1

Low-cloud, boundary layer, and sea ice interactions over the Southern Ocean

2

during winter

3



4

Casey J. Wall, Tsubasa Kohyama, and Dennis L. Hartmann

5

Department of Atmospheric Sciences, University of Washington, Seattle, Washington

6



7

Corresponding Author:

8

Casey J. Wall, [email protected]

9

Department of Atmospheric Sciences, University of Washington

10

408 ATG Building, Seattle, WA 98195-1640

11



12

Journal of Climate

13

Submitted June 30, 2016

14



15



16

Revision submitted January 19, 2017



1

17

Abstract

18



19

layer structure across the Antarctic sea ice edge is seen in ship-based measurements

20

and in active satellite retrievals from Cloud-Aerosol Lidar and Infrared Pathfinder

21

Satellite Observation (CALIPSO), which provide an unprecedented view of polar

22

clouds during winter. Sea ice inhibits heat and moisture transport from the ocean to

23

the atmosphere, and, as a result, the boundary layer is cold, stable and clear over sea

24

ice, and warm, moist, well-mixed and cloudy over open water. The mean low-cloud

25

fraction observed by CALIPSO is roughly 0.7 over open water and 0.4-0.5 over sea

26

ice, and the low-cloud layer is deeper over open water. Low-level winds in excess of

27

10 ms-1 are common over sea ice. Cold advection off of the sea ice pack causes

28

enhanced low-cloud fraction over open water, and thus an enhanced longwave

29

cloud radiative effect at the surface. Quantitative estimates of the surface longwave

30

cloud radiative effect contributed by low-clouds are presented. Finally, ten state-of-

31

the-art global climate models with satellite simulators are compared to

32

observations. Near the sea ice edge, seven out of ten models simulate cloudier

33

conditions over open water than over sea ice. Most models also underestimate low-

34

cloud fraction both over sea ice and over open water.

35

1. Introduction

During austral winter, a sharp contrast in low-cloud fraction and boundary

36

Sea ice, low-clouds, and the atmospheric boundary layer modulate the

37

climate of the Southern Ocean by influencing surface heat fluxes. During winter, sea

2

38

ice insulates the ocean from the cold atmosphere above, reducing the rate of ocean

39

heat loss at the surface by a factor 10-100 [Gordon, 1991]. Low-clouds and moisture

40

emit longwave (LW) radiation downward and heat the surface, and low-level winds

41

control the surface turbulent heat and moisture fluxes. When sea ice forms, brine is

42

rejected, adding salt to the near-surface waters. These processes modify the

43

buoyancy of surface waters and are responsible for deep and intermediate water

44

formation. Roughly two-thirds of the deep water in the global ocean is formed in the

45

Southern Ocean [Johnson, 2008], making it a region of critical importance for the

46

global overturning circulation of the ocean [Marshall and Speer, 2012; Talley, 2013].

47

Surface fluxes of heat and moisture in the polar regions are intimately linked to the

48

atmospheric boundary layer and to sea ice, are poorly observed, and are a topic of

49

high priority for improving our understanding of polar climate and climate change

50

[Bourassa et al., 2013].

51

Interactions between sea ice and boundary layer clouds have previously

52

been studied, but focus on this topic has generally been on the Arctic. Across the

53

Arctic basin during summer and early fall, low-clouds are more abundant and

54

optically thicker over open water than over sea ice when viewed from active

55

satellite remote sensing products [Kay and Gettelman, 2009; Palm et al., 2010] and

56

from surface observers [Eastman and Warren, 2010]. On the other hand, Schweiger

57

et al. [2008] used passive satellite retrievals and found that, during fall, regions of

58

low sea ice concentration coincide with enhanced mid-level cloudiness and reduced

59

low-cloud cover. Barton et al. [2012] found that the sensitivity of Arctic low-cloud

60

fraction to variations in sea ice concentration depends on synoptic regime. For



3

61

stable regimes, which support low-clouds, a significant but weak covariance

62

between sea ice concentration and cloud properties occurs during most seasons

63

[Taylor et al., 2015]. Near the sea ice edge, cold, off-ice advection is known to cause

64

enhanced low-cloud cover, but, due to a lack of observations, previous work has

65

focused on case studies of extreme events [e.g. Walter, 1980; Renfrew and Moore,

66

1999; Petersen and Renfrew, 2007]. It has also been argued that boundary layer

67

moisture, or lack thereof, triggers the onset of sea ice melt and freeze-up when

68

advected over Arctic sea ice [Kapsch et al., 2013] and that a warmer, moister

69

atmospheric boundary layer has amplified Arctic sea ice decline in recent years

70

[Serreze et al., 2009; Screen and Simmonds, 2010; Boisvert and Stroeve, 2015].

71

Unlike the Arctic, interactions between Antarctic sea ice and boundary layer

72

clouds have been given relatively little attention. Bromwich et al. [2012] first

73

pointed out that the total cloud fraction observed from active satellite retrievals is

74

about 0.1-0.2 lower in sea ice covered regions of the Southern Ocean than over open

75

water to the north. Fitzpatrick and Warren [2007] used ship-based measurements

76

of downwelling solar radiation over the Southern Ocean to show that, during austral

77

spring and summer, clouds tend to be optically thicker over open water than over

78

sea ice. It is not clear if the relationship between low-clouds and sea ice in the Arctic

79

is similar to that in the Southern Ocean.

80

In this study we describe the boundary layer properties and low-cloud

81

fraction across the Southern Ocean during winter, but with an emphasis on the

82

marginal sea ice zone. We use satellite-based active retrievals of clouds, which

83

provide an unprecedented view of polar clouds during winter, as well as ship-based



4

84

measurements. We also use a radiative transfer model to compute estimates of the

85

downward flux of LW radiation near the sea ice edge and its sensitivity to low-level

86

warm and cold advection. Finally, we evaluate ten state-of-the-art climate models.

87

This paper is organized as follows: datasets and the methodology of the radiative

88

transfer calculations are described in Section 2, results are given in Section 3, and

89

conclusions are given in Section 4.

90

2. Data and Methods

91

a. Datasets

92

Cloud observations are taken from the Cloud-Aerosol Lidar with Orthogonal

93

Polarization (CALIOP) instrument onboard the Cloud-Aerosol Lidar and Infrared

94

Pathfinder Satellite Observation (CALIPSO) satellite. CALIOP is a lidar that measures

95

high vertical resolution profiles of backscatter from which estimates of cloud

96

properties are derived [Winker et al., 2007]. Because CALIOP is an active

97

instrument, retrievals are not affected by lack of sunlight or near-surface inversions

98

– two conditions that are common at high latitudes during winter and are

99

problematic for passive satellite retrievals of low-clouds. As an example of the

100

challenge of cloud detection with passive instruments over the polar regions, Liu et

101

al. [2004] found that, during polar night, about 40% of all clouds went undetected

102

by the cloud mask algorithm of the moderate-resolution imaging spectroradiometer

103

(MODIS) used at that time. The algorithm has since been improved [Baum et al.,

104

2012], but detection of low-clouds over the polar regions remains a major challenge



5

105

for passive instruments [Ackerman et al., 2008]. Unlike passive instruments, the

106

signal-to-noise ratio of CALIOP is maximized in the absence of sunlight, making it

107

well suited for studying clouds during polar night.

108

We use the General Circulation Model-Oriented CALIPSO Cloud Product

109

(CALIPSO-GOCCP) version 2.9 [Chepfer et al., 2010; CALIPSO, 2015]. CALIPSO-

110

GOCCP provides cloud fraction on a 2° longitude, 2° latitude and 480 m height grid.

111

It also provides low (below 3.2 km), middle (3.2-6.5 km) and high (above 6.5 km)

112

cloud fraction, and estimates for how they are partitioned between liquid and ice.

113

For a given grid box and time interval, cloud fraction is defined as the number of

114

scenes in which a cloud was positively identified divided by the number of scenes in

115

which the lidar was not fully attenuated in the grid box. The lidar beam becomes

116

fully attenuated at an optical depth of ~3 [Winker et al., 2007], so CALIOP often does

117

not measure the bottom portion of low-clouds [Cesana et al., 2016]. The vertical

118

resolution of CALIOP is 30 m below 8 km and 60 m above 8 km, with a total of 583

119

vertical levels. Using the relatively coarse GOCCP vertical grid, which has only 40

120

levels, significantly increases the signal-to-noise ratio and provides a grid that is

121

better for comparison with global climate models [Chepfer et al., 2010].

122



123

National Oceanic and Atmospheric Administration/National Snow and Ice Data

124

Center Climate Data Record of Passive Microwave Sea Ice Concentration dataset

125

[Peng et al., 2013; NOAA/NSIDC, 2015], and cloud liquid water path from the Multi-

126

Sensor Advanced Climatology of Liquid Water Path dataset [Elsaesser et al., 2015].

127

Cloud liquid water path is defined as the total mass of cloud liquid water above a



We also use satellite-based observations of sea ice concentration from the

6

128

unit area of the earth’s surface. The latitude of the sea ice edge, which we define as

129

the northernmost point at which the sea ice concentration is 0.35, is computed from

130

the sea ice data. Our main conclusions are not sensitive to choosing a threshold of

131

0.50 or 0.25 for this definition. We also use temperature, specific humidity, and

132

wind fields from the European Center for Medium-Range Weather Forecasts

133

Reanalysis (ERA Interim; [Dee et al., 2011; ECMWF, 2015]). Boundary layer fields in

134

reanalysis products are poorly constrained by observations over the Southern

135

Ocean, and should be interpreted with caution. However, ERA-Interim is

136

consistently ranked among the most reliable reanalysis products in the high

137

southern latitudes [e.g. Bromwich et al., 2011; Bracegirdle and Marshall, 2012; Jones

138

et al., 2016].

139

All satellite and reanalysis data are analyzed on either monthly- or daily-

140

mean timescales during the months of June, July and August (JJA) from 2006

141

through 2014. ERA-Interim reanalysis data are available on monthly-mean and

142

instantaneous 6-hourly time resolutions, and daily-means are computed from the

143

instantaneous 6-hourly data. Monthly-mean fields are analyzed unless stated

144

otherwise.

145



146

speed from soundings, and cloud-base height measured by a ceilometer.

147

Measurements were made on two cruises that traversed the Weddell Sea during

148

June-August 2013 and May-August 1992 [König-Langlo et al., 2006; König-Langlo,

149

2013; König-Langlo, 2005]. The cruise tracks are shown in Figure 1. On the 2013

150

cruise, soundings were launched once per day and have a vertical resolution of



Additionally, we use ship-based observations of air temperature and wind

7

151

about 30 m in the lower troposphere, while, on the 1992 cruise, soundings were

152

launched four times per day and have a vertical resolution of about 60 m. Sounding

153

data are linearly interpolated to a vertical grid with a spacing of 30 m and 60 m on

154

the 2013 cruise and 1992 cruise, respectively. A total of 57 and 161 soundings were

155

taken poleward of 55°S on the 2013 cruise and 1992 cruise, respectively. Wind

156

speed measurements used global positioning system (GPS) technology on the 2013

157

cruise and the OMEGA radio navigation system on the 1992 cruise. As a result,

158

higher quality wind speed measurements were made on the 2013 cruise [König-

159

Langlo et al., 2006]. Measurements of cloud-base height were made on the 1992

160

cruise but not the 2013 cruise. These ship-based observations complement the

161

satellite observations because the soundings can resolve the vertical structure of the

162

boundary layer and the ceilometer measurements can reliably detect the cloud-base

163

height. The main weakness of the ship-based observations is that measurements are

164

sparse.

165



166

Coupled Model Intercomparison Project Phase 5 (CMIP5) [Taylor et al., 2012],

167

including output from the CALIPSO simulator [Chepfer et al., 2008]. Models in fully-

168

coupled and atmosphere-only configurations are evaluated, and the first ensemble

169

member for each model is used.

170

b. Radiative transfer modeling

171

Finally, we use output from ten global climate models that participated in the

One goal of this study is to quantify the downward flux of LW radiation at the

172

ocean surface, which we call “surface 𝐿𝑊↓ ”, and its dependence on low-cloud cover.

173

Because direct observations of surface 𝐿𝑊↓ are not available, we compute it using a

8

174

radiative transfer model. The advantage of using a radiative transfer model

175

compared to using reanalysis data is that we are able to vary low-clouds while

176

holding middle- and high-cloud cover fixed to zero. This method isolates the

177

contribution to surface 𝐿𝑊↓ made by low-clouds.

178

We use the Rapid and Accurate Radiative Transfer Model for Global Climate

179

Models [Mlawer et al., 1997; Clough et al., 2005; Iacono et al., 2008]. This model is

180

one-dimensional and takes vertical profiles of temperature, humidity, cloud liquid

181

and ice water content as inputs and computes surface 𝐿𝑊↓ . The temperature and

182

humidity profiles are taken from the ERA-Interim reanalysis, and for each profile

183

the surface 𝐿𝑊↓ is computed with a clear-sky and with low-cloud completely

184

obscuring the sky. We refer to these values as 𝐿𝑊↓,!"#$% and 𝐿𝑊↓,!"#$%&'( ,

185

respectively. A best estimate for the true value of the flux of downward LW

186

radiation at the surface, which we call 𝐿𝑊↓,!""!!"# , is also computed: 𝐿𝑊↓,!""!!"# = 𝐿𝑊↓,!"#$% (1 − 𝐿𝐶𝐹) + 𝐿𝑊↓,!"#$%&'( 𝐿𝐶𝐹

187

where 𝐿𝐶𝐹 is the low-cloud fraction observed by CALIPSO. Daily-mean data are used

188

because it is the shortest time resolution for which CALIPSO-GOCCP cloud

189

observations are available.

190

The radiative transfer calculations are done only for regions of open water

191

near the sea ice edge. We focus on scenes between 1°-3° equatorward of the ice

192

edge, where the sea ice concentration is approximately zero. This restriction is made

193

because microwave, satellite-based retrievals of cloud liquid water path are not

194

available over sea ice. Although the domain is restricted to open water scenes only,



9

195

the surface heat budget over open water is of interest because extreme air-sea heat

196

fluxes can occur there [e.g. Papritz et al., 2014].

197

In the radiative transfer calculations, several assumptions about low-clouds

198

are made that are based on observations presented in Table 1. First, low-clouds are

199

assumed to consist entirely of supercooled liquid. CALIPSO-GOCCP observations of

200

cloud fraction partitioned by phase are presented in Table 1 and show that low-

201

clouds over the Southern Ocean are frequently composed of liquid, consistent with

202

previous studies [Hu et al., 2010; Morrison et al., 2011]. Cesana et al. [2016] suggest

203

that low-clouds with unclassified phase in the CALIPSO-GOCCP dataset are mostly

204

mixed-phase clouds. Even so, at least three-quarters of the low-clouds detected by

205

CALIPSO are liquid (low-cloud fraction is 0.50 for liquid clouds and 0.69 for all

206

phases). Second, all cloud liquid water is assumed to reside in the lowest 3.2 km of

207

the atmosphere. Under this assumption the column-integrated cloud liquid water

208

path is equal to the total liquid water contained in low-clouds. This assumption is

209

justified by the CALIPSO phase observations (Table 1), which show that liquid

210

clouds are usually found at low levels. Third, the liquid water path is assumed to be

211

60 gm-2, which is near the median of satellite observations (Table 1). Fourth, the

212

height of low-cloud base and top are set to 500 m and 1000 m, respectively. The

213

cloud-base height value comes from measurements from the 1992 cruise taken

214

when surface observers reported open water or open pack ice near the ship (Table

215

1). During this cruise, observed cloud-base height was distributed nearly uniformly

216

between the surface and 1000 m. Fifth, low-clouds are assumed to have a droplet



10

217

effective radius of 16 𝜇m, which is close to the observed wintertime mean over the

218

Southern Ocean [McCoy et al., 2014, their Figure 9].

219



220

about low-cloud properties. In order to test the sensitivity of the radiative transfer

221

calculations to the assumptions about low-cloud properties, runs were performed

222

with cloud-base height set to 0 m and 1000 m, with liquid water path halved to 30

223

gm-2 and with cloud effective radius doubled to 32 𝜇m and halved to 8 𝜇m. The

224

results are not sensitive to modest changes in liquid water path because liquid

225

clouds are nearly opaque to LW radiation for liquid water path values greater than

226

~20 g/m2 [Hartmann, 2016], which is much lower than observed values (Table 1).

227

In fact, each modification of the assumptions resulted in a change in 𝐿𝑊↓,!"#$%&'( of

228

about 7 Wm-2 or less, which is small compared to the contribution to surface 𝐿𝑊↓

229

made by low-clouds during overcast conditions (~80 Wm-2 – discussed in the

230

Results section). In other words, to leading order, low-cloud fraction controls the

231

surface LW radiative effect from low-clouds. The model error is less than 1 Wm-2

232

[Mlawer et al., 1997], which is much smaller than the uncertainty due to the five

233

assumptions made about cloud properties.

234

3. Results

235

a. Satellite observations of sea ice and low-cloud over the Southern Ocean

236



237

shows July-average sea ice concentration over the Southern Ocean. Two contours of



Our results are insensitive to modest changes in these five assumptions



We start with a brief description of Antarctic sea ice during winter. Figure 1

11

238

sea ice concentration are shown: 0.35 and 0.95. These contours can be thought of as

239

marking the boundaries between open water, fragmented sea ice, and a sea ice pack

240

that covers the surface nearly completely. Throughout most of the Eastern

241

Hemisphere, sea ice concentration rarely exceeds 0.95. This could be a result of the

242

coastline extending equatorward and forcing the sea ice closer to the Antarctic

243

Circumpolar Current. In regions where the coastline cuts poleward, like the Weddell

244

and Ross Seas, sea ice concentrations greater than 0.95 are much more common.

245

Average sea ice concentrations in June and August are similar in this regard (not

246

shown). Wadhams et al. [1987] describe the winter sea ice pack in the Weddell Sea

247

as observed from a cruise. They found the marginal sea ice zone to be a ~250 km

248

band of fragmented pancake ice with pockets of exposed seawater. Farther south,

249

they found sea ice organized into vast floes that covered the ocean surface nearly

250

completely. We recommend viewing photographs of these features in Wadhams et

251

al. [1987, their Figure 12].

252

Now, consider low-cloud fraction over the Southern Ocean. The 2006-2014

253

winter climatology of low-cloud fraction and the latitude of the sea ice edge are

254

shown in Figure 2. The interannual standard deviation of the latitude of the sea ice

255

edge ranges between about 0.5° to 1.5° latitude. One standard deviation on either

256

side of the mean position of the sea ice edge is shaded in Figure 2 to show that the

257

effects of interannual variability of the location of the ice edge are likely small.

258

Throughout the Southern Ocean, cloudier conditions are seen over open water than

259

over sea ice. Near the sea ice edge, low-cloud fraction is about 0.7 over open water

260

and 0.5 over sea ice. The gradient of low-cloud fraction across the sea ice edge is



12

261

weakest in the Southern Indian and Western Pacific Oceans (20°E – 160°E). This

262

weak gradient is likely because the sea ice pack is more fragmented in this region

263

than elsewhere in the Southern Ocean (Figure 1). In this region, the low-cloud

264

fraction is more variable over sea ice than over open water because gaps in the sea

265

ice pack are found throughout the ice pack, but little sea ice is found equatorward of

266

the sea ice edge. In the Weddell and Ross Seas, where the sea ice pack covers the

267

surface nearly completely, the low-cloud fraction is about 0.4 or less and the

268

gradient in low-cloud fraction across the sea ice edge is sharp.

269

The relationship between sea ice concentration and low-cloud properties

270

near the sea ice edge is made clearer by stratifying the observations based on

271

distance from the sea ice edge. For each grid point and time (monthly-means from

272

June, July and August between 2006-2014 are considered), the meridional distance

273

between the grid point and the ice edge is computed. Data are then composited by

274

meridional distance from the ice edge, using a bin width of 0.5° latitude, and

275

averaged. We analyze data from the Weddell and Ross Seas (defined as 50°W-0°E

276

and 130°W-170°E, respectively; Figure 1), two regions where the sea ice pack

277

covers the surface nearly completely and where the sea ice edge is located far

278

offshore (Figure 1). This procedure was also done on the JJA-mean of each year, and

279

the main conclusions are the same using either monthly- or seasonal-averages.

280



281

troposphere over the Weddell Sea as a function of meridional distance from the sea

282

ice edge, and Figure 3d is similar but for the Ross Sea. On average, low-clouds

283

extend deeper and are more prevalent equatorward of the sea ice edge. The mean



Figure 3a shows the vertical profile of mean cloud fraction in the lower

13

284

low-cloud fraction and sea ice concentration are shown in Figure 3b and 3e, and the

285

method for deriving the confidence interval for the mean is described in the

286

Appendix. The domain can be split into three regions based on sea ice

287

concentration: an “ice” zone where sea ice concentration is ~1 that is located

288

poleward of 2° south of the ice edge, an “open water” zone where sea ice

289

concentration is ~0 that is located equatorward of 1° north of the ice edge, and a

290

“transition” zone between. Within the ice zone the mean low-cloud fraction is nearly

291

uniform at around 0.5, and within the open water zone the mean low-cloud fraction

292

is nearly uniform at around 0.7. The mean low-cloud fraction is significantly larger

293

in the open water zone than the ice zone. From south to north across the transition

294

zone, the low-cloud fraction increases smoothly as sea ice concentration decreases.

295

Figures 3c and 3f show vertical profiles of mean potential temperature and

296

specific humidity from reanalysis data as a function of meridional distance from the

297

ice edge. In current reanalysis data, the surface heat budget and the atmospheric

298

boundary layer over the Southern Ocean are poorly constrained by observations,

299

and therefore these data should be interpreted with caution. Nevertheless, the data

300

suggest several differences between the boundary layer over sea ice and over open

301

water. The lower troposphere is more stable over sea ice than over open water, as

302

can be seen by the vertical spacing in the potential temperature contours. Over open

303

water, near-surface temperatures are close to the freezing temperature of seawater,

304

and across the sea ice edge, near-surface temperatures drop rapidly. Boundary layer

305

specific humidity values are also nearly a factor of two larger over open water than

306

over sea ice.



14

307

b. Boundary layer structure from ship-based observations

308



309

further insight into the physical processes at work. In this section, sounding data are

310

represented by probability distributions. For each height measured by the

311

soundings, the probability distributions are computed by binning the data,

312

computing the number of observations in each bin, and normalizing by the total

313

number of soundings. Data are composited into measurements made between 55°-

314

65°S and poleward of 65°S. Because the sea ice edge is typically located between

315

60°-65°S in the Weddell Sea during winter, it is likely that most of the soundings

316

poleward of 65°S were taken over consolidated pack ice. Meanwhile, soundings

317

between 55°-65°S are likely a mixture of some taken over consolidated pack ice and

318

some taken where open water was exposed to the atmosphere.

319

Soundings resolve the vertical structure of the boundary layer and provide

Figure 4 shows the probability distribution of temperature at each height

320

between 10-1500 m. The 2013 and 1992 cruises are shown separately in Figure 4a-

321

b and 4c-d, respectively, because the cruises used different sounding technologies

322

[König-Langlo et al., 2006] and had different times between successive launches.

323

Two boundary layer regimes are seen: a warm and a cold mode. The warm mode is

324

characterized by having near-surface temperatures close to the freezing

325

temperature of seawater and by a moist adiabatic lapse rate above. In this regime

326

the boundary layer is well-mixed and moist. The cold mode is characterized by

327

typical near-surface temperatures of about -15°C to -25°C and by a low-level

328

inversion. Poleward of 65°S, the cold mode dominates (Figure 4b and 4d). Between

329

55°-65°S, both the warm and the cold modes are seen, albeit with different



15

330

likelihoods between the two cruises (Figure 4a and 4c). Differences in the relative

331

occurrence of the warm and cold mode in the 55°-65°S composite between the two

332

cruises could be a result of different weather events. The latitudinal distribution of

333

the warm and cold modes suggests that the cold mode forms over consolidated pack

334

ice, and the warm mode forms over open ocean or gaps in the sea ice.

335

The soundings also measured wind speed, and this is shown in Figure 5. The

336

probability distribution of wind speed as a function of height is shown for all

337

soundings taken poleward of 55°S. Sounding data are not composited by latitude

338

here, but doing so results in composites that resemble Figure 5 but are noisier (not

339

shown). The soundings reveal that wind speeds of 10 ms-1 or more are common at

340

heights of 200-600 m. For both cruises, the average wind speed between 200-600 m

341

is 10 ms-1 or more for 60-70% of the soundings. For the 2013 cruise, the strong low-

342

level winds are often associated with a low-level jet. On this cruise, the modal value

343

of wind speed is ~12-15 ms-1 at heights of 200-400 m and decreases with height to

344

~8 ms-1 at heights of 800-1000 m (Figure 5a). Data from both cruises show that

345

strong low-level winds are common during winter.

346

Low-level jets are of interest because they indicate the presence of a stable

347

boundary layer. Low-level jets exist at the top of stable boundary layers and, at least

348

in temperate latitudes, are initiated when the boundary layer transitions from

349

convective to stable. During this transition, the sudden shoaling of the boundary

350

layer causes a reduction in drag from turbulent momentum flux, and therefore a

351

sudden increase in wind speed, at heights above the stable boundary layer but

352

below the top of the former convective boundary layer. The stable boundary layer



16

353

limits drag on the winds above and allows the jet to persist and follow an inertial

354

oscillation [Blackadar 1957; Thorpe and Guymer, 1977]. The mechanisms that

355

initiate low-level jets over Antarctic sea ice during winter are not fully understood.

356

One possible mechanism is warm advection from open water to sea ice covered

357

regions, which temporarily deepens the boundary layer and then allows a new jet to

358

form when the boundary layer collapses to a stable profile [Andreas et al., 2000].

359

Another possible mechanism is motions arising from baroclinic instability

360

associated with the thermal contrast between sea ice and open ocean.

361

We emphasize that a weakness of this study is the short time span of

362

sounding data. Soundings were taken over a total of 50 days between 55°-65°S and

363

59 days poleward of 65°S. Despite this drawback, the main conclusions are robust:

364

in both cruises, a warm and a cold boundary layer regime are seen, and low-level

365

wind speeds in excess of 10 ms-1 are common.

366

c. Advection across the sea ice edge

367



368

sea ice edge, and vice versa? We start with an investigation of low-cloud fraction and

369

its sensitivity to advection across the sea ice edge. Cold air outbreaks, in which air is

370

advected from a cold land or ice surface to a warmer ocean, are known to cause the

371

development of low-clouds [e.g. Walter, 1980; Liu et al., 2006]. When the cold air

372

mass is heated from below by the warm ocean surface, convection occurs and low-

373

clouds form. Once formed, low-clouds are maintained by radiative cooling at cloud-

374

top, radiative heating at cloud-base, and the moisture source of the ocean.



How do clouds respond when cold air is advected equatorward, across the

17

375



376

distance from the sea ice edge, stratified by low-level advection across the sea ice

377

edge. As a metric for low-level advection across the sea ice edge, the meridional

378

wind at 1000 hPa is linearly interpolated to the latitude of the sea ice edge. We refer

379

to this value as 𝑣!"# !"#! . Data are composited into scenes in which 𝑣!"# !"#! is less

380

than −0.5𝜎 ≈ −3 ms-1 and greater than 0.5𝜎 ≈ 3 ms-1, where 𝜎 is the standard

381

deviation of 𝑣!"# !"#! . These composites correspond to on-ice flow and off-ice flow,

382

respectively. These composites are made using daily-mean data over the Weddell

383

Sea. The mean low-cloud fraction equatorward of the sea ice edge is significantly

384

larger during periods of off-ice flow than periods of on-ice flow. The peak in low-

385

cloud fraction during periods of off-ice flow is located at about 2° equatorward of

386

the ice edge, suggesting that low-clouds formed by cold advection can persist well

387

away from the sea ice edge. The fact that the peak in low-cloud fraction is about 2°

388

latitude equatorward of the sea ice edge may be a result of the predominant low-

389

cloud type transitioning from roll clouds near the sea ice edge to cellular convection

390

downstream [Walter, 1980]. This hypothesis is also consistent with the composites

391

in Figure 3a and Figure 3d, which show that, near the sea ice edge, the low-cloud

392

layer deepens toward the equator. Finally, for latitudes 2° south of the ice edge and

393

poleward, where sea ice covers the surface nearly completely (Figure 3b), there is

394

either no significant difference, or a very small difference, in low-cloud fraction

395

between the on-ice flow and off-ice flow composites. This result suggests that low-

396

clouds over open water are coupled to the surface and require the moisture source

397

of the open ocean to exist, and therefore dissipate when separated from open water.



Figure 6 shows the mean low-cloud fraction as a function of meridional

18

398

d. Impact of low-level advection on the surface heat budget

399

We have seen evidence of a warm and a cold boundary layer regime, and that

400

cold, low-level advection off of the sea ice pack causes low-clouds to form over open

401

water. How do low-level advection and the resulting boundary layer and low-cloud

402

changes impact the surface heat budget? To address this question we use a radiative

403

transfer model to compute surface 𝐿𝑊↓ over open water near the sea ice edge and to

404

estimate the contribution made by low-clouds. The Weddell Sea is again used as the

405

region of study. Recall that estimates of surface 𝐿𝑊↓ are computed for a clear sky,

406

with low-cloud completely covering the sky, and using low-cloud fraction observed

407

by CALIPSO. These values will be called 𝐿𝑊↓,!"#$% , 𝐿𝑊↓,!"#$%&'( , and 𝐿𝑊↓,!""!!"#

408

respectively. The 𝐿𝑊↓,!""!!"# values are the best estimate for the real world, while

409

the 𝐿𝑊↓,!"#$% and 𝐿𝑊↓,!"#$%&'( values help with interpretation. Also, recall that there

410

are no middle- or high-clouds in these calculations, so the radiative effects of low-

411

clouds are isolated here.

412

First, consider the average values of surface 𝐿𝑊↓ . The average values of

413

𝐿𝑊↓,!"#$% and 𝐿𝑊↓,!"#$%&'( are about 210 Wm-2 and 290 Wm-2, respectively. In other

414

words, if a point at the ocean surface was located under a clear sky, and a low-cloud

415

passed overhead, then the downward flux of LW radiation would suddenly increase

416

by about 80 Wm-2, a 40% increase from the clear-sky value. The average value of

417

𝐿𝑊↓,!""!!"# is around 270 Wm-2. The surface LW cloud radiative effect, defined as

418

𝐿𝑊↓,!""!!"# − 𝐿𝑊↓,!"#$% , is about 50-60 Wm-2. During winter, low-clouds warm the

419

ocean surface by about 50-60 Wm-2 on average.



19

420

Furthermore, surface 𝐿𝑊↓ depends on the strength of warm or cold

421

advection at low-levels. In the calculations of 𝐿𝑊↓,!"#$% and 𝐿𝑊↓,!"#$%&'( , temperature

422

and specific humidity are varied but low-cloud fraction is held fixed, and therefore

423

the surface LW cloud radiative effect is nearly constant. In the calculation of

424

𝐿𝑊↓,!""!!"# , temperature, humidity, and low-cloud fraction are all varied. Thus, by

425

comparing data from the 𝐿𝑊↓,!""!!"# , 𝐿𝑊↓,!"#$% and 𝐿𝑊↓,!"#$%&'( calculations, the

426

sensitivity of surface 𝐿𝑊↓ to low-cloud variations can be separated from the effects

427

of temperature and humidity variations.

428

Figure 7 shows surface 𝐿𝑊↓ plotted as a function of 𝑣!"# !"#! . In the 𝐿𝑊↓,!"#$%

429

and 𝐿𝑊↓,!"#$%&'( calculations, where low-cloud fraction is held fixed, the data are

430

anticorrelated with 𝑣!"# !"#! (𝑟 = −0.61 and 𝑟 = −0.64, respectively). This happens

431

because air masses that form over the sea ice pack are cold and have low specific

432

humidity, and these features of the air cause it to radiate relatively weakly to the

433

surface when advected over open water (when low-cloud fraction is held fixed).

434

However, in the 𝐿𝑊↓,!""!!"# calculation, where low-cloud fraction is varied according

435

to CALIPSO observations, data are weakly anticorrelated with 𝑣!"# !"#! (scatterplot

436

in Figure 7). When the 𝐿𝑊↓,!""!!"# data are binned by 𝑣!"# !"#! and averaged, the

437

result agrees well with a linear regression (compare the blue and red lines in Figure

438

7). Because cold advection causes cloudy conditions, 𝐿𝑊↓,!"!!!"# data approach the

439

𝐿𝑊↓,!"#$%&'( regression line for large positive values of 𝑣!"# !"#! . The surface LW

440

cloud radiative effect, seen in Figure 7 by the difference between 𝐿𝑊↓,!""!!"# and

441

𝐿𝑊↓,!"#$% , increases by 1.1 ± 0.1 Wm-2 per 1 ms-1 increase in 𝑣!"# !"#! . Put another



20

442

way, typical values of the average surface LW cloud radiative effect, estimated by

443

the regression, range from 43 Wm-2 to 65 Wm-2 for 𝑣!"# !"#! = −10 ms-1 to

444

𝑣!"# !"#! = 10 ms-1, respectively. As a result, the regression coefficient of 𝐿𝑊↓,!""!!"#

445

on 𝑣!"# !"#! (−0.7 ± 0.1 Wm-2 per 1 ms-1 increase in 𝑣!"# !"#! ) is significantly smaller

446

in magnitude than the regression coefficients of 𝐿𝑊↓,!"#$% and 𝐿𝑊↓,!"#$%&'( (compare

447

slopes in Figure 7). Therefore, when low-cloud fraction, temperature, and humidity

448

are all allowed to vary, as they are in the real world, then surface 𝐿𝑊↓ is much less

449

sensitive to warm or cold advection than when low-cloud fraction is held fixed. Low-

450

clouds warm the surface most strongly during cold advection events, and therefore

451

act to reduce the sensitivity of surface 𝐿𝑊↓ to cold advection.

452

We emphasize that these calculations are only able to capture one term of the

453

surface heat budget: surface 𝐿𝑊↓ . Surface turbulent heat fluxes are likely very

454

important as well. Over open water in the Southern Ocean during winter, average

455

values of surface turbulent fluxes of sensible and latent heat are around 30 Wm-2

456

and 50 Wm-2 respectively – on the order of the average surface LW cloud radiative

457

effect – but surface turbulent heat fluxes can be several hundred watts per square

458

meter during extreme cold air outbreaks [Papritz et al., 2014]. These turbulent heat

459

flux values come from reanalysis data, and should therefore be interpreted with

460

caution. Our work is progress towards constraining the Southern Ocean surface heat

461

budget, but a complete understanding also requires knowledge of the surface

462

turbulent heat flux.



21

463

e. Evaluation of Global Climate Models

464

We have seen that the surface LW cloud radiative effect from low-clouds is

465

about 50-60 Wm-2 on average. Therefore, low-cloud biases in global climate models

466

(GCMs) have the potential to significantly bias the modeled surface heat budget.

467

Accurate representation of marine boundary layer clouds and their radiative effects

468

are a major challenge for GCMs [Dufresne and Bony, 2008; Trenberth and Fasullo,

469

2010], and low-clouds are especially poorly represented in GCMs during polar night

470

[Karlsson and Svensson, 2011]. Here we evaluate the climatology of low-cloud

471

fraction near the sea ice edge in ten GCMs. Low-cloud fraction in the models is

472

computed by a CALIPSO simulator that estimates what CALIPSO would observe if it

473

were flying above the atmosphere in the model. Both fully-coupled and atmosphere-

474

only (AMIP) configurations are analyzed between 1990-2004 [Taylor et al., 2012;

475

Gates, 1992]. Fully-coupled models have prescribed atmospheric greenhouse gas

476

concentrations from observations, while atmosphere-only models have prescribed

477

sea surface temperature and sea ice concentration from observations. Most

478

importantly, interactions between the ocean, sea ice, low-clouds and the

479

atmospheric boundary layer are active in fully-coupled configuration and are

480

suppressed in atmosphere-only configuration. Model climatologies are compared to

481

the observed climatology between 2006-2014.

482

Figure 8 shows the July-mean low-cloud fraction observed by CALIPSO and

483

simulated by six GCMs that have output from both atmosphere-only and fully-

484

coupled configuration. Model behavior is quite diverse. Starting with the

485

atmosphere-only runs, HadGEM2, MIROC5 and MRI-CGCM3 have cloudier



22

486

conditions over open water than over sea ice. However, the gradient in low-cloud

487

fraction across the sea ice edge is weak in MIROC5. The IPSL models and MRI-

488

CGCM3 significantly underestimate low-cloud cover over open water. Compared to

489

the atmosphere-only runs, models in fully-coupled configuration generally have

490

similar low-cloud fractions, and the MIROC5 and MPI-ESM-LR models have a sea ice

491

edge that is located too close to the coastline. The models do not capture the weaker

492

low-cloud fraction gradient across the sea ice edge in the Eastern Hemisphere that is

493

seen in observations (Figure 2). Other than that, the modeled low-cloud fraction

494

does not have any systematic geographic bias.

495

Model bias in low-cloud fraction is quantified and shown in Figure 9. As a

496

metric for the low-cloud fraction near the sea ice edge, the mean low-cloud fraction

497

is computed over open water and sea ice, again defined as 1°-3° equatorward and

498

2°-4° poleward of the ice edge, respectively. These will henceforth be referred to as

499

LCFopen water and LCFsea ice. One common feature between models is a bias of too little

500

low-cloud fraction over open water, consistent with previous work [e.g. Zhang et al.,

501

2005]. In fact, over open water, nine out of ten models underestimate low-cloud

502

fraction, one model (GFDL-CM3) agrees with observations at 95% confidence, and

503

no models overestimate low-cloud fraction. This can be seen by noting that LCFopen

504

water for all of the models appears to the left of the observations in Figure 9a. The

505

magnitude of the largest model bias is about 0.35 (IPSL models), meaning that low-

506

cloud fraction in these models is roughly half of the observed value. Model bias in

507

LCFsea ice is more diverse and ranges from +.15 to -.26. However, out of the eight

508

fully-coupled models, all but one model underestimate LCFsea ice, one model agrees



23

509

with observations (MPI-ESM-LR), and no models overestimate LCFsea ice. This can be

510

seen by noting that all but one of the fully-coupled models are below the

511

observations in Figure 9a. Again, models that underestimate low-cloud fraction the

512

most have low-cloud fractions that are roughly half of the observed value. Most

513

models underestimate low-cloud fraction over open water and over sea ice.

514

Having established the mean low-cloud fraction bias in the models, we now

515

examine the difference of low-cloud fraction between the open water and sea ice

516

regions, which is shown in Figure 9b. Seven out of ten models have significantly

517

cloudier conditions over open water than over sea ice. Although most models

518

capture the correct sign of LCF!"#$ !"#$% − LCF!"# !"# , the magnitude varies between

519

0.38 in the MIROC-ESM and MIROC-ESM-CHEM models and -0.11 in the IPSL

520

models, while the observed value is LCF!"#$ !"#$% − LCF!"# !"# = 0.17 ± 0.01.

521

Finally, it is interesting to consider how low-clouds change with model

522

resolution and configuration. Models that have both atmosphere-only and fully-

523

coupled output provide an opportunity to compare low-clouds when the ocean and

524

sea ice are prescribed and when they are interactive. Of these six models that have

525

both atmosphere-only and fully-coupled output, three have little or no difference

526

between low-cloud fraction in fully-coupled and atmosphere-only configuration

527

(IPSL-CM5A-LR, IPSL-CM5A-MR and MRI-CGCM3). However, these models also

528

underestimate low-cloud fraction the most (Figure 8 and Figure 9a). The MIROC5

529

and MPI-ESM-LR models have more realistic values of LCF!"#$ !"#$% − LCF!"# !"# in

530

the atmosphere-only configuration than in the fully-coupled configuration (Figure

531

9b), but this may be due to the fact that the sea ice edge in these models is much



24

532

closer to shore, and therefore more exposed to cold continental air, in fully-coupled

533

configuration (Figure 8). Finally, the IPSL models make for an interesting

534

comparison because they differ only in resolution: IPSL-CM5A-LR and IPSL-CM5A-

535

MR have horizontal resolutions of 1.9°×3.75° and 1.25°×2.5°, respectively. In this

536

model, finer resolution does not improve the low-cloud bias.

537

4. Summary and conclusions

538

During austral winter, active satellite retrievals from CALIPSO and ship-

539

based measurements show a strong contrast in low-cloud fraction and boundary

540

layer structure over Antarctic sea ice and the adjacent open ocean. Low-cloud

541

fraction is roughly 0.7 over open water and 0.4-0.5 over sea ice, and the low-cloud

542

layer is much deeper over open water. The boundary layer is cold, stable, dry and

543

clear over consolidated sea ice and warm, moist, cloudy and well-mixed over open

544

water. At heights of 200-600 m, wind speeds in excess of 10 ms-1 are common over

545

sea ice, and are often associated with a low-level jet. During periods of cold, off-ice

546

advection, low-cloud fraction and the surface LW cloud radiative effect are

547

enhanced over open water. This enhanced cloud radiative effect acts to substantially

548

slow the rate of LW cooling of the ocean mixed layer compared to what would

549

happen if low-cloud fraction were uncorrelated with warm or cold advection. Low-

550

cloud fraction over sea ice is similar for on-ice and off-ice advection conditions,

551

indicating that low-clouds that form over the open ocean are coupled to the surface

552

and do not survive when separated from the moisture source provided by open

553

water.



25

554

These results support the hypothesis of two-way interactions between the

555

ocean surface and the atmospheric boundary layer during winter over polar oceans.

556

Regions of open water have relatively warm surface temperatures and large surface

557

fluxes of heat and moisture to the atmosphere. Moist and warm boundary layers

558

with a strong greenhouse effect form over open water and thus reinforce the warm

559

surface temperatures there. If such a region were to become covered by sea ice then

560

the surface heat and moisture fluxes would reduce and the boundary layer would

561

cool, dry, and become less cloudy, causing the greenhouse effect of the boundary

562

layer to weaken. This would reinforce the cool surface temperatures and help the

563

sea ice persist.

564



565

output were examined. Seven out of ten models simulate a larger low-cloud fraction

566

over open water than over sea ice. Nine out of ten models underestimate low-cloud

567

fraction over open water, and seven out of eight fully-coupled models

568

underestimate low-cloud fraction over sea ice. The observed low-cloud and

569

boundary layer properties shown in this work can be used as a test in future model

570

intercomparison projects when CALIPSO simulator output for more models is

571

available.



Additionally, ten state-of-the-art climate models with CALIPSO simulator



26

572

Acknowledgements

573



574

Terra and Aqua Science. Tsubasa Kohyama was supported by the National Science

575

Foundation (NSF) under grant AGS-1549579 and AGS-0960497, and the Takenaka

576

Scholarship Foundation. We are grateful to Steve Warren for helpful discussion,

577

Peter Blossey for sharing code for the radiative transfer model, Gregory Elsaesser

578

for providing the MAC-LWP data, and three anonymous reviewers for their

579

thorough and constructive feedback. We also thank the CALIPSO team and the

580

authors of [Chepfer et al., 2010] for creating the CALIPSO-GOCCP dataset. We

581

acknowledge the World Climate Research Programme's Working Group on Coupled

582

Modelling, which is responsible for CMIP, and we thank the climate modeling groups

583

for producing and making available their model output. For CMIP the U.S.

584

Department of Energy's Program for Climate Model Diagnosis and Intercomparison

585

provides coordinating support and led development of software infrastructure in

586

partnership with the Global Organization for Earth System Science Portals.



This research was conducted with support from NASA grant NNX14AJ26G,

27

587

Appendix

588

Derivation of 95% confidence interval for the mean low-cloud fraction

589

Figure 3b and Figure 3e show the mean low-cloud fraction as a function of

590

meridional distance from the sea ice edge in the Weddell Sea and Ross Sea,

591

respectively. The 95% confidence interval for mean low-cloud fraction is computed

592

assuming that low-cloud fraction measurements at each grid-cell and time are

593

independent. The low-cloud fraction data is available as monthly averages on a 2°

594

longitude by 2° latitude grid. To justify the assumption that measurements of low-

595

cloud fraction at each grid cell and time are independent we must assess spatial and

596

temporal autocorrelation.

597



598 •

Serial correlation in the meridional dimension

599

Our goal is to bin the data by meridional distance from the ice edge and

600

compute the mean and 95% confidence interval for the mean of each bin. The

601

CALIPSO grid is resolved in 2° latitude grid cells, and when compositing by

602

meridional distance from the ice edge we use bins of width 0.5° latitude. Therefore,

603

for any given time, no two grid cells of the same longitude and neighboring latitudes

604

can be assigned to the same bin. We therefore do not need to consider serial

605

correlation in the meridional dimension when estimating the effective degrees of

606

freedom of each bin.

607



608 •

Serial correlation in the zonal dimension



28

609

We estimate a lower bound for the number of degrees of freedom in the

610

zonal dimension by computing the correlation length scale 𝐿 for each latitude and

611

time, and comparing it to the resolution of the CALIPSO grid. Following [Taylor,

612

1921; Keller, 1935], we define the correlation length scale as

613



𝐿=

𝑟 𝜏 𝑑𝜏 !

614



615

where 𝑟(𝜏) is the spatial autocorrelation function of low-cloud fraction in the zonal

616

dimension and 𝜏 is the separation distance. The distance between independent

617

points in the zonal dimension can be thought of as 2𝐿. We computed 2𝐿 for each

618

time and latitude within 10 degrees of the sea ice edge, and the maximum value was

619

an order of magnitude smaller than the 2° longitude resolution of the CALIPSO grid.

620

We therefore treat each grid cell as an independent measurement of low-cloud

621

fraction in the zonal dimension.

622



623 •

Serial correlation in the time dimension

624

Each longitude and latitude grid-cell contains a timeseries of low-cloud

625

fraction observations. The number of effective degrees of freedom of the low-cloud

626

fraction timeseries (𝑁!"" ) is related to the number observations of low-cloud

627

fraction (𝑁) and the lag-1 autocorrelation of low-cloud fraction (𝑟! ) by the following

628

equation [Bretherton et al., 1999]:

629





29

𝑁!"" 1 − 𝑟! = 𝑁 1 + 𝑟! 630



631

Our goal is to bin the observations of low-cloud fraction by their meridional distance

632

from the ice edge and then to use this equation to estimate the effective degrees of

633

freedom for each bin. For each longitude and latitude grid cell we compute 𝑟! over

634

the entire timeseries. Then, for a given bin, say “bin A,” and a given grid cell, say

635

“grid cell B,” we keep track of the number of times that grid cell B is assigned to bin

636

A and then compute an estimate for the number of effective degrees of freedom

637

contributed to bin A by grid cell B by scaling by the right hand side of the above

638

equation. This procedure is done for every grid cell and every bin.

639

The total number of effective degrees of freedom for each bin estimated by this

640

procedure is slightly greater than if one were to assume each observation of low-

641

cloud fraction is independent in the time dimension. This happens because the lag-1

642

autocorrelation of low-cloud fraction over the domain tends to be slightly negative,

643

probably because of random sampling variability. We therefore assume that each

644

estimate of low-cloud fraction is independent in the time dimension.



30

645

References

646

Ackerman, S., et al. (2008). Cloud detection with MODIS. Part II: Validation. Journal of Atmospheric and

647

Oceanic Technology, 25(7), 1073–1086. doi:10.1175/2007JTECHA1053.1

648

Andreas, E. L., Claffey, K. J., & Makshtas, a. P. (2000). Low-level atmospheric jets and inversions over the

649

western Weddell Sea. Boundary-Layer Meteorology, 97(3), 459–486. doi:10.1023/A:1002793831076

650

Barton, N. P., Klein, S. a., Boyle, J. S., & Zhang, Y. Y. (2012). Arctic synoptic regimes: Comparing

651

domain-wide Arctic cloud observations with CAM4 and CAM5 during similar dynamics. Journal of

652

Geophysical Research Atmospheres, 117(15), 1–19. doi:10.1029/2012JD017589

653

Baum, B., et al. (2012). MODIS cloud-top property refinements for collection 6. Journal of Applied

654 655

Meteorology and Climatology, 51(6), 1145–1163. doi:10.1175/JAMC-D-11-0203.1

Blackadar, A. K. (1957). Boundary Layer Wind Maxima and their Significance for the Growth of

656 657

Nocturnal Inversions, Bull. Amer. Meteorol. Soc. 38, 283–290.

Boisvert, L. N., & Stroeve, J. C. (2015). The Arctic is becoming warmer and wetter as revealed by the

658

Atmospheric Infrared Sounder, Geophysical Research Letters 4439–4446.

659

doi:10.1002/2015GL063775

660

Bourassa, M. A., et al. (2013). High-latitude ocean and sea ice surface fluxes: Challenges for climate

661

research. Bulletin of the American Meteorological Society, 94(3), 403–423. doi:10.1175/BAMS-D-

662

11-00244.1

663

Bracegirdle, T. J., & Marshall, G. J. (2012). The Reliability of Antarctic Tropospheric Pressure and

664

Temperature in the Latest Global Reanalyses. Journal of Climate, 25(20), 7138–7146.

665

doi:10.1175/JCLI-D-11-00685.1



31

666

Bretherton, C. S., Widmann, M., Dymnikov, V. P., Wallace, J. M., & Blade, I. (1999). Effective number of

667

degrees of freedom of a spatial field. Journal of Climate, 12(1969), 1990–2009. doi:10.1175/1520-

668

0442(1999)012<1990:Tenosd>2.0.Co;2

669

Bromwich, D. H., Nicolas, J. P., & Monaghan, A. J. (2011). An Assessment of Precipitation Changes over

670

Antarctica and the Southern Ocean since 1989 in Contemporary Global Reanalyses. Journal of

671

Climate, 24(16), 4189–4209. doi:10.1175/2011JCLI4074.1

672

Bromwich, D. H., et al. (2012). Tropospheric clouds in Antarctica. Reviews of Geophysics, 50(1), 1–40.

673 674

doi:10.1029/2011RG000363

CALIPSO, 2015: GOCCP, Subset: Monthly, June, 2006-August, 2014. CFMIP-Obs: Cloud Observations

675

for Model Evaluation. Accessed 1 January, 2016. [available online at

676

http://climserv.ipsl.polytechnique.fr/cfmip-obs/]

677

Cesana, G., et al. (2016). Using in situ airborne measurements to evaluate three cloud phase products

678

derived from CALIPSO. Journal of Geophysical Research: Atmospheres, 121(10), 5788–5808.

679

doi:10.1002/2015JD024334

680

Chepfer, H., et al. (2008). Use of CALIPSO lidar observations to evaluate the cloudiness simulated by a

681 682

climate model. Geophysical Research Letters, 35(15), 1–6. doi:10.1029/2008GL034207

Chepfer, H., et al. (2010). The GCM-oriented CALIPSO cloud product (CALIPSO-GOCCP). Journal of

683 684

Geophysical Research Atmospheres, 115(5), 1–13. doi:10.1029/2009JD012251

Clough, S.A., et al. (2005). Atmospheric radiative transfer modeling: a summary of the AER codes, J.

685 686

Quant. Spectrosc. Radiat. Transfer., 91, 233-244

Dee, D. P., et al. (2011). The ERA-Interim reanalysis: Configuration and performance of the data

687

assimilation system. Quarterly Journal of the Royal Meteorological Society, 137(April), 553–597.

688

doi:10.1002/qj.828



32

689

Dufresne, J. L., & Bony, S. (2008). An assessment of the primary sources of spread of global warming

690

estimates from coupled atmosphere-ocean models. Journal of Climate, 21(19), 5135–5144.

691

doi:10.1175/2008JCLI2239.1

692

Eastman, R., & Warren, S. G. (2010). Interannual variations of arctic cloud types in relation to sea ice.

693 694

Journal of Climate, 23(15), 4216–4232. doi:10.1175/2010JCLI3492.1

ECMWF, 2015: ERA-Interim Project. Subset: Monthly, June 2006-August 2014. Accessed 1 January,

695

2016. [available online at http://apps.ecmwf.int/datasets/]

696

Elsaesser, G., C. O'Dell, and J. Teixeira (2015). Algorithm Theoretical Basis Document (ATBD) Version

697

1: Multi-Sensor Advanced Climatology of Liquid Water Path (MAC-LWP). An Enhanced Climate

698

Data Record (CDR) for MEaSUREs Project “A Data Record of the Cloudy Boundary Layer”. 20

699

pps.

700

Fitzpatrick, M. F., & Warren, S. G. (2007). The Relative Importance of Clouds and Sea Ice for the Solar

701 702

Energy Budget of the Southern Ocean. Journal of Climate, 20(6), 941–954. doi:10.1175/JCLI4040.1

Gates, W. L. (1992), AMIP: The Atmospheric Model Intercomparison Project. Bull. Amer. Meteor. Soc.,

703 704

73, 1962-1970

Gordon, A. L. (1991), Two stable modes of Southern Ocean winter stratification, in Deep Convection and

705

Deep Water Formation in the Oceans, Elsevier Oceanography Series, vol. 57, edited by P. C. Chu

706

and J. C. Gascard, pp. 17–35, Elsevier, New York.

707

Hartmann, D. L. (2016). Global Physical Climatology, Second Ed., 77 pp., Academic Press, Amsterdam

708

Hu, Y., et al. (2010). Occurrence, liquid water content, and fraction of supercooled water clouds from

709

combined CALIOP/IIR/MODIS measurements. Journal of Geophysical Research: Atmospheres,

710

115(19), 1–13. doi:10.1029/2009JD012384



33

711

Iacono, M.J., J.S. Delamere, E.J. Mlawer, M.W. Shephard, S.A. Clough, and W.D. Collins, (2008).

712

Radiative forcing by long-lived greenhouse gases: Calculations with the AER radiative transfer

713

models, J. Geophys. Res., 113, D13103, doi:10.1029/2008JD009944

714

Johnson, G. C. (2008). Quantifying Antarctic Bottom Water and North Atlantic Deep Water volumes.

715

Journal of Geophysical Research: Oceans, 113(5), 1–13. doi:10.1029/2007JC004477

716

Jones, R. W., I. A. Renfrew, A. Orr, B. G. M. Webber, D. M. Holland, and M. A. Lazzara (2016),

717

Evaluation of four global reanalysis products using in situ observations in the Amundsen Sea

718

Embayment, Antarctica, J. Geophys. Res. Atmos., 121, doi:10.1002/2015JD024680.

719

Kapsch, M.-L., Graversen, R. G., & Tjernström, M. (2013). Springtime atmospheric energy transport and

720

the control of Arctic summer sea-ice extent. Nature Climate Change, 3(8), 744–748.

721

doi:10.1038/nclimate1884

722

Karlsson, J., & Svensson, G. (2011). The simulation of Arctic clouds and their influence on the winter

723

surface temperature in present-day climate in the CMIP3 multi-model dataset. Climate Dynamics,

724

36(3), 623–635. doi:10.1007/s00382-010-0758-6

725

Kay, J. E., & Gettelman, A. (2009). Cloud influence on and response to seasonal Arctic sea ice loss.

726 727

Journal of Geophysical Research Atmospheres, 114(18), 1–18. doi:10.1029/2009JD011773

Keller, L. B., 1935: Expanding of limit theorems of probability theory on functions with continuous

728 729

arguments (in Russian). Works Main Geophys. Observ., 4, 5–19. Kikkawa,

König-Langlo, Gert (2005): Meteorological observations during POLARSTERN cruise ANT-X/4. Alfred

730

Wegener Institute, Helmholtz Center for Polar and Marine Research, Bremerhaven,

731

doi:10.1594/PANGAEA.269526 [available online at

732

https://pangaea.de/PHP/CruiseReports.php?b=Polarstern]



34

733

König-Langlo, Gert (2013): Meteorological observations during POLARSTERN cruise ANT-XXIX/6

734

(AWECS). Alfred Wegener Institute, Helmholtz Center for Polar and Marine Research,

735

Bremerhaven, doi:10.1594/PANGAEA.819610 [available online at

736

https://pangaea.de/PHP/CruiseReports.php?b=Polarstern]

737

König-Langlo, G., Loose, B., & Bräuer, B. (2006). 25 Years of Polarstern Meteorology. WDC-MARE

738

Reports. doi:10.2312/wdc-mare.2006.4

739

Liu, Y., Key, J. R., Frey, R. a., Ackerman, S. a., & Menzel, W. P. (2004). Nighttime polar cloud detection

740

with MODIS. Remote Sensing of Environment, 92(2), 181–194. doi:10.1016/j.rse.2004.06.004

741

Liu, a. Q., Moore, G. W. K., Tsuboki, K., & Renfrew, I. a. (2006). The effect of the sea-ice zone on the

742

development of boundary-layer roll clouds during cold air outbreaks. Boundary-Layer Meteorology,

743

118(3), 557–581. doi:10.1007/s10546-005-6434-4

744

Marshall, J., & Speer, K. (2012). Closure of the meridional overturning circulation through Southern Ocean

745 746

upwelling. Nature Geoscience, 5(3), 171–180. doi:10.1038/ngeo1391

Mccoy, D. T., Hartmann, D. L., & Grosvenor, D. P. (2014). Observed Southern Ocean Cloud Properties

747

and Shortwave Reflection Part I : Calculation of SW flux from observed cloud. Journal of Climate,

748

27, 8836–8857. doi:10.1175/JCLI-D-14-00287.1

749

Mlawer, E.J., S.J. Taubman, P.D. Brown, M.J. Iacono and S.A. Clough (1997). RRTM, a validated

750 751

correlated-k model for the longwave. J. Geophys. Res., 102, 16,663-16,682

Morrison, A. E., Siems, S. T., & Manton, M. J. (2011). A three-year climatology of cloud-top phase over

752

the Southern Ocean and North Pacific. Journal of Climate, 24, 2405–2418.

753

doi:10.1175/2010JCLI3842.1

754

Nicolas, J. P., & Bromwich, D. H. (2011). Climate of West Antarctica and influence of marine air

755

intrusions. Journal of Climate, 24(1), 49–67. doi:10.1175/2010JCLI3522.1



35

756

NOAA/NSIDC Climate Data Record of Passive Microwave Sea Ice Concentration, Version 2, 2015.

757

Subset: Monthly, June 2006-August, 2014. National Snow and Ice Data Center, accessed 1 January,

758

2016. [available online at https://nsidc.org/data]

759

Palm, S. P., Strey, S. T., Spinhirne, J., & Markus, T. (2010). Influence of Arctic sea ice extent on polar

760

cloud fraction and vertical structure and implications for regional climate. J. Geophys. Res.,

761

115(D21), D21209–. doi:10.1029/2010JD013900

762

Papritz, L., Pfahl, S., Sodemann, H., & Wernli, H. (2014). A climatology of cold air outbreaks and their

763

impact on air-sea heat fluxes in the high-latitude South Pacific. Journal of Climate, 28(1), 342–364.

764

doi:10.1175/JCLI-D-14-00482.1

765

Peng, G., Meier, W. N., Scott, D. J., & Savoie, M. H. (2013). A long-term and reproducible passive

766

microwave sea ice concentration data record for climate studies and monitoring. Earth System

767

Science Data, 5(2), 311–318. doi:10.5194/essd-5-311-2013

768

Peterson, G. N., & Renfrew, I. A. (2007). Aircraft-based observations of air-sea fluxes over Denmark Strait

769

and the Irminger Sea during high wind speed Conditions. Quarterly Journal of the Royal

770

Meteorological Society, 133(October), 937–948. doi:10.1002/qj.355

771

Renfrew, I. a., & Moore, G. W. K. (1999). An Extreme Cold-Air Outbreak over the Labrador Sea: Roll

772

Vortices and Air–Sea Interaction. Monthly Weather Review, 127(10), 2379–2394. doi:10.1175/1520-

773

0493(1999)127<2379:AECAOO>2.0.CO;2

774

Schweiger, A. J., Lindsay, R. W., Vavrus, S., & Francis, J. a. (2008). Relationships between Arctic sea ice

775

and clouds during autumn. Journal of Climate, 21(18), 4799–4810. doi:10.1175/2008JCLI2156.1

776

Taylor, G. I., 1921: Diffusion by continuous movement. Proc. London Math. Soc., 20 (2), 196–212.

777

Taylor, K. E., Stouffer, R. J., & Meehl, G. a. (2012). An overview of CMIP5 and the experiment design.

778

Bulletin of the American Meteorological Society, 93(4), 485–498. doi:10.1175/BAMS-D-11-00094.1



36

779

Taylor, P. C., Kato, S., Xu, K. M., & Cai, M. (2015). Covariance between Arctic sea ice and clouds within

780

atmospheric state regimes at the satellite footprint level. Journal of Geophysical Research:

781

Atmospheres, 120(24), 12656–12678. doi:10.1002/2015JD023520

782

Talley, L.D. (2013). Closure of the global overturning circulation through the Indian, Pacific, and Southern

783

Oceans: Schematics and transports. Oceanography 26(1):80–97,

784

http://dx.doi.org/10.5670/oceanog.2013.07.

785

Thorpe, A. J. and Guymer, T. H. (1977). ‘The Nocturnal Jet’, Quart. J. Roy. Meteorol. Soc. 103, 633–653.

786

Trenberth, K. E., & Fasullo, J. T. (2010). Simulation of present-day and twenty-first-century energy

787

budgets of the southern oceans. Journal of Climate, 23, 440–454. doi:10.1175/2009JCLI3152.1

788

Wadhams, P., Lange, M. a, & Ackley, S. F. (1987). The ice thickness distribution across the Atlantic sector

789

of the Antarctic Ocean in midwinter. Journal of Geophysical Research: Oceans (1978--2012),

790

92(C13), 14535–14552.

791

Walter, B. A. (1980). Wintertime observations of roll clouds over the Bering Sea. Monthly Weather

792 793

Review.

Winker, D. M., Hunt, W. H., & McGill, M. J. (2007). Initial performance assessment of CALIOP.

794 795

Geophysical Research Letters, 34(19), 1–5. doi:10.1029/2007GL030135

Zhang, M. H., et al. (2005). Comparing clouds and their seasonal variations in 10 atmospheric general

796

circulation models with satellite measurements. Journal of Geophysical Research D: Atmospheres,

797

110(15), 1–18. doi:10.1029/2004JD005021



37

798

Table Captions

799

Table 1. Summary of observations used to guide the radiative transfer calculations.

800

(top) Cloud fraction partitioned by cloud height and cloud thermodynamic phase

801

from CALIPSO-GOCCP observations. (middle) Summary of the distribution of cloud

802

liquid water path from MAC-LWP observations. Cloud phase and liquid water path

803

are from satellite observations taken during JJA from 2006-2014 over open water

804

near the sea ice edge in the Weddell Sea. (bottom) Cloud-base height measurements

805

from a cloud ceilometer onboard the 1992 cruise. Statistics of cloud-base height are

806

shown for all scenes in which a cloud with a base below 1500 m was detected and

807

surface observers reported open water or open pack ice near the ship. The

808

probability density function shown in parenthesis was computed by dividing the

809

probability of each bin by the width of the bin.

810

Figure Captions

811

Figure 1. Map of Antarctica and the Southern Ocean showing average sea ice

812

concentration during July from 2006-2014 from satellite observations. Two

813

contours of sea ice concentration are shown: 0.35, which we use as a metric for the

814

sea ice edge, and 0.95. Boundaries for the Weddell (50°W-0°) and Ross (130°W-

815

170°E) Seas are shown in the black dashed lines, and tracks for the cruises to the

816

Weddell Sea are shown as colored lines. The cruises started near 0° longitude and

817

finished near the Antarctic Peninsula.



38

818

Figure 2. 2006-2014 climatology of June, July and August low-cloud fraction (color)

819

from CALIPSO-GOCCP observations and the position of the sea ice edge. The red line

820

shows the average position of the sea ice edge, and the red shading shows one

821

standard deviation on either side of the mean.

822

Figure 3. Wintertime cloud fraction, temperature and humidity in the lower

823

troposphere plotted as a function of meridional distance from the sea ice edge. (a)

824

Vertical profile of mean cloud fraction, (b) mean sea ice concentration and low-

825

cloud fraction, with error bars showing the 95% confidence interval of the mean,

826

and (c) mean potential temperature (contours) and specific humidity (color) over

827

the Weddell Sea. (d-f) as in (a-c) but for the Ross Sea. Cloud and sea ice fields come

828

from satellite observations, and temperature and humidity come from ERA Interim

829

reanalysis data. The boundaries for the Weddell Sea and Ross Sea are shown in

830

Figure 1.

831

Figure 4. Temperature profile of the lower troposphere over the Weddell Sea from

832

soundings. For each height, color shows the probability density function of air

833

temperature. Data are composited into soundings taken poleward of 65°S and

834

between 55°-65°S. The number of days in which soundings were collected is shown

835

in the top right corner of each panel. Data from the 2013 cruise are shown in (a-b),

836

and from the 1992 cruise in (c-d). Bins of width 2°C are used in the calculation. The

837

black dashed line shows a profile with a surface temperature of -1.8°C, which is

838

about the freezing temperature of seawater in the Southern Ocean, and a moist

839

adiabatic lapse rate. Note that two boundary layer regimes are seen: a warm mode

840

with near-surface temperatures close to the freezing temperature of seawater and



39

841

with a most adiabatic lapse rate, and a cold mode with near-surface temperatures

842

from -15°C to -25°C and with a low-level inversion.

843

Figure 5. Vertical profile of wind speed from soundings in the Weddell Sea region

844

poleward of 55°S. For each height, color shows the probability density function of

845

wind speed. Bins of width 3 ms-1 are used in the calculation. Data from the 2013 and

846

1992 cruises are shown in the top and bottom panels, respectively. Data were

847

collected over 53 days on both cruises. Note that wind speeds of 10 ms-1 or more are

848

common at heights between 200-600 m, and that the signature of a low-level jet can

849

be seen in the measurements from the 2013 cruise.

850

Figure 6. Mean low-cloud fraction observed by CALIPSO as a function of meridional

851

distance from the sea ice edge, and its dependence on low-level warm or cold

852

advection. Observations are composited into periods of poleward, on-ice flow at

853

low-levels (𝑣!"# !"#! < −0.5𝜎 ≈ −3 ms-1, where 𝜎 is the standard deviation of

854

𝑣!"# !"#! ) and periods of equatorward, off-ice flow at low-levels (𝑣!"# !"#! > 0.5𝜎 ≈ 3

855

ms-1). Averages are computed using daily-mean data, and error bars show the 95%

856

confidence interval of the mean. Over open ocean, cloudier conditions are seen

857

during periods of off-ice advection. Over sea ice, low-cloud fraction is similar for

858

periods of on-ice and off-ice advection.

859

Figure 7. Computed surface 𝐿𝑊↓ over open water near the sea ice edge plotted as a

860

function of near-surface meridional wind at the sea ice edge (𝑣!"# !"#! ). The dots

861

show individual 𝐿𝑊↓,!""!!"# values, and the blue line shows 𝐿𝑊↓,!""!!"# binned by

862

𝑣!"# !"#! and averaged. Error bars on the blue line are the 95% confidence interval of

863

the mean. The red, solid black, and dashed black lines show linear regressions of

40

864

𝐿𝑊↓,!""!!"# , 𝐿𝑊↓,!"#$%&'( , and 𝐿𝑊↓,!"#$% on 𝑣!"# !"#! , respectively. The red shading is

865

the 95% confidence interval for the regression slope of 𝐿𝑊↓,!""!!"# . Note that surface

866

LW cloud radiative effect, seen in the figure as the difference between 𝐿𝑊↓,!""!!"#

867

and 𝐿𝑊↓,!"#$% , increases with 𝑣!"# !"#! . As a result, surface LW cloud radiative effect

868

is largest during periods of strong off-ice flow.

869

Figure 8. July-mean low-cloud fraction and latitude of the sea ice edge in climate

870

models and observations. Low-cloud fraction is shown in blue, and the sea ice edge

871

is shown in red (the red line shows the mean, and red shading shows one standard

872

deviation on either side of the mean). Each model is shown with output from fully-

873

coupled and atmosphere-only configuration. Low-cloud fraction in the models was

874

computed by a CALIPSO simulator, and observations are from CALIPSO-GOCCP.

875

Figure 9. Evaluation of wintertime low-cloud fraction near the sea ice edge in

876

climate models. (a) CALIPSO observed and simulated mean low-cloud fraction over

877

open water (1°-3° equatorward of the sea ice edge) and sea ice (2°-4° poleward of

878

the sea ice edge). The gray plus sign shows the 95% confidence interval for the

879

observed value. Models in both atmosphere-only and fully-coupled configurations

880

are shown. Six models have both atmosphere-only and fully-coupled output, and

881

these models are labeled with bold text in the legend. For these models, the

882

atmosphere-only and fully-coupled data points are connected by dashed lines, but

883

for some models the difference is small and the dashed line is not visible. Note that

884

all but one model (GFDL-CM3) underestimate low-cloud fraction over open water,

885

and all but one fully-coupled model (MPI-ESM-LR) underestimate low-cloud fraction



41

886

over sea ice. (b) Difference between mean low-cloud fraction over open water and

887

over sea ice near the sea ice edge. Error bars show the 95% confidence interval.

888







42

889

Tables Cloud Phase





Level



Cloud Fraction









Total

Liquid

Ice

Unclassified







High

0.29

0

0.28

0.02







Middle

0.27

0.09

0.16

0.04







Low

0.69

0.50

0.05

0.14





Liquid Water Path





Percentile

5

25

50

75

95





Liquid

43.1

54.1

61.4

70.1

79.1



water path (g/m2) Cloud-Base Height Height Range

0-50

(m) Counts (Probability

12 (.10)

50-

100-

200-

100

200

300

11

22

36

(0.09)

(0.09)

(0.15)

300-600

58 (0.08)

600-

1000-

1000

1500

67

34

(0.07)

(0.03)

density ×10!! (m-1)) 890



891

Table 1. Summary of observations used to guide the radiative transfer calculations.

892

(top) Cloud fraction partitioned by cloud height and cloud thermodynamic phase



43

893

from CALIPSO-GOCCP observations. (middle) Summary of the distribution of cloud

894

liquid water path from MAC-LWP observations. Cloud phase and liquid water path

895

are from satellite observations taken during JJA from 2006-2014 over open water

896

near the sea ice edge in the Weddell Sea. (bottom) Cloud-base height measurements

897

from a cloud ceilometer onboard the 1992 cruise. Statistics of cloud-base height are

898

shown for all scenes in which a cloud with a base below 1500 m was detected and

899

surface observers reported open water or open pack ice near the ship. The

900

probability density function shown in parenthesis was computed by dividing the

901

probability of each bin by the width of the bin.

902



44

903

Figures Weddell! Sea



0.95 90°W

90°E 60°S

0.35

70°S

sea ice! concentration

50°W

130°W

Ross! Sea

904

170°E



905

Figure 1. Map of Antarctica and the Southern Ocean showing average sea ice

906

concentration during July from 2006-2014 from satellite observations. Two

907

contours of sea ice concentration are shown: 0.35, which we use as a metric for the

908

sea ice edge, and 0.95. Boundaries for the Weddell (50°W-0°) and Ross (130°W-

909

170°E) Seas are shown in the black dashed lines, and tracks for the cruises to the

910

Weddell Sea are shown as colored lines. The cruises started near 0° longitude and

911

finished near the Antarctic Peninsula.

912







45

June

July

August 0.7 0.6 0.5

914

70°

70°

70°

60°

60°

60°

50°

50°

50°

0.4

low-cloud fraction

913



915

Figure 2. 2006-2014 climatology of June, July and August low-cloud fraction (color)

916

from CALIPSO-GOCCP observations and the position of the sea ice edge. The red line

917

shows the average position of the sea ice edge, and the red shading shows one

918

standard deviation on either side of the mean.





46

Ross Sea 0.3 .3

2 2

0.25

1.5 1.5

0.2 .2

1 1

0.15

0.5 0.5

0.1 .1

-4

distance from ice edge (° lat)

920

0

2

2

4

6 4

0.675 0.675 0.6 0.6

0.45 0.45 -6 -6

-4 -4

-2 -2

0 0

2 2

33

280

2.5

22

900 265

1.5

270

0.5 0.5 0.25 0.25

11

260 -6

0.75 0.75

3.5

275

950

1 1

0 6 0 6

850

south

6

4 4

Potential Temp./q

800

1000

north

0

0.525 0.525

f

south

-2

LCF/SIC LCF/SIC

750

c

-2

sea ice conc.

e

LowCloud CloudFraction Fraction Low

low-cloud frac.

sea ice conc.

b

-4

0.75 0.75

Pressure (hPa)

pressure (hPa)

low-cloud frac.

-6 -6

SeaIce IceConcentration Concentration Sea

d

cloud fraction

Cloud Incidence Cloud Incidence

2.5 2.5

a

Height(km) (km) Height

height (km)

Weddell Sea

specific! humidity (g/kg)

919

-4

-2

0

2

4

Distance from Ice Edge (degrees)

6

north

distance from ice edge (° lat)



921

Figure 3. Wintertime cloud fraction, temperature and humidity in the lower

922

troposphere plotted as a function of meridional distance from the sea ice edge. (a)

923

Vertical profile of mean cloud fraction, (b) mean sea ice concentration and low-

924

cloud fraction, with error bars showing the 95% confidence interval of the mean,

925

and (c) mean potential temperature (contours) and specific humidity (color) over

926

the Weddell Sea. (d-f) as in (a-c) but for the Ross Sea. Cloud and sea ice fields come

927

from satellite observations, and temperature and humidity come from ERA Interim

928

reanalysis data. The boundaries for the Weddell Sea and Ross Sea are shown in

929

Figure 1.



47

poleward of 65°S

65°-55°S (25)

b

(28)

c

(34)

d

(22)

probability

1992

Height (m)

2013

a

930

Temperature (°C)

931

Figure 4. Temperature profile of the lower troposphere over the Weddell Sea from

932

soundings. For each height, color shows the probability density function of air

933

temperature. Data are composited into soundings taken poleward of 65°S and

934

between 55°-65°S. The number of days in which soundings were collected is shown

935

in the top right corner of each panel. Data from the 2013 cruise are shown in (a-b),

936

and from the 1992 cruise in (c-d). Bins of width 2°C are used in the calculation. The

937

black dashed line shows a profile with a surface temperature of -1.8°C, which is

938

about the freezing temperature of seawater in the Southern Ocean, and a moist

939

adiabatic lapse rate. Note that two boundary layer regimes are seen: a warm mode

940

with near-surface temperatures close to the freezing temperature of seawater and

941

with a most adiabatic lapse rate, and a cold mode with near-surface temperatures

942

from -15°C to -25°C and with a low-level inversion.





48

2013

height (m)

1400

1000

600 0.3

5

10

15

20

1992

0.2

0.1

probability

200

1400

height (m)

0 1000

600

200 5

10

15

20

wind speed (ms-1)

943



944

Figure 5. Vertical profile of wind speed from soundings in the Weddell Sea region

945

poleward of 55°S. For each height, color shows the probability density function of

946

wind speed. Bins of width 3 ms-1 are used in the calculation. Data from the 2013 and

947

1992 cruises are shown in the top and bottom panels, respectively. Data were

948

collected over 53 days on both cruises. Note that wind speeds of 10 ms-1 or more are

949

common at heights between 200-600 m, and that the signature of a low-level jet can

950

be seen in the measurements from the 2013 cruise.





49

low-cloud fraction

0.8 0.8

off-ice flow

0.7 0.7

0.6 0.6

on-ice flow

0.5 0.5

-6 -6

-4 -4

-2 -2

0 0

2 2

4

south distance from ice edge (° latitude)

951

6 6

north



952

Figure 6. Mean low-cloud fraction observed by CALIPSO as a function of meridional

953

distance from the sea ice edge, and its dependence on low-level warm or cold

954

advection. Observations are composited into periods of poleward, on-ice flow at

955

low-levels (𝑣!"# !"#! < −0.5𝜎 ≈ −3 ms-1, where 𝜎 is the standard deviation of

956

𝑣!"# !"#! ) and periods of equatorward, off-ice flow at low-levels (𝑣!"# !"#! > 0.5𝜎 ≈ 3

957

ms-1). Averages are computed using daily-mean data, and error bars show the 95%

958

confidence interval of the mean. Over open ocean, cloudier conditions are seen

959

during periods of off-ice advection. Over sea ice, low-cloud fraction is similar for

960

periods of on-ice and off-ice advection.



50

surface LW↓ (W/m2)

on-ice! flow

961

vice edge (ms-1)

off-ice! flow



962

Figure 7. Computed surface 𝐿𝑊↓ over open water near the sea ice edge plotted as a

963

function of near-surface meridional wind at the sea ice edge (𝑣!"# !"#! ). The dots

964

show individual 𝐿𝑊↓,!""!!"# values, and the blue line shows 𝐿𝑊↓,!""!!"# binned by

965

𝑣!"# !"#! and averaged. Error bars on the blue line are the 95% confidence interval of

966

the mean. The red, solid black, and dashed black lines show linear regressions of

967

𝐿𝑊↓,!""!!"# , 𝐿𝑊↓,!"#$%&'( , and 𝐿𝑊↓,!"#$% on 𝑣!"# !"#! , respectively. The red shading is

968

the 95% confidence interval for the regression slope of 𝐿𝑊↓,!""!!"# . Note that surface

969

LW cloud radiative effect, seen in the figure as the difference between 𝐿𝑊↓,!""!!"#

970

and 𝐿𝑊↓,!"#$% , increases with 𝑣!"# !"#! . As a result, surface LW cloud radiative effect

971

is largest during periods of strong off-ice flow.



51

IPSL-CM5A-LR

MPI-ESM-LR

MRI-CGCM3

IPSL-CM5A-MR

MIROC5

atmosphere-! only

fully-! coupled

HadGEM2

fully-! coupled

Observations

0.6

atmosphere-! only

0.5 0.4 70°

low-cloud fraction

0.7

0.3

60°

972



973

Figure 8. July-mean low-cloud fraction and latitude of the sea ice edge in climate

974

models and observations. Low-cloud fraction is shown in blue, and the sea ice edge

975

is shown in red (the red line shows the mean, and red shading shows one standard

976

deviation on either side of the mean). Each model is shown with output from fully-

977

coupled and atmosphere-only configuration. Low-cloud fraction in the models was

978

computed by a CALIPSO simulator, and observations are from CALIPSO-GOCCP.



52

LCFsea ice

0.7 0.7

1:1

a

0.7

0.6 0.6 0.5 0.5

0.6

0.4 0.4 0.3 0.3

0.5

0.3

0.3

0.4

0.4

0.5

0.5

0.6

0.6

0.4

0.3

b

0.3

0.2 0.2 0.3

0.4

0.5

0.6

0.1

00 -0.1 -0.2 -0.2

HH aadd GG IIPP EEMM SSLL 22 --CC MM IIPP 55AA SSLL --LL RR --CC MM 55AA --MM RR MM IIRR O O MM CC5 PPII 5 --EE SSMM --LL MM R RRI I--CC R GGC CM MM M IIRR 33 OOC C-EE S SMM --CC HHE EMM MM IIRR OOC C-EE SSMM CCN NR RM M-CC MM5 GGF 5 FDD LL-CCM M3 3

979

0.7

0.7

0.4 0.4

oobb ssee rrvv aatt ioio nnss

LCFopen water - LCFsea ice

LCFopen water

HadGEM2 IPSL-CM5A-LR IPSL-CM5A-MR MIROC5 MPI-ESM-LR MRI-CGCM3 MIROC-ESM-CHEM MIROC-ESM CNRM-CM5 GFDL-CM3 observations atmosphere-only fully-coupled



980

Figure 9. Evaluation of wintertime low-cloud fraction near the sea ice edge in

981

climate models. (a) CALIPSO observed and simulated mean low-cloud fraction over

982

open water (1°-3° equatorward of the sea ice edge) and sea ice (2°-4° poleward of

983

the sea ice edge). The gray plus sign shows the 95% confidence interval for the

984

observed value. Models in both atmosphere-only and fully-coupled configurations

985

are shown. Six models have both atmosphere-only and fully-coupled output, and

986

these models are labeled with bold text in the legend. For these models, the

987

atmosphere-only and fully-coupled data points are connected by dashed lines, but

988

for some models the difference is small and the dashed line is not visible. Note that

989

all but one model (GFDL-CM3) underestimate low-cloud fraction over open water,

990

and all but one fully-coupled model (MPI-ESM-LR) underestimate low-cloud fraction



53

0.7

991

over sea ice. (b) Difference between mean low-cloud fraction over open water and

992

over sea ice near the sea ice edge. Error bars show the 95% confidence interval.



54

1 Low-cloud, boundary layer, and sea ice interactions over the ... - dust

cloud radiative effect contributed by low-clouds are presented. Finally, ten ..... The domain can be split into three regions based on sea ice. 286 ... computing the number of observations in each bin, and normalizing by the total. 312 ...... Remote Sensing of Environment, 92(2), 181–194. doi:10.1016/j.rse.2004.06.004. 740.

8MB Sizes 2 Downloads 171 Views

Recommend Documents

boundary layer
RH and OH not only depends on the strength and spatial distribution of the. RH emissions ..... Any present-day numerical model is only capable of sim- ulating 2 ...

boundary layer climates oke pdf
Sign in. Loading… Whoops! There was a problem loading more pages. Whoops! There was a problem previewing this document. Retrying... Download. Connect ...

Distinct CCN activation kinetics above the marine boundary layer ...
Aug 7, 2009 - boundary layer along the California coast, Geophys. Res. Lett., ... 3Department of Earth and Atmospheric Sciences, Georgia Institute of.

Internal boundary layer model for the evolution of ...
Feb 5, 2012 - and ranging (lidar) data we computed z02 =10−2 m (Supplementary ... portantly, the analytical expression was not fit to sand flux data; all ...

ePub Boundary-Layer Theory, 8th ed. Full Online
Issuu is a digital publishing platform that makes it simple to publish magazines catalogs newspapers books and more online Easily share your publications and ...

Metadata of Village Boundary - 2001 layer -
Dec 31, 2011 - Enumerated domain value. "name" : "Assam", "code" : "AS", "name" : "Chandigarh", "code" : "CH" ,"name" : "Dadara &. Nagar Havelli", "code" : "DN", "name" : "Meghalaya", "code" : "ML" ,. "name" : "Manipur", "code" : "MN" , "name" : "Miz

Comparison of Planetary Boundary Layer Model Winds with ...
with the unique characteristic of the PBL model being able to account for the nonlinear effects of organized large eddies. .... event off France in 1999. This ability to predict .... sure analyses are corrected with the surface pressure observations 

boundary layer theory schlichting pdf free download
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. boundary layer ...

Estimating the Contribution of Sea Ice Response to Climate Sensitivity ...
Nov 15, 2014 - defined as the initial tropopause energy imbalance divided. Denotes Open Access ... sea ice cover (i.e., cloud or water vapor changes). Thus,.

Designing Interactions - Moggridge - Chapter 1 - The Mouse and the ...
There was a problem previewing this document. Retrying. ... Designing Interactions - Moggridge - Chapter 1 - The Mouse and the Desktop.pdf. Designing ...

Surface temperatures of sea ice and ice sheets from ...
Arctic surface fluxes. • Assimilation: Ocean and Sea Ice models (Hycom + Cice). • Ice managing. • Sea ice break up forecasts. • Servicing the Oil industry for oil exploration. • Coupled climate models. • Improve Greenland ice sheet modell

pdf-0712\interactions-mosaic-silver-edition-interactions-1-low ...
... the apps below to open or edit this item. pdf-0712\interactions-mosaic-silver-edition-interaction ... iate-reading-class-audio-cd-by-elaine-kirn-pamela-h.pdf.

climate change inquiry sea ice lab.pdf
climate change inquiry sea ice lab.pdf. climate change inquiry sea ice lab.pdf. Open. Extract. Open with. Sign In. Main menu.

Unprecedented low twentieth century winter sea ice ...
Jun 9, 2009 - The climatic impacts of sea ice extent can reach systems located a long distance ... Eurasia and North America; (b) extended periods of low sea.

The Role of Azopolymer/Dendrimer Layer-by-Layer Film Architecture ...
The Role of Azopolymer/Dendrimer Layer-by-Layer Film Architecture in Photoinduced Birefringence and the Formation of Surface-Relief. Gratings. David S. dos Santos, Jr.,*,† Marcos R. Cardoso,‡ Fabio L. Leite,‡,§ Ricardo F. Aroca,†. Luiz H. C.

interactions 1 reading pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. interactions 1 ...

Interactions of the Sulfonylurea Receptor 1 Subunit in ...
system did not lead to glibenclamide binding activity, although .... These data indicate that there are strong interactions ... U2010 spectrophotometer (Hitachi).

Two New Elementary Boundary Letter December 2016 (1) (1).pdf ...
Page 1 of 1. Two New Elementary Boundary Letter December 2016 (1) (1).pdf. Two New Elementary Boundary Letter December 2016 (1) (1).pdf. Open. Extract.

Interactions of the Sulfonylurea Receptor 1 Subunit in ...
N- or C-terminal domain in a baculovirus expression system did not lead to glibenclamide binding activity, although studies with green fluorescent protein fusion.

the influence of the southern annular mode (sam) over the sea surface ...
Apr 28, 2006 - SAM can influence the SST fields on different time and space scales ... point and used to construct a synthetic map of the spatial correlation.