a change in the valley flow field direction when the diabatic effect of melting 563

snow is considered. Although all schemes produced a slightly different timing. 564

Further sensitivity studies with different precipitation rates led to the 565

following conclusions: 566

• The timing of the flow reversal as well as the depth of the layer in 567

which it was produced was comparable to the radar data for the control 568

simulation. 569

• All precipitation rates tested produced a flow reversal/stagnation but 570

at different rates. However, the time to produce it is much longer for 571

lower precipitation rates. That means that calm synoptic conditions 572

need to be present of up to 8 h to observe a flow reversal produced by 573

a 1 mm h−1 snowfall rate. Since upwind evolution is key, adjusting 574

snow rate can only affect the timing of reversal, not its existence. Even 575

very light snowfall will eventually cool and stratify the air, leading to 576

reversal. However, in reality, only weak long lasting synoptic conditions 577

could lead to change in the flow direction in the valley. 578

• The speed of the rain-snow boundary traveling down the mountain 579

(26)

increases linearly with increasing the time for the valley flow to change 580

direction as the initial precipitation rate increases. It generally reaches 581

the base of the mountain before the flow reversal has completely fill up 582

the initial depth of the melting layer. 583

Further study should be conducted to verify the results with a full atmo-584

spheric model. It would be useful to study the impact of three-dimensional 585

valley geometry and the surrounding mountains on the timing of the flow 586

reversal. Also, an idealized study could be conducted to examine the rela-587

tive timing of large-scale warm air advection and cooling due to melting of 588

snow. For instance, given a synoptic forcing and a snowfall rate, one could 589

determine if the flow will reverse or not. 590

Overall, this study showed the importance that microphysical processes 591

can have on mesoscale flow and the conditions at the surface. This was 592

exemplified by the challenging prediction of local weather conditions during 593

the Vancouver 2010 Winter Olympics which was critical for safe and fair 594

competition. 595

Acknowledgments 596

We would like to thank the Natural Sciences and Engineering Research Coun-597

cil of Canada (NSERC) and the Fond Quebecois de la recherche sur la nature 598

et les technologies (FRQNT) for the financial support needed to accomplish 599

this work. The research made by N. Sarkadi and I. Geresdi was supported 600

by Hungarian Scientific Research Found (number: 109679). We would like to 601

acknowledge the contribution to constructive discussion with participants, in 602

particular Roy Rasmussen, at the 8th International Cloud Modeling Work-603

(27)

shop held in Warsaw, Poland in July 2012. The authors would like to ac-604

knowledge the SNOW-V10 principal and co-investigators for sharing the data 605

used in this study. 606

6. References 607

Barnett, T. P., Adam, J. C., Lettenmaier, D. P., 2005. Potential impacts of a 608

warming climate on water availability in snow-dominated regions. Nature 609

438, 303–309. 610

Durran, D. R., Klemp, J. B., 1982. On the effects of moisture on the brunt-v 611

aisa l a frequency. J. Atmos. Sci. 39 (10), 2152–2158. 612

Elsner, M., Cuo, L., Voisin, N., Deems, J., Hamlet, A., Vano, J., Mickel-613

son, K., Lee, S.-Y., Lettenmaier, D. P., 2010. Implications of 21st century 614

climate change for the hydrology of Washington State. Climatic Change 615

102 (1-2), 225–260. 616

Findeisen, C., 1940. The formation of the 0◦C isothermal layer and fractocu-617

mulus under nimbostratus. Meteorol. Z. 54, 49–54. 618

Geresdi, I., 1998. Idealized simulation of the colorado hail storm case: com-619

parison of bulk and detailed microphysics. Atmospheric Research 45, 237– 620

252. 621

Geresdi, I., Sarkadi, N., Thompson, G., 2014. Effect of the accretion by water 622

drops on the melting of snow flakes. Atmospheric Research, in press. 623

(28)

Henson, W., Stewart, R., Kochtubajda, B., Th´eriault, J., Sep. 2011. 624

The 1998 Ice Storm: Local flow fields and linkages to precipitation. 625

Atmospheric Research 101 (4), 852–862. 626

URL http://linkinghub.elsevier.com/retrieve/pii/S0169809511001645 627

Huppert, H., Miles, J., 1969. Lee waves in a stratified flow. Part 3: Semi-628

elliptical obstacle. Journal of Fluid Mechanics 35 (3), 481–496. 629

Isaac, G. A., Joe, P., Mailhot, J., Bailey, M., B´elair, S., Boudala, F., Brug-630

man, M., Campos, E., Carpenter, R., Crawford, R. W., Cober, S., Denis, 631

B., Doyle, C., Reeves, H., Gultepe, I., Haiden, T., Heckman, I., Huang, L., 632

Milbrandt, J., Mo, R., Rasmussen, R., Smith, T., Stewart, R. E., Wang, 633

D., 2014. Science of Nowcasting Olympic Weather for Vancouver 2010 634

(V10): A World Weather Research Programme project. SNOW-635

V10 Special Issue Pure and Appl. Geophys. 171, 1–24. 636

Jiang, Q., 2003. Moist dynamics and orographic precipitation. Tellus Series 637

A-dynamic Meteorology and Oceanography 55 (4), 301–316. 638

Joe, P., Scott, B., Doyle, C., Isaac, G. A., Cober, S., Mirecki, F., Gultepe, 639

I., Campos, E., Donaldson, N., Hudak, D., Stewart, R. E., 2014. The 640

monitoring network of the Vancouver 2010 Olympics. SNOW-V10 Special 641

Issue Pure and Appl. Geophys. 171, 25–58. 642

Klemp, J. B., Dudhia, J., Hassiotis, A. D., 2008. An upper gravity-wave 643

absorbing layer for nwp applications. Monthly Weather Review 136, 3987– 644

4004. 645

(29)

Lin, C. A., Stewart, R. E., 1986. Mesoscale circulations initiated by melting 646

snow. J. Geophys. Res. 91, 13299–13302. 647

Medina, S., Smull, B. F., R. A. Houze, M. S., 2005. Cross-barrier flow during 648

orographic precipitation events: Results from MAP and IMPROVE. J. 649

Atmos. Sci. 62. 650

Milbrandt, J., Th´eriault, J., Mo, R., 2014. Modeling the phase transition 651

associated with melting snow in a 1d kinematic framework: Sensitivity to 652

the microphysics. Pure and Applied Geophysics 171 (1-2), 303–322. 653

Milbrandt, J. A., Yau, M. K., 2005b. A multi-moment bulk microphysics 654

parameterization. Part II: A proposed three-moment closure and scheme 655

description. J. Atmos. Sci. 62, 3065–3081. 656

Minder, J. R., Durran, D. R., Roe, G. H., 2011. Mesoscale controls on the 657

mountainside snow line. J. Atmos. Sci. 68, 2107–2127. 658

Minder, J. R., Kingsmill, D. E., 2014/09/19 2013. Mesoscale variations of the 659

atmospheric snow line over the Northern Sierra Nevada: Multiyear statis-660

tics, case study, and mechanisms. Journal of the Atmospheric Sciences 661

70 (3), 916–938. 662

Pierrehumbert, R. T., Wyman, B., 1985. Upstream effects of mesoscale 663

mountains. J. Atmos. Sci. 42 (10), 977–1003. 664

Skamarock, W., Klemp, J., Dudhia, J., Gill, D., Barker, D., Duda, M., 665

Huang, X.-Y., Powers, J., 2008. A description of the advanced research 666

wrf version 3. Tech. rep., NCAR Techn. Note NCAR/TN-475+STR, 113 667

pp. 668

(30)

Smith, R., 1988. Linear theory of stratified flow past an isolated mountain 669

in isosteric coordinates. J. Atmos. Sci. 45, 3889–3896. 670

Smith, R., 1989. Mountain-induced stagnation points in hydrostatic flow. 671

Tellus 41a, 270–274. 672

Smolarkiewicz, P. K., Rotunno, R., 1990. Low froude-number flow past 3-673

dimensional obstacles .2. upwind flow reversal zone. J. Atmos. Sci. 47 (12), 674

1498–1511. 675

Steinacker, R., 1983. Diagnose und prognose der schneefallgrenze (diagnosing 676

and predicting the snowline). Wetter Leven 35, 81–90. 677

Steiner, M., O. Bousquet, R. A. H., Smull, B. F., Mancini, M., 2003. Air-678

flow within major alpine river valleys under heavy rainfall. Quart. J. Roy. 679

Meteor. Soc. 129, 411–431. 680

Szyrmer, W., Zawadzki, I., 1999. Modeling of the melting layer. Part I: dy-681

namics and microphysics. J. Atmos. Sci. 56, 3573–3592. 682

Th´eriault, J. M., Rasmussen, R., Smith, T., Mo, R., Milbrandt, J. A., Brug-683

man, M. M., Joe, P., Isaac, G. A., Mailhot, J., Denis, B., 2012. A case 684

study of processes impacting precipitation phase and intensity during the 685

vancouver 2010 winter olympics. Wea. Forecasting 27 (6), 1301–1325. 686

Thompson, G., Field, P. R., Rasmussen, R. M., Hall, W. D., 2008. Ex-687

plicit forecasts of winter precipitation using an improved bulk microphysics 688

scheme. Part II: Implementation of a new snow parameterization. Monthly 689

Weather Review 136 (12), 5095–5115. 690

(31)

Thompson, G., Rasmussen, R. M., Manning, K., 2004. Explicit forecasts of 691

winter precipitation using an improved bulk microphysics shceme. Part I: 692

Description and sensitivity analysis. Mon. Wea. Rev. 132, 519–542. 693

Unterstrasser, S., Z¨angl, G., 2006. Cooling by melting precipitation in alpine 694

valleys: An idealized numerical modelling study. Quarterly Journal of the 695

Royal Meteorological Society 132 (168), 1489–1508. 696

White, A. B., Gottas, D. J., Strem, E. T., Ralph, F. M., Neiman, P. J., 2002. 697

An automated brightband height detection algorithm for use with doppler 698

radar spectral moments. Journal of Atmospheric and Oceanic Technology 699

19, 687–697. 700

Z¨angl, G., 2007. Reversed flow in the south-alpine Toce valley during MAP-701

IOP 8: Further analysis of latent cooling effects. Q. J. R. Meteorol. Soc. 702

133, 1717–1729. 703

(32)

VOD

VOC

VVO

(a) (b)

VOT

Figure 1: (a) Western North America and (b) the Whistler area British Columbia. The Callaghan Valley is located at VOD, the radar was located at VVO and the soundings were launched from VOC. The precipitation rate shown in Fig. 2b was measured at VOT because no precipitation sensors were installed in the Callaghan Valley. Figure adapted from Th´eriault et al. (2012).

(33)

2100 UTC 13 Feb 0000 UTC 14 Feb 0300 UTC 14 Feb 0600 UTC 14 Feb ï5 ï2.5 0 2.5 5 Temperature ( o C)

(a) Temperature at VOD

2100 UTC 13 Feb0 0000 UTC 14 Feb 0300 UTC 14 Feb 0600 UTC 14 Feb

1 2 3 4 Precipitation Rate (mm h ï 1 ) Time (UTC)

(b) Precipitation rate at VOT

Figure 2: (a) The surface temperature evolution at VOD in the Callaghan Valley shows a rapid decrease in temperature. (b) The precipitation rate measured by FD12P at VOT, which is the base of Whistler Mountain. The onset of precipitation in the Whistler Area

(34)

!"# $"# %&'()*+,#-,.)/0,#(1#223#4567# "# "# !# $# 8#45 6 7# 23%# "# !# $# 9# 8#45 6 7# 23%# 9# !"# $"# %&'()*+,#-,.)/0,#(1#223#4567# "# "# !# $# 8#45 6 7# "# !# $# 9# 8#45 6 7# 23%# 9# 23%# !"# $"# %&'()*+,#-,.)/0,#(1#223#4567# "# "# !# $# 8#45 6 7# 23%# "# !# $# 9# 8#45 6 7# 23%# 9# !"# $"# %&'()*+,#-,.)/0,#(1#223#4567# "# "# !# $# 8#45 6 7# 23%# "# !# $# 9# 8#45 6 7# 23%# 9#

(a) 2230 UTC 13 Feb 2010 (b) 2300 UTC 13 Feb 2010 (c) 2330 UTC 13 Feb 2010 (d) 0100 UTC 14 Feb 2010

Legend: Radial Velocity

Reflectivity -20 -10 0 10 20 30 40 50 dBZ

Figure 3: The radar range-height indicator (RHI) of radar reflectivity (top panels) and radial doppler velocity (bottom panels). (a) is before precipitation started (b) is 30 min after the onset of precipitation (c) is when the valley flow starts to reverse and (d) is when the valley flow has completely change direction. This is a north-south cross-section looking northward from VVO into the Callaghan Valley.

(35)

ï60 ï40 ï200 0 20 40 60 80 100 120 140 2 4 6 8 10 12 14 16 18 20 22 T = 0oC (a)

Distance Relative to VVO (km)

Z (km) 0.5 0 10 20 30 1 1.5 2 2.5 VVO VOD T = 0oC Real IC ï60 ï40 ï20 0 (b) T (oC) 0 40 80 RH (%) (c) Real IC ï10 0 10 20 30 (d) Wind Speed (m/s)

Figure 4: An overview of the experimental design. (a) is the domain chosen and the real (blue) and smoother (black) topography. (b)-(d) are the observations measured by the sounding launched from VOC at 0000 UTC 14 February 2010 (fig 5a). (b) is the temperature, (c) is the relative humidity and (d) is the wind speed. The blue lines are the observed value and black are the WRF initial conditions.

(36)

−5 0 5 10 500 600 700 800 900 1000 Pressure hPa Temperature C (a) 0000 UTC 14 February 2010

−5 0 5 10 500 600 700 800 900 1000 Pressure hPa Temperature C (b) 0600 UTC 14 February 2010

Figure 5: The skew-T from Whistler area (VOC) at (a) 0000 UTC 14 February 2010 and at (b) 0600 14 February 2010.

(37)

Distance relative to VV0 (km) Z (km) 0oC −5oC −10oC 0.5 0.5 0.75 1 1 (a) VOD VVO −400 −30 −20 −10 0 10 20 30 40 1 2 3 4 m/s−15 −12 −9 −6 −3 0 3 6 9 12 15 18 21 24 27 −400 −30 −20 −10 0 10 20 30 1 2 3 4 5 6 7 Distance relative to VV0 (km) Precipitation Rate (mm/h) (b) R S G Total

Figure 6: Atmospheric conditions associated with the control run at 60 min. (a) The horizontal wind speed (filled contour), the temperature (white lines) and mass mixing

ratio of snow (g kg−1) (black lines) fields across the full horizontal domain but only up to

8 km above sea level. (b) The precipitation rate of snow, rain and graupel at the surface across the same domain.

(38)

Z (km) VOD (a) 0 5 10 15 20 0 1 2 3 m/s −15−12 −9 −6 −3 0 3 6 9 12 15 18 21 24 27 VOD (c) Distance relative to VV0 (km) Z (km) 0 5 10 15 20 0 1 2 3 VOD (b) Z (km) 0 5 10 15 20 0 1 2 3 VOD (d) Distance relative to VV0 (km) Z (km) 0 5 10 15 20 0 1 2 3 t =210 min t =120 min

Figure 7: The horizontal wind speed using the same color bar as the radar radial velocity to allow direct comparison. This is a sub-domain of Fig. 6 of comparable size to the radar image (Fig. 3) to facilitate the comparison of the fields. The top panels (a and b) are the control run, which includes all microphysical processes. The bottom panels (c and d) include the experiment suppressing the effect of cooling due to melting snow. The radar

is located at VVO at 0 km horizontal distance. The black line indicates the initial 0◦C

line. The left panels are the time associated with a change in the valley flow direction and the right panels are the time when the valley flow field has completely change direction.

(39)

t  =  120  min   t  =  240  min   −2 0 2 T′ (° C) 30 40 50 60 70 1 1.5 2 2.5 3 z (km) CTL56−1 zoom: 120 min −0.4 −0.2 0 0.2 w (m s − 1) 30 40 50 60 70 1 1.5 2 2.5 3 z (km) −0.2 −0.1 0 0.1 p (Pa) 30 40 50 60 70 1 1.5 2 2.5 3 z (km) −2 0 2 4 u (m s − 1) 30 40 50 60 70 1 1.5 2 2.5 3 x (km) z (km) −2 0 2 T′ (° C) 30 40 50 60 70 1 1.5 2 2.5 3 z (km) CTL56−1 zoom: 240 min −0.4 −0.2 0 0.2 w (m s − 1) 30 40 50 60 70 1 1.5 2 2.5 3 z (km) −0.2 −0.1 0 0.1 p (Pa) 30 40 50 60 70 1 1.5 2 2.5 3 z (km) −2 0 2 4 u (m s − 1) 30 40 50 60 70 1 1.5 2 2.5 3 x (km) z (km) θe   T=0°C   θe   T=0°C   qc  =  1x10-­‐6     θe   T=0°C   θe   u  =  0  m  s-­‐1   ‘   ‘   (a)   (b)   (c)   (d)   (e)   (f)   (g)   (h)  

Figure 8: Vertical sections showing perturbation fields from the control experiment. (a)-(d) show results at t = 120min. (e)-(h) show results at t=240min. (a) and (e) show

temperature perturbations with respect to the initial T (z, t0) profile. (b) and (f) show

vertical velocity. (c) and (g) show pressure perturbations with respect to the current

upwind profile p(z, t, xo). (d) and (h) show velocity perturbations with respect to the

initial u(z, t0) profile and a bold contour denotes regions of flow reversal. All panels

(40)

−2 −1 0 1 2 3 T′ (° C) 30 40 50 60 70 1 1.5 2 2.5 3 z (km) EXP56−2 zoom: 120 min −0.4 −0.2 0 0.2 w (m s − 1) 30 40 50 60 70 1 1.5 2 2.5 3 z (km) −0.2 −0.1 0 0.1 p (Pa) 30 40 50 60 70 1 1.5 2 2.5 3 z (km) −2 0 2 4 u (m s − 1) 30 40 50 60 70 1 1.5 2 2.5 3 x (km) z (km) −3 −2 −1 0 1 2 3 T′ (° C) 30 40 50 60 70 1 1.5 2 2.5 3 z (km) EXP56−2 zoom: 240 min −0.4 −0.2 0 0.2 w (m s − 1) 30 40 50 60 70 1 1.5 2 2.5 3 z (km) −0.2 −0.1 0 0.1 p (Pa) 30 40 50 60 70 1 1.5 2 2.5 3 z (km) −2 0 2 4 u (m s − 1) 30 40 50 60 70 1 1.5 2 2.5 3 x (km) z (km) t  =  120  min   t  =  240  min   θe   T=0°C   θe   T=0°C   qc  =  1x10-­‐6     θe   T=0°C   θe   u  =  0  m  s-­‐1   ‘   ‘   (a)   (b)   (c)   (d)   (e)   (f)   (g)   (h)  

Figure 9: As in Fig. 8, but showing perturbation fields from the no temperature feedbacks from melting snow simulation.

(41)

−3−2−1 0 1 2 3 4 0.6 0.8 1 1.2 1.4 (a) z (km) T (C) 0 0.5 1 (b) qs (g kg−1) 3.5 4 4.5 5 (c) qv (g kg−1) 279 282 285 288 (d) θ (K) 288 291 294 297 (e) θ e (K) −2 0 2 4 x 10−4 (f) Nm2 (s−2) −3−2−1 0 1 2 3 4 0.6 0.8 1 1.2 1.4 (g) z (km) T (C) 0 0.5 1 (h) qs (g kg−1) 3.5 4 4.5 5 (i) qv (g kg−1) 279 282 285 288 (j) θ (K) 288 291 294 297 (k) θ e (K) −2 0 2 4 x 10−4 (l) Nm2 (s−2)

Figure 10: Lower atmospheric profiles taken from upwind boundary. Profiles are plotted every 10 min from t = 0-240 min. Profiles from t=0 are red and successive profilers progress towards blue. (a)-(f) shows results from the control simulation. (g)-(l) show results from the no temperature feedbacks from melting snow simulation.

(42)

0 30 60 90 120 150 180 210 240 ï3 ï2 ï1 0 1 2 3 Time (min) Wind Speed (m/s) (a)

Tmelting & 0 Tmelting = 0

0 30 60 90 120 150 180 210 240 ï1 0 1 2 3 (c) Temperature ( o C) Time (min) 0 30 60 90 120 150 180 210 240 0 1 2 3 4 5 6 Time (min) Precipitation Rate (mm/h) (e) R S G 0 30 60 90 120 150 180 210 240 ï3 ï2 ï1 0 1 2 3 (b) Wind Speed (m/s) Time (min) 0 30 60 90 120 150 180 210 240 ï1 0 1 2 3 (d) Temperature ( o C) Time (min) 0 30 60 90 120 150 180 210 240 0 1 2 3 4 5 6 (f) Precipitation Rate (mm/h) Time (min)

Figure 11: The comparison of the surface (a) and (b) wind speed, (c) and (d) temperature as well as (e) and (f) precipitation rates of rain(R), snow (S) and graupel (G) at the base of the mountain (VVO). The results are obtained using the MY2 scheme (left column). The left columns are the results from the THOM microphysics scheme and the GERBIN micro-physics scheme (bold lines). The simulations with full microphysical processes (Tmelting 6= 0) and with suppressing the effect of cooling from melting snow (Tmelting = 0) are shown for MY and THOM but only the (Tmelting 6= 0) for the GERBIN.

(43)

0 1 2 3 4 5 6 7 8 9 10 0 60 120 180 240 300 360 420 480 R (mm hï1) Time (min) (a) Start to reverse Completely reversed 0 1 2 3 4 5 6 7 8 9 10 0 60 120 180 240 300 360 420 480 R (mm hï1) Time (min) (b) 0oCïline Rain/Snow

Figure 12: (a) The time needed for the valley flow field to start changing direction (Start to reverse) and to have changed direction throughout the depth of the melting layer at t = 0 (Completely Reversed) varying with the initial snow field aloft. (b) Time associated

(44)

0 60 120 180 240 300 360 420 480 0 60 120 180 240 300 360 420 480 Timing of 6 U (min)

Timing of rainïsnow boundary (min) Completely reversed Start to reverse

Figure 13: The relationship between the evolution of the valley flow field and the 0◦C

line reaching the base of the mountain. The dashed-line-circle is associated with the time needed for the valley flow field to start changing direction and the dashed-line-star is the time needed for the valley flow field to change direction throughout the depth of the melting layer at t = 0.

Figure

Figure 3: The radar range-height indicator (RHI) of radar reflectivity (top panels) and radial doppler velocity (bottom panels)
Figure 4: An overview of the experimental design. (a) is the domain chosen and the real (blue) and smoother (black) topography
Figure 5: The skew-T from Whistler area (VOC) at (a) 0000 UTC 14 February 2010 and at (b) 0600 14 February 2010.
Figure 6: Atmospheric conditions associated with the control run at 60 min. (a) The horizontal wind speed (filled contour), the temperature (white lines) and mass mixing ratio of snow (g kg −1 ) (black lines) fields across the full horizontal domain but on
+7

Références

Documents relatifs

during the collaborative building of an agricultural Community Information System - the Hien Valley experiment -1. a participative building approach adopted in the Hien

The main properties of this inner magnetospheric region are briefly recalled. The dynamical response of the plasma s phere to enhancements of geomagnetic activity

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

souligne aussi les limites. Un survey ne peut pas remplacer une fouille. Outre leur conservation différenciée qui crée d’inévitable biais, beaucoup de catégories de matériel

In summary, the axisymmetric squirmer model presented above predicts very well the behavior of the quadrupolar symmetry of the experimental flow field, captures the amplitude of

The well-posedness of a phase-field approximation to the Willmore flow with area and volume constraints is established when the functional approximating the area has no critical

As the rotation rate is further increased, this lifetime increases until a threshold is reached (the percolation threshold) where they finally form permanent turbulent spirals